content
stringlengths 1
15.9M
|
---|
\section{Introduction and overview}
This paper develops a method for parsimoniously summarizing the shared dependence of many individual response variables upon a common set of predictor variables drawn at random. The focus is on multivariate Gaussian linear models where an analyst wants to find, among $p$ available predictors $X$, a subset which work well for predicting $q > 1$ response variables $Y$. The multivariate normal linear model assumes that a set of responses $\{ Y_{j} \}_{j=1}^{q}$ are linearly related to a shared set of covariates $\{ X_{i} \}_{i=1}^{p}$ via
\begin{equation}\label{modelfirst}
\begin{split}
Y_{j} &= \beta_{j1}X_{1} + \cdots + \beta_{jp}X_{p} + \epsilon_{j}, \;\;\;\;\; \boldsymbol{\epsilon} \sim \mbox{N}(0, \Psi),
\end{split}
\end{equation}
where $\Psi$ is a non-diagonal covariance matrix.
Bayesian variable selection in (single-response) linear models is the subject of a vast literature, from prior specification on parameters \citep{Berger12} and models \citep{ScottBerger06} to efficient search strategies over the model space \citep{GeorgeandMcCulloch, hans2007shotgun}. For a more complete set of references we refer the reader to the reviews of \cite{Clyde04} and \cite{HahnCarvalho}. By comparison, variable selection has not been widely studied in concurrent regression models, perhaps because it is natural simply to apply existing variable selection methods to each univariate regression individually. Indeed, such joint regression models go by the name ``seemingly unrelated regressions'' (SUR) in the Bayesian econometrics literature, reflecting the fact that the regression coefficients from each of the separate regressions can be obtained in isolation from one another (i.e., conducting estimation as if $\Psi$ were diagonal). However, allowing non-diagonal $\Psi$ can lead to more efficient estimation \citep{zellner1962efficient} and can similarly impact variable selection \citep{brown1998multivariate, wangSUR}.
This paper differs from \cite{brown1998multivariate} and \cite{wangSUR} in that we focus on the case where the predictor variables (the regressors, or covariates) are treated as random as opposed to fixed.Our goal will be to summarize codependence among multiple responses in {\em subsequent} periods, making the uncertainty in future realizations highly central to our selection objective. This approach is natural in many contexts (e.g., macroeconomic models) where the purpose of selection is inherently forward-looking. To our knowledge, no existing variable selection methods are suitable in this context. The new approach is based on the sparse summary perspective outlined in \cite{HahnCarvalho}, which applies Bayesian decision theory to summarize complex posterior distributions. By using a utility function that explicitly rewards sparse summaries, a high dimensional posterior distribution is collapsed into a more interpretable sequence of sparse point summaries.
A related approach to variable selection in multivariate Gaussian models is the Gaussian graphical model framework \citep{jones2005experiments,dobra2004sparse,wang2009bayesian}. In that approach, the full conditional distributions are represented in terms of a sparse $(p+q)$-by-$(p+q)$ precision matrix. By contrast, we partition the model into response and predictor variable blocks, leading to a distinct selection criterion that narrowly considers the $p$-by-$q$ covariance between $Y$ to $X$.
\subsection{Methods overview}\label{overview} Posterior summary variable selection consists of three phases: {\em model specification and fitting}, {\em utility specification}, and {\em graphical summary}. Each of these steps is outlined below. Additional details of the implementation are described in Section \ref{DSS} and the Appendix.
\subsubsection*{Step 1: Model specification and fitting}
The statistical model may be described compositionally as $p(Y,X) = p(Y \vert X)p(X)$. For $(Y,X) \sim \mbox{N}(\mu,\Sigma)$, the regression model (\ref{modelfirst}) implies $\Sigma$ has the following block structure:
\begin{align}\label{model2}
\Sigma
=
\left[
\begin{array}{c|c}
{\boldsymbol \beta}^{T}\Sigma_{x}{\boldsymbol\beta} + \Psi & (\Sigma_{x}{\boldsymbol\beta})^{T} \\
\hline
\Sigma_{x}{\boldsymbol\beta}
&
\Sigma_{x} \\
\end{array}
\right].
\end{align}
We denote the unknown parameters for the full joint model as $\Theta = \{\mu_{x},\mu_{y},\Sigma_{x},\boldsymbol{\beta},\Psi\}$ where $\mu = (\mu_y^T, \mu_x^T)^T$ and $\Sigma_x = \mbox{cov}(X)$.
For a given prior choice $p(\Theta)$, posterior samples of all model parameters are computed by routine Monte Carlo methods, primarily Gibbs sampling. Details of the specific modeling choices and associated posterior sampling strategies are described in the Appendix.
A notable feature of our approach is that {\it steps 2} (and {\it 3}) will be unaffected by modeling choices made in {\it step 1} except insofar as they lead to different posterior distributions $p(\Theta \vert \mathbf{Y}, \mathbf{X})$. In short, {\it step 1} is ``obtain a posterior distribution''; posterior samples then become inputs to {\it step 2}.
\subsubsection*{Step 2: Utility specification}
For our utility function we use the log-density of the regression $p(Y \vert X)$ above. It is convenient to work in terms of negative utility, or loss:
\begin{equation}
\begin{split}
\mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}) = \frac{1}{2}( \tilde{Y} - \boldsymbol{\gamma}\tilde{X} )^{T} \Omega ( \tilde{Y} - \boldsymbol{\gamma}\tilde{X} ),
\end{split}
\end{equation}
where $\Omega = \Psi^{-1}$. Note that this log-density is being used in a descriptive capacity, not an inferential one; that is, all posterior inferences are based on the posterior distribution from {\it step 1}. The ``action'' $\boldsymbol{\gamma}$ is regarded as a point estimate of the regression parameters $\boldsymbol{\beta}$, which would be a good fit to {\em future} data $(\tilde{Y}, \tilde{X})$ drawn from the same model as the observed data.
Taking expectations over the posterior distribution of all unknowns
\begin{equation}
\begin{split}
p(\tilde{Y},\tilde{X}, \Theta \vert \textbf{Y}, \textbf{X}) = p(\tilde{Y} \vert \tilde{X}, \Theta) p(\tilde{X} \vert \Theta) p(\Theta \vert \textbf{Y}, \textbf{X}),
\end{split}
\end{equation}
yields expected loss
\begin{equation}
\mathcal{L}(\boldsymbol{\gamma}) \equiv \mathbb{E}[ \mathcal{L}(\tilde{Y},\tilde{X}, \Theta, \boldsymbol{\gamma}) ] =\text{tr}[ M \boldsymbol{\gamma} S \boldsymbol{\gamma}^{T} ] - 2\text{tr}[A\boldsymbol{\gamma}^{T}] + \mbox{constant},
\end{equation}where $A=\mathbb{E}[\Omega\tilde{Y}\tilde{X}^{T}]$, $S=\mathbb{E}[\tilde{X}\tilde{X}^{T}] = \overline{\Sigma_{x}}$, and $M=\overline{\Omega}$, the overlines denote posterior means, and the final term is a constant with respect to $\boldsymbol{\gamma}$.
Finally, we add an explicit penalty, reflecting our preference for sparse summaries:
\begin{equation}\label{ex_loss}
\mathcal{L}_{\lambda}(\boldsymbol{\gamma}) \equiv \text{tr}[ M \boldsymbol{\gamma} S \boldsymbol{\gamma}^{T} ] - 2\text{tr}[A\boldsymbol{\gamma}^{T}] + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0},
\end{equation}
where $\norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0}$ counts the number of non-zero elements in $\boldsymbol{\gamma}$. In practice, we will use an approximation to this utility based on the $\ell_1$ penalty; optimal actions under this approximation will still be sparse.
\subsubsection*{Step 3: Graphical summary}
Traditional applications of Bayesian decision theory derive {\em point-estimates} by minimizing expected loss for certain loss functions. The present goal is not an {\em estimator} per se, but a parsimonious summary of information contained in a complicated, high dimensional posterior distribution. This distinction is worth emphasizing because we have not one, but rather a continuum of loss functions, indexed by the penalty parameter $\lambda$. This class of loss functions can be used to organize the posterior distribution as follows.
Using available convex optimization techniques, expression (\ref{ex_loss}) can be optimized efficiently for a range of $\lambda$ values simultaneously. Posterior graphical summaries consist of two components. First, graphs depicting which response variables have non-zero $\boldsymbol{\gamma}_{\lambda}^*$ coefficients on which predictor variables can be produced for any given $\lambda$. Second, posterior distributions of the quantity
\begin{equation}
\Delta_{\lambda} = \mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}_{\lambda}^*) - \mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}^*)
\end{equation}
can be used to gauge the impact $\lambda$ has on the descriptive capacity of $\gamma_{\lambda}^*$. Here, $\boldsymbol{\gamma}^* = \gamma_{\lambda=0}^*$ is the unpenalized optimal solution to the minimization of loss (\ref{ex_loss}).
\section{Posterior summary variable selection}\label{DSS}
The statistical model is given in equations (\ref{modelfirst}) and (\ref{model2}); prior specification and model fitting details can be found in the Appendix. Alternatively, the models described in \cite{brown1998multivariate} or \cite{wangSUR} could be used. In this section, we flesh out the details of {\it steps 2} and {\it 3}, which represent the main contributions of this paper.
\subsection{Deriving the sparsifying expected utility function}
Define the optimal posterior summary as the $\boldsymbol{\gamma}^*$ minimizing some expected loss $\mathcal{L}_{\lambda}(\boldsymbol{\gamma}) = \mathbb{E}[\mathcal{L}_{\lambda}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma})]$. Here, the expectation is taken over the joint posterior predictive and posterior distribution: $p(\tilde{Y},\tilde{X}, \Theta \mid \textbf{Y}, \textbf{X})$.
As described in the previous section, our loss takes the form of a penalized log conditional distribution:
\begin{equation}
\begin{split}
\mathcal{L}_{\lambda}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}) \equiv \frac{1}{2}( \tilde{Y} - \boldsymbol{\gamma}\tilde{X} )^{T} \Omega ( \tilde{Y} - \boldsymbol{\gamma}\tilde{X} ) + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0}, \label{newlossstoch}
\end{split}
\end{equation}where $\Omega = \Psi^{-1}$, $\norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0} = \sum_{j}\mathds{1}\left(\text{{\bf vec}}(\boldsymbol{\gamma}) \neq 0\right)$, and $\text{{\bf vec}}(\boldsymbol{\gamma})$ is the vectorization of the action matrix $\boldsymbol{\gamma}.$ The first term of this loss measures the distance (weighted by the precision $\Omega$) between the linear predictor $\boldsymbol{\gamma}\tilde{X}$ and a future response $\tilde{Y}$. The second term promotes a sparse optimal summary, $\boldsymbol{\gamma}$. The penalty parameter $\lambda$ determines the relative importance of these two components. Expanding the quadratic form gives:
\small
\begin{equation}\label{modnew1}
\begin{split}
\mathcal{L}_{\lambda}(\tilde{Y},\tilde{X}, \Theta, \boldsymbol{\gamma}) &= \frac{1}{2}\left(\tilde{Y}^{T} \Omega \tilde{Y} - 2\tilde{X}^{T}\boldsymbol{\gamma}^{T} \Omega \tilde{Y} + \tilde{X}^{T}\boldsymbol{\gamma}^{T} \Omega \boldsymbol{\gamma} \tilde{X}\right) + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0} \\
&= \left( \tilde{X}^{T}\boldsymbol{\gamma}^{T} \Omega \boldsymbol{\gamma} \tilde{X} -2\tilde{X}^{T}\boldsymbol{\gamma}^{T} \Omega \tilde{Y}\right) + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0} + \mbox{constant}.
\end{split}
\end{equation}
\normalsize
Integrating over $(\tilde{Y},\tilde{X},\Theta \mid \textbf{Y}, \textbf{X})$ (and dropping the constant) gives:
\begin{equation}\label{almostlassoform}
\begin{split}
\mathcal{L}_{\lambda}(\boldsymbol{\gamma}) &= \mathbb{E}[ \mathcal{L}_{\lambda}(\tilde{Y},\tilde{X}, \Theta, \boldsymbol{\gamma}) ]\\
&= \mathbb{E}\left[ \text{tr}[ \boldsymbol{\gamma}^{T} \Omega \boldsymbol{\gamma} \tilde{X} \tilde{X}^{T}] \right] - 2\mathbb{E}\left[ \text{tr}[\boldsymbol{\gamma}^{T} \Omega \tilde{Y}\tilde{X}^{T}] \right] + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0}, \\
&= \mathbb{E}\left[ \text{tr}[ \boldsymbol{\gamma}^{T} \Omega \boldsymbol{\gamma} S] \right] - 2\text{tr}[A\boldsymbol{\gamma}^{T}] + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0},\\
&= \text{tr}[ M \boldsymbol{\gamma} S \boldsymbol{\gamma}^{T} ] - 2\text{tr}[A\boldsymbol{\gamma}^{T}] + \lambda \norm{\text{{\bf vec}}(\boldsymbol{\gamma})}_{0},
\end{split}
\end{equation}where
\begin{equation}\label{moments}
\begin{split}
A &\equiv\mathbb{E}[\Omega\tilde{Y}\tilde{X}^{T}],\\
S&\equiv\mathbb{E}[\tilde{X}\tilde{X}^{T}] = \overline{\Sigma}_{x},\\
M&\equiv\overline{\Omega},
\end{split}
\end{equation}
and the overlines denote posterior means. Define the Cholesky decompositions $M = LL^{T}$ and $S = QQ^{T}$. To make the optimization problem tractable we replace the $\ell_0$ norm with the $\ell_1$ norm, leading to an expression that can be formulated in the form of a standard penalized regression problem:
\begin{equation}\label{lasso_form}
\mathcal{L}_{\lambda}(\boldsymbol{\gamma}) = \norm{ \left[Q^{T} \otimes L^{T}\right]\text{\bf vec}(\boldsymbol{\gamma}) - \text{\bf vec}(L^{-1}AQ^{-T}) }_{2}^{2} + \lambda\norm{ \text{\bf vec}(\boldsymbol{\gamma})}_{1},
\end{equation}
with covariates $Q^{T} \otimes L^{T}$, ``data" $L^{-1}AQ^{-T}$, and regression coefficients $\boldsymbol{\gamma}$ (see the Appendix for details). Accordingly, (\ref{lasso_form}) can be optimized using existing software such as the {\tt lars} R package of \cite{Efron} and still yield sparse solutions.
\subsection{Sparsity-utility trade-off plots}
Rather than attempting to determine an ``optimal'' value of $\lambda$, we advocate displaying plots that reflect the utility attenuation due to $\lambda$-induced sparsification. We define the ``loss gap'' between a $\lambda$-sparse solution, $\mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}_{\lambda}^*)$, and the optimal unpenalized (non-sparse, $\lambda = 0$) summary, $\mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}^*)$ as
\begin{equation}
\Delta_{\lambda} = \mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}_{\lambda}^*) - \mathcal{L}(\tilde{Y},\tilde{X},\Theta,\boldsymbol{\gamma}^*).
\end{equation}
As a function of $(\tilde{Y},\tilde{X}, \Theta)$, $\Delta_{\lambda}$ is itself a random variable which we can sample by obtaining posterior draws from $p(\tilde{Y},\tilde{X}, \Theta \mid \textbf{Y}, \textbf{X})$. The posterior distribution(s) of $\Delta_{\lambda}$ (for various $\lambda$) therefore reflects the deterioration in utility attributable to ``sparsification''. Plotting these distributions as a function of $\lambda$ allows one to visualize this trade-off. Specifically, $\pi_{\lambda} \equiv \mbox{Pr}(\Delta_{\lambda} < 0 \mid \textbf{Y}, \textbf{X})$ is the (posterior) probability that the $\lambda$-sparse summary is no worse than the non-sparse summary. Using this framework, a useful heuristic for obtaining a single sparse summary is to report the sparsest model (associated with the highest $\lambda$) such that $\pi_{\lambda}$ is higher than some pre-determined threshold, $\kappa$; we adopt this approach in our application section.
We propose summarizing the posterior distribution of $\Delta_{\lambda}$ via two types of plots. First, one can examine posterior means and credible intervals of $\Delta_{\lambda}$ for a sequence of models indexed by $\lambda$. Similarly, one can plot $\pi_{\lambda}$ across the same sequence of models. Also, for a fixed value of $\lambda$, one can produce graphs where nodes represent predictor variables and response variables and an edge is drawn between nodes whenever the corresponding element of $\gamma^*_{\lambda}$ is non-zero. All three types of plots are exhibited in Section \ref{apps}.
\subsection{Relation to previous methods}
Loss function (\ref{lasso_form}) is similar in form to the univariate \textit{DSS} (decoupled shrinkage and selection) strategy developed by \cite{HahnCarvalho}. Our approach generalizes \cite{HahnCarvalho} by optimizing over the matrix $\boldsymbol{\gamma} \in \mathbb{R}^{qxp}$ rather than a single vector of regression coefficients, extending the sparse summary utility approach to seemingly unrelated regression models \citep{brown1998multivariate, wangSUR}.
Additionally, the present method considers random predictors, $\tilde{X}$, whereas \cite{HahnCarvalho} considered only a matrix of fixed design points. The impact of accounting for random predictors on the posterior summary variable selection procedure is examined in more detail in the application section.
An important difference between the sparse summary utility approach and previous approaches is in the role played by the posterior distribution. Many Bayesian variable importance metrics are based on the posterior distribution of an indicator variable that records if a given variable is non-zero (included in the model). The model we will use in our application utilizes such an indicator vector, which is called $\alpha$.
For example, a widely-used model selection heuristic is to examine the ``inclusion probability'' of predictor $i$, defined as the posterior mean of component $\alpha_{i}$. However, any approach based on the posterior mean of $\alpha$ necessarily ignores information about the codependence between its elements, which can be substantial in cases of collinear predictors. Our method focuses instead on the expected log-density of future predictions, which synthesizes information from all parameters simultaneously in gauging how important they are in terms of future predictions.
\section{Applications}\label{apps}
In this section, the sparse posterior summary method is applied to a data set from the finance (asset pricing) literature. A key component of our analysis will be a comparison between the posterior summaries obtained when the predictors are drawn at random versus when they are assumed fixed.
The response variables are returns on 25 tradable portfolios and our predictor variables are returns on 10 other portfolios thought to be of theoretical importance. In the asset pricing literature \cite{Ross}, the response portfolios represent assets to be priced (so-called {\em test assets}) and the predictor portfolios represent distinct sources of variation (so-called {\em risk factors}). More specifically, the test assets $Y$ represent trading strategies based on company size (total value of stock shares) and book-to-market (the ratio of the company's accounting valuation to its size); see \cite{FF3} and \cite{FF5} for details. Roughly, these assets serve as a lower-dimensional proxy for the stock market. The risk factors are also portfolios, but ones which are thought to represent {\em distinct} sources of risk. What constitutes a distinct source of risk is widely debated, and many such factors have been proposed in the literature \citep{cochrane2011presidential}. We use monthly data from July 1963 through February 2015, obtained from Ken French's website:
\begin{center}
{\tt http://mba.tuck.dartmouth.edu/pages/faculty/ken.french/}.
\end{center}
Our analysis investigates which subset of risk factors are most relevant (as defined by our utility function). As our initial candidates, we consider factors known in previous literature as: market, size, value, direct profitability, investment, momentum, short term reversal, long term reversal, betting against beta, and quality minus junk. Each factor is constructed by cross-sectionally sorting stocks by various characteristics of a company and forming linear combinations based on these sorts. For example, the value factor is constructed using the book-to-market ratio of a company. A high ratio indicates the company's stock is a ``value stock" while a low ratio leads to a ``growth stock" assessment. Essentially, the value factor is a portfolio built by going long stocks with high book-to-market ratio and shorting stocks with low book-to-market ratio. For detailed definitions of the first five factors, see \cite{FF5}. In the figures to follow, each is labeled as, for example, ``Size2 BM3," to denote the portfolio buying stocks in the second quintile of size and the third quintile of book-to-market ratio.
Recent related work includes \cite{ericsson2004choosing} and \cite{harvey2015lucky}. \cite{ericsson2004choosing} follow a Bayesian model selection approach based off of inclusion probabilities, representing the preliminary inference step of our methodology. \cite{harvey2015lucky} take a different approach that utilizes multiple hypothesis testing and bootstrapping.
\subsection{Results}
As described in Section \ref{overview}, the first step of our analysis consists of fitting a Bayesian model. We fit model (\ref{modelfirst}) using a variation of the well-known stochastic search variable selection algorithm of \cite{GeorgeandMcCulloch} and similar to \cite{brown1998multivariate} and \cite{wangSUR}. Details are given in the Appendix.
In the subsections to follow, we will show the following two figures. First, we plot the expectation of $\Delta_{\lambda}$ (and associated posterior credible interval) across a range of $\lambda$ penalties. Recall, $\Delta_{\lambda}$ is the ``loss gap'' between a sparse summary and the best non-sparse (saturated) summary, meaning that smaller values are ``better''. Additionally, we plot the probability that a given model is no worse than the saturated model $\pi_{\lambda}$ on this same figure, where ``no worse'' means $\Delta_{\lambda} < 0$. Note that even for very weak penalties (small $\lambda$), the distribution of $\Delta_{\lambda}$ will have non-zero variance and therefore even if it is centered about zero, some mass can be expected to fall above zero; practically, this means that $\pi_{\lambda} > 0.5$ is a very high score.
Second, we display a summary graph of the selected variables for the $\kappa=12.5\%$ threshold. Recall that this is the highest penalty (sparsest graph) that is no worse than the saturated model with $12.5\%$ posterior probability. For these graphs, the response and predictor variables are colored gray and white, respectively. A test asset label of, for example, ``Size2 BM3," denotes the portfolio that buys stocks in the second quintile of size and the third quintile of book-to-market ratio. The predictors without connections to the responses under the optimal graph are not displayed.
\newpage
These two figures are shown in four scenarios:
\begin{enumerate}
\item Random predictors.
\item Fixed predictors.
\item Random predictors under alternative prior.
\item Fixed predictors under alternative prior.
\end{enumerate}The ``alternative prior" scenario serves to show the impact of the statistical modeling comprising {\it step 1}. Specifically, we use the same Monte Carlo model fitting procedure as before (described in the Appendix) but fix $\alpha$ to the identity vector. That is, we omit the point-mass component of the priors for the elements of $\boldsymbol{\beta}$.
\subsubsection{Random predictors}
This section introduces our baseline example where the risk factors (predictors) are random. We evaluate the set of potential models by analyzing plots such as figure \ref{Lossgraph25}. This shows $\Delta_{\lambda}$ and $\pi_{\lambda}$ evaluated across a range of $\lambda$ values. Additionally, we display the posterior uncertainty in the $\Delta_{\lambda}$ metric with gray vertical uncertainty bands: these are the centered $P\%$ posterior credible intervals where $\kappa = (1-P)/2$. As the accuracy of the sparsified solution increases, the posterior of $\Delta_{\lambda}$ concentrates around zero by construction, and the probability of the model being no worse than the saturated model, $\pi_{\lambda}$, increases. We choose the sparsest model such that its corresponding $\pi_{\lambda} > \kappa = 12.5\%$. This model is displayed in figure \ref{graph25} and is identified by the black dot in figure \ref{Lossgraph25}.
The selected set of factors in graph \ref{graph25} are the market (Mkt.RF), value (HML), and size (SMB). This three factor model is no worse than the saturated model with $12.5\%$ posterior probability where all test assets are connected to all risk factors. Note also that in our selected model almost every test asset is distinctly tied to one of either value or size and the market factor. These are the three factors of Ken French and Eugene Fama's pricing model developed in \cite{FF3}. They are known throughout the finance community as being ``fundamental dimensions" of the financial market, and our procedure is consistent with this widely held belief at a small $\kappa$ level.
The characteristics of the test assets in graph \ref{graph25} are also important to highlight. The test portfolios that invest in small companies (``Size1" and ``Size2") are primarily connected to the SMB factor which is designed as a proxy for the risk of small companies. Similarly, the test portfolios that invest in high book-to-market companies (``BM4" and ``BM5") have connections to the HML factor which is built on the idea that companies whose book value exceeds the market's perceived value should generate a distinct source of risk. As previously noted, all of the test portfolios are connected to the market factor suggesting that it is a relevant predictor even for the sparse $\kappa=12.5\%$ selection criterion.
In figure \ref{graphseq25}, we examine how different choices of the $\kappa$ threshold change the selected set of risk factors. In this analysis, there is a tradeoff between the posterior probability of being ``close" to the saturated model and the utility's preference for sparsity. When the threshold is low ($\kappa=2,4,$ and $12.5$\%) the summarization procedure selects relatively sparse graphs with up to three factors (Mkt.RF, HML, and SMB). The market (Mkt.RF) and size (SMB) factors appear first, connected to a small number of the test assets ($\kappa=2$\%). As the threshold is increased, the point summary becomes denser and correspondingly more predictively accurate (as measured by the utility function). The value factor (HML) enters at $\kappa=12.5$\% and quality minus junk (QMJ), investment (CMA), and profitability (RMW) factors enter at $\kappa=32.5$\%. The graph for $\kappa=32.5$\% excluding QMJ is essentially the new five factor model proposed by \cite{FF5}. The five Fama-French factors (plus OMJ with three connections) persist up to the $\kappa=47.5\%$ threshold. This indicates that, up to a high posterior probability, the five factor model of \cite{FF5} does no worse than an asset pricing model with all ten factors connected to all test assets.
Notice also that our summarization procedure displays the specific relationship between the factors and test assets through the connections. Using this approach, the analyst is able to identify which predictors drive variation in which responses and at what thresholds they may be relevant. This feature is significant for summarization problems where individual characteristics of the test portfolios and their joint dependence on the risk factors is may be a priori unclear.
As $\kappa$ approaches the $50\%$ threshold ($\kappa=49.75\%$ in figure \ref{graphseq25}), the model summary includes all ten factors. Requesting a summary with this level of certainty results in little sparsification. However, compared to the nearby $\kappa=47.5\%$ model with only six factors, we also now know that the remaining four contribute little to our utility. These factors are betting against beta (BAB), momentum (Mom), long term reversal (LTR), and short term reversal (STR). Sparse posterior summarization applied in this context allows an analyst to study the impact of risk factors on pricing while taking uncertainty into account. Coming to a similar conclusion via common alternative techniques (e.g., component-wise ordinary least squares combined with thresholding by $t$-statistics) is comparatively ad hoc; our method is simply a perspicuous summary of a posterior distribution. Likewise, applying sparse regression techniques based on $\ell_1$ penalized likelihood methods would not take into account the residual correlation $\Psi$, nor would that approach naturally accommodate random predictors.
\begin{figure}[H]
\centering
\includegraphics[scale=0.47]{deltapilambdagraph-25PortGenPsi}
\caption{Evaluation of $\Delta_{\lambda}$ and $\pi_{\lambda}$ along the solution path for the 25 size/value portfolios modeled by the 10 factors. An analyst may use this plot to select a particular model. Uncertainty bands are 75\% posterior intervals on the $\Delta_{\lambda}$ metric. The large black dot represents the model selected in Figure \ref{graph25}.}
\label{Lossgraph25}
\includegraphics[scale=.65]{graph-25PortGenPsi}
\vspace{-10mm}
\caption{The selected model for 25 size/value portfolios modeled by the 10 factors. The responses and predictors are colored in gray and white, respectively. Edges represent nonzero components of the optimal action, $\boldsymbol{\gamma}$.}\label{graph25}
\end{figure}
\begin{figure}[H]
\centerline{
\includegraphics[scale=.8]{graphseqquant-25PortGenPsi.pdf}}
\caption{Sequence of selected models for varying threshold level $\kappa$ under the assumption of \textbf{random predictors}.}
\label{graphseq25}
\end{figure}
\subsubsection{Fixed predictors}
In this section, we consider posterior summarization with the loss function derived under the assumption of \textit{fixed predictors}. The analogous loss function when the predictor matrix is fixed is:
\begin{equation}\label{Lossfix}
\begin{split}
\mathcal{L}_{\lambda}(\boldsymbol{\gamma}) &= \norm{ \left[Q_{f}^{T} \otimes L^{T}\right]\text{\bf vec}(\boldsymbol{\gamma}) - \text{\bf vec}(L^{-1}A_{f}Q_{f}^{-T}) }_{2}^{2} + \lambda\norm{ \text{\bf vec}(\boldsymbol{\gamma})}_{1},\\
\end{split}
\end{equation}
with $Q_{f}Q_{f}^{T} = \textbf{X}^{T}\textbf{X}$, $A_{f}=\mathbb{E}[\Omega\tilde{\textbf{Y}}^{T}\textbf{X}]$, and $M=\overline{\Omega}=LL^{T}$; compare to (\ref{moments}) and (\ref{lasso_form}). The derivation of (\ref{Lossfix}) is similar to the presentation in Section \ref{DSS} and may be found in the Appendix.
The corresponding version of the loss gap is
\begin{equation}
\Delta_{\lambda} = \mathcal{L}(\tilde{\textbf{Y}},\textbf{X},\Theta,\boldsymbol{\gamma}_{\lambda}^*) - \mathcal{L}(\tilde{\textbf{Y}},\textbf{X},\Theta,\boldsymbol{\gamma}^*).
\end{equation}
which has distribution induced by the posterior over $(\tilde{\textbf{Y}}, \Theta)$ rather than $(\tilde{Y}, \tilde{X}, \Theta)$ as before. By fixing $\textbf{X}$, the posterior of $\Delta_{\lambda}$ has smaller dispersion
which results in denser summaries for the same level of $\kappa$. For example, compare how dense Figure \ref{graph25XFIX} is relative to Figure \ref{graph25}. The denser graph in Figure \ref{graph25XFIX} contains nine out of ten potential risk factors compared to just three in Figure \ref{graph25}, which correspond to the Fama-French factors described in \cite{FF3}. Recall, both graphs represent the sparsest model such that the probability of being no worse than the saturated model is greater than $\kappa = 12.5\%$ --- the difference is that one of the graphs defines ``worse-than'' in terms of a fixed set of risk factor returns while the other acknowledge that those returns are themselves uncertain in future periods.
Figure \ref{graphseq25XFIX} demonstrates this problem for several choices of the uncertainty level. Regardless of the uncertainty level chosen, the selected models contain most of the ten factors and many edges. In fact, it is difficult to distinguish even the $\kappa=2\%$ and $\kappa=49.75\%$ models.
\begin{figure}[H]
\centering
\includegraphics[scale=.47]{deltapilambdagraph-25PortXFIXGenPsi}
\caption{Evaluation of $\Delta_{\lambda}$ and $\pi_{\lambda}$ along the solution path for the 25 size/value portfolios modeled by the 10 factors assuming \textbf{fixed predictors (factors)}. An analyst may use this plot to select a particular model. Uncertainty bands are 75\% posterior intervals on the $\Delta_{\lambda}$ metric. The large black dot represents the model selected in Figure \ref{graph25XFIX}.}
\label{Lossgraph25XFIX}
\includegraphics[scale=.65]{graph-25PortXFIXGenPsi}\label{graph25XFIX}
\vspace{-10mm}
\caption{The selected model for 25 size/value portfolios modeled by the 10 factors when \textbf{uncertainty in future factor returns is not taken into account}. The responses and predictors are colored in gray and white, respectively. Edges represent nonzero components of the optimal action, $\boldsymbol{\gamma}$.}\label{graph25XFIX}
\end{figure}
\begin{figure}[H]
\centerline{
\includegraphics[scale=.8]{graphseqquant-25PortXFIXGenPsi.pdf}}
\vspace{-12mm}
\caption{Sequence of selected models for varying threshold level $\kappa$ under the assumption of \textbf{fixed predictors}.}
\label{graphseq25XFIX}
\end{figure}
\subsubsection{Alternative prior analysis}
Here, we consider how our posterior summaries change as a function of using a different posterior, based on a different choice of prior. Specifically, in this section we do not employ model selection point-mass priors on the elements of $\boldsymbol{\beta}$ as we did in the above analysis. These results are displayed in figures \ref{Lossgraph25MF} and \ref{graph25MF}. Broadly, the same risk factors are flagged as important ---
the market factor followed by the size (SMB) and value (HML) factors. One notable difference is that the quality minus junk (QMJ), investment (CMA), and profitability (RMW) factors appear at smaller levels of $\kappa$. This result is intuitive in the sense that point-mass priors demand stronger evidence for a variable to impact the posterior means defining the loss function. Without the strong shrinkage imposed by the point-mass priors, these risk factors show up more strongly in the posterior and hence in the posterior summary. In each case, the three Fama and French factors from \cite{FF3} predictably appear and seem to be the only relevant factors for pricing these 25 portfolios.
Similarly, the weaker shrinkage model in the fixed predictor version (figures \ref{Lossgraph25MFXFIX} and \ref{graph25MFXFIX}) yields yet denser summaries (for a given level of $\kappa$).
\begin{figure}[H]
\centering
\includegraphics[scale=.47]{deltapilambdagraph-25PortMODELFIXEDGenPsi}
\caption{Evaluation of $\Delta_{\lambda}$ and $\pi_{\lambda}$ along the solution path for the 25 size/value portfolios modeled by the 10 factors with alternative prior. An analyst may use this plot to select a particular model. Uncertainty bands are 75\% posterior intervals on the $\Delta_{\lambda}$ metric. The large black dot represents the model selected in Figure \ref{graph25MF}.}
\label{Lossgraph25MF}
\includegraphics[scale=.65]{graph-25PortMODELFIXEDGenPsi}
\vspace{-20mm}
\caption{The selected model for 25 size/value portfolios modeled by the 10 factors with alternative prior. The responses and predictors are colored in gray and white, respectively. Edges represent nonzero components of the optimal action, $\boldsymbol{\gamma}$.}
\label{graph25MF}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[scale=.47]{deltapilambdagraph-25PortMODELFIXEDXFIXGenPsi}
\caption{Evaluation of $\Delta_{\lambda}$ and $\pi_{\lambda}$ along the solution path for the 25 size/value portfolios modeled by the 10 factors assuming \textbf{fixed predictors (factors)} and with alternative prior. An analyst may use this plot to select a particular model. Uncertainty bands are 75\% posterior intervals on the $\Delta_{\lambda}$ metric. The large black dot represents the model selected in Figure \ref{graph25MFXFIX}.}
\label{Lossgraph25MFXFIX}
\includegraphics[scale=.65]{graph-25PortMODELFIXEDXFIXGenPsi}
\vspace{-10mm}
\caption{The selected model for 25 size/value portfolios modeled by the 10 factors when \textbf{uncertainty in future factor returns is not taken into account} and with alternative prior. The responses and predictors are colored in gray and white, respectively. Edges represent nonzero components of the optimal action, $\boldsymbol{\gamma}$.}
\label{graph25MFXFIX}
\end{figure}
\subsubsection{Comparison of four scenarios at fixed $\kappa$}
The selected summary graphs for the four scenarios are displayed together for comparison in figure \ref{summarygraph}. Observe that graphs (c) and (d) selected under the alternative prior are marginally denser than their counterparts (a) and (b) under the point-mass model selection prior. However, the assumption of random predictors results in notably sparser summaries -- graphs (a) and (c) are much sparser than (b) and (d). These comparisons emphasize the impact that incorporating random predictors may have on a variable selection procedure; especially the present approach where we extract point summaries from a posterior by utilizing uncertainty in all unknowns $(\tilde{Y},\tilde{X},\Theta)$.
\begin{figure}[H]
\centering
\subfigure[Random predictors]{\includegraphics[scale=.35]{graph-25PortGenPsi}}
\subfigure[Fixed predictors]{\includegraphics[scale=.35]{graph-25PortXFIXGenPsi}}
\subfigure[Random predictors and alternative prior]{\includegraphics[scale=.35]{graph-25PortMODELFIXEDGenPsi}}
\subfigure[Fixed predictors and alternative prior]{\includegraphics[scale=.35]{graph-25PortMODELFIXEDXFIXGenPsi}}
\caption{Comparison of selected models under four scenarios.}\label{summarygraph}
\end{figure}
\section{Conclusion}
In this paper, we propose a general model selection procedure for multivariate linear models when future realizations of the predictors are unknown. Such models are widely used in many areas of science and economics, including genetics and asset pricing. Our utility-based sparse posterior summary procedure is a multivariate extension of the ``decoupling shrinkage and selection" methodology of \cite{HahnCarvalho}. The approach we develop has three steps: (\textit{i}) fit a Bayesian model, (\textit{ii}) specify a utility function with a sparsity-inducing penalty term and optimize its expectation, and (\textit{iii}) graphically summarize the posterior impact (in terms of utility) of the sparsity penalty. Our utility function is based on the kernel of the conditional distribution responses given the predictors and can be formulated as a tractable convex program. We demonstrate how our procedure may be used in asset pricing under a variety of modeling choices.
The remainder of this discussion takes a step back from the specifics of the seemingly unrelated regressions model and considers a broader role for utility-based posterior summaries.
A paradox of applied Bayesian analysis is that posterior distributions based on relatively intuitive models like the SUR model are often just as complicated as the data itself. For Bayesian analysis to become a routine tool for practical inquiry, methods for summarizing posterior distributions must be developed apace with the models themselves. A natural starting point for developing such methods is decision theory, which suggests developing loss functions specifically geared towards practical posterior summary. As a matter of practical data analysis, articulating an apt loss function has been sorely neglected relative to the effort typically lavished on the model specification stage, specifically prior specification. Ironically (but not surprisingly) our application demonstrates that one's utility function has a dominant effect on the posterior summaries obtained relative to which prior distribution is used.
This paper makes two contributions to this area of ``utility design''. First, we propose that the likelihood function has a role to play in posterior summary apart from its role in inference. That is, one of the great practical virtues of likelihood-based statistics is that the likelihood serves to summarize the data by way of the corresponding point estimates. By using the log-density as our utility function applied to {\em future} data, we revive the fundamental summarizing role of the likelihood. Additionally, note that this approach allows three distinct roles for parameters. First, all parameters of the model appear in defining the posterior predictive distribution. Second, some parameters appear in {\em defining} the loss function; $\Psi$ plays this role in our analysis. Third, some parameters define the action space. In this framework there are no ``nuisance'' parameters that vanish from the estimator as soon as a marginal posterior is obtained. Once the likelihood-based utility is specified, it is a natural next step to consider augmenting the utility to enforce particular features of the desired point summary. For example, our analysis above was based on a utility that explicitly rewards sparsity of the resulting summary. A traditional instance of this idea is the definition of high posterior density regions, which are defined as the {\em shortest, contiguous} interval that contains a prescribed fraction of the posterior mass.
Our second contribution is to consider not just one, but a range, of utility functions and to examine the posterior distributions of the corresponding posterior loss. Specifically, we compare the utility of a sparsified summary to the utility of the optimal non-sparse summary. Interestingly, these utilities are random variables themselves (defined by the posterior distribution) and examining their distributions provides a fundamentally Bayesian way to measure the extent to which the sparsity preference is driving one's conclusions. The idea of comparing a hypothetical continuum of decision-makers based on the posterior distribution of their respective utilities represents a principled Bayesian approach to exploratory data analysis. This is an area of ongoing research.
|
\section{Introduction}
\label{sec:intro}
The main motivation behind all de Finetti type theorems is to reduce
the study of permutation-invariant scenarios to that of i.i.d.~ones,
which are often much easier to understand.
In many information theoretic situations, the problem is posed in such
a way that one almost directly sees that the solution is (or is without
loss of generality) permutation-invariant. Furthermore, in many
scenarios one needs only to upper bound (and not to accurately approximate)
a permutation-invariant object by i.i.d.~ones. The seminal \textit{de Finetti
reduction} (aka \textit{post-selection lemma}) of Christandl, K\"{o}nig and Renner \cite{CKR}
was precisely designed for that: for any permutation-invariant
state $\rho$ on $\mathcal{H}^{\otimes n}$, with $d=|\mathcal{H}|$ the
``local'' Hilbert space dimension,
\begin{equation}
\label{eq:CKR}
\rho \leq (n+1)^{d^2} \int_{\sigma\in\mathcal{D}(\mathcal{H})} \sigma^{\otimes n} \,\mathrm{d}\sigma,
\end{equation}
where $\mathrm{d}\sigma$ is a universal probability measure over the set of mixed states $\mathcal{D}(\mathcal{H})$ on $\mathcal{H}$,
and the inequality refers to the matrix order ($A\leq B$ meaning
that $B-A$ is positive semidefinite).
The beauty of this statement is that on the right hand side we have a
universal object: one and the same convex combination provides the
upper bound to all permutation-invariant states.
At the same time, though, its very universality can be a drawback:
every permutation-invariant state (quantum or classical) is upper bounded
by the same convex combination of tensor power states, so that any
other a priori information (apart from its permutation-symmetry),
that one may have on it, is lost. In~\cite[Appendix~B]{DSW}, it was shown that
at the sole cost of slightly increasing the polynomial pre-factor in
front of the upper bounding de Finetti operator, it is actually possible
to make it depend on the state of interest, or on some property
that this state has, including in the integral on the right hand side
of equation~(\ref{eq:CKR}) a fidelity term between $\rho$ and the
i.i.d.~state $\sigma^{\otimes n}$.
In~\cite{DSW}, this \emph{constrained de Finetti reduction}
was applied to prove a coding
theorem in a setting with adversarially chosen channel.
In~\cite{LaW2} another application to parallel repetition
of no-signalling games was given.
In Section~\ref{sec:finite-de-finetti}, we first review the constrained de
Finetti reduction of~\cite[Appendix~B]{DSW}, for the sake both of completeness and of presenting its proof in a slightly alternative way
(Subsection~\ref{subsec:finite-general}). We then show that certain \emph{linear} constraints lead to very simple
and at the same time useful forms of the de Finetti reduction, such
that certain ``unwanted'' contributions in the integral on the right hand
side of equation~(\ref{eq:CKR}) are either completely absent or exponentially
suppressed (Subsection~\ref{subsec:linear-constraints}).
Next, in Sections~\ref{sec:sep1} and \ref{sec:sep2} we study in depth
the case of separability, a convex constraint. In particular we show
that there are several essentially equivalent ways of thinking about
the exponential decay of the fidelity term.
Inspired by separability, in Section~\ref{sec:general} we present an
axiomatic treatment of a wider class of convex constraints.
Finally, in Section~\ref{sec:infinite} we move to de Finetti reductions
in the infinite-dimensional case.
\section{Flexible de Finetti reductions for finite-dimensional symmetric quantum systems}
\label{sec:finite-de-finetti}
\subsection{A general constrained de Finetti reduction}
\label{subsec:finite-general}
Before getting into more specific statements, let us fix once and for
all some definitions and notation that we shall use throughout the
whole paper. Consider $\mathcal{H}$ a finite-dimensional Hilbert space, and denote by $\{\ket{1},\ldots,\ket{d}\}$ an orthonormal basis of $\mathcal{H}$, where $d=|\mathcal{H}|<+\infty$.
Next, for any natural number $n$ and permutation $\pi\in\mathcal{S}_n$,
define $U_{\pi}$ as the associated permutation unitary on
$\mathcal{H}^{\otimes n}$, characterized by
\[
\forall\ 1\leq j_1,\ldots, j_n\leq d,\ U_{\pi}\ket{j_{1}}\otimes\cdots\otimes\ket{j_{n}}
= \ket{j_{\pi(1)}}\otimes\cdots\otimes\ket{j_{\pi(n)}}.
\]
Note that this definition is independent of the basis.
The $n$-symmetric subspace of $\mathcal{H}^{\otimes n}$ can then be defined as
the simultaneous $+1$-eigenspace of all $U_\pi$'s,
\begin{align*}
\Sym^n(\mathcal{H})
:=& \left\{ \ket{\psi}\in\mathcal{H}^{\otimes n}
\ : \ \forall\ \pi\in\mathcal{S}_n,\ U_{\pi}\ket{\psi}=\ket{\psi} \right\} \\
=& \Span\left\{ \ket{v_{j_1,\ldots,j_n}}=\sum_{\pi\in\mathcal{S}_n}
\ket{j_{\pi(1)}}\otimes\cdots\otimes\ket{j_{\pi(n)}} \ : \ 1\leq j_1\leq\cdots\leq j_n\leq d \right\}.
\end{align*}
The orthogonal projector onto $\Sym^n(\mathcal{H})$ may thus be written as
\[
P_{\Sym^n(\mathcal{H})}
= \sum_{1\leq j_1\leq\cdots\leq j_n\leq d} \proj{\psi_{j_1,\ldots,j_n}}
= {n+d-1 \choose n} \int_{\ket{\psi}\in S_{\mathcal{H}}} \proj{\psi}^{\otimes n}\mathrm{d}\psi,
\]
where for each $1\leq j_1\leq\cdots\leq j_n\leq d$, $\ket{\psi_{j_1,\ldots,j_n}}$
denotes the unit vector having the same direction as $\ket{v_{j_1,\ldots,j_n}}$,
and where $\mathrm{d}\psi$ stands for the uniform probability measure
on the unit sphere $S_{\mathcal{H}}$ of $\mathcal{H}$. The second line is due to Schur's Lemma, since $\Sym^n\left(\mathcal{H}\right)$ is an irreducible representation (irrep) of the commutant action of $\{U_\pi:\pi\in S_n\}$,
the local unitaries $V^{\otimes n}$, $V\in SU(\mathcal{H})$ (see e.g.~\cite{Harrow} for more details).
A state $\rho$ on $\mathcal{H}^{\otimes n}$ is then called permutation-invariant (or simply symmetric) if
$U_{\pi}\rho U_{\pi}^{\dagger} = \rho$ for all $\pi\in\mathcal{S}_n$.
This can be expressed equivalently by saying that there exists a
unit vector $\ket{\psi}\in\Sym^n(\mathcal{H}\otimes\mathcal{H}')$ such that
$\rho = \Tr_{\mathcal{H}'^{\otimes n}} \proj{\psi}$.
Going from rigid to more flexible de Finetti reductions relies essentially on
the so-called ``pinching trick'', which we state formally as Lemma \ref{lemma:pinching} below. This is a generalization of results appearing in \cite{Hayashi} and \cite{HO}.
\begin{lemma}
\label{lemma:pinching}
Let $\mathcal{H}$ be a Hilbert space and $M_1,\ldots,M_r$ be operators on $\mathcal{H}$.
Then, for any state $\rho$ on $\mathcal{H}$,
\[
\sum_{i,j=1}^r M_i\rho M_j^{\dagger} \leq r\sum_{i=1}^r M_i\rho M_i^{\dagger}.
\]
\end{lemma}
\begin{proof}
To prove that Lemma \ref{lemma:pinching} holds for any state on $\mathcal{H}$, it is
sufficient to prove that it holds for any pure state on $\mathcal{H}$. Let therefore
$\ket{\psi}$ be a unit vector in $\mathcal{H}$. Then, for any unit vector $\ket{\varphi}$ in $\mathcal{H}$,
we have by the Cauchy-Schwarz inequality
\[\begin{split}
\bra{\varphi} \left(\sum_{i,j=1}^r M_i\proj{\psi} M_j^{\dagger}\right) \ket{\varphi}
&= \left| \sum_{i=1}^r \bra{\varphi}M_i\ket{\psi} \right|^2 \\
&\leq r \sum_{i=1}^r \left|\bra{\varphi}M_i\ket{\psi} \right|^2 \\
&= \bra{\varphi} \left( r\sum_{i=1}^r M_i\proj{\psi} M_i^{\dagger}\right) \ket{\varphi},
\end{split}\]
which concludes the proof.
\end{proof}
With this tool at hand, we are ready to get, first of all,
the pure state version of the flexible de Finetti reduction.
\begin{proposition}
\label{prop:ps-pure}
Any unit vector $\ket{\theta}\in\Sym^n\left(\mathcal{H}\right)$ satisfies
\[
\proj{\theta} \leq {n+d-1 \choose n}^3 \int_{\ket{\psi}\in S_{\mathcal{H}}}
\left|\langle\theta|\psi^{\otimes n}\rangle\right|^2 \proj{\psi}^{\otimes n}\mathrm{d}\psi.
\]
\end{proposition}
\begin{proof}
Let $\ket{\theta}\in\Sym^n\left(\mathcal{H}\right)$ be a unit vector. Then,
\[
\proj{\theta} = P_{\Sym^n(\mathcal{H})} \proj{\theta} P_{\Sym^n(\mathcal{H})}^{\dagger}
= {n+d-1 \choose n}^2 \int_{\ket{\psi},\ket{\varphi}\in S_{\mathcal{H}}}
\proj{\psi}^{\otimes n} \proj{\theta}
\proj{\varphi}^{\otimes n} \,\mathrm{d}\psi\,\mathrm{d}\varphi.
\]
Now observe, setting $r={n+d-1 \choose n}^2$, that the span of $\left\{ \proj{\psi}^{\otimes n},\ \ket{\psi}\in S_{\mathcal{H}} \right\}$, subject to the condition of having trace $1$, has dimension $r-1$. So by Caratheodory's theorem, we know that there exist $\{p_1,\ldots,p_r\}$, a convex combination, and $\{\psi_1,\ldots,\psi_r\}$, a set of unit vectors in $\mathcal{H}$, such that
\begin{equation} \label{eq:cara} \int_{\ket{\psi}\in S_{\mathcal{H}}} \proj{\psi}^{\otimes n}\mathrm{d}\psi = \sum_{i=1}^r p_i \proj{\psi_i}^{\otimes n}. \end{equation}
We can therefore rewrite
\[ \proj{\theta} = r \sum_{i,j=1}^r p_ip_j \proj{\psi_i}^{\otimes n}\proj{\theta}\proj{\psi_j}^{\otimes n} \leq r^2 \sum_{i=1}^r p_i^2\left|\braket{\theta}{\psi_i^{\otimes n}}\right|^2 \proj{\psi_i}^{\otimes n} \leq r^{3/2} \sum_{i=1}^r p_i\left|\braket{\theta}{\psi_i^{\otimes n}}\right|^2 \proj{\psi_i}^{\otimes n}, \]
where the next to last inequality is by Lemma \ref{lemma:pinching}, and the last inequality is because, for each $1\leq i\leq r$, $p_i\leq 1/\sqrt{r}$ (which can be seen by contracting both sides of equation \eqref{eq:cara} with $\bra{\psi_i^{\otimes n}}\cdot\ket{\psi_i^{\otimes n}}$). And consequently, since this holds for any ensemble $\{p_i,\,\psi_i\}_{1\leq i\leq r}$ satisfying equation \eqref{eq:cara}, we have by convex combination
\[ \proj{\theta} \leq r^{3/2} \int_{\ket{\psi}\in S_{\mathcal{H}}} \left|\langle\theta|\psi^{\otimes n}\rangle\right|^2 \proj{\psi}^{\otimes n} \,\mathrm{d}\psi, \]
which is precisely the advertised result.
\end{proof}
From Proposition \ref{prop:ps-pure}, we can now easily derive the general
mixed state version of our flexible de Finetti reduction, which was originally obtained in \cite{DSW} by a slightly different route.
\begin{theorem}[Cf.~{\cite[Lemma 18]{DSW}}]
\label{th:ps-mixed}
Any symmetric state $\rho$ on $\mathcal{H}^{\otimes n}$ satisfies
\[
\rho \leq {n+d^2-1 \choose n}^3 \int_{\ket{\psi}\in S_{\mathcal{H}\otimes\mathcal{H}'}}
F\left(\rho,\sigma(\psi)^{\otimes n}\right)^2 \sigma(\psi)^{\otimes n} \,\mathrm{d}\psi,
\]
where for a unit vector $\ket{\psi}\in\mathcal{H}\otimes\mathcal{H}'$,
$\sigma(\psi) = \Tr_{\mathcal{H}'}\proj{\psi}$ is the reduced state of $\proj{\psi}$ on $\mathcal{H}$.
\end{theorem}
\begin{proof}
As noted before, there exists a unit vector $\ket{\theta}\in\Sym^n\left(\mathcal{H}\otimes\mathcal{H}'\right)$
such that $\rho=\Tr_{\mathcal{H}'^{\otimes n}}\proj{\theta}$.
By Proposition \ref{prop:ps-pure}, we have
\[
\proj{\theta} \leq {n+d^2-1 \choose n}^3 \int_{\ket{\psi}\in S_{\mathcal{H}\otimes\mathcal{H}'}}
\left|\langle\theta|\psi^{\otimes n}\rangle\right|^2
\proj{\psi}^{\otimes n} \,\mathrm{d}\psi.
\]
Thus, after partial tracing over $\mathcal{H}'^{\otimes n}$, we obtain
\[
\rho\leq {n+d^2-1 \choose n}^3 \int_{\ket{\psi}\in S_{\mathcal{H}\otimes\mathcal{H}'}}
\left|\langle\theta|\psi^{\otimes n}\rangle\right|^2
\sigma(\psi)^{\otimes n} \,\mathrm{d}\psi.
\]
To get the announced result, we then just have to notice that,
by monotonicity of the fidelity under the CPTP map $\Tr_{\mathcal{H}'^{\otimes n}}$,
we have for each $\ket{\psi}\in\mathcal{H}\otimes\mathcal{H}'$,
\[
\left|\langle\theta|\psi^{\otimes n}\rangle\right|
= F\left(\proj{\theta}, \proj{\psi}^{\otimes n} \right)
\leq F\left(\rho,\sigma(\psi)^{\otimes n}\right). \qedhere
\]
\end{proof}
\subsection{Linear constraints}
\label{subsec:linear-constraints}
Let $\rho$ be a symmetric state on $\mathcal{H}^{\otimes n}$.
What Theorem \ref{th:ps-mixed} tells us is that there exists a
probability measure $\mu$ over the set of states on $\mathcal{H}$ such that
\begin{equation}
\label{eq:ps-fidelity}
\rho \leq (n+1)^{3d^2} \int_{\sigma\in\in\mathcal{D}(\mathcal{H})}
F\left(\rho,\sigma^{\otimes n}\right)^2
\sigma^{\otimes n} \,\mathrm{d}\mu(\sigma).
\end{equation}
It may be pointed out that $\mu$ is in fact the uniform probability measure over the set of mixed states on $\mathcal{H}$ (with respect to the Hilbert--Schmidt distance), since the latter is equivalently characterized as the partial trace over an environment $\mathcal{H}'$ having same dimension as $\mathrm{H}$ of uniformly distributed pure states on $\mathcal{H}\otimes\mathcal{H}'$ (see \cite{ZS}).
Observe that, contrary to the original de Finetti reduction, where the upper bound is the same for every symmetric state, we here have a highly
state-dependent upper bound, where only states which have a high
fidelity with the state of interest $\rho$ are given an important weight.
This is especially useful when one knows that $\rho$ satisfies some
additional property. Indeed, one would then expect that, amongst states
of the form $\sigma^{\otimes n}$, only those approximately satisfying
this same property should have a non-negligible fidelity weight. There
are at least two archetypical cases where this intuition can easily be seen to be true.
\begin{corollary}[Cf.~{\cite[Lemma 18]{DSW}}]
\label{cor:product-image}
Let $\mathcal{N}:\mathcal{L}(\mathcal{H})\rightarrow\mathcal{L}(\mathcal{K})$ be a quantum channel, with $d=|\mathcal{H}|<+\infty$.
Assume that $\rho$ is a symmetric state on $\mathcal{H}^{\otimes n}$, which
is additionally satisfying $\mathcal{N}^{\otimes n}(\rho)=\tau_0^{\otimes n}$,
for some given state $\tau_0$ on $\mathcal{K}$. Then,
\[
\rho \leq (n+1)^{3d^2} \int_{\sigma\in\mathcal{D}(\mathcal{H})}
F\left(\tau_0,\mathcal{N}(\sigma)\right)^{2n} \sigma^{\otimes n}\,\mathrm{d}\mu(\sigma).
\]
\end{corollary}
\begin{proof}
This follows directly from inequality \eqref{eq:ps-fidelity}
by monotonicity of the fidelity under the CPTP map $\mathcal{N}$,
and by multiplicativity of the fidelity on tensor products.
\end{proof}
This especially implies that, under the hypotheses of
Corollary \ref{cor:product-image}, we have: for any $0<\delta<1$,
setting
$\mathcal{K}_{\delta}=\left\{ \sigma\in\mathcal{D}(\mathcal{H}) \ : \ F\left(\tau_0,\mathcal{N}(\sigma)\right)\geq 1-\delta\right\}$,
\[
\rho \leq (n+1)^{3d^2} \left( \int_{\sigma\in\mathcal{K}_{\delta}} \sigma^{\otimes n}\mathrm{d}\mu(\sigma)
+ (1-\delta)^{2n}\int_{\sigma\notin\mathcal{K}_{\delta}} \sigma^{\otimes n}\mathrm{d}\mu(\sigma) \right).
\]
Such flexible de Finetti reduction, for states which satisfy the constraint of
being sent to a certain tensor power state by a certain tensor power
CPTP map, has already been fruitfully applied, for instance in the context of
zero-error communication via quantum channel~\cite{DSW}.
\medskip
Another linear constraint is that of a fixed point equation:
\begin{corollary}
\label{cor:fixed-point}
Let $\mathcal{N}:\mathcal{L}(\mathcal{H})\rightarrow\mathcal{L}(\mathcal{H})$ be a quantum channel, with $d=|\mathcal{H}|<+\infty$.
Assume that $\rho$ is a symmetric state on $\mathcal{H}^{\otimes n}$, which
is additionally satisfying $\mathcal{N}^{\otimes n}(\rho)=\rho$. Then,
\[
\rho \leq (n+1)^{3d^2} \int_{\sigma\in\mathcal{D}(\mathcal{H})}
F\left(\rho,\mathcal{N}(\sigma)^{\otimes n}\right)^2 \mathcal{N}(\sigma)^{\otimes n}\mathrm{d}\mu(\sigma).
\]
\end{corollary}
\begin{proof}
Apply $\mathcal{N}^{\otimes n}$ on both sides of inequality \eqref{eq:ps-fidelity},
and use once more the monotonicity of the fidelity under the CPTP map $\mathcal{N}$.
\end{proof}
This means that, under the assumptions of Corollary \ref{cor:fixed-point}, there actually exists a probability measure $\widetilde{\mu}$ over the set of states on $\mathcal{H}$ which belong to the range of $\mathcal{N}$ such that
\begin{equation}
\label{eq:fixed-point}
\rho \leq (n+1)^{3d^2} \int_{\sigma\in\mathrm{Range}(\mathcal{N})}
F\left(\rho,\sigma^{\otimes n}\right)^2 \sigma^{\otimes n}\mathrm{d}\widetilde{\mu}(\sigma).
\end{equation}
A case of particular interest for equation \eqref{eq:fixed-point} is the following. Let $G$ be a subgroup of the unitary group on $\mathcal{H}$, equipped with its Haar measure $\mu_G$ (unique normalised left and right invariant measure over $G$). Its associated twirl is the quantum channel $\mathcal{T}_G:\mathcal{L}(\mathcal{H})\rightarrow\mathcal{L}(\mathcal{H})$ defined by
\[ \mathcal{T}_G:\sigma\mapsto\int_{U\in G}U\sigma U^{\dagger}\mathrm{d}\mu_G(U). \]
The range of $\mathcal{T}_G$ is then precisely the set of states on $\mathcal{H}$ in the commutant of $G$, i.e.~\[ \mathcal{K}_G=\left\{\sigma\in\mathcal{D}(\mathcal{H}) \ : \ \forall\ U\in G,\ [\sigma,U]=0 \right\}. \]
Hence, there exists a probability measure $\widetilde{\mu}$ over $\mathcal{K}_G$ such that, if $\rho$ is a symmetric state on $\mathcal{H}^{\otimes n}$ satisfying $\mathcal{T}_G^{\otimes n}(\rho)=\rho$, then
\[
\rho \leq (n+1)^{3d^2} \int_{\sigma\in\mathcal{K}_G}
F\left(\rho,\sigma^{\otimes n}\right)^2 \sigma^{\otimes n}\mathrm{d}\widetilde{\mu}(\sigma).
\]
Another situation where equation \eqref{eq:fixed-point} might be especially useful is when $\mathcal{N}$ is a quantum-classical channel, so that its range can be identified with the set of classical probability distributions. We get in that case the corollary below.
\begin{corollary}
\label{cor:probability}
Let $\mathcal{X}$ be a finite alphabet and let $P_{\mathcal{X}^n}$ be a
symmetric probability distribution on $\mathcal{X}^n$. There exists
a universal probability measure $\mathrm{d}Q_{\mathcal{X}}$ over the set of
probability distributions on $\mathcal{X}$ such that
\[
P_{\mathcal{X}^n} \leq (n+1)^{3|\mathcal{X}|^2} \int_{Q_{\mathcal{X}}}
F\left(P_{\mathcal{X}^n},Q_{\mathcal{X}}^{\otimes n}\right)^2
Q_{\mathcal{X}}^{\otimes n}\mathrm{d}Q_{\mathcal{X}},
\]
where the inequality sign signifies point-wise inequality between probability
distributions on $\mathcal{X}^n$.
\end{corollary}
\begin{proof}
This is a special case of Corollary \ref{cor:fixed-point}.
Indeed, we know that we can make the identification $\mathcal{X}\equiv\{1,\ldots,d\}$,
where $d=|\mathcal{X}|$. So let $\mathcal{H}$ be a $d$-dimensional Hilbert space, and denote by $\{\ket{1},\ldots,\ket{d}\}$ an orthonormal basis of $\mathcal{H}$. We can then define the ``classical'' state $\rho$ on $\mathcal{H}^{\otimes n}$ by
\[
\rho= \sum_{1\leq x_1,\ldots,x_n\leq d} P(x_1,\ldots,x_n)\proj{x_1\otimes\cdots\otimes x_n},
\]
and the quantum-classical channel $\mathcal{N}:\mathcal{L}(\mathcal{H})\rightarrow\mathcal{L}(\mathcal{H})$ by
\[
\mathcal{N}:\sigma \mapsto \sum_{1\leq x\leq d} Q_{\sigma}(x)\proj{x}
= \sum_{1\leq x\leq d} \proj{x}\sigma\proj{x}.
\]
By assumption on $P$, $\rho$ is a symmetric state on $\mathcal{H}^{\otimes n}$, which is additionally, by construction, a fixed point of $\mathcal{N}^{\otimes n}$. Hence, by Corollary \ref{cor:fixed-point},
\[ \rho \leq (n+1)^{3d^2} \int_{\sigma\in\mathcal{D}(\mathcal{H})} F\left(\rho,\mathcal{N}(\sigma)^{\otimes n}\right)^2 \mathcal{N}(\sigma)^{\otimes n}\mathrm{d}\mu(\sigma). \]
By the way $\rho$ and $\mathcal{N}$ have been designed, this actually translates into the point-wise inequality
\[ \forall\ 1\leq x_1,\ldots,x_n\leq d,\ P(x_1,\ldots,x_n) \leq (n+1)^{3d^2} \int_{\sigma\in\mathcal{D}(\mathcal{H})} F\left(P,Q_{\sigma}^{\otimes n}\right)^2 Q_{\sigma}(x_1)\cdots Q_{\sigma}(x_n)\mathrm{d}\mu\left(Q_{\sigma}\right), \]
which is exactly the announced result.
\end{proof}
This flexible de Finetti reduction for probability distributions turns out
to be especially useful when studying the parallel repetition of multi-player
non-local games, as exemplified in \cite{LaW2}.
\begin{remark}
Note that all these results generalize to non-normalized permutation
invariant positive semidefinite operators on finite-dimensional spaces (or positive
distributions on finite alphabets). One just has to extend the usual
definition of the fidelity by setting $F(M,N)=\|\sqrt{M}\sqrt{N}\|_1$
for any positive semidefinite operators (or positive distributions) $M,N$.
\end{remark}
\subsection{On to convex constraints?}
\label{subsec:convex-constraints}
We just saw that, in the case where the symmetric state $\rho$ under
consideration is additionally known to satisfy certain linear constraints,
it is possible to upper bound it by a de Finetti operator where
either no or exponentially small weight is given to tensor power
states which do not satisfy this same constraint. But what about
the case where the a priori information on $\rho$ is that it
belongs or not to a certain convex subset of states? This is the
question we investigate in the sequel, focussing first in
Sections \ref{sec:sep1} and \ref{sec:sep2} on the paradigmatic example
of the set of separable states, and then describing in
Section \ref{sec:general} the general setting in which similar conclusions hold.
\section{Exponential decay and concentration of $h_{sep}$ via de Finetti reduction approach}
\label{sec:sep1}
As we just mentioned, we will now be interested for a while in the case where the underlying Hilbert space is a tensor product Hilbert space $\mathcal{H} = \mathrm{A}\otimes \mathrm{B}$, and the kind of symmetric states on $\mathcal{H}^{\otimes n}$ that we will look at are those which additionally satisfy the (convex but non-linear) constraint of being separable across the bipartite cut $\mathrm{A}^{\otimes n}{:}\mathrm{B}^{\otimes n}$. For such a state $\rho$, one can of course still write down a de Finetti reduction of the form
\[ \rho \leq (n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \int_{\sigma\in\mathcal{D}(\mathrm{A}\otimes\mathrm{B})}
F\left(\rho,\sigma^{\otimes n}\right)^2 \sigma^{\otimes n}\,\mathrm{d}\mu(\sigma). \]
And what we would like to understand is whether it is possible to argue that only the states $\sigma^{\otimes n}$ which are such that $\sigma$ is separable across the bipartite cut $\mathrm{A}{:}\mathrm{B}$ are given a non exponentially small weight in this integral representation. As we shall see, this question is especially relevant when analysing the multiplicative behaviour of the support function of the set of biseparable states.
So let us specify a bit what we have in mind. Given a positive operator $M$ on $\mathrm{A}\otimes \mathrm{B}$, its maximum overlap with states which are
separable across the bipartite cut $\mathrm{A}{:}\mathrm{B}$, which we denote by $\mathcal{S}(\mathrm{A}{:}\mathrm{B})$, is defined as
\[
h_{sep}(M) = \sup_{\sigma\in\mathcal{S}(\mathrm{A}{:}\mathrm{B})} \Tr(M\sigma).
\]
Here, we are interested in understanding how this quantity behaves
under tensoring. Concretely, this means that we want to know,
for any $n\in\mathbf{N}$, how $h_{sep}(M^{\otimes n})$ relates to $h_{sep}(M)$
(where the former quantity is defined as the maximum overlap of $M^{\otimes n}$ with states which are separable across the bipartite cut $\mathrm{A}^{\otimes n}{:}\mathrm{B}^{\otimes n}$). Because $h_{sep}$ is linear homogeneous in its argument, we can always rescale $M$ by a positive constant such that $0\leq M \leq \openone$, meaning that $M$ can be interpreted as a POVM element of the binary test with operators $(M,\openone-M)$. We shall make this assumption throughout from now on.
Then, it is easy to see that, for any $n\in\mathbf{N}$, we have the inequalities
\begin{equation}
\label{eq:h_SEP-h_SEPn}
h_{sep}(M)^n \leq h_{sep}(M^{\otimes n}) \leq h_{sep}(M) \leq 1.
\end{equation}
But in the case where $h_{sep}(M)<1$, the gap between the lower and upper bounds in
equation \eqref{eq:h_SEP-h_SEPn} grows exponentially with $n$, making these inequalities very little
informative.
This problem is interesting in itself, but also because it connects to plethora of others, some of them even outside the purely quantum information range of applications. The reader is referred to \cite{HM} for a full list of problems which are exactly or approximately equivalent to estimating $h_{sep}$. Two notable applications of $h_{sep}$ arise in quantum computing and in quantum Shannon theory: The first is to $\mathrm{QMA}(2)$, the class of quantum Merlin-Arthur interactive proof systems with two unentangled provers. The setting is that a verifier requires states $\alpha$ and $\beta$ from separate provers which are assumed to be computationally unlimited, and then performs a binary test with POVM $(M,\openone-M)$ on the separable state $\alpha\otimes\beta$. The maximum probability of passing the test that the provers can achieve, evidently equals precisely $h_{sep}(M)$. For complexity theoretic considerations (in particular the so-called soundness gap amplification) it is important to understand how well many instances of the same test, performed in parallel, can be passed -- either all $n$, leading to $h_{sep}(M^{\otimes n})$, or $t$ out of $n$, where $t > n h_{sep}(M)$. The second application appears in the problem of minimum output entropies of quantum channels, and their asymptotic behaviour. Namely, a quantum channel $\mathcal{N} : \mathcal{L}(\mathrm{A}) \rightarrow\mathcal{L}(\mathrm{B})$ can be represented in Stinespring form $\mathcal{N}(\rho) = \Tr_{\mathrm{E}} (V \rho V^{\dagger})$, with an isometry $V:\mathrm{A}\hookrightarrow \mathrm{B} \otimes \mathrm{E}$. Its minimum output R\'{e}nyi $p$-entropy is given by
\[ \widehat{S}_p(\mathcal{N}) = \min_{\rho\in\mathcal{D}(\mathrm{B})} S_p(\mathcal{N}(\rho)),\ \text{where}\ \forall\ \sigma\in\mathcal{D}(\mathrm{A}),\ S_p(\sigma) = \frac{1}{1-p}\log \Tr\sigma^p. \]
For $p=1$, taking the limit, we recover the von Neumann entropy, while for $p=\infty$, $S_{\infty}(\sigma) = -\log\| \sigma \|_\infty$. From this, it is not hard to see that, with $M = VV^{\dagger}$ the projector onto the range of $V$, i.e.~the subspace $V(\mathrm{A}) \subset \mathrm{B}\otimes \mathrm{E}$, we have $\widehat{S}_{\infty}(\mathcal{N}) = -\log h_{sep}(M)$. In quantum Shannon theory, the asymptotic behaviour of $\widehat{S}_p(\mathcal{N}^{\otimes n})$ is of great interest.
\subsection{Some general facts about ``filtered by measurements'' distance measures}
We need to introduce first a few definitions and properties regarding ``filtered by measurements'' distance measures.
Let $\mathcal{H}$ be a Hilbert space and let $\mathbf{M}$ be a set of POVMs on $\mathcal{H}$. For any states $\rho,\sigma$ on $\mathcal{H}$, we define their measured by $\mathbf{M}$ trace-norm distance as
\[ D_{\mathbf{M}}(\rho,\sigma)=\sup_{\mathcal{M}\in\mathbf{M}}\frac{1}{2}\left\|\mathcal{M}(\rho)-\mathcal{M}(\sigma)\right\|_1, \]
and their measured by $\mathbf{M}$ fidelity distance as
\[ F_{\mathbf{M}}(\rho,\sigma)=\inf_{\mathcal{M}\in\mathbf{M}}F\left(\mathcal{M}(\rho),\mathcal{M}(\sigma)\right). \]
We have the well-known relations between these two distances (see e.g.~\cite{NC}, Chapter 9)
\begin{equation} \label{eq:trace-fidelity}
1-F_{\mathbf{M}}\leq D_{\mathbf{M}} \leq \left(1-F_{\mathbf{M}}^2\right)^{1/2}.
\end{equation}
We further define, for any set of states $\mathcal{K}$ on $\mathcal{H}$, the measured by $\mathbf{M}$ trace-norm distance of $\rho$ to $\mathcal{K}$ as
\[ D_{\mathbf{M}}\left(\rho,\mathcal{K}\right)=\inf_{\sigma\in\mathcal{K}}D_{\mathbf{M}}(\rho,\sigma), \]
and the measured by $\mathbf{M}$ fidelity distance of $\rho$ to $\mathcal{K}$ as
\[ F_{\mathbf{M}}\left(\rho,\mathcal{K}\right)=\sup_{\sigma\in\mathcal{K}}F_{\mathbf{M}}(\rho,\sigma). \]
In the sequel, we shall consider the case where $\mathcal{H}=\mathrm{A}\otimes \mathrm{B}$ is a tensor product Hilbert space, with $|\mathrm{A}|,|\mathrm{B}|<+\infty$. In this setting, we denote by $\mathcal{S}$ the set of separable states and by $\mathbf{SEP}$ the set of separable POVMs on $\mathcal{H}$ (in the bipartite cut $\mathrm{A}{:}\mathrm{B}$).
\begin{lemma} \label{lemma:F_SEP}
Let $\mathrm{A}_1,\mathrm{B}_1,\mathrm{A}_2,\mathrm{B}_2$ be Hilbert spaces, and let $\rho_1$ be a state on $\mathrm{A}_1\otimes \mathrm{B}_1$, $\rho_2$ be a state on $\mathrm{A}_2\otimes \mathrm{B}_2$. Then,
\[ F\big(\rho_1\otimes\rho_2,\mathcal{S}(\mathrm{A}_1\mathrm{A}_2{:}\mathrm{B}_1\mathrm{B}_2)\big) \leq F_{\mathbf{SEP}}\big(\rho_1,\mathcal{S}(\mathrm{A}_1{:}\mathrm{B}_1)\big) F\big(\rho_2,\mathcal{S}(\mathrm{A}_2{:}\mathrm{B}_2)\big). \]
\end{lemma}
\begin{proof}
The proof is directly inspired from \cite{Piani}, adapted here to the case of fidelities rather than relative entropies.
Let $\mathcal{M}_1\equiv \big(M_1^{(i)}\big)_{i\in I}\in\mathbf{SEP}(\mathrm{A}_1{:}\mathrm{B}_1)$. Then, by monotonicity of the fidelity under the CPTP map $\mathcal{M}_1\otimes\mathcal{I}_2$, we have
\begin{align*} \sup_{\sigma_{12}\in\mathcal{S}(\mathrm{A}_1\mathrm{A}_2{:}\mathrm{B}_1\mathrm{B}_2)} F\left(\rho_1\otimes\rho_2,\sigma_{12}\right) \leq & \sup_{\sigma_{12}\in\mathcal{S}(\mathrm{A}_1\mathrm{A}_2{:}\mathrm{B}_1\mathrm{B}_2)} F\left(\mathcal{M}_1\otimes\mathcal{I}_2(\rho_1\otimes\rho_2),\mathcal{M}_1\otimes\mathcal{I}_2(\sigma_{12})\right)\\
= & F\left(\mathcal{M}_1\otimes\mathcal{I}_2(\rho_1\otimes\rho_2),\mathcal{M}_1\otimes\mathcal{I}_2(\widetilde{\sigma}_{12})\right), \end{align*}
for some $\widetilde{\sigma}_{12}\in\mathcal{S}(\mathrm{A}_1\mathrm{A}_2{:}\mathrm{B}_1\mathrm{B}_2)$. And,
\[ F\left(\mathcal{M}_1\otimes\mathcal{I}_2(\rho_1\otimes\rho_2), \mathcal{M}_1\otimes\mathcal{I}_2(\widetilde{\sigma}_{12})\right) = \sum_{i\in I}\sqrt{\Tr\left(M_1^{(i)}\rho_1\right)}\sqrt{\Tr\left(M_1^{(i)}\widetilde{\sigma}_1\right)} F\left(\rho_2,\widetilde{\sigma}_2^{(i)}\right), \]
where $\widetilde{\sigma}_1=\Tr_{\mathrm{A}_2\mathrm{B}_2}\left(\widetilde{\sigma}_{12}\right)\in\mathcal{S}(\mathrm{A}_1{:}\mathrm{B}_1)$ and for all $i\in I$, $\widetilde{\sigma}_2^{(i)}= \Tr_{\mathrm{A}_1\mathrm{B}_1}\big(M_1^{(i)}\otimes\openone_2 \widetilde{\sigma}_{12}\big)/\Tr_{\mathrm{A}_1\mathrm{B}_1}\big(M_1^{(i)}\widetilde{\sigma}_1\big)\in\mathcal{S}(\mathrm{A}_2{:}\mathrm{B}_2)$. Hence, for all $i\in I$, $F\big(\rho_2,\widetilde{\sigma}_2^{(i)}\big)\leq \sup_{\sigma_2\in\mathcal{S}(\mathrm{A}_2{:}\mathrm{B}_2)} F\left(\rho_2,\sigma_2\right)$, and subsequently
\begin{align*}
\sum_{i\in I}\sqrt{\Tr\left(M_1^{(i)}\rho_1\right)}\sqrt{\Tr\left(M_1^{(i)}\widetilde{\sigma}_1\right)} F\left(\rho_2,\widetilde{\sigma}_2^{(i)}\right) \leq & \left(\sup_{\sigma_2\in\mathcal{S}_{\mathrm{A}_2:\mathrm{B}_2}} F\left(\rho_2,\sigma_2\right) \right) F\left(\mathcal{M}_1(\rho_1),\mathcal{M}_1(\widetilde{\sigma}_1)\right)\\
\leq & \left(\sup_{\sigma_2\in\mathcal{S}(\mathrm{A}_2{:}\mathrm{B}_2)} F\left(\rho_2,\sigma_2\right)\right) \left(\sup_{\sigma_1\in\mathcal{S}(\mathrm{A}_1{:}\mathrm{B}_1)} F\left(\mathcal{M}_1(\rho_1),\mathcal{M}_1(\sigma_1)\right)\right).
\end{align*}
We thus have shown that, for any $\mathcal{M}_1\in\mathbf{SEP}(\mathrm{A}_1{:}\mathrm{B}_1)$,
\[ F\left(\rho_1\otimes\rho_2,\mathcal{S}(\mathrm{A}_1\mathrm{A}_2{:}\mathrm{B}_1\mathrm{B}_2)\right) \leq \left(\sup_{\sigma_1\in\mathcal{S}(\mathrm{A}_1{:}\mathrm{B}_1)} F\left(\mathcal{M}_1(\rho_1),\mathcal{M}_1(\sigma_1)\right) \right) F\left(\rho_2,\mathcal{S}(\mathrm{A}_2{:}\mathrm{B}_2)\right). \]
Taking the infimum over $\mathcal{M}_1\in\mathbf{SEP}(\mathrm{A}_1{:}\mathrm{B}_1)$, we get precisely the statement in Lemma \ref{lemma:F_SEP}.
\end{proof}
\begin{theorem}
\label{th:F_SEP}
Let $\mathrm{A},\mathrm{B}$ be Hilbert spaces, and let $\rho$ be a state on $\mathrm{A}\otimes\mathrm{B}$.
Then, for any $n\in\mathbf{N}$,
\[
F\big(\rho^{\otimes n},\mathcal{S}(\mathrm{A}^n{:}\mathrm{B}^n)\big)
\leq F_{\mathbf{SEP}}\big(\rho,\mathcal{S}(\mathrm{A}{:}\mathrm{B})\big)^n.
\]
\end{theorem}
\begin{proof}
Theorem \ref{th:F_SEP} is a direct corollary of Lemma \ref{lemma:F_SEP}, obtained by iterating the latter.
\end{proof}
\subsection{Weak multiplicativity of $h_{sep}$}
With these facts prepared, we can now derive our main theorem.
\begin{theorem}
\label{th:h_sep-n}
Let $M$ be an operator on the tensor product Hilbert space $\mathrm{A}\otimes \mathrm{B}$, satisfying $0 \leq M \leq \openone$, and set $r=\|M\|_2$. If $h_{sep}(M)\leq 1-\delta$, for some $0<\delta<1$,
then for any $n\in\mathbf{N}$,
\[
h_{sep}(M^{\otimes n}) \leq \left(1-\frac{\delta^2}{5r^2}\right)^n.
\]
\end{theorem}
\begin{proof}
Let $\rho\in\mathcal{S}(\mathrm{A}^{n}{:}\mathrm{B}^{n})$. Our goal will be first of all to show that $\Tr\left(M^{\otimes n}\rho\right)\leq 2(n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2}\left(1-\delta^2/2r^2\right)^n$.
Now, observe that
\[ \Tr\left(M^{\otimes n}\rho\right) = \Tr\left( \left(\frac{1}{n!}\sum_{\pi\in\mathcal{S}_n}U_{\pi}M^{\otimes n}U_{\pi}^{\dagger}\right) \rho\right) = \Tr\left(M^{\otimes n} \left(\frac{1}{n!}\sum_{\pi\in\mathcal{S}_n}U_{\pi}^{\dagger}\rho U_{\pi}\right) \right), \]
the first equality being by $n$-symmetry of $M^{\otimes n}$ and the second one by cyclicity
of the trace. Hence, for our purposes, we may actually assume without loss of generality
that $\rho\in\mathcal{S}(\mathrm{A}^{n}{:}\mathrm{B}^{n})$ is $n$-symmetric.
Yet, if $\rho$ is an $n$-symmetric state on $\left(\mathrm{A}\otimes\mathrm{B}\right)^{\otimes n}$,
we know by Theorem \ref{th:ps-mixed} that there exists a probability measure
$\mu$ on the set of states on $\mathrm{A}\otimes\mathrm{B}$ such that
\[
\rho \leq (n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \int_{\sigma\in\mathcal{D}(\mathrm{A}\otimes\mathrm{B})}
F\left(\rho,\sigma^{\otimes n}\right)^2 \sigma^{\otimes n}\,\mathrm{d}\mu(\sigma).
\]
So, by multiplicativity of the trace on tensor products, we get in that case
\[
\Tr\left(M^{\otimes n}\rho\right)
\leq (n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \int_{\sigma\in\mathcal{D}(\mathrm{A}\otimes\mathrm{B})}
F\left(\rho,\sigma^{\otimes n}\right)^2 \Tr\left(M\sigma\right)^n\,\mathrm{d}\mu(\sigma).
\]
Consequently, for any $0<\epsilon<1$, setting $\mathcal{K}_{\epsilon}=\left\{\sigma\in\mathcal{D}(\mathrm{A}\otimes\mathrm{B}) \ : \ \left\|\sigma-\mathcal{S}(\mathrm{A}{:}\mathrm{B})\right\|_2\leq \epsilon/r\right\}$, we have, upper bounding either $F\left(\rho,\sigma^{\otimes n}\right)$ or $\Tr\left(M\sigma\right)$ by $1$,
\[ \Tr\left(M^{\otimes n}\rho\right) \leq\, (n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \left( \int_{\sigma\in\mathcal{K}_{\epsilon}} \Tr\left(M\sigma\right)^n \mathrm{d}\mu(\sigma) + \int_{\sigma\notin\mathcal{K}_{\epsilon}} F\left(\rho,\sigma^{\otimes n}\right)^2 \mathrm{d}\mu(\sigma)\right). \]
Now, if $\sigma\in\mathcal{K}_{\epsilon}$, this means that there exists $\tau\in\mathcal{S}(\mathrm{A}{:}\mathrm{B})$ such that $\|\sigma-\tau\|_2\leq\epsilon/r$, so that
\[
\Tr(M\sigma) = \Tr(M\tau)+\Tr(M(\sigma-\tau))
\leq \Tr(M\tau)+ \|M\|_2\|\sigma-\tau\|_2
\leq 1-\delta+\epsilon.
\]
The next to last inequality is simply by the Cauchy--Schwarz inequality,
while the last one is by assumption on $M$, $\tau$, $\sigma$. And if
$\sigma\notin\mathcal{K}_{\epsilon}$, then
\[ F\left(\rho,\sigma^{\otimes n}\right) \leq F\left(\sigma^{\otimes n}, \mathcal{S}(\mathrm{A}^{n}{:}\mathrm{B}^{n})\right) \leq F_{\mathbf{SEP}}\left(\sigma, \mathcal{S}(\mathrm{A}{:}\mathrm{B})\right)^n \leq \left(1-\frac{\epsilon^2}{4r^2}\right)^{n/2}. \]
The first inequality is because $\rho\in\mathcal{S}(\mathrm{A}^{n}{:}\mathrm{B}^{n})$, the second one is by Theorem \ref{th:F_SEP}, and the third one is obtained by combining equation \eqref{eq:trace-fidelity} with the known lower bound $\left\|\sigma-\mathcal{S}(\mathrm{A}{:}\mathrm{B})\right\|_{\mathbf{SEP}}\geq \left\|\sigma-\mathcal{S}(\mathrm{A}{:}\mathrm{B})\right\|_2$ (see e.g.~\cite{LaW1}).
Putting everything together, we obtain in the end that for any $0<\epsilon<1$,
\begin{equation} \label{eq:epsilon} \Tr\left(M^{\otimes n}\rho\right) \leq\, (n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \left( (1-\delta+\epsilon)^n + \left(1-\frac{\epsilon^2}{4r^2}\right)^n \right). \end{equation}
In particular, choosing $\epsilon=2r^2\left((1+\delta/r^2)^{1/2}-1\right)$ in equation \eqref{eq:epsilon}, so that $\epsilon^2/4r^2=\delta-\epsilon\geq \delta^2/5r^2$, we get $\Tr\left(M^{\otimes n}\rho\right) \leq 2(n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \left(1-\delta^2/5r^2\right)^n$. And consequently
\begin{equation}
\label{eq:h_SEPn}
h_{sep}(M^{\otimes n}) \leq 2(n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2}\left(1-\frac{\delta^2}{5r^2}\right)^n.
\end{equation}
In order to conclude, we just need to remove the polynomial pre-factor in equation \eqref{eq:h_SEPn}.
Assume that there exists a constant $C>0$ such that
$h_{sep}(M^{\otimes N})\geq C\left(1-\delta^2/5r^2\right)^N$ for some $N\in\mathbf{N}$.
Then, we would have for any $n\in\mathbf{N}$,
\[
h_{sep}\left(M^{\otimes Nn}\right) \geq h_{sep}(M^{\otimes N})^n
\geq C^n\left(1-\frac{\delta^2}{5r^2}\right)^{Nn}.
\]
On the other hand, equation \eqref{eq:h_SEPn} says that we also have
\[
h_{sep}\left(M^{\otimes Nn}\right) \leq 2(Nn+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \left(1-\frac{\delta^2}{5r^2}\right)^{Nn}.
\]
Letting $n$ grow, we see that the only option to make these two inequalities
compatible is to have $C\leq 1$, which is precisely what we wanted to show.
\end{proof}
The conclusion of Theorem \ref{th:h_sep-n} had already been obtained via completely
different techniques than the one presented here (and even with slightly better constants).
However, the good thing about the de Finetti reduction approach is that it
gives, almost for free, not only this exponential decay result for the behaviour
of $h_{sep}$ under tensoring, but also some kind of concentration statement.
To be precise, assume that $M$ is an operator on $\mathrm{A}\otimes\mathrm{B}$, satisfying $0 \leq M \leq \openone$
and $h_{sep}(M)\leq 1-\delta$ for some $0<\delta<1$.
Then, $M$ can be identified with a binary test that a separable state is
guaranteed to pass only with probability $h_{sep}(M)\leq 1-\delta$,
while there exists some (entangled) state that would pass it with probability $h_{all}(M)=\|M\|_{\infty}$,
which may be $1$.
Hence, a natural question would be: performing this test $n$ times in parallel,
what is the probability that a separable state passes a certain fraction $t/n$ of them?
Such maximum probability is nothing else than $h_{sep}\left(M^{(t/n)}\right)$, where the operator $M^{(t/n)}$ on $(\mathrm{A}\otimes\mathrm{B})^{\otimes n}$ is defined as
\[ M^{(t/n)} = \sum_{I\subset[n],\,|I|\geq t}M^{\otimes I}\otimes\openone^{\otimes I^c} . \]
Obviously, if $t < (1-\delta)n$ then the answer is asymptotically $1$,
whereas for $t=n$ the answer is $h_{sep}(M^{\otimes n})$, which decays
exponentially fast with $n$ as established in Theorem \ref{th:h_sep-n}.
But is such exponential amplification of the failing probability already
true for $t$ just slightly above $(1-\delta)n$? Theorem \ref{th:h_sep-t/n}
answers this question affirmatively.
\begin{theorem}
\label{th:h_sep-t/n}
Let $M$ be an operator on the tensor product Hilbert space $\mathrm{A}\otimes \mathrm{B}$,
satisfying $0 \leq M \leq \openone$, and set $r=\|M\|_2$.
If $h_{sep}(M)\leq 1-\delta$ for some $0<\delta<1$, then for any
$n,t\in\mathbf{N}$ with $t\geq(1-\delta+\alpha)n$ for some $0<\alpha\leq\delta$, we have
\[
h_{sep}\left(M^{(t/n)}\right)\leq\exp\left(-n\frac{\alpha^2}{5r^2}\right).
\]
\end{theorem}
\begin{proof}
Following the exact same lines as in the proof of Theorem \ref{th:h_sep-n}, we now have in place of equation \eqref{eq:epsilon}
\begin{equation} \label{eq:epsilon'} h_{sep}\left(M^{(t/n)}\right) \leq\, (n+1)^{3|\mathrm{A}|^2|\mathrm{B}|^2} \left( \exp\left[-2n(\alpha-\epsilon)^2\right] + \exp\left[-n\frac{\epsilon^2}{4r^2}\right] \right). \end{equation}
This is indeed a consequence of Hoeffding's inequality (and of the fact that $e^{-x}\geq 1-x$ for any $x>0$). So in particular, choosing $\epsilon=\alpha\left(1-(\sqrt{2}-1)/(8r^2-1)\right)$ in equation \eqref{eq:epsilon'}, so that $\epsilon^2/4r^2=2(\alpha-\epsilon)^2\geq\alpha^2/5r^2$, and removing the polynomial pre-factor by the same trick as in the proof of Theorem \ref{th:h_sep-n}, we get as announced
\[ h_{sep}\left(M^{(t/n)}\right)\leq\exp\left(-n\frac{\alpha^2}{5r^2}\right). \qedhere \]
\end{proof}
\section{Exponential decay and concentration of $h_{sep}$ via entanglement measure approach}
\label{sec:sep2}
\subsection{Quantifying the disturbance induced by measurements}
Let us state first a few technical lemmas that we will need later on to establish our main result. In what follows, we will use a few standard definitions from quantum Shannon theory, which we recall here: The entropy of a state $\rho$ is defined as $S(\rho)=-\Tr(\rho\log\rho)$. From there, one can define the mutual information of a bipartite state $\rho_{\mathrm{A}\mathrm{B}}$ and the conditional mutual information of a tripartite state $\rho_{\mathrm{A}\mathrm{B}\mathrm{C}}$ as, respectively,
\begin{align*}
& I(\mathrm{A}:\mathrm{B})_{\rho}=S(\rho_{\mathrm{A}})+S(\rho_{\mathrm{B}})-S(\rho_{\mathrm{A}\mathrm{B}}),\\
& I(\mathrm{A}:\mathrm{B}|\mathrm{C})_{\rho}= S(\rho_{\mathrm{A}\mathrm{C}})+S(\rho_{\mathrm{B}\mathrm{C}})-S(\rho_{\mathrm{C}})-S(\rho_{\mathrm{A}\mathrm{B}\mathrm{C}}).
\end{align*}
Finally, the relative entropy between states $\rho$ and $\sigma$ is defined as $D(\rho\|\sigma)=\Tr(\rho(\log\rho-\log\sigma))$.
\begin{lemma}
\label{lemma:post-POVM}
Let $\rho$ be a state on $\mathrm{U}\otimes\mathrm{V}$ and let $T$ be an operator on $\mathrm{U}$, satisfying $0\leq T\leq\openone$. Next, define $p=\Tr_{\mathrm{U}\mathrm{V}}\left[\left(T_{\mathrm{U}}\otimes\openone_{\mathrm{V}}\right)\rho_{\mathrm{U}\mathrm{V}}\right]$ as the probability of obtaining the first outcome when the two-outcome POVM $\left(T_{\mathrm{U}}\otimes\openone_{\mathrm{V}}, (\openone_{\mathrm{U}}-T_{\mathrm{U}})\otimes\openone_{\mathrm{V}}\right)$ is performed on $\rho_{\mathrm{U}\mathrm{V}}$, and $\tau_{\mathrm{V}}=\Tr_{\mathrm{U}}\left[\left(T_{\mathrm{U}}\otimes\openone_{\mathrm{V}}\right)\rho_{\mathrm{U}\mathrm{V}}\right]/p$ as the corresponding post-measurement state on $\mathrm{V}$. Also, denote by $\rho_{\mathrm{V}}=\Tr_{\mathrm{U}}\left[\rho_{\mathrm{U}\mathrm{V}}\right]$ the reduced state of $\rho_{\mathrm{U}\mathrm{V}}$ on $\mathrm{V}$. Then,
\[ D\left(\tau_{\mathrm{V}}\big\|\rho_{\mathrm{V}}\right) \leq - \log p.\]
\end{lemma}
\begin{proof}
Note that $\rho_{\mathrm{V}}= p\tau_{\mathrm{V}} +(1-p)\sigma_{\mathrm{V}}$, where $\sigma_{\mathrm{V}}=\Tr_{\mathrm{U}}\left[\left((\openone_{\mathrm{U}}-T_{\mathrm{U}})\otimes\openone_{\mathrm{V}}\right) \rho_{\mathrm{U}\mathrm{V}}\right]/(1-p)$. We therefore have the operator inequality $p\tau_{\mathrm{V}}\leq\rho_{\mathrm{V}}$. And hence,
\[ D\left(\tau_{\mathrm{V}}\big\|\rho_{\mathrm{V}}\right) = \Tr\left[\tau_{\mathrm{V}}\left(\log\tau_{\mathrm{V}}-\log\rho_{\mathrm{V}}\right)\right] \leq \Tr\left[\tau_{\mathrm{V}}\left(\log\tau_{\mathrm{V}}-\log (p\tau_{\mathrm{V}})\right)\right] = -\log p, \]
the next to last inequality being because $\log$ is an operator monotone function.
\end{proof}
Let us recall the definition of the squashed entanglement $E_{sq}$, introduced in \cite{CW}:
\[ E_{sq}\left(\rho_{\mathrm{A}\mathrm{B}}\right) = \inf\left\{ \frac{1}{2} I(\mathrm{A}:\mathrm{B}|\mathrm{E})_{\rho} \ : \ \Tr_{\mathrm{E}}\left(\rho_{\mathrm{A}\mathrm{B}\mathrm{E}}\right)=\rho_{\mathrm{A}\mathrm{B}} \right\}. \]
\begin{lemma} \label{lemma:E_sq}
Let $M_{\mathrm{A}\mathrm{B}}$ be an operator on the tensor product Hilbert space $\mathrm{A}\otimes\mathrm{B}$, satisfying $0\leq M_{\mathrm{A}\mathrm{B}}\leq \openone$, and let $\alpha_{\mathrm{A}^n},\beta_{\mathrm{B}^n}$ be states on $\mathrm{A}^{\otimes n},\mathrm{B}^{\otimes n}$ respectively. Next, fix $1\leq k\leq n-1$, and define
\[ p_k=\Tr_{\mathrm{A}^n\mathrm{B}^n}\left[\left(M_{\mathrm{A}\mathrm{B}}^{\otimes k}\otimes\openone_{\mathrm{A}\mathrm{B}}^{\otimes n-k}\right)\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}\right]\ \text{and}\ \tau(k)_{\mathrm{A}^{n-k}\mathrm{B}^{n-k}}=\frac{1}{p_k}\Tr_{\mathrm{A}^k\mathrm{B}^k}\left[\left(M_{\mathrm{A}\mathrm{B}}^{\otimes k}\otimes\openone_{\mathrm{A}\mathrm{B}}^{\otimes n-k}\right)\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}\right].\]
Then,
\[ \sum_{j=k+1}^n E_{sq}\left(\tau(k)_{\mathrm{A}_j\mathrm{B}_j}\right) \leq \frac{1}{2}\log \frac{1}{p_k} .\]
\end{lemma}
\begin{proof}
By Lemma \ref{lemma:post-POVM}, with $\mathrm{U}=\mathrm{A}^{\otimes k}\otimes\mathrm{B}^{\otimes k}$, $\mathrm{V}=\mathrm{A}^{\otimes n-k}\otimes\mathrm{B}^{\otimes n-k}$, $T_{\mathrm{U}}=M_{\mathrm{A}\mathrm{B}}^{\otimes k}$ and $\rho_{\mathrm{U}\mathrm{V}}=\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}$, we have
\[ D\left(\tau(k)_{\mathrm{A}^{n-k}\mathrm{B}^{n-k}}\big\|\alpha_{\mathrm{A}^{n-k}}\otimes\beta_{\mathrm{B}^{n-k}}\right) \leq \log \frac{1}{p_k}. \]
Now, observe that
\begin{align*}
D\left(\tau(k)_{\mathrm{A}^{n-k}\mathrm{B}^{n-k}}\big\|\alpha_{\mathrm{A}^{n-k}}\otimes\beta_{\mathrm{B}^{n-k}}\right) \geq & \, D\left(\tau(k)_{\mathrm{A}^{n-k}\mathrm{B}^{n-k}}\big\|\tau(k)_{\mathrm{A}^{n-k}}\otimes\tau(k)_{\mathrm{B}^{n-k}}\right) \\
= & \, I\left(\mathrm{A}_{k+1}\ldots \mathrm{A}_n:\mathrm{B}_{k+1}\ldots \mathrm{B}_n\right)_{\tau(k)}\\
= & \sum_{j=k+1}^n I\left(\mathrm{A}_j:\mathrm{B}_{k+1}\ldots \mathrm{B}_n|\mathrm{A}_{k+1}\ldots \mathrm{A}_{j-1}\right)_{\tau(k)}\\
\geq & \sum_{j=k+1}^n I\left(\mathrm{A}_j:\mathrm{B}_j|\mathrm{A}_{k+1}\ldots \mathrm{A}_{j-1}\right)_{\tau(k)} \\
\geq & \sum_{j=k+1}^n 2\,E_{sq}\left(\tau(k)_{\mathrm{A}_j\mathrm{B}_j}\right).
\end{align*}
The first inequality is due to the fact that, given a bipartite state $\tau_{\mathrm{U}\mathrm{V}}$ on $\mathrm{U}\otimes\mathrm{V}$, for any states $\rho_{\mathrm{U}}$, $\rho_{\mathrm{V}}$ on $\mathrm{U}$, $\mathrm{V}$ respectively, $D(\tau_{\mathrm{U}\mathrm{V}}\|\rho_{\mathrm{U}}\otimes\rho_{\mathrm{V}})\geq D(\tau_{\mathrm{U}\mathrm{V}}\|\tau_{\mathrm{U}}\otimes\tau_{\mathrm{V}})$.
The third equality and the fourth inequality follow from the chain rule and the monotonicity under discarding of subsystems, respectively, for the quantum mutual information.
And the last inequality is by definition of the squashed entanglement.
\end{proof}
\begin{remark}
\label{remark:E_I}
Observe that under the assumptions of Lemma \ref{lemma:E_sq},
we actually have the stronger conclusion
\[
\sum_{j=k+1}^n E_{I}\left(\tau(k)_{\mathrm{A}_j\mathrm{B}_j}\right) \leq \frac{1}{2}\log \frac{1}{p_k},
\]
where $E_I$ is the \emph{conditional entanglement of mutual information (CEMI)}
introduced in \cite{HWY}:
\[
E_I\left(\rho_{\mathrm{A}\mathrm{B}}\right) = \inf \left\{ \frac{1}{2}\big[ I(\mathrm{A}\mathrm{A}':\mathrm{B}\mathrm{B}')_{\rho}
- I(\mathrm{A}':\mathrm{B}')_{\rho}\big]
\ : \ \Tr_{\mathrm{A}'\mathrm{B}'}\left(\rho_{\mathrm{A}\mathrm{A}'\mathrm{B}\mathrm{B}'}\right) =\rho_{\mathrm{A}\mathrm{B}} \right\}.
\]
CEMI is always at least as large as squashed entanglement: for any state $\rho_{\mathrm{A}\mathrm{B}}$, $E_I(\rho_{\mathrm{A}\mathrm{B}}) \geq E_{sq}(\rho_{\mathrm{A}\mathrm{B}})$. But the precise relation between these two entanglement measures is unknown.
In~\cite{HWY} it was furthermore shown that, like squashed entanglement, CEMI is additive,
and more generally super-additive in the sense that
\[
E_I(\rho_{\mathrm{A}_1\mathrm{A}_2{:}\mathrm{B}_1\mathrm{B}_2}) \geq E_I(\rho_{\mathrm{A}_1{:}\mathrm{B}_1}) + E_I(\rho_{\mathrm{A}_2{:}\mathrm{B}_2}).
\]
However, unlike squashed entanglement, there is no simple proof of
monogamy of CEMI, and it may well not hold in general.
\end{remark}
\begin{lemma}[Cf.~\cite{Holenstein}, Lemma 8.6]
\label{lemma:recursivity}
Let $0<\nu<1$ and $c>0$. Let also $n\in\mathbf{N}$ and assume that $(p_k)_{1\leq k\leq n}$ is a sequence of numbers satisfying $1> p_1\geq\cdots\geq p_n >0$ and
\[ \forall\ 1\leq k\leq n-1,\ p_{k+1} \leq p_k \left( \sqrt{\frac{c}{n-k}\log\frac{1}{p_k}} + \nu \right). \]
Then, for any $0<\gamma<1-\nu$ such that $p_1\leq \nu+\gamma$, we have
\[ \forall\ 1\leq k\leq n,\ p_k\leq (\nu+\gamma)^{\min(k,k_0)},\ \text{where}\ k_0=\frac{\gamma^2}{c\log[1/(\nu+\gamma)]+\gamma^2}(n+1). \]
\end{lemma}
\begin{proof}
To prove Lemma \ref{lemma:recursivity}, we only have to show that
\begin{equation} \label{eq:rec} \forall\ 1\leq k\leq k_0,\ p_k\leq (\nu+\gamma)^k. \end{equation}
Indeed, the case $k>k_0$ then directly follows from the assumption that the sequence $(p_k)_{1\leq k\leq n}$ is non-increasing, so that $p_k\leq p_{k_0} \leq (\nu+\gamma)^{k_0}$.
Let us establish \eqref{eq:rec} by recursivity. The statement obviously holds for $k=1$ since $p_1\leq \nu+\gamma$ by hypothesis. So assume next that it holds for some $k\leq k_0-1$. If $p_k\leq(\nu+\gamma)^{k+1}$, then clearly $p_{k+1}\leq p_k\leq(\nu+\gamma)^{k+1}$. Otherwise, by the way $p_{k+1}$ is related to $p_k$, we then have $p_{k+1}\leq (\nu+\gamma)^k\big(\sqrt{c(k+1)\log[1/(\nu+\gamma)]/(n-k)}+\nu\big)$, and the latter quantity is smaller than $(\nu+\gamma)^{k+1}$ if $(k+1)/(n-k)\leq \gamma^2/\left(c\log[1/(\nu+\gamma)]\right)$, which can be checked to be equivalent to $k+1\leq k_0$. Hence in both cases, the statement holds for $k+1$.
\end{proof}
\begin{corollary}[Cf. \cite{Holenstein}, Lemma 8.6]
\label{cor:recursivity}
Let $(p_k)_{1\leq k\leq n}$ be a sequence of numbers satisfying the assumptions of Lemma \ref{lemma:recursivity}, and the additional condition $p_1\leq 1-(1-\nu)/2$. Then,
\[ p_n\leq \left(1-\frac{(1-\nu)^2}{8 c}\right)^n. \]
\end{corollary}
\begin{proof}
Corollary \ref{cor:recursivity} follows from applying Lemma \ref{lemma:recursivity} in the particular case $\gamma=(1-\nu)/2$. Indeed, we then have $\nu+\gamma=1-(1-\nu)/2=(1+\nu)/2$, so that
\[ k_0= \frac{(1-\nu)^2}{4c\log[2/(1+\nu)]+(1-\nu)^2}(n+1) \geq \frac{(1-\nu)^2}{4c\log[2^{1-\nu}]+1}(n+1) \geq \frac{1-\nu}{4c}n, \]
And consequently,
\[ p_n \leq \left(1-\frac{1-\nu}{2}\right)^{n(1-\nu)/(4c)} \leq \left(1-\frac{(1-\nu)^2}{8 c}\right)^n, \]
which is exactly the announced upper bound for $p_n$.
\end{proof}
\subsection{Weak multiplicativity of $h_{sep}$}
Our approach in this section, to prove the multiplicative behaviour of $h_{sep}$, is directly inspired from the seminal angle of attack to the parallel repetition problem for classical non-local games: our Theorem \ref{th:h_sep-h_qext} is an analogue of the exponential decay results by Raz \cite{Raz} and Holenstein \cite{Holenstein}, while our Theorem \ref{th:w_sep-h_qext} is an analogue of the concentration bound result by Rao \cite{Rao}. Indeed, here in the same spirit as theirs, we want to make precise the following intuition: if the initial state of a system on $(\mathrm{A}\otimes\mathrm{B})^{\otimes n}$ is product across the cut $\mathrm{A}^n{:}\mathrm{B}^n$, then performing a measurement $(M,\openone-M)$ on a few subsystems $\mathrm{A}\otimes\mathrm{B}$ only should not create too much correlations in the post-measurement state on the remaining subsystems.
Before we prove the main result of this section,
we need to recall one last definition: For any $q\in\mathbf{N}$,
a state $\rho_{\mathrm{A}\mathrm{B}}$ on a bipartite Hilbert space $\mathrm{A}\otimes \mathrm{B}$
is said to be $q$-extendible with respect to $\mathrm{B}$
if there exists a state $\rho_{\mathrm{A}\mathrm{B}^q}$ on $\mathrm{A}\otimes \mathrm{B}^{\otimes q}$
that is invariant under any permutation of the $\mathrm{B}$-subsystems and such
that $\rho_{\mathrm{A}\mathrm{B}}=\Tr_{\mathrm{B}^{q-1}}\rho_{\mathrm{A}\mathrm{B}^q}$.
We shall denote by $\mathcal{E}_q(\mathrm{A}{:}\mathrm{B})$ the set of
$q$-extendible states with respect to $\mathrm{B}$ on $\mathrm{A}\otimes \mathrm{B}$, and by $h_{q-ext}$ its associated support function.
\begin{theorem}
\label{th:h_sep-h_qext}
Let $M$ be an operator on the tensor product Hilbert space $\mathrm{A} \otimes \mathrm{B}$, satisfying $0 \leq M \leq \openone$. Then, for any $q\in\mathbf{N}$,
\begin{equation}
\label{eq:qext-bound}
h_{sep}\left(M^{\otimes n}\right) \leq \left( 1-\frac{\left(1-h_{q-ext}(M)\right)^2}{8\ln 2\,q^2}\right)^n . \end{equation}
And consequently, if $h_{sep}(M)\leq 1-\delta$ for some $0<\delta<1$, then
\begin{equation}
\label{eq:sep-bound}
h_{sep}\left(M^{\otimes n}\right) \leq \left(1 -\frac{\delta^4}{512\ln 2\,d^4}\right)^n,
\end{equation}
assuming $|\mathrm{A}|=|\mathrm{B}|=d$.
\end{theorem}
\begin{proof}
To establish the first statement \eqref{eq:qext-bound}, we have to show that,
\[
\forall\ \rho_{\mathrm{A}^n\mathrm{B}^n}\in\mathcal{S}(\mathrm{A}^{n}{:}\mathrm{B}^{n}),\ \Tr\left(M_{\mathrm{A}\mathrm{B}}^{\otimes n}\rho_{\mathrm{A}^n\mathrm{B}^n} \right) \leq \left( 1-\frac{\left(1-h_{q-ext}(M_{\mathrm{A}\mathrm{B}})\right)^2}{8\ln 2\,q^2}\right)^n .
\]
Note that, with this aim in view, we can without loss of generality focuss only on states which are extremal in $\mathcal{S}(\mathrm{A}^{n}{:}\mathrm{B}^{n})$, namely on states which are product across the cut $\mathrm{A}^{\otimes n}{:}\mathrm{B}^{\otimes n}$. So let $\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}$ be such a state, and set $p_0=1$, $\tau(0)_{\mathrm{A}^n\mathrm{B}^n}=\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}$. In the sequel, we will use the following notation: given $I_k\subset[n]$ with $|I_k|=k$, define $M^{(I_k)}_{\mathrm{A}^n\mathrm{B}^n}$ as
\[ M^{(I_k)}_{\mathrm{A}^n\mathrm{B}^n} = M_{\mathrm{A}\mathrm{B}}^{\otimes I_k}\otimes\openone_{\mathrm{A}\mathrm{B}}^{\otimes I_k^c}. \]
Then, for each $1\leq k\leq n$, construct recursively
\begin{align*}
& p_{k}= \Tr_{\mathrm{A}^n\mathrm{B}^n} \left[M^{(I_k)}_{\mathrm{A}^n\mathrm{B}^n}\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}\right], \\
& \tau(k)_{\mathrm{A}_{I_k^c}\mathrm{B}_{I_k^c}}= \frac{1}{p_k} \Tr_{\mathrm{A}_{I_k}\mathrm{B}_{I_k}} \left[M^{(I_k)}_{\mathrm{A}^n\mathrm{B}^n}\alpha_{A^n}\otimes\beta_{B^n}\right],
\end{align*}
with $i_{k}$ chosen in $I_{k-1}^c$ such that
\[ E_{sq}\left(\tau(k-1)_{\mathrm{A}_{i_k}\mathrm{B}_{i_k}}\right)\leq \frac{1}{n-k+1}\,\frac{1}{2}\log \frac{1}{p_{k-1}}. \]
We know that this is possible. Indeed, assuming that $p_{k-1}$, $\tau(k-1)_{\mathrm{A}_{I_{k-1}^c}\mathrm{B}_{I_{k-1}^c}}$ have been constructed, Lemma \ref{lemma:E_sq} guarantees that
\[ \frac{1}{n-k+1}\sum_{j=1}^{n-k+1}E_{sq}\left(\tau(k-1)_{\mathrm{A}_{I_{k-1}^c}\mathrm{B}_{I_{k-1}^c}}\right)\leq \frac{1}{n-k+1}\,\frac{1}{2}\log \frac{1}{p_{k-1}}, \]
so that there necessarily exists an index $i\in I_{k-1}^c$ such that $E_{sq}\big(\tau(k-1)_{\mathrm{A}_{i}\mathrm{B}_{i}}\big)$ is smaller than the quantity on the right-hand-side of the average upper bound above.
Now, notice that the $p_k$, $0\leq k\leq n$, are related by the recursion formula
\[ \forall\ 0\leq k\leq n-1,\ p_{k+1} = p_k\Tr_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\left(M_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}} \tau(k)_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\right), \]
where, by the way the $\tau(k)$, $0\leq k\leq n$, are built
\[ E_{sq}\left(\tau(k)_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\right) \leq \frac{1}{n-k}\frac{1}{2}\log \frac{1}{p_k}. \]
Yet, we know from \cite{LiW} that this implies that
\begin{equation} \label{eq:E_sq-q-ext} \exists\ \sigma_{\mathrm{A}\mathrm{B}}\in\mathcal{E}_q(\mathrm{A}{:}\mathrm{B}):\ \left\| \tau(k)_{\mathrm{A}\mathrm{B}} - \sigma_{\mathrm{A}\mathrm{B}} \right\|_1 \leq \sqrt{2\ln 2}(q-1)\sqrt{\frac{1}{n-k}\frac{1}{2}\log \frac{1}{p_k}} \leq \sqrt{\frac{\ln 2\, q^2}{n-k}\log \frac{1}{p_k}}. \end{equation}
And therefore,
\[ p_{k+1} \leq p_k \left(\left\|M_{\mathrm{A}\mathrm{B}}\right\|_{\infty}\left\| \tau(k)_{\mathrm{A}\mathrm{B}} - \sigma_{\mathrm{A}\mathrm{B}} \right\|_1 + \Tr\left(M_{\mathrm{A}\mathrm{B}}\sigma_{\mathrm{A}\mathrm{B}}\right) \right) \leq p_k \left(\sqrt{\frac{\ln 2\, q^2}{n-k}\log \frac{1}{p_k}} + h_{q-ext}\left(M_{\mathrm{A}\mathrm{B}}\right) \right). \]
With this upper bound, and because we also clearly have $1> p_1\geq\cdots\geq p_n> 0$ as well as the requirement $p_1\leq h_{sep}(M)\leq h_{q-ext}(M)\leq 1-(1-h_{q-ext})/2$, it follows from Corollary \ref{cor:recursivity} that
\[ \Tr\left(M_{\mathrm{A}\mathrm{B}}^{\otimes n}\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}\right)=p_n \leq \left( 1-\frac{\left(1-h_{q-ext}(M_{\mathrm{A}\mathrm{B}})\right)^2}{8\ln 2\,q^2}\right)^n, \]
which is precisely what we wanted to prove.
From there, the second statement \eqref{eq:sep-bound} easily follows.
Indeed, in the case where $|\mathrm{A}|=|\mathrm{B}|=d$, we know from \cite{CKMR} that,
for any $q\in\mathbf{N}$, $\rho_{\mathrm{A}\mathrm{B}}\in\mathcal{E}_q(\mathrm{A}{:}\mathrm{B})$ implies that
there exists $\sigma_{\mathrm{A}\mathrm{B}}\in\mathcal{S}(\mathrm{A}{:}\mathrm{B}):\ \|\rho_{\mathrm{A}\mathrm{B}}-\sigma_{\mathrm{A}\mathrm{B}}\|_1\leq 2d^2/q$,
so that $h_{q-ext}\leq h_{sep}+2d^2/q$. Hence, if $h_{sep}(M_{\mathrm{A}\mathrm{B}})\leq 1-\delta$,
making the choice $q=4d^2/\delta$, in order to have $h_{q-ext}(M_{\mathrm{A}\mathrm{B}})\leq 1-\delta/2$,
yields, after a straightforward computation, exactly the announced exponential decay result.
\end{proof}
The scaling as $(\delta/d)^4$ in the upper bound provided by equation \eqref{eq:sep-bound} of Theorem \ref{th:h_sep-h_qext} is much worse than the scaling as $(\delta/d)^2$ in the upper bound provided by Theorem \ref{th:h_sep-n}. However, equation \eqref{eq:qext-bound} of Theorem \ref{th:h_sep-h_qext}, which relates $h_{sep}(M^{\otimes n})$ to $h_{q-ext}(M)$, may be of interest in some specific cases, namely when $M$ has a maximum overlap with $q$-extendible states which is already of the same order as its maximum overlap with separable states for $q \ll d^2$.
\begin{remark}
By Remark \ref{remark:E_I}, we see that we could also have done the recursive construction described in the proof of Theorem \ref{th:h_sep-h_qext} by imposing instead that, for each $0\leq k\leq n-1$,
\[ p_{k+1} = p_k\Tr_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\left(M_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}} \tau(k)_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\right),\ \text{with}\ E_I\left(\tau(k)_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\right) \leq \frac{1}{n-k}\frac{1}{2}\log \frac{1}{p_k}. \]
Now, it is an open question to determine whether there exists a dimension independent constant $C>0$ such that
\begin{equation} \label{eq:E_I-conjecture} E_I\left(\rho_{\mathrm{A}\mathrm{B}}\right)\leq \epsilon\ \Rightarrow\ \exists\ \sigma_{\mathrm{A}\mathrm{B}}\in\mathcal{S}(\mathrm{A}{:}\mathrm{B}) \ : \ \left\| \rho_{\mathrm{A}\mathrm{B}} - \sigma_{\mathrm{A}\mathrm{B}} \right\|_1 \leq C\sqrt{\epsilon}. \end{equation}
If Conjecture \eqref{eq:E_I-conjecture} indeed held, this would imply that the $(p_k)_{1\leq k\leq n}$ satisfy
\[
\forall\ 0\leq k\leq n-1,\
p_{k+1} \leq p_k \left(\sqrt{\frac{C^2/2}{n-k}\log \frac{1}{p_k}} + h_{sep}\left(M_{\mathrm{A}\mathrm{B}}\right) \right).
\]
And hence eventually, the following dimension-free exponential decay result for $h_{sep}$:
\[
h_{sep}(M)\leq 1-\delta\ \Rightarrow\ h_{sep}\left(M^{\otimes n}\right) \leq \left( 1-\frac{\delta^2}{4C^2}\right)^n . \]
And in fact, if a more general variant of Conjecture \eqref{eq:E_I-conjecture} held, with $C\sqrt{\epsilon}$ replaced by $\varphi(\epsilon)$ for $\varphi$ a (universal) non-decreasing function such that $\varphi(0)=0$, then one could prove analogously that
\[ h_{sep}(M)\leq 1-\delta\ \Rightarrow\ h_{sep}\left(M^{\otimes n}\right) \leq \left(1-\frac{\varphi^{-1}(\delta)}{4}\right)^n. \]
The way property \eqref{eq:E_sq-q-ext} of strong faithfulness of squashed entanglement with respect to $q$-extendible states, is proved in \cite{LiW} is relying on the breakthrough result by Fawzi and Renner \cite{FR} that small conditional mutual information does imply approximate recoverability. Now, in an even stronger manner than $E_{sq}(\rho)$ being small means that the conditional mutual information of any extension of $\rho$ is small, $E_I(\rho)$ being small is a condition that is expressible as a bunch of conditional mutual information of extensions of $\rho$ being simultaneously small. So it could be that recoverability results (in particular the best one up-to-date \cite{JRSWW}, which carries the advantage over the original one \cite{FR} of being universal and explicit) would help in an attempt to prove a strong faithfulness property of CEMI with respect to separable states such as \eqref{eq:E_I-conjecture}.
\end{remark}
\begin{theorem}
\label{th:w_sep-h_qext}
Let $M$ be an operator on the tensor product Hilbert space $\mathrm{A}\otimes \mathrm{B}$, satisfying $0 \leq M \leq \openone$. If $h_{sep}(M)\leq 1-\delta$ for some $0<\delta<1$, then for any $n,t\in\mathbf{N}$ with $t\geq(1-\delta+\alpha)n$ for some $0<\alpha\leq\delta$, we have
\[ h_{sep}\left(M^{(t/n)}\right)\leq\left(1-\frac{\alpha^5}{2048\ln 2\, d^4\,(2\delta-\alpha)}\right)^n, \]
assuming $|A|=|B|=d$.
\end{theorem}
\begin{proof}
The proof of this theorem follows a very similar route to that of
Theorem \ref{th:h_sep-h_qext}: For any given state $\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}$ which is product across the cut $\mathrm{A}^{\otimes n}{:}\mathrm{B}^{\otimes n}$, we want to show that the probability that it passes at least $t$ amongst $n$ tests defined by $M_{\mathrm{A}\mathrm{B}}$ is upper bounded as
\[
P_t(\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}) \leq \left(1-\frac{\alpha^5}{2048\ln 2\, d^4\,(2\delta-\alpha)}\right)^n.
\]
In that aim, we start by defining the following deterministic set, number and state: $I_0=\emptyset$, $p_{I_0}=1$ and $\tau(I_0)_{\mathrm{A}^n\mathrm{B}^n}=\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}$. Then, for each $1\leq k\leq n$, we construct recursively the following random set, number and state: pick $i_k$ uniformly at random in $I_{k-1}^c$, and define
\begin{align*}
& I_k=I_{k-1}\cup\{i_k\}, \\
& p_{I_k}= \Tr_{\mathrm{A}^n\mathrm{B}^n} \left[M^{(I_k)}_{\mathrm{A}^n\mathrm{B}^n} \alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}\right],\\
& \tau(I_k)_{\mathrm{A}_{I_k^c}\mathrm{B}_{I_k^c}}= \frac{1}{p_{I_k}}\Tr_{\mathrm{A}_{I_k}\mathrm{B}_{I_k}} \left[M^{(I_k)}_{\mathrm{A}^n\mathrm{B}^n} \alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}\right].
\end{align*}
Lemma \ref{lemma:E_sq} guarantees that, on average, for each $0\leq k\leq n-1$,
\[
E_{sq}\left(\overline{\tau}(I_k)_{\mathrm{A}_{i_{k+1}}\mathrm{B}_{i_{k+1}}}\right)
\leq \frac{1}{n-k}\frac{1}{2}\log \frac{1}{\overline{p}_{I_k}},
\]
so that, on average, for any $q\in\mathbf{N}$,
\[ \overline{p}_{I_{k+1}} \leq \overline{p}_{I_k} \left(\sqrt{\frac{\ln 2\, q^2}{n-k}\log \frac{1}{\overline{p}_{I_k}}} + h_{q-ext}\left(M_{\mathrm{A}\mathrm{B}}\right) \right). \]
In particular, we can make the choice $q=8d^2/\alpha$, in order to have $h_{q-ext}(M_{\mathrm{A}\mathrm{B}})\leq 1-\delta+\alpha/4$. And we thus get from Lemma \ref{lemma:recursivity}, after computation, that on average,
\[ \overline{p}_{I_{k_0}} \leq \left(1-\delta+\frac{\alpha}{2}\right)^{k_0},\ \text{where}\ k_0= \frac{\alpha^4}{1024\ln 2\,d^4\,\log[1/(1-\delta+\alpha/2)]+\alpha^4}(n+1)\geq \frac{\alpha^4}{1024\ln 2\,d^4\,(2\delta-\alpha)}\,n. \]
To finish off the proof, we just have to observe (Cf. \cite{Rao}, Section 8) that
\[ P_t(\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}) \leq \sum_{I_{k_0}\subset[n],\,|I_{k_0}|=k_0}\frac{1}{{(1-\delta+\alpha)n \choose k_0}}p_{I_{k_0}} \leq \frac{{n \choose k_0}}{{(1-\delta+\alpha)n \choose k_0}} \overline{p}_{I_{k_0}}\leq \left(\frac{n-k_0+1}{(1-\delta+\alpha)n-k_0+1}\right)^{k_0}\left(1-\delta+\frac{\alpha}{2}\right)^{k_0}, \]
where the last inequality follows from the fact that $\prod_{i=0}^{l-1}(a+i)/(b+i)\leq (a/b)^l$, combined with the upper bound on $\overline{p}_{I_{k_0}}$. In the end, we can therefore conclude that
\[ P_t(\alpha_{\mathrm{A}^n}\otimes\beta_{\mathrm{B}^n}) \leq \left(1-\frac{\alpha}{2}\right)^{k_0}\leq \left(1-\frac{\alpha^5}{2048\ln 2\, d^4\,(2\delta-\alpha)}\right)^n, \]
where the first inequality follows from the fact that $(1-\delta+\alpha/2)/(1-\delta+\alpha)\leq 1-\alpha/2$, while the second inequality is a consequence of the lower bound on $k_0$.
\end{proof}
\begin{remark}
Here again, we see by Remark \ref{remark:E_I} that, if
Conjecture \eqref{eq:E_I-conjecture} held, then we could have
obtained in the proof of Theorem \ref{th:w_sep-h_qext} that, on average
\[
\overline{p}_{I_{k_0}} \leq \left(1-\delta+\frac{\alpha}{2}\right)^{k_0},\
\text{where}\
k_0= \frac{\alpha^2}{8\log[1/(1-\delta+\alpha/2)]+\alpha^2}(n+1)\geq \frac{\alpha^2}{8(2\delta-\alpha)}\,n.
\]
And hence eventually, the following dimension-free concentration result for $h_{sep}$:
\[
h_{sep}(M)\leq 1-\delta\ \Rightarrow\ \forall\ 0<\alpha<\delta,\ \forall\ t\geq(1-\delta+\alpha)n,\
h_{sep}\left(M^{(t/n)}\right) \leq \left( 1-\frac{\alpha^3}{16C^2(2\delta-\alpha)}\right)^n .
\]
\end{remark}
\section{Equivalence between weak multiplicativity of support functions and of maximum fidelities}
\label{sec:general}
In the previous Sections \ref{sec:sep1} and \ref{sec:sep2}, we studied in great depth one particular example of convex constraint on quantum states, namely the separability one. We showed in this specific case that there is a strong connection between the (weakly) multiplicative behaviour under tensoring of either the support function $h_{sep}$ or the maximum fidelity $F(\cdot,\mathcal{S})$. We would now like to describe, more generally, which kind of convex sets of states exhibit a similar feature.
So let us fix $d\in\mathbf{N}$, $\mathcal{H}$ a $d$-dimensional Hilbert space, and assume that we have a sequence of convex sets
of states $\mathcal{K}^{(n)}$ on $\mathcal{H}^{\otimes n}$, $n\in\mathbf{N}$, with
the following stability properties (under permutation and partial trace):
\begin{equation}
\label{eq:sym-stability}
\rho\in\mathcal{K}^{(n)}\ \Rightarrow\ \forall\ \pi\in\mathcal{S}_n,\ U_{\pi}\rho U_{\pi}^{\dagger}\in\mathcal{K}^{(n)}\ \text{and}\
\Tr_{\mathcal{H}}\rho\in\mathcal{K}^{(n-1)}.
\end{equation}
Note that requirement \eqref{eq:sym-stability} implies in particular that,
if $\rho^{\otimes n}\in\mathcal{K}^{(n)}$, then $\rho\in\mathcal{K}^{(1)}$. In view of our subsequent discussion, it would be meaningless not to impose that the opposite holds as well, i.e.~that, if $\rho\in\mathcal{K}^{(1)}$, then $\rho^{\otimes n}\in\mathcal{K}^{(n)}$. This means in other words that, for each $n\in\mathbf{N}$, $\mathcal{K}^{(n)}$ is assumed to contain the so-called $n^{\text{th}}$ projective tensor power of $\mathcal{K}^{(1)}$, which is defined as
\[ \left(\mathcal{K}^{(1)}\right)^{\hat{\otimes} n} := \mathrm{conv}\left\{\rho_1\otimes\cdots\otimes\rho_n,\ \rho_1,\ldots,\rho_n\in\mathcal{K}^{(1)} \right\}. \]
\subsection{Exponential decay and concentration of $h_{\mathcal{K}}$ from multiplicativity of $F(\cdot,\mathcal{K})$}
Given an operator $M$ on $\mathcal{H}$, satisfying $0\leq M\leq\openone$, define the support function of $\mathcal{K}^{(n)}$ at $M^{\otimes n}$ as
\[ h_{\mathcal{K}^{(n)}}\left(M^{\otimes n}\right) = \underset{\sigma\in\mathcal{K}^{(n)}}{\sup}\Tr\left(M^{\otimes n}\sigma\right). \]
Define also more generally, for any $0\leq t\leq n$, $h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)$ as the maximum probability for a state in $\mathcal{K}^{(n)}$ to pass a fraction $t/n$ of $n$ binary tests $(M,\openone-M)$ performed in parallel. The question we are next interested in is to understand how $h_{\mathcal{K}^{(n)}}(M^{\otimes n})$ and $h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)$ relate to $h_{\mathcal{K}^{(1)}}(M)$.
Hence, assume also that these sets $\mathcal{K}^{(n)}$ satisfy the following condition: there exists a non-decreasing function $f:\epsilon\in]0,1[\mapsto f(\epsilon)\in]0,1[$ such that, for any state $\rho$ on $\mathbf{C}^d$ and any $0<\epsilon<1$,
\begin{equation} \label{eq:exp-decay-F} \left\|\rho-\mathcal{K}^{(1)}\right\|_2\geq\epsilon\ \Rightarrow\ F\left(\rho^{\otimes n},\mathcal{K}^{(n)} \right)^2\leq\left(1-f(\epsilon)\right)^n. \end{equation}
Then, under assumption \eqref{eq:exp-decay-F} for the sets $\mathcal{K}^{(n)}$, the following holds: for any operator $M$ on $\mathcal{H}$, satisfying $0\leq M\leq\openone$, and with $\|M\|_2=r$ for some $0\leq r\leq \sqrt{d}$, and any $0<\alpha\leq\delta<1$,
\[ h_{\mathcal{K}^{(1)}}(M)\leq 1-\delta\ \Rightarrow\ h_{\mathcal{K}^{(n)}}(M^{\otimes n})\leq \left(1-g(\delta,r)\right)^n\ \text{and}\ \forall\ t\geq (1-\delta+\alpha)n,\ h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)\leq e^{-ng'(\alpha,r)}, \]
where $g(\delta,r)=f(\epsilon(\delta,r))$ for $0<\epsilon(\delta,r)<1$ the solution of the equation $f(\epsilon)=\delta-r\epsilon$ and $g'(\alpha,r)=f(\epsilon(\alpha,r))$ for $0<\epsilon(\alpha,r)<1$ the solution of the equation $f(\epsilon)=2(\alpha-r\epsilon)^2$.
To come to these statements, the strategy is entirely analogous to the one adopted in the proofs of Theorems \ref{th:h_sep-n} and \ref{th:h_sep-t/n}. It is therefore only sketched below. First of all, when looking for a state $\rho\in\mathcal{K}^{(n)}$ maximizing $\Tr\left(M^{\otimes n}\rho\right)$, one can in fact assume without loss of generality that $\rho$ is $n$-symmetric. And for such state $\rho$, reasoning as in the proof of Theorem \ref{th:h_sep-n}, we know that we have
\[ \Tr\left(M^{\otimes n}\rho\right) \leq (n+1)^{3d^2} \int_{\sigma\in\mathcal{D}(\mathcal{H})} F\left(\rho,\sigma^{\otimes n}\right)^2 \Tr\left(M\sigma\right)^n \mathrm{d}\mu(\sigma). \]
Hence, we get as a consequence of hypothesis \eqref{eq:exp-decay-F} that, for any $0<\epsilon<1$,
\[ \Tr\left(M^{\otimes n}\rho\right) \leq (n+1)^{3d^2} \left( \left(1-\delta+r\epsilon\right)^n + \left(1-f(\epsilon)\right)^n\right). \]
So choosing $\epsilon$ such that $f(\epsilon)=\delta-r\epsilon$ and $\rho$ such that $\Tr\left(M^{\otimes n}\rho\right)= h_{\mathcal{K}^{(n)}}(M^{\otimes n})$ yields in particular
\[ h_{\mathcal{K}^{(n)}}(M^{\otimes n}) \leq 2(n+1)^{3d^2}\left(1-g(\delta,r)\right)^n. \]
Similarly, it follows from hypothesis \eqref{eq:exp-decay-F} as well that, for any $0<\epsilon<1$,
\[ h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)\leq (n+1)^{3d^2}\left( \exp\left[-2n(\alpha-r\epsilon)^2\right] + \exp\left[-nf(\epsilon)\right] \right), \]
so that choosing $\epsilon$ such that $f(\epsilon)=2(\alpha-r\epsilon)^2$ gives
\[ h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)\leq 2(n+1)^{3d^2} e^{-ng'(\alpha,r)}. \]
In both cases the polynomial pre-factor $2(n+1)^{3d^2}$ can then be removed by the exact same argument as in the proof of Theorem \ref{th:h_sep-n}.
\subsection{Weak multiplicativity of $F(\cdot,\mathcal{K})$ from exponential decay and concentration of $h_{\mathcal{K}}$}
We would now like to go in the other direction. Namely, let us assume this time that these sets $\mathcal{K}^{(n)}$ satisfy the following condition: there exists a function $f:(\alpha,d)\in]0,1[\times\mathbf{N}\mapsto f(\alpha,d)\in]0,1[$, non-decreasing in $\alpha$ and non-increasing in $d$, such that, for any operator $M$ on $\mathcal{H}$, satisfying $0\leq M\leq\openone$, and any $0<\alpha\leq\delta<1$,
\begin{equation} \label{eq:exp-decay-h} h_{\mathcal{K}^{(1)}}(M)\leq 1-\delta\ \Rightarrow\ \forall\ t\geq (1-\delta+\alpha)n,\ h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)\leq e^{-nf(\alpha,d)}. \end{equation}
Then, under assumption \eqref{eq:exp-decay-h} for the sets $\mathcal{K}^{(n)}$, the following holds: for any state $\rho$ on $\mathcal{H}$ and any $0<\epsilon<1$,
\[ \frac{1}{2}\left\|\rho-\mathcal{K}^{(1)}\right\|_1 \geq \epsilon\ \Rightarrow\ F\left(\rho^{\otimes n},\mathcal{K}^{(n)}\right) \leq 2e^{-ng(\epsilon,d)}, \]
where $g(\epsilon,d)=f(\alpha(\epsilon,d),d)/2$ for $0<\alpha(\epsilon,d)<1$ the solution of the equation $f(\alpha,d)/2=(\alpha-\epsilon)^2$.
Here is the strategy to derive such result: Imagine you are given a state on $\mathcal{H}^{\otimes n}$, which you know is either $\rho^{\otimes n}$ or in $\mathcal{K}^{(n)}$, and you want to decide between these two hypotheses. For that, you can design a binary test $(T_+,T_-)$ such that outcome $+$ is obtained with a high probability $p$ if the state was $\rho^{\otimes n}$ and outcome $-$ is obtained with a high probability $q$ if the state was in $\mathcal{K}^{(n)}$. Then, clearly
\[ F\left(\rho^{\otimes n},\mathcal{K}^{(n)}\right) \leq F\left((p,1-p),(q,1-q)\right) \leq \sqrt{1-p}+\sqrt{1-q}. \]
Therefore, if both error probabilities $1-p$ and $1-q$ are exponentially small, the conclusion follows.
In the present case, the fact that $\left\|\rho-\mathcal{K}^{(1)}\right\|_1=2\epsilon$, implies that there exist $0\leq M\leq \openone$ and $\epsilon<\eta<1$ such that $\Tr(M\rho)=1-\eta+\epsilon$ whereas $h_{\mathcal{K}^{(1)}}(M)= 1-\eta$. So consider the binary POVM $(M_0,M_1)=(M,\openone-M)$ performed $n$ times in parallel, and the corresponding binary test $(T_+,T_-)$ with $+$ being the event ``outcome $0$ is obtained more than $(1-\eta+\alpha)n$ times'' and $-$ being the event ``outcome $0$ is obtained less than $(1-\eta+\alpha)n$ times'', for some $0<\alpha<\epsilon$ to be chosen later. Define next, for each $1\leq i\leq n$, the random variable $X_i$, respectively $Y_i$, as the outcome of measurement number $i$ given that the state was $\rho^{\otimes n}$, respectively in $\mathcal{K}^{(n)}$. Then,
\[ 1-p = \P\left(-\big|\rho^{\otimes n}\right) = \P\left(\sum_{i=1}^n X_i < (1-\eta+\alpha)n\right)\ \text{and}\ 1-q = \P\left(+\big|\mathcal{K}^{(n)}\right) = \P\left(\sum_{i=1}^n Y_i > (1-\eta+\alpha)n\right). \]
Yet on the one hand, $X_1,\ldots,X_n$ are independent Bernoulli random variables with expectation $1-\eta+\epsilon$, so by Hoeffding's inequality
\[ \P\left(\sum_{i=1}^n X_i < (1-\eta+\alpha)n\right) \leq e^{-2n(\epsilon-\alpha)^2}. \]
While on the other hand, for any $0\leq t\leq n$, $\P\left(\sum_{i=1}^n Y_i > t\right) = h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)$, so assumption \eqref{eq:exp-decay-h} guarantees that
\[ \P\left(\sum_{i=1}^n Y_i > (1-\eta+\alpha)n\right) \leq e^{-nf(\alpha,d)}. \]
Hence, putting everything together, we eventually obtain that, for any $0<\alpha<\epsilon$,
\[ F\left(\rho^{\otimes n},\mathcal{K}^{(n)}\right) \leq e^{-n(\epsilon-\alpha)^2} + e^{-nf(\alpha,d)/2}, \]
which yields the wanted result after choosing $\alpha$ such that $f(\alpha,d)/2=(\epsilon-\alpha)^2$.
\begin{remark}
Note that requirement \eqref{eq:sym-stability} is clearly fulfilled by the sets $\mathcal{S}_{\mathrm{A}^n:\mathrm{B}^n}$ of biseparable states on $\left(\mathrm{A}\otimes\mathrm{B}\right)^{\otimes n}$. Furthermore, they satisfy requirements \eqref{eq:exp-decay-F} and \eqref{eq:exp-decay-h} as well, with $f(\epsilon)=\epsilon^2/4$ and $f(\alpha,d^2)=\alpha^2/5d^2$.
It may also be worth emphasizing that conditions \eqref{eq:exp-decay-F} and \eqref{eq:exp-decay-h} are just strengthened and quantitative versions of the following stability property for the sets $\mathcal{K}^{(n)}$: $\rho\notin\mathcal{K}^{(1)}\ \Rightarrow\ \rho^{\otimes n}\notin\mathcal{K}^{(n)}$, i.e.~ equivalently $\rho^{\otimes n}\in\mathcal{K}^{(n)}\ \Rightarrow\ \rho\in\mathcal{K}^{(1)}$.
\end{remark}
\subsection{One simple example}
Let us look at what the previous discussion becomes in the case of the simplest possible sequence $\{\mathcal{K}^{(n)},\ n\in\mathbf{N}\}$ satisfying requirement \eqref{eq:sym-stability}, namely when there exists a set of states $\mathcal{K}$ on $\mathcal{H}$ such that, for each $n\in\mathbf{N}$, $\mathcal{K}^{(n)}$ is exactly the $n^{\text{th}}$ projective tensor power of $\mathcal{K}$, i.e.~\[ \mathcal{K}^{(n)}= \mathcal{K}^{\hat{\otimes} n} := \mathrm{conv}\left\{\rho_1\otimes\cdots\otimes\rho_n,\ \rho_1,\ldots,\rho_n\in\mathcal{K} \right\}. \]
Then, assumption \eqref{eq:exp-decay-h} is clearly satisfied, in the following way: for any operator $M$ on $\mathcal{H}$, satisfying $0\leq M\leq \openone$, and any $0<\alpha\leq\delta<1$,
\[ h_{\mathcal{K}^{(1)}}(M)\leq 1-\delta\ \Rightarrow\ \forall\ t\geq (1-\delta+\alpha)n,\ h_{\mathcal{K}^{(n)}}\left(M^{(t/n)}\right)\leq e^{-n2\alpha^2}. \]
This is a consequence of Hoeffding's inequality, following an argument similar to the one detailed in the previous subsection. And by the result established in the latter, this implies that: for any state $\rho$ on $\mathcal{H}$ and any $0<\epsilon<1$,
\[ F\left(\rho,\mathcal{K}^{(1)}\right)\leq e^{-\epsilon}\ \Rightarrow\ F\left(\rho^{\otimes n},\mathcal{K}^{(n)}\right) \leq 2e^{-n\epsilon^2/8}. \]
This is because $F(\rho,\mathcal{K}^{(1)})\leq e^{-\epsilon}\ \Rightarrow\ \|\rho-\mathcal{K}^{(1)}\|_1/2 \geq 1-e^{-\epsilon}\geq \epsilon/2$.
In connection with the discussion developed in Sections \ref{sec:sep1} and \ref{sec:sep2}, we see that we are actually facing the following interesting open question: how differently do $\mathcal{S}(\mathrm{A}^n{:}\mathrm{B}^n)$ and $\mathcal{S}(\mathrm{A}{:}\mathrm{B})^{\hat{\otimes} n}$ behave, from the (more or less equivalent) points of view of support functions and maximum fidelity functions?
\section{De Finetti reductions for infinite-dimensional symmetric quantum systems}
\label{sec:infinite}
All quantum de Finetti theorems and reductions require a bound on the dimension of the involved
Hilbert spaces. So what can be said about symmetric states on $\mathcal{H}^{\otimes n}$ when $\mathcal{H}$ is an infinite-dimensional Hilbert space? What extra assumptions do we need on them in order to be able to reduce their study to that of states in some de Finetti form? The original de Finetti reduction of \cite{CKR} was especially designed to prove the security of QKD protocols against general attacks. Yet, showing security of continuous variable QKD is also a major issue. This was the motivation behind the infinite-dimensional de Finetti type theorem of \cite{CR}. Our ultimate goal here is the same, which we rather try to achieve via a de Finetti reduction under constraints.
\subsection{Infinite-dimensional post-selection lemma}
Let $\mathcal{H}$ be an infinite-dimensional Hilbert space, and let
$\bar{\mathcal{H}}\subset\mathcal{H}$ be a (finite) $d$-dimensional subspace of $\mathcal{H}$.
Denote by $\{\ket{j}\}_{j\in\mathbf{N}}$ an orthonormal basis of $\mathcal{H}$, chosen such
that $\{\ket{j}\}_{1\leq j\leq d}$ is an orthonormal basis of $\bar{\mathcal{H}}$.
Then, for any $n,k\in\mathbf{N}$, the $(n+k)$-symmetric subspace of
$\bar{\mathcal{H}}^{\otimes n}\otimes\mathcal{H}^{\otimes k}$ is defined as
\[ \Sym^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right) := \Span\left\{ \sum_{\pi\in\mathcal{S}_{n+k}} \ket{j_{\pi(1)}}\otimes\cdots\otimes\ket{j_{\pi(n+k)}} \ : \ j_1\leq\cdots\leq j_{n+k}\in\mathbf{N},\ \forall\ 1\leq q\leq n,\ j_q\leq d\right\}. \]
Note that, denoting by $\bar{\mathcal{H}}_{\perp}\subset\mathcal{H}$ the orthogonal complement of $\bar{\mathcal{H}}$, i.e.~$\mathcal{H}=\bar{\mathcal{H}}\oplus\bar{\mathcal{H}}_{\perp}$, we have
\[ \Sym^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right) \subset \mathrm{V}^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right) := \bigoplus_{\underset{|I|\geq n}{I\subset[n+k]}} \bar{\mathcal{H}}_{\perp}^{\otimes I^c}\otimes\Sym\left(\bar{\mathcal{H}}^{\otimes I}\right). \]
\begin{lemma}
\label{lemma:ps-infinite}
Let $\mathcal{H}$ be an infinite-dimensional Hilbert space, and let $\bar{\mathcal{H}}\subset\mathcal{H}$ be a (finite) $d$-dimensional subspace of $\mathcal{H}$. Let also $n,k\in\mathbf{N}$. Then, any unit vector $\ket{\theta}\in\Sym^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right)$ satisfies
\[ \proj{\theta} \leq \left[\sum_{q=0}^k{n+k \choose q} {n+d-1 \choose n}^3 \right] \sum_{\underset{|I|\geq n}{I\subset[n+k]}} \int_{\ket{x}\in S_{\bar{\mathcal{H}}}} \epsilon(\theta_x)_{\bar{\mathcal{H}}_{\perp}^{I^c}}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes I} \mathrm{d}x , \]
where for all $0\leq q\leq k$ and all unit vector $\ket{x}\in \bar{\mathcal{H}}$, $\epsilon(\theta_x)_{\bar{\mathcal{H}}_{\perp}^{k-q}}=\Tr_{\bar{\mathcal{H}}^{ n+q}}\big[\big(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes k-q}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes n+q}\big)\proj{\theta}\big]$ is a sub-normalized state on $\bar{\mathcal{H}}_{\perp}^{\otimes k-q}$.
\end{lemma}
\begin{proof}
Since $\Sym^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right)\subset \mathrm{V}^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right)$, any unit vector $\ket{\theta}\in\Sym^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right)$ satisfies
\begin{align*}
\proj{\theta} = & P_{\mathrm{V}^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right)} \proj{\theta} P_{\mathrm{V}^{n+k}\left(\bar{\mathcal{H}},\mathcal{H}\right)}^{\dagger}\\
= & {n+d-1 \choose n}^2 \sum_{\underset{|I|,|J|\geq n}{I,J\subset[n+k]}} \int_{\ket{x},\ket{y}\in S_{\bar{\mathcal{H}}}} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes I^c}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes I}\right) \proj{\theta} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes J^c}\otimes\proj{y}_{\bar{\mathcal{H}}}^{\otimes J}\right)^{\dagger} \mathrm{d}x\mathrm{d}y.
\end{align*}
Now, by Lemma \ref{lemma:pinching} (and using the same Caratheodory argument as in the proof of Proposition \ref{prop:ps-pure}), we have
\begin{align*}
& \sum_{\underset{|I|,|J|\geq n}{I,J\subset[n+k]}} \int_{\ket{x},\ket{y}\in S_{\bar{\mathcal{H}}}} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes I^c}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes I}\right) \proj{\theta} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes J^c}\otimes\proj{y}_{\bar{\mathcal{H}}}^{\otimes J}\right)^{\dagger} \mathrm{d}x\mathrm{d}y\\
& \ \ \ \leq \left[\sum_{q=0}^k{n+k \choose q} {n+d-1 \choose n}\right] \sum_{\underset{|I|\geq n}{I\subset[n+k]}} \int_{\ket{x}\in S_{\bar{\mathcal{H}}}} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes I^c}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes I}\right) \proj{\theta} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes I^c}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes I}\right)^{\dagger} \mathrm{d}x.\\
\end{align*}
We then just have to notice that, for any $0\leq q\leq k$ and any unit vector $\ket{x}\in \bar{\mathcal{H}}$,
\[ \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes k-q}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes n+q}\right) \proj{\theta} \left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes k-q}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes n+q}\right)^{\dagger} = \Tr_{\bar{\mathcal{H}}^{ n+q}}\left[\left(\openone_{\bar{\mathcal{H}}_{\perp}}^{\otimes k-q}\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes n+q}\right)\proj{\theta}\right]\otimes\proj{x}_{\bar{\mathcal{H}}}^{\otimes n+q}, \]
in order to actually get the advertised result.
\end{proof}
\subsection{An application}
One application of Lemma \ref{lemma:ps-infinite} is to the security analysis of quantum cryptographic schemes, when there is no a priori bound on the dimension of the information carriers. This problem was originally investigated in \cite{CR} via a de Finetti theorem specifically designed for it. It was shown there that, under experimentally verifiable conditions, it is possible to ensure the security of quantum key distribution (QKD) protocols with continuous variables against general attacks. We show here that similar conclusions can be reached using the de Finetti reduction of Lemma \ref{lemma:ps-infinite}.
We look at things from the exact same point of view as the one adopted in \cite{CR}. Let $\mathcal{H}$ be an infinite-dimensional Hilbert space and let $X,Y$ be two canonical operators on $\mathcal{H}$. Then, denote by $\Lambda=X^2+Y^2$ the corresponding Hamiltonian, fix $\lambda_0>0$, and define
\[ \bar{\mathcal{H}}:=\left\{ \ket{\theta}\in\mathcal{H} \ : \ \Lambda\ket{\theta}=\lambda\ket{\theta},\ \lambda\leq\lambda_0\right\}, \]
finite-dimensional subspace of $\mathcal{H}$ spanned by the eigenvectors of $\Lambda$ with associated eigenvalue at most $\lambda_0$.
Let $n,k\in\mathbf{N}$, with $n\geq 2k$, and let $\rho^{(n+2k)}$ be a $(n+2k)$-symmetric state on $\mathcal{H}^{\otimes n+2k}$. Next, define the two events $\mathcal{A}$ and $\mathcal{B}$ as
\begin{align*}
\mathcal{A}\ = & \ \text{``}\ \forall\ 1\leq q\leq k,\ \Tr\left(\Lambda\rho_q^{(1)}\right) \leq \lambda_0\ \text{''}\\
\mathcal{B}\ = & \ \text{``}\ \exists\ \ket{\theta^{(n+k)}}\in\Sym^{n+k}\left(\bar{\mathcal{H}}\otimes\bar{\mathcal{H}}',\mathcal{H}\otimes\mathcal{H}'\right):\ \rho^{(n+k)}=\Tr_{\mathcal{H}'^{n+k}} \proj{\theta^{(n+k)}}\ \text{''},
\end{align*}
where for all $1\leq q\leq k$, $\rho_q^{(1)}=\Tr_{\mathcal{H}^{n+2k}\setminus\mathcal{H}_q}\rho^{(n+2k)}$, and $\rho^{(n+k)}=\Tr_{\mathcal{H}_{k+1}\cdots\mathcal{H}_{2k}}\rho^{(n+2k)}$. We know from \cite[Lemma III.3]{CR} that there exist universal constants $C_0,c>0$ such that, whenever $\lambda_0\geq C_0\log (n/k)$, we have
\[ \P\left(\mathcal{A}\wedge\neg\mathcal{B}\right) \leq e^{-ck^3/n^2} . \]
In words, this means the following. Fix a threshold $\lambda_0\geq C_0\log (n/k)$, and assume that when measuring the energy $\Lambda$ on the $k$ first subsystems of $\rho^{(n+2k)}$, only values below $\lambda_0$ are obtained. Then, with probability greater than $1-e^{-ck^3/n^2}$, the remaining $n+k$ subsystems of $\rho^{(n+2k)}$ have a purification which is the symmetrization of a state with more than $n$ subsystems supported in $\bar{\mathcal{H}}\otimes\bar{\mathcal{H}}'$.
Now, let $\widetilde{\rho}$ be a state on $\mathcal{H}^{\otimes n+k}$ such that $\widetilde{\rho}=\Tr_{\mathcal{H}'^{n+k}} \ket{\widetilde{\theta}}\bra{\widetilde{\theta}}$ for some unit vector $\ket{\widetilde{\theta}} \in \Sym^{n+k}\left(\bar{\mathcal{H}}\otimes\bar{\mathcal{H}}',\mathcal{H}\otimes\mathcal{H}'\right)$.
Denoting by $d$ the dimension of $\bar{\mathcal{H}}$, we have by Lemma \ref{lemma:ps-infinite} that $\ket{\widetilde{\theta}}$ satisfies
\[ \proj{\widetilde{\theta}} \leq \left[\sum_{q=0}^k {n+k \choose q}{n+d^2-1 \choose n}^3\right] \sum_{\underset{|I|\geq n}{I\subset[n+k]}} \int_{\ket{x}\in S_{\bar{\mathcal{H}}\otimes\bar{\mathcal{H}}'}} \epsilon(\widetilde{\theta}_x)_{(\mathcal{H}\cH')^{ I^c}}\otimes\proj{x}_{\bar{\mathcal{H}}\bar{\mathcal{H}}'}^{\otimes I} \mathrm{d}x. \]
And hence, after partial tracing over $\mathcal{H}'^{\otimes n+k}$, we finally get
\[ \widetilde{\rho} \leq \left[\sum_{q=0}^k {n+k \choose q}{n+d^2-1 \choose n}^3\right] \sum_{\underset{|I|\geq n}{I\subset[n+k]}} \int_{\ket{x}\in S_{\bar{\mathcal{H}}\otimes\bar{\mathcal{H}}'}} \varepsilon(\widetilde{\theta}_x)_{\mathcal{H}^{I^c}}\otimes\sigma(x)_{\bar{\mathcal{H}}}^{\otimes I} \mathrm{d}x, \]
where for all $0\leq q\leq k$ and all unit vector $\ket{x}\in \bar{\mathcal{H}}\otimes\bar{\mathcal{H}}'$, $\sigma(x)_{\bar{\mathcal{H}}}=\Tr_{\bar{\mathcal{H}}'}\proj{x}_{\bar{\mathcal{H}}\bar{\mathcal{H}}'}$ is the reduced state of $\proj{x}$ on $\bar{\mathcal{H}}$, and
$\varepsilon(\widetilde{\theta}_x)_{\mathcal{H}^{k-q}}=\Tr_{\mathcal{H}'^{k-q}}\epsilon(\widetilde{\theta}_x)_{(\mathcal{H}\cH')^{k-q}}$ is the reduced sub-normalized state of $\epsilon(\widetilde{\theta}_x)$ on $\mathcal{H}^{\otimes k-q}$.
Putting everything together, we get the following: Let $n\in\mathbf{N}$ and $k=\lfloor n^{\alpha} \rfloor$ for a given $\alpha$ fulfilling $2/3<\alpha<1$. Let also $\lambda_0$ be a threshold such that on the one hand $\lambda_0\geq C_0\log n$, where $C_0>0$ is a universal constant, and on the other hand $d\leq n^{\beta}$ for a given $\beta$ fulfilling $0<\beta<1/2$. Suppose next that $\rho^{(n+2k)}$ is a $(n+2k)$-symmetric state on $\mathcal{H}^{\otimes n+2k}$ such that event $\mathcal{A}$ holds. Then, with probability greater than $1-e^{-cn^{3\alpha-2}}$, where $c>0$ is a universal constant, the reduced state $\rho^{(n+k)}$ of $\rho^{(n+2k)}$ on $\mathcal{H}^{\otimes n+k}$ satisfies
\[ \rho^{(n+k)} \leq (Cn)^{n^{\alpha}+n^{2\beta}} \sum_{\underset{|I|\geq n}{I\subset[n+k]}} \int_{\sigma_{\bar{\mathcal{H}}}\in\mathcal{D}(\bar{\mathcal{H}})} \varepsilon(\rho,\sigma)_{\mathcal{H}^{I^c}}\otimes\sigma_{\bar{\mathcal{H}}}^{\otimes I} \mathrm{d}\mu(\sigma_{\bar{\mathcal{H}}}), \]
where $C>0$ is a universal constant, $\mu$ is a probability measure on the set of states on $\bar{\mathcal{H}}$, and for each $0\leq q\leq k$ and each state $\sigma$ on $\bar{\mathcal{H}}$, $\varepsilon(\rho,\sigma)_{\mathcal{H}^{k-q}}$ is a sub-normalized state on $\mathcal{H}^{\otimes k-q}$.
Now, let $\mathcal{N}:\mathcal{L}(\mathcal{H})\rightarrow\mathcal{L}(\mathcal{K})$ be a quantum channel, and assume that there exists some $0<\delta<1$ such that
\[ \sup\left\{ \left\|\mathcal{N}(\sigma)\right\|_1 \ : \ \sigma\in\mathcal{D}(\bar{\mathcal{H}})\right\} \leq \delta. \]
This implies in particular that, for any $0\leq q\leq k$, and any states $\varepsilon$ on $\mathcal{H}^{\otimes k-q}$, $\sigma$ on $\bar{\mathcal{H}}$, we have
\[ \left\|\mathcal{N}^{\otimes n+k}\left(\varepsilon\otimes\sigma^{\otimes n+q}\right)\right\|_1 = \left\|\mathcal{N}^{\otimes k-q}\left(\varepsilon\right)\right\|_1\left\| \mathcal{N}(\sigma)\right\|_1^{n+q} \leq \delta^{n+q}. \]
And subsequently, by what precedes, we obtain the following: For any state $\rho^{(n+2k)}$ on $\mathcal{H}^{\otimes n+2k}$ such that event $\mathcal{A}$ holds, denoting by $\rho^{(n+k)}$ its reduced state on $\mathcal{H}^{\otimes n+k}$, we have with probability greater than $1-e^{-cn^{3\alpha-2}}$,
\begin{align*} \left\|\mathcal{N}^{\otimes n+k}\left(\rho^{(n+k)}\right)\right\|_1 \leq & \, (Cn)^{n^{\alpha}+n^{2\beta}} \sum_{q=0}^{k}{n+k \choose q} \sup\left\{ \left\|\mathcal{N}^{\otimes n+k}\left(\varepsilon\otimes\sigma^{\otimes n+ q}\right)\right\|_1 \ : \ \varepsilon\in\mathcal{D}(\mathcal{H})^{\otimes k-q},\ \sigma\in\mathcal{D}(\bar{\mathcal{H}}) \right\}\\
\leq & \, (Cn)^{n^{\alpha}+n^{2\beta}}\sum_{q=0}^{k}{n+k \choose q}\delta^{n+q}\\
\leq & \, (C'n)^{n^{\alpha}+n^{2\beta}}\, \delta^n, \end{align*}
where $C'>0$ is a universal constant. By the way $\alpha,\beta$ have been chosen, this means that, for any $\tilde{\delta}>\delta$ and $n\geq n_{\tilde{\delta}}$, we have with high probability
\[
\sup\left\{ \left\|\mathcal{N}^{\otimes n+k}\left(\rho^{(n+k)}\right)\right\|_1 \ : \ \rho^{(n+k)}=\Tr_{\mathcal{H}^{\otimes k}}\rho^{(n+2k)}\ \text{with}\ \rho^{(n+2k)}\in\mathcal{D}(\mathcal{H})^{\otimes n+2k}\ \text{such that}\ \mathcal{A}\ \text{holds} \right\}
\leq \tilde{\delta}^n.
\]
\section{Conclusion and outlook}
We have reviewed (and given a new proof) of the constrained de Finetti
reduction of \cite{DSW}. We have demonstrated its adaptability to various situations
where one would like to impart a known constraint satisfied by a
permutation-invariant state onto the i.i.d.~states occurring in the
operator with which to compare it. We have seen that our technique works especially well in the case of linear constraints (see \cite{DSW} and \cite{LaW2} for two developed such applications).
We have then spent considerable effort on a particularly interesting
convex constraint, separability. Apart from the obvious relevance
to entanglement theory, the constrained de Finetti reduction provides a
very natural framework in which to derive bounds on the success
probability of parallel repetitions of tests, and has immediate
applications in the parallel repetition of $\mathrm{QMA}(2)$, quantum
Merlin-Arthur interactive proof systems with two unentangled provers (see \cite{HM} for further details).
Conversely, we showed that certain progress in entanglement theory
(on the conjectured faithfulness properties of the CEMI entanglement
measure for instance) would imply even stronger, dimension-independent bounds, which
would show in particular that the soundness gap of $\mathrm{QMA}(2)$ can
be amplified exponentially by parallel repetition, without any other
devices. It is curious to see that the progress on questions like this
can depend on the properties of a simple, but little-understood
entanglement measure such as CEMI, and we would like to recommend
its study to the reader's attention. Indeed, it seems to be the best candidate so far for a \emph{magical}, or even \emph{supercalifragilistic} entanglement measure \cite{Winter}. The latter is defined as one which has the post-selection property with respect to an initial product state and measurement on a separate subsystem (cf.~Lemma \ref{lemma:post-POVM}), is super-additive, and satisfies a universal faithfulness bound with respect to the trace-norm distance (cf.~Conjecture \eqref{eq:E_I-conjecture}).
We have also presented a more abstract framework of convex
constraints, that allows us to demonstrate in greater generality
the interplay between the multiplicative behaviour
of $(i)$ the support function and $(ii)$ the maximum fidelity function. The way $(i)$ is derived from $(ii)$ is via our de Finetti reduction with fidelity weight in the upper bounding operator. And $(ii)$ is obtained from $(i)$ by constructing a test whose failure probability decays exponentially under parallel repetition.
Finally, seeing that the de Finetti reductions had been so far always
limited by the finite dimensionality of the system
involved, we have made first steps towards an extension of
the main technical tool to infinite-dimensional systems under
suitable constraints. It remains to be seen how widely it or a variation can be applied to quantum cryptography in continuous variable systems \cite{CR,CGPLR}, or similar problems.
\section*{Acknowledgments}
This research was supported by the European Research Council (Advanced Grant IRQUAT,
ERC-2010-AdG-267386), the European Commission (STREP RAQUEL, FP7-ICT-2013-C-323970),
the Spanish MINECO (project FIS2013-40627-P), with the support of FEDER
funds, the Generalitat de Catalunya (CIRIT project 2014-SGR-966), and the French CNRS (ANR
projects OSQPI 11-BS01-0008 and Stoq 14-CE25-0033).
\addcontentsline{toc}{section}{References}
|
\section{Introduction}
\paragraph{Motivation}
This paper deals with the question of why, and whether, a model
of interacting strategic agents converges to equilibrium. We
study this question in Fisher markets, under conditions where
market equilibrium is unique and finding it is computationally tractable. Over
the years and in particular recently, several game and market
dynamics have been studied, but they fall short of modeling
the key scenario which we attempt to address.
In particular, in game theory,
dynamics are studied in the context of repeated games. Extensive
form solution concepts such as subgame perfect or sequential
equilibria assume that the agents unravel the entire evolution
of the game and choose in advance their entire play optimally.
This is likely to be computationally infeasible (e.g.~\cite{BorgsCIKMP10},
but see in contrast~\cite{HalpernPS14}). The strategies
are unrealistically prescient of the distant future, contradicting
experience and hindering on-the-fly adaptation to unexpected
changes. Just as importantly, since the entire play is determined a-priori
and an equilibrium is played throughout, such concepts do not
capture out-of-equilibrium behavior (that may lead to equilibrium
over time), so they are in fact a static notion.
Walrasian
{\em t\^atonnement}, and more generally game theoretic
{\em learning} dynamics (a.k.a. no-regret dynamics), are an
alternative approach. These are truly dynamic, out-of-equilibrium frameworks, that
can be shown in many cases to converge to an attractive
solution concept. However, the reactions of the agents have
to be damped carefully for a desirable outcome to materialize;
such reactions lack strategic justification
(see~\cite{BlumM07} and the references therein, e.g.,~\cite{HartM00}).
Closer to our work, various formulations of bounded rationality
have provided a rich basis for progress in game theory and, over
the last two decades, in its algorithmic aspects. The most basic
approach in this vein is {\em best-response} dynamics. Agents
play myopically an optimal move at each round, assuming that the
other agents will not deviate from their
existing strategy.
\footnote{The situations where best-response is known to lead to
an attractive outcome are tightly connected to the concept of
{\em potential games}.
See~\cite{MondererS96,AwerbuchAEMS08,ChienS11}. For a damped
version, {\em logit} dynamics, see~\cite{AulettaFPPP15}. For a general
discussion of best-response and the related {\em fictitious play}
dynamics, see~\cite{ShohamL09}.}
A strategy which is somewhat
more sophisticated than best-response is limited-depth exploration
of an extensive form game tree. This is an approach to complex
games that was developed in the early days of AI (the exploration
depth is sometimes called the {\em ply} of a search). Essentially
the same concept is known in control theory as {\em receding-horizon
control}. This is in contrast with the full-rationality approach underlying
solution concepts such as the aforementioned subgame perfect or
sequential equilibria.
In game theory, the idea that people compete by pursuing limited-lookahead situational analysis goes under the rubric of the {\em level $k$ model}, initiated
by~\cite{stahlW94,stahlW95} and~\cite{nagel95}; related ideas are also known as
{\em cognitive hierarchy}, {\em higher-order rationality}, and
{\em bounded depth of reasoning}.
The idea has been subjected to many experimental
tests---see~\cite{hoCW98,cgCB01,crawford03,camererHC04,cgC06,crawfordI07a,crawfordI07b}---and has emerged with considerable support. For recent theoretical work on the model see~\cite{strzalecki14,kneeland,dCSS,gorelkina}; for a survey see~\cite{ccgi13}.
In view of the above, it is important to study the dynamics and
stability of markets composed of agents each of whom performs
some limited lookahead and plays optimally against that forecast.
Limited lookahead means that each agent $j$ has a mental model
of each other agent $k$, where $k$ looks ahead some constant number
of steps, and based on that chooses an optimal action (according to $j$'s
perception). Based on this model, $j$ chooses a move that is optimal
conditional on those other imagined actions. The paradox of endless
self-reference is obvious here, and is precisely the point of the exercise:
such a model does not make sense for infinitely-intelligent agents who
possess perfect common knowledge of the properties of the market.
But such agents do not exist. Instead, the model is consistent with
experience that markets are composed of many agents who, despite
having limited ability to predict the actions of others, do their best to
make such a prediction and then respond optimally to their own prediction.
This is a very different approach to agent choice than the ``solution concept"
notion on which game theory rests: Nash equilibria, correlated equilibria,
the core, and so forth. In particular, one difference is that in contrast with
full rationality, in the limited lookahead case the beliefs of the agents are
not necessarily consistent with each other and with reality. In fact, they may
even be self-inconsistent across time steps. From a purely mathematical perspective these inconsistencies might appear to be a fatal flaw. We hold differently, that this is part of the challenge of modeling out-of-equilibrium strategic play. The market is out of equilibrium
{\em because} players do not have perfect models of each other, or because they are uncertain about exogenous factors a few steps into the future. We further hold that the predictive power in experiments of the {\em level $k$ model} is ample reason to study its dynamics.
That is what we do here (and for a more general notion of
best-response with lookahead).
\paragraph{Our results}
This paper is devoted to studying the dynamics and stability
of markets where the agents model their
peers as using limited lookahead. We focus on one of the best-understood
cases of general equilibrium theory, namely Fisher markets that
consist of sellers of goods and buyers endowed with budgets.
In fact, we develop a general framework that shows convergence
of dynamics based on limited lookahead, only requiring certain abstract conditions on the updates of the players. A concrete special case is a Fisher market
in which the buyers generate demand due to utilities that exhibit
{\em constant elasticity of substitution} (CES), in the {\em weak
gross substitutes} (WGS) regime. (Rigorous definitions await
Sections~\ref{sec: preliminaries} and~\ref{sec: concrete}.)
CES utilities were chosen because this setting is very well
understood in the context of discrete-time t\^atonnement
(see~\cite{CCD13}), and this gives us some comparative
perspective. The restriction to WGS was imposed because
otherwise even staying at a market equilibrium cannot be
reasoned by individual sellers best-responding to the equilibrium.
Our dynamic model focuses on the sellers. Each seller
controls and sets the price of a unique single good. The buyers
are assumed to react instantly and myopically to current prices
by adjusting their demand to optimize their utilities subject to
their budgets. This assumption can be justified, for instance, by
assuming that there is a large number of buyers, each contributing
negligibly to the demand.
Each seller
is assumed to form a belief on the next move of each of the
other sellers, and then to choose a price that optimizes its own
profit based on this belief. We analyze a rather general
belief formation process that includes, as a special case,
beliefs based on assuming that the other sellers use limited
lookahead. We assume neither consistency among the
beliefs formed by different sellers, nor consistency
among the beliefs formed by the same seller at different
times. This includes as a special case, but is considerably
more general than, {\em level $k$} choices. See Figure~\ref{genl-fict-play}.
We refer to dynamics of this sort as
{\em best-response with lookahead} (abbreviated BRL) dynamics.
\begin{figure}
\includegraphics[height=150mm]{genl-fict-play}
\caption{Various collective mental models for one round of play in a 3-seller market. A leaf ({\em level $1$}) denotes best-response dynamics. In {\em level $2$} dynamics everyone
best-responds to everyone's best-response to current prices. Players' beliefs can be far more complex. In the last example C plays by a {\em level $2$} model while A and B have more elaborate mental models.}
\label{genl-fict-play}
\end{figure}
BRL dynamics, and even the special case of best-response
(with no lookahead), can be quite volatile,
as compared with usual t\^atonnement processes, because
of the absence of any damping factor. Despite the volatility and the potential
inconsistency of beliefs, we show that
regardless of the specifics of the beliefs formed by the agents,
the dynamic converges rapidly to market equilibrium. More precisely, we analyze two
versions of our process. In the synchronous case, all sellers
update prices simultaneously. In this case, the distance
to equilibrium decays exponentially in the number of steps
(a.k.a.\ {\em linear convergence}).
In the asynchronous case, at each time step only a subset
of one or more sellers update prices. In this case, the distance
to equilibrium decays exponentially in the number of epochs,
where an epoch consists of time intervals in which all the sellers
update at least once.
To the best of our knowledge, convergence, and definitely linear
convergence, was not previously demonstrated even for the simplest
version of our process, namely best-response. Our proof of convergence
relies on showing that in a judiciously chosen metric (the Thompson
metric), the BRL dynamics form a contraction map.
\paragraph{Related work}
General equilibrium theory is the principal framework through
which economists understand the operation of markets
(see~\cite{McKenzie02,Mukherji02}). It is one of the great
achievements of economic theory in general and of mathematical
modeling of microeconomics in particular. The theory is
largely responsible for the governing paradigm that a state
of {\em equilibrium} which the participants in economic
exchange do not wish to deviate from individually
is under mild conditions attainable~\cite{ArrowD54,McKenzie54}
(see also~\cite{Hildenbrand98}),
and that this is normally roughly the state of the economy.
This is a paradigm that can be observed ``in the field''
and also reproduced in controlled experiments, and it
lends credence and concreteness to the famed
{\em invisible hand} metaphor.
In contrast, there is less agreement
on an effective explanation as to {\em why} markets tend to
reach a state of equilibrium. This is a question about the
{\em stability} or {\em out-of-equilibrium} behavior of
markets. It is important because in reality economic
conditions are not static. They vary continually and
suffer serious ``shocks'' occasionally. So justifying an
equilibrium outcome requires a dynamic that moves
an economy at disequilibrium back to a new
equilibrium, and does so sufficiently quickly that the
periods of disequilibrium due to fluctuations are relatively
negligible (see~\cite{Dix90}). The classical mechanism
proposed to explain general equilibrium is Walrasian
t\^atonnement~\cite{Walras74}, a process that reacts
to excess demand by raising the price and to excess
supply by reducing the price. Variants of t\^atonnement
are known to converge to equilibrium, at least in some
classes of markets including those we consider here
(e.g.~\cite{Samuelson41,ABH59,CCD13}). However,
the classical view of
t\^atonnement posits the existence of an imaginary
``auctioneer'' who controls the process by announcing
prices. Recent work on the convergence of discrete-time
t\^atonnement in Fisher markets attempts to present it
as an in-market process in the context of the so-called
{\em ongoing markets}~\cite{CF08,CCR12}. However, even
this attempt requires a somewhat careful choice of the
magnitude of the price adjustment which is not motivated
by any agent considerations (aside from a common
inexplicable passion to equilibrate the economy).
Thus, the difficulty is in formulating
a theory of out-of-equilibrium behavior that makes sense
in terms of the incentives of the participants.
In is well-known that market equilibria in Fisher markets
with CES utilities can be expressed as solutions to a
convex program, first proposed by Eisenberg and Gale
(see~\cite{JainV10}).
We observe that best-response dynamics (i.e., the simplest
example of our setting) can, in fact, be explained as a specific
implementation of coordinate descent (in the dual program).
The convergence of coordinate descent
was established in~\cite{tseng2001convergence},
without bounds on the rate. Recently, \cite{saha2013nonasymptotic}
established a sublinear convergence rate (the distance to the optimum
decays linearly with the number of iterations), if the objective function
satisfies some conditions. We note that the objective
function of the dual Eisenberg-Gale program satisfies these conditions.
Our general result shows a linear convergence rate (the
distance to equilibrium decays exponentially in the number of
iterations), and this holds in particular in the case of best-response.
To the best of our knowledge, this is not implied by previous results.
Two recent papers consider market dynamics under strategic
behavior. Both bound the fraction of optimal welfare that is guaranteed.
In~\cite{BabaioffLNP14}, strategic buyers play a Nash (or Bayesian)
equilibrium in a market in which the sellers' prices are determined by
Walrasian t\^atonnement; note that here the t\^atonnement
is part of the {\em mechanism} defining the game, rather than the agents' strategies.
In~\cite{BabaioffPS15}, sellers engage in best-response
dynamics. In this setting the market does not actually have an equilibrium,
but a fraction of the optimal welfare can be extracted by the dynamic.
In both papers the market model is quite different from ours.
In the game
theory setting (as opposed to markets), best-response dynamics have
been studied extensively in recent years, mostly concerning bounds on
the quality of the play and conditions
that imply or prevent convergence to a Nash equilibrium~\cite{Mirrokni2004,Roughgarden15,FanelliFM12,EngelbergFSW13}.
The paper~\cite{NisanSVZ11} investigates conditions under which
best-response is a fully rational strategy.
\section{Preliminaries}\label{sec: preliminaries}
\paragraph{The market model}
We consider a Fisher market with $n$ perfectly divisible goods
and $m$ buyers. Each good is initially owned by a unique seller
that controls its price, and its quantity is scaled to $1$.
The utility of that seller is the price times $\min\{\text{demand},1\}$.
The buyers
respond instantly and myopically to price changes. Thus their role
in the process is to specify in a convenient way the demands for
the goods at any given assignment of prices to those goods. This
is done as follows. Each buyer $i$ is endowed with a positive
budget $b_i$ and a utility function $u_i$ over baskets of goods
$x$.
For a price vector $p$, we write $p > 0$ to indicate that all
the prices are strictly positive. Similarly for price vectors $p,q$,
we write $p>q$ (resp.\ $p \geq q$) if $p_j>q_j$ (resp.\ $p_j\geq q_j$)
for all $j$.
Given a price vector $p > 0$, the demands for the goods are
determined as
follows. Every buyer $i$ chooses $x_i$ to optimize the utility
function $u_i(x)$, subject to the budget constaint
\begin{align}
\sum_j p_j x_{ij}\leq b_i.\label{eq:Budget}
\end{align}
We denote the utility maximizing allocations for prices $p$
by $x(p)$. The demand for each good $j$ at prices $p$
is $\sum_i x_{ij}(p)$.
\paragraph{Price updates}
In general, a market dynamic is based on an update rule for
each seller that determines its new price. The rules can then
by applied synchronously to all sellers, or serially to one
seller at a time in some order. We will discuss these variations
later. For now, we focus on the update rules. An update rule
can take into account some or all of the dynamic history leading
to the current state (including the current prices), and also
some internal state of the seller that takes other factors into
account. We are interested in update rules that depend on
the current price vector (and any other parameters),
and are {\em monotone}, {\em sub-homogeneous},
{\em price-bounded}, and {\em positive} with respect to that price vector.
To define these properties formally, let $F^{\iota}_j(p)$ denote the
new price of seller $j$, given current prices $p$, and $\iota$
encoding all the other relevant parameters (if any). Then,
\begin{definition}[monotonicity, sub-homogeneity, price-boundedness, positivity]
We say that:
\begin{itemize}
\item $F^{\iota}_j$ is monotone if for all pairs
of price vectors $p,q$ such that $p\ge q$ coordinate-wise,
$F^{\iota}_j(p)\ge F^{\iota}_j(q)$;
\item $F^{\iota}_j$ is sub-homogeneous if for all
price vectors $p$ and for all $\lambda\in (0,1)$,
$F^{\iota}_j(\lambda p)\ge \lambda F^{\iota}_j(p)$,
also $F^{\iota}_j$ is strictly sub-homogeneous iff
the inequality is strict for all $p > 0$;
\item $F^{\iota}_j$ is
$[p_{\min},p_{\max}]$-price-bounded if
for all price vectors $p\in [p_{\min},p_{\max}]^n$,
$F^{\iota}_j(p)\in [p_{\min},p_{\max}]$.
\item $F^{\iota}_j$ is positive if $p_{\min} > 0$.
\end{itemize}
\end{definition}
For a price vector $p$ and price updates $F_j$
for all $j\in [n]$, we denote by $F(p)$ the price
vector derived by applying the updates simultaneously
to $p$. We say that $F$ has a property (e.g.,
is monotone) if all of its components have this property.
\begin{lemma}\label{lm: composition}
Fix $p_{\min},p_{\max}$, and suppose that
$F:[p_{\min},p_{\max}]^n\rightarrow [p_{\min},p_{\max}]^n$
and $g:[p_{\min},p_{\max}]^n\rightarrow [p_{\min},p_{\max}]$
are both monotone, sub-homogeneous, and
$[p_{\min},p_{\max}]$-price bounded updates. Then
so is $g\circ F$. Moreover, if $g$ is strictly sub-homogeneous
and $F$ is positive, then $g\circ F$ is also strictly sub-homogeneous.
\end{lemma}
\begin{proof}
By the monotonicity of $F$, if $p \ge q$ coordinate-wise,
then $F(p) \ge F(q)$ coordinate-wise. Therefore, by
the monotonicity of $g$, we have that $g(F(p))\ge g(F(q))$.
Next,
$g(F(\lambda p)) \ge g(\lambda F(p))\ge \lambda g(F(p))$,
where the first inequality uses the
monotonicity of $g$ and the
sub-homogeneity of $F$,
and the second inequality uses the
sub-homogeneity of $g$. Moreover, if $g$ is strictly
sub-homogeneous, then the second inequality is strict
if $F(p) > 0$, which is implied when $p > 0$ by
the assumption that $F$ is positive.
Finally, using the price boundedness of both $F$ and $g$,
if $p\in [p_{\min},p_{\max}]^n$, then
$F(p)\in [p_{\min},p_{\max}]^n$, so $g(F(p))\in [p_{\min},p_{\max}]$.
\end{proof}
\paragraph{Belief formation}
We consider dynamics where each seller updates its
price according to a belief of which prices the other sellers
will set in the next step. The beliefs that are formed by
different sellers or by the same seller at different
times need not be consistent.
We show that despite this inconsistency, the dynamics still
converge to equilibrium, assuming that the ingredients
satisfy certain properties.
In general, a belief
$\pi$ is a function that maps a pair $(p,\iota)$,
where $p$ is the current price vector and $\iota$
is the internal state of the seller,
to the believed price vector.
We now discuss a rather general framework of forming
such beliefs. This framework in particular enables the
sellers to form {\em level $k$} best-response beliefs, and
more general best-response beliefs. We haven't yet formally
defined ``best-response'', but for now it
suffices to assume that there is at our disposal a price
update called best-response.
The details of best-response update
are discussed in Section~\ref{sec: concrete}. Also,
in order to get some intuition on the following explanation,
it might be useful
to visualize the trees in the bottom example in
Figure~\ref{genl-fict-play}.
We explain how seller $j$ forms a belief $\pi = \pi^{\iota}$.
The idea is that seller $j$ has, for every other
seller $k$, a mental model $\iota_k$ of the update rule
that $k$ employs, and $\pi_k$ is simply the
price that $\iota_k$ generates. Of course, in order to
form $\iota_k$, seller $j$ must also imagine
seller $k$'s mental models for all $k'\ne k$ (this
includes $j$). So, we define inductively a set
of possible mental models of seller updates, and
seller $j$ simply picks each $\iota_k$ from this set.
The set ${\cal M}$ of mental models consists of levels;
${\cal M} = \bigcup_{s=0}^{\infty}{\cal M}_s$.
They are defined inductively as follows. The base
case, level $0$, is ${\cal M}_0$ that contains a single model
of staying put at the current price.
Inductively, a mental model or belief $\iota$ for player $j$ is formed by selecting any $\iota_{k_1},\ldots,\iota_{k_n}$ (but there is no $\iota_j$); the price update defined by this mental model is player $j$'s best-response to the prices generated by all the other players if they act with the assigned mental models on the basis of the current prices. The \textit{level} of $\iota$ is one more than the maximum level of $\iota_{k_1},\ldots,\iota_{k_n}$.
\suppress{
\ljs{Previous text is suggested replacement for the rest of this OLD TEXT:}
The inductive step
includes in ${\cal M}_s$ all the models $\iota_k$ that can
be generated as follows. For every $k'\ne k$, pick a
model $\iota_{k,k'}\in\bigcup_{r=0}^{s-1}{\cal M}_{r}$.
This defines $\iota_k$, which is $j$'s model of $k$;
according to $j$'s model, seller $k$ forms a belief
$\pi^{\iota_{k,\cdot}}$ based on the models
$\iota_{k,k'}$ for all $k'\ne k$, and updates by
best-responding to this belief.
Notice that there is an intuitive notion of ``depth'' that
can be associated with a mental model $\iota$, which
is the minimum $s$ such that for all $k$,
$\iota_k\in \bigcup_{r=0}^{s}{\cal M}_{r}$. Notice
that the depth of $j$'s model $\iota$ is greater than the
depth of any model $\iota_{k,\cdot}$ that $j$ assumes
the other sellers have.
\ljs{OLD TEXT TILL HERE}
}
We note in passing that beliefs thus formed, implicitly model
sellers with epistemic assumptions that they are a bit smarter
than their peers---every seller $j$ updates with one extra step
beyond the maximum number of steps used in $j$'s mental
model $\iota$. Of course, such beliefs cannot possibly be
consistent among sellers (unless they are children in Lake
Wobegon).
The following lemma states the desired properties of
belief formation.
\begin{lemma}\label{lm: belief-based updates}
Fix $p_{\min},p_{\max}$. Suppose that best-response
is monotone, sub-homogeneous,
$[p_{\min},p_{\max}]$-price bounded and positive. Further suppose
that seller $j$ uses a monotone, strictly sub-homogeneous,
and $[p_{\min},p_{\max}]$-price bounded update function
$F_j$, and given current prices $p$, updates to
$F^{\iota}_j(p) = F_j(\pi^{\iota}(p))$.
Then, $F^{\iota}_j$ is monotone, strictly sub-homogeneous,
and $[p_{\min},p_{\max}]$-price-bounded.\footnote{In fact, the
conclusion of Lemma~\ref{lm: belief-based updates} holds even
for beliefs that are formed by a set of monotone, sub-homogeneous,
price bounded, and positive price updates, instead of a single
such update. In the tree view of belief formation, such a set
is used by choosing, for each node of the tree, an arbitrary
member of the set as the modeled action. To simplify the exposition,
we do not elaborate on this generalization.}
\end{lemma}
\begin{proof}
Since staying put at the current price is monotone, sub-homogeneous, and
$[p_{\min},p_{\max}]$-price-bounded, a simple induction
on $s$ using Lemma~\ref{lm: composition} shows that
$\pi^{\iota}$ is monotone, sub-homogeneous, and
$[p_{\min},p_{\max}]$-price-bounded. One more application
of Lemma~\ref{lm: composition} gives the desired properties
of $F^{\iota}_j$.
\end{proof}
\section{Concrete case of CES-WGS markets}\label{sec: concrete}
\paragraph{CES utilities}
These are utility functions of the form
\begin{align}
u_i(x) = \left(\sum_j \left(c_{ij}x_{ij}\right)^{\rho}\right)^{\frac{1}{\rho}},\label{eq:Util}
\end{align}
where $x_{ij}$ denotes the quantity of good $j$ that buyer $i$
purchased. The parameter $\rho\in (-\infty,0)\cup (0,1)$ is assumed,
for simplicity, to be uniform for all buyers. These utility functions
are known as {\em constant elasticity of substitution} (CES) utilities.
When $\rho\in (0,1)$, the goods are
{\em weak gross substitutes} (WGS). In the case of $\rho\in (-\infty,0)$,
the goods are complementary.
For CES utilities, the utility-maximizing
allocations are given explicitly by the equation
\begin{align}
x_{ij}(p) = \frac{b_i}{p_j}\cdot\frac{\left(c_{ij}/p_j\right)^{\epsilon}}
{\sum_k \left(c_{ik}/p_k\right)^{\epsilon}},\label{eq:Optx}
\end{align}
where $\epsilon = \frac{\rho}{1-\rho}$.
Notice that if $\rho\in (0,1)$, then $\epsilon \in (0,\infty)$.
In this case, the demand satisfies the following property.\footnote{The
proof of this fundamental fact is rather trivial, but we are not aware
of a good reference. Notice that the same proof
shows that if the CES utilities are complementary ($\epsilon < 0$), then
the spending on good $j$ is monotonically increasing in $p'_j$. This
is the motivation for considering only WGS utilities.}
\begin{lemma}\label{lm: profit monotonicity} Let the utilities be CES in the WGS regime.
Fix a price vector $p$ and a good $j$. Consider all price
vectors $p'$ with the property that for all $k\ne j$, $p'_k = p_k$.
Among these price vectors, the total desired spending
$\sum_i x_{ij}(p')\cdot p'_j$ on good $j$ is monotonically
decreasing in $p'_j$.
\end{lemma}
\begin{proof}
Using Equation~\eqref{eq:Optx},
the total desired spending on $j$ is given by the equation
$$
\sum_i x_{ij}(p')\cdot p'_j =
\sum_i b_i\cdot\frac{\left(c_{ij}/p'_j\right)^{\epsilon}}
{\left(c_{ij}/p'_j\right)^{\epsilon}+\sum_{k\ne j} \left(c_{ik}/p_k\right)^{\epsilon}}.
$$
The derivative of the right-hand side with respect to $p'_j$
is
$$
-\epsilon\cdot \sum_i \frac{b_i}{p'_j}\cdot\left(\left(\frac{\left(c_{ij}/p'_j\right)^{\epsilon}}
{\left(c_{ij}/p'_j\right)^{\epsilon}+\sum_{k\ne j} \left(c_{ik}/p_k\right)^{\epsilon}}\right)
- \left(\frac{\left(c_{ij}/p'_j\right)^{\epsilon}}
{\left(c_{ij}/p'_j\right)^{\epsilon}+\sum_{k\ne j} \left(c_{ik}/p_k\right)^{\epsilon}}\right)^2\right).
$$
This expression is negative, because
$\frac{\left(c_{ij}/p'_j\right)^{\epsilon}}
{\left(c_{ij}/p'_j\right)^{\epsilon}+\sum_{k\ne j} \left(c_{ik}/p_k\right)^{\epsilon}} < 1$.
\end{proof}
\begin{corollary}\label{cor: profit maximality}
Using the same notation as in Lemma~\ref{lm: profit monotonicity},
the profit $\min\left\{\sum_i x_{ij}(p'),1\right\}\cdot p'_j$ of seller $j$
is maximized at the price $p'_j$ for which the demand $\sum_i x_{ij}(p')$
equals $1$.
\end{corollary}
\begin{proof}
By Equation~\eqref{eq:Optx}, the demand $\sum_i x_{ij}(p')$ decreases
monotonically in $p'_j$. By Lemma~\ref{lm: profit monotonicity}, also
the desired spending decreases monotonically in $p'_j$. Therefore,
the profit is maximized at the lowest price for which the demand is
at most $1$ (lowering the price further will not increase the quantity
sold beyond the initial endowment).
\end{proof}
\paragraph{Best-response updates}
In standard best-response dynamics, each seller updates its price
to maximize its revenue given the current
prices of the other players. In the particular setting of demand
that is generated by CES utilities in the WGS regime, a seller $j$
maximizes profit by setting the price $p_j$ to clear the market for
good $j$ (by Corollary~\ref{cor: profit maximality}).
I.e., if the current price vector is $p$, the seller chooses
a new price $F_j(p)$ for good $j$, so that
\begin{align}
\sum_{i=1}^m x_{ij}(p')=1,\label{eq:Seller}
\end{align}
where $p'_j = F_j(p)$, and for all $k\ne j$, $p'_k = p_k$.
More explicitly, seller $j$ best-responds by solving
for $p'_j$ the equation
\begin{align}
p'_j = \sum_{i=1}^m b_i\cdot\frac{\left(c_{ij}/p'_j\right)^{\epsilon}}
{\sum_{k\neq j} \left(c_{ik}/p_k\right)^{\epsilon}+\left(c_{ij}/p'_j\right)^{\epsilon}}. \label{eq:BestResponse}
\end{align}
Notice that the right-hand side of Equation~\eqref{eq:BestResponse}
is simply the total spending of all the buyers on good $j$ when this
good's price is $p'_j$ and the other prices are given by the vector
$p$.
\begin{lemma}\label{lm: best response properties}
For CES utilities with $\rho\in (0,1)$,
best-response updates $F_j$ are monotone, strictly
sub-homogeneous, positive, and $[p_{\min},p_{\max}]$-price-bounded
for some $p_{\min} = p_{\min}(b,c,\rho)>0$ and
$p_{max} = p_{\max}(b)$.
\end{lemma}
\begin{proof}
We begin with monotonicity. Consider two price vectors $p \ge q$,
and let $p'_j = F_j(p)$ and let $q'_j = F_j(q)$. Consider the function
$$
g(\alpha,p) = \alpha - \sum_{i=1}^m b_i\cdot\frac{\left(c_{ij}/\alpha\right)^{\epsilon}}
{\sum_{k\neq j} \left(c_{ik}/p_k\right)^{\epsilon}+\left(c_{ij}/\alpha\right)^{\epsilon}}.
$$
In other words, $g(\alpha,p)$ is $\alpha$ minus the total spending of all the
buyers on good $j$ when the price of good $j$ is $\alpha$ and the other prices
are given by the vector $p$ (thus, $g(\alpha,p) = 0$ iff $\alpha = F_j(p)$).
We have that $g(p'_j,p) = g(q'_j,q) = 0$. Notice that for any $k$, increasing $p_k$
decreases $g(\alpha,p)$. On the other hand, $g(\alpha,p)$ increases as $\alpha$
increases (an immediate consequence of
Lemma~\ref{lm: profit monotonicity}). Thus, $g(p'_j,q)\geq 0$. If
$g(\alpha,q) = 0 \le g(p'_j,q)$, then it must be that $\alpha\le p'_j$.
Thus $q'_j\le p'_j$.
Next we prove sub-homogeneity. Let $p'_j=F_j(p)$. Then,
$$
g(\lambda p'_j,\lambda p) =
\lambda p'_j - \sum_i b_i \cdot \frac{\left(c_{ij}/p'_j\right)^{\epsilon}}
{\sum_{k\neq j} \left(c_{ik}/p_k\right)^{\epsilon} +
\left(c_{ij}/p'_j\right)^{\epsilon}} = g(p'_j,p) - (1-\lambda)\cdot p'_j \leq 0.
$$
Thus, by the monotonicity of $g$ in $\alpha$,
if $g(\alpha,\lambda p) = 0\ge g(\lambda p'_j,\lambda p)$,
then $\alpha \geq \lambda p'_j$. Finally, if all the entries of $p$
are strictly positive, then $p'_j > 0$, so
$g(\lambda p'_j,\lambda p) < 0$ and therefore
$g(\alpha,\lambda p) = 0$ implies that $\alpha > \lambda p'_j$.
Finally, we show price-boundedness.
As before, let $p'_j=F_j(p)$.
For the upper bound, notice that if
$p_j > \sum_i b_i$ then the total demand for good $j$ must be
less than $1$, regardless of the other prices. So we can simply
set $p_{\max} = \sum_i b_i$. For the lower bound, consider any buyer
$i$ with $c_{ij} > 0$. Consider the situation where all the entries
of $p$ are at least $p_{\min}$, and the price of good $j$ is
$q_j < p_{\min}$ (we will specify $p_{\min}$ shortly). The
demand that $i$ has for good $j$ in this situation is
$$
\frac{b_i}{q_j}\cdot\frac{\left(c_{ij}/q_j\right)^{\epsilon}}
{\sum_{k\neq j} \left(c_{ik}/p_k\right)^{\epsilon}+\left(c_{ij}/q_j\right)^{\epsilon}}
> \frac{C_{ij}}{p_{\min}},
$$
where
$C_{ij} = \frac{b_i\cdot c_{ij}^{\epsilon}}{\sum_k c_{ik}^{\epsilon}}$.
Set $p_{\min} = \min\{1/C_{ij}:\ c_{ij} > 0\}$. The
demand that $i$ alone generates for good $j$ is more than $1$,
so $q_j\ne p'_j$. Thus, we must have $p'_j \ge p_{\min}$. Given the value of $p_{\min}$, this shows also positivity.
\end{proof}
\paragraph{Best-response with lookahead (BRL) beliefs}
In these dynamics, each seller best-responds to
a belief (a mental model) of what the other sellers plan to do.
We already discussed a general framework of forming
beliefs. Now that we have defined best-response updates,
Lemma~\ref{lm: belief-based updates} immediately implies
the following corollary.
\begin{corollary}\label{cor: iterative best response}
Let $p_{\min},p_{\max}$ be as stipulated by
Lemma~\ref{lm: best response properties}.
Suppose that a price update $F_j$ that seller $j$ applies
to a price vector $p\in [p_{\min},p_{\max}]^n$ is a
best-response (as defined in this section) to a belief
$\pi = \pi^{\iota}(p)$ that
was generated by a mental model $\iota\in {\cal M}^{n-1}$
(using best-response as defined in this section).
Then $F_j$ is monotone, strictly sub-homogeneous, and
$[p_{\min},p_{\max}]$-price-bounded.
\end{corollary}
\section{Synchronous Dynamics}
\newcommand{\br}[1]{\left({#1}\right)}
\newcommand{\norm}[1]{\left\| {#1} \right\|}\newcommand{\abr}[1]{\left | {#1} \right |}
\newcommand{\expb}[1]{\exp\left({#1} \right)}
Our main tool for proving convergence of BRL
dynamics is the following theorem. Before stating the
theorem, we require a definition.
\begin{definition}
Consider the set ${\mathbb{R}}_{++}^n\subset {\mathbb{R}}^n$ of vectors
with strictly positive coordinates. The
Thompson metric $d$ on ${\mathbb{R}}_{++}^n$ (see~\cite{NonlinearPerron})
is defined as follows.
For $x,y\in{\mathbb{R}}_{++}^n$,
\[
d(x,y) =\max_i \left| \log \frac{x_i}{y_i} \right| =
\| \log x - \log y\|_\infty,\]
where $\log x$ means the vector of logarithms of the entries of $x$.
\end{definition}
The following Lemma shows how this metric is related to the standard $\ell_2$ and $\ell_\infty$ metrics.
\begin{lemma}\label{lm:MetricComparison}
Let $p^a,p^b \in [p_{\min},p_{\max}]^n$ with $0 < p_{\min} < p_{\max}$. Then, we have
\begin{align*}
& \norm{p^a-p^b}_{\infty}\leq \frac{\br{p_{\max}}^2}{p_{\min}}d\br{p^a,p^b} \\
& \norm{p^a-p^b}_{2}\leq \sqrt{n}\frac{\br{p_{\max}}^2}{p_{\min}}d\br{p^a,p^b}
\end{align*}
\end{lemma}
\begin{proof}
For each $i$, we have
\[\abr{p^a_i-p^b_i} \leq p_{\max}\abr{\frac{p^a_i}{p^b_i}-1}\leq p_{\max} \br{\expb{d\br{p^a,p^b}}-1}\]
The function $f\br{t}=\expb{t}-1-\kappa t$ is non-increasing on the interval $[0,\log\br{\kappa}]$ for every $\kappa>0$ and evaluates to $0$ at $t=0$. Hence $\expb{t}-1 \leq \kappa t$ for every $t \in [0,\log\br{\kappa}]$.
Since $d\br{p^a,p^b}\leq \log\br{\frac{p_{\max}}{p_{\min}}}$, we can choose $\kappa=\frac{p_{\max}}{p_{\min}}$ and conclude that
\[\expb{d\br{p^a,p^b}}-1 \leq \frac{p_{\max}}{p_{\min}} d\br{p^a,p^b}\]
Thus, we have
\[\abr{p^a_i-p^b_i} \leq p_{\max}\frac{p_{\max}}{p_{\min}}d\br{p^a,p^b}=\frac{\br{p_{\max}}^2}{p_{\min}}d\br{p^a,p^b}\]
Since the bound holds for each $i$, it holds for the $\infty$ norm as well. The $2$-norm bound simply uses the fact that the $2$ norm is at most $\sqrt{n}$ times the infinity norm.
\end{proof}
\begin{theorem}\label{thm: contraction}
Let $0 < p_{\min}\le p_{\max} < \infty$.
If $F:[p_{\min},p_{\max}]^n\rightarrow [p_{\min},p_{\max}]^n$ is
monotone and strictly sub-homogeneous,
then it is a contraction
with respect to the Thompson metric $d$.
Also, if we require merely sub-homogeneity,
rather than strict sub-homogeneity,
then $F$ is non-expanding with respect to $d$.
\end{theorem}
\begin{proof}
By assumption $F$ is $[p_{\min},p_{\max}]$-price-bounded.
Fix $p,q\in [p_{\min},p_{\max}]^n$. Let $\eta = e^{d(p,q)}$.
Then, we have $p \geq \frac{1}{\eta}\cdot q$ and $q\geq \frac{1}{\eta}\cdot p$.
We get that
$$
F(p) \geq F\left(\frac{1}{\eta}\cdot q\right) > \frac{1}{\eta}\cdot F(q),
$$
where the first inequality uses monotonicity and the second
inequality uses strict sub-homogeneity. Similarly,
$$
F(q) > \frac{1}{\eta}\cdot F(p).
$$
Thus, $d(F(p),F(q)) < \log \eta = d(p,q)$ for all $p,q\in [p_{\min},p_{\max}]^n$.
Define
$$
h(\xi) = \sup\{d(F(p),F(q)) - \xi\cdot d(p,q):\ p,q\in [p_{\min},p_{\max}]^n\}.
$$
Since $[p_{\min},p_{\max}]^n$ is a compact set, $h(1) < 0$. As $h$ is continuous,
there exists $\xi\in (0,1)$ such that $h(\xi) < 0$. Thus, $F$ is a contraction
mapping with a contraction constant $\xi < 1$.
Finally, if we replace strict sub-homogeneity by sub-homogeneity,
then all the strict inequalities above become weak inequalities,
so we get that $d(F(p),F(q))\le d(p,q)$, as stipulated.
\end{proof}
By Lemma~\ref{lm: best response properties}, we immediately
get the following corollary.
\begin{corollary}\label{cor: best response contraction}
An update that consists of all sellers best-responding
to the current price $p$ is a contraction. We denote
its contraction constant by $\xi_{\best} < 1$.
\end{corollary}
We say that a set ${\cal F}$ of price vector updates is {\em contracting}
if all its elements are contractions and there is a uniform upper
bound $<1$ on the contraction constants. We have the following
corollary of Theorem~\ref{thm: contraction}.
\begin{corollary}\label{cor: belief based best response contraction}
Let ${\cal F}$ be a set of price vector updates where each
$F\in {\cal F}$ is generated by all the sellers forming beliefs
that satisfy the conditions of Lemma~\ref{lm: belief-based updates},
then best-responding to those beliefs. Then, ${\cal F}$ is contracting, with
uniform bound $\le \xi_{\best}$.
\end{corollary}
\begin{proof}
Clearly, combining Lemmas~\ref{lm: belief-based updates}
and~\ref{lm: best response properties} with Theorem~\ref{thm: contraction}
proves that every $F\in {\cal F}$ is a contraction. So in order to
complete the proof, we need to show that the contraction constant
is at most $\xi_{\best}$. Let $B_j$ denote the best-response
price update of seller $j$. Notice that by the definition of $\xi_{\best}$,
for every $p,q\in [p_{\min},p_{\max}]^n$, $p\ne q$, and for every
$i,j\in [n]$,
$$
\frac{|\log(B_j(p) / B_j(q))|}{|\log(p_i/q_i)|}\le\xi_{\best}.
$$
Let $\pi^j$ denote the belief (a mapping between price vectors) that
is used by $j$ in the update $F_j$. I.e., $F_j(p) = B_j(\pi^j(p))$. Notice
that by our assumptions on $\pi^j$ and Theorem~\ref{thm: contraction},
$\pi^j$ is non-expanding. In particular, for every
$p,q\in [p_{\min},p_{\max}]^n$, $p\ne q$, and for every
$i,k\in [n]$,
$$
\frac{|\log(\pi^j_k(p) / \pi^j_k(q))|}{|\log(p_i/q_i)|}\le 1.
$$
Thus, we have
that for every $p,q\in [p_{\min},p_{\max}]^n$, $p\ne q$, and for every
$i,j,k\in [n]$,
$$
\frac{|\log(F_j(p) / F_j(q))|}{|\log(p_i/q_i)|} =
\frac{|\log(B_j(\pi^j(p)) / B_j(\pi^j(q)))|}{|\log(\pi^j_k(p)/\pi^j_k(q))|}\cdot
\frac{|\log(\pi^j_k(p) / \pi^j_k(q))|}{|\log(p_i/q_i)|}\le \xi_{\best}\cdot 1 = \xi_{\best},
$$
which completes the proof.
\end{proof}
We need an extra property to guarantee convergence
to equilibrium.
\begin{definition}[local stability]
We say that a set of price vector updates ${\cal F}$
is {\em stable} if the following holds for every $F\in {\cal F}$:
If $p^\ast$ is a vector of equilibrium prices, then $F(p^\ast) = p^\ast$.
\end{definition}
Our main result is the following convergence theorem.
\begin{theorem}\label{thm: synchronous main}
Fix $p_{\min},p_{\max}$.
Let ${\cal F}$ be a contracting and stable set of price
vector updates, where all $F\in {\cal F}$
are monotone, strictly sub-homogeneous, and
$[p_{\min},p_{\max}]$-price-bounded.
Consider the dynamic $p^{t+1} = F^t(p^t)$,
where the choice of $F^t \in {\cal F}$ is arbitrary.\footnote{In
particular, if ${\cal F}$ is a product set, then
for all $j$, $F^t_j$ can be chosen arbitrarily by seller
$j$. Also, the choices can depend on the entire
history of the process, including, but not limited to, the
current prices.}
Then, with initial price vector $p^0\in [p_{\min},p_{\max}]^n$,
the dynamic converges to an equilibrium point (which must
be unique in this case). Moreover, the rate of convergence
is linear (i.e., the distance to equilibrium decays
exponentially fast in the number of time steps).
\end{theorem}
\begin{proof}
Let $d$ be the Thompson metric on $[p_{\min},p_{\max}]^n$.
Consider an equilibrium point $p^*$. By the stability of
${\cal F}$, for all $t$, $F^t(p^*) = p^*$. By the fact that ${\cal F}$
is contracting, there exists $\xi_{\max} < 1$ such that for
all $t$, $F^t$ is a $\xi_{\max}$-contraction.
Therefore,
$$
d(p^{t+1},p^*) = d(F^t(p^t),F^t(p^*))\le
\xi_{\max}\cdot d(p^t,p^*).
$$
Inductively, for every $T\ge 0$,
$$
d(p^T,p^*)\le \left(\xi_{\max}\right)^T d(p^0,p^*).
$$
This completes the proof.
\end{proof}
\begin{corollary}
If the buyer utilities are CES with $\rho > 0$, then
synchronous best-response dynamics, as well as BRL
dynamics that satisfy the conditions of
Corollary~\ref{cor: iterative best response}, both converge
to equilibrium at a linear rate.
\end{corollary}
\begin{corollary}
Let $\mathcal{F}$ satisfy the assumptions of theorem \ref{thm: synchronous main}.
Let $p^\ast$ denote the vector of equilibrium prices and $p^0$ the initial price vector. After $T$ steps of synchronous price updates, let $p^T$ denote the resulting price vector. Then,
\[{\| p^T-p^\ast \|}_{\infty} = \max_i |p_i - p^\ast_i| \leq \frac{\br{p_{\max}}^2}{p_{\min}} d\br{p^0,p^{\ast}}\br{\zeta_{\max}}^T\]
\[{\| p^T-p^\ast \|}_{2} \leq \sqrt{n} \frac{\br{p_{\max}}^2}{p_{\min}} d\br{p^0,p^{\ast}}\br{\zeta_{\max}}^T\]
\end{corollary}
\begin{proof}
Apply lemma \ref{lm:MetricComparison} to theorem \ref{thm: synchronous main}.
\end{proof}
\section{Asynchronous Dynamics}
We consider dynamics where each seller updates its price
at its own varying rate. Thus, at any given time, only a subset
of the sellers update their price. Adopting the notation from
the previous section, we can formally define the dynamic
by allowing some, but not all, of the coordinates of the
price vector updates $F^t$ to be the identity map.
The other coordinates are required to satisfy the same
conditions that are stated in Theorem~\ref{thm: synchronous main}.
We further require that no seller stays put with no update forever.
Thus we can partition the time line into epochs.
An epoch ends when all the prices are updated at least once,
and a new epoch begins in the next time step.
\begin{theorem}\label{thm: asynchronous main}
Under the assumptions stated above, the dynamic, starting
at initial state $p^0\in [p_{\min},p_{\max}]^n$,
converges to the unique equilibrium point. The rate of
convergence, measured by the number of epochs, is
linear, i.e., the distance to
equilibrium decays exponentially fast in the number
of epochs.
\end{theorem}
\begin{proof}
In an epoch, we can think of the last update of
each seller as being applied to the price vector in the
beginning of the epoch, and based on a belief that takes
into account all the previous updates in the epoch. So
replace the asynchronous process by a synchronous
process where time steps are epochs and price updates
map the prices in the beginning of an epoch to the
prices at the end of an epoch. The claim
follows by applying Lemma~\ref{lm: belief-based updates}
and Theorem~\ref{thm: synchronous main}.
\end{proof}
\begin{corollary}
If the buyer utilities are CES with $\rho > 0$, then
asynchronous best-response dynamics, as well as
asynchronous BRL dynamics, both converge
to equilibrium at a linear rate, when measured against
the number of epochs. By lemma \ref{lm:MetricComparison}, the linear convergence holds in the Thompson metric $d$, the $\ell_2$ and the $\ell_\infty$ metrics.
\end{corollary}
|
\section{Introduction}
Despite centuries of study, many of the fundamentals of lightning physics are poorly-understood. Much recent study has been devoted to measurements such as return stroke peak currents (e.g. \citet{Schoene2010}) or impulsive charge moment changes (e.g. \citet{cummer2004lcm}), and such descriptive results are extremely useful. A comprehensive review including references to such descriptive results can be found in Chapters 4, 5, and 9 of \citet{Rakov2007}. However, such descriptions do not necessarily help understand the fundamental physics of lightning. The difficulty lies in capturing the fundamental physics, for example electron detachment and attachment rates, within a framework capable of reproducing lightning behavior. Doing so involves a cascade of physical scales ranging from sub-millimeter-scale electron avalanches governed by micro-physical processes with measurable cross sections to 100 km-scale plasma channels governed by complex aggregate physical properties such as conductivity and charge density. The general goal of this work is to describe and apply a large-scale simulation technique to connect the observed properties of lightning to physical properties of the channel that can be directly compared to existing understanding of the micro-physics. The work here will not completely bridge the gap, but the techniques described herein can be broadly applied to many problems of lightning research.
The specific process considered in this paper is the step-wise extension of the lightning channel. Though the exact mechanism is not known, the lightning leader channel tends to grow in length by impulsive steps whereby the channel lengthens suddenly, jumping forward into space ahead of the existing channel (see Chapter 4 of \citet{Rakov2007} for an overview). Such steps range in length from tens to hundreds of meters and occur on timescales of order 10 $\mu$s. One possible mechanism of such stepping is gradual heating of the gas near the channel tip by corona and streamer discharges. This heating eventually reaches the point of thermally ionizing the gas, drastically increasing the conductivity. An increase in conductivity results in an increase in current which acts to further heat the gas, further increasing the conductivity in a positive feedback effect, rapidly forming a new segment of conducting channel. This development of the new segment of channel may occur slightly displaced from the end of the existing channel to form a disconnected channel dubbed the ``space stem.'' Recent observations include \citet{Hill2011}, who describe high-speed video observations of this stepping process and report the presence of such space stems ahead of the leader channel, while \citet{Winn2011} report balloon-borne electric field observations associated with lightning mapping data that are not completely consistent with such a connection process. As such, at present the details of the stepping mechanism are not understood. Once charges flow onto the narrow newly-heated channel, the charge density results in an outward electric force that drives excess charge outward to fill a ``corona sheath'' surrounding the channel. This outward motion allows continued current flow onto the new segment, so the overall electrodynamics consist of a rapid increase in current as the new segment heats followed by a slower decrease in current as the corona sheath fills with charge. These current and charge motions can be detected by electric field change meters as short pulses.
The goal of this paper is to understand and reproduce such pulses from preliminary breakdown in natural lightning as detected by the Huntsville Alabama Marx Meter Array (HAMMA) \citep{bitzer2013characterization}. Preliminary breakdown here refers to electromagnetic emissions from natural lightning prior to the first return stroke. These emissions are produced by a lightning channel growing within the cloud before it reaches the ground. We approach this goal by self-consistently simulating the electrodynamics of the extension of an evolving channel with approximate treatment of channel plasma behavior. Our work thus differs from most lightning modeling work, which most often either focuses on the return stroke where the channel is already highly-conducting or loses self-consistency by driving the channel with an assumed current source. A review of return stroke literature can be found in \citet{baba2007emo}. Existing stepped leader and preliminary breakdown models take a variety of approaches. \citet{Karunarathne2014} approach preliminary breakdown pulses with a variety of modified transmission line models of preexisting channels driven by fitting the parameters of an assumed current source. Note 20 of \citet{Baum1999} treats the leader step as a continuous extension at a given velocity within the framework of a segmented nonlinear transmission line model. \citet{Gallimberti2002} characterizes each part of the system (leader, corona, space stem, etc.) and determines the time evolution of the characteristics as corona starts and stops and as the leader extends. \citet{Kumar2000} describes the leader as a quasi-static system of charges and currents exhibiting RLC circuit behavior. \citet{Larigaldie1992} considers the time domain electrodynamics and uses a similar treatment as in our approach, but focuses on strikes involving aircraft. Our approach also distinctly differs from the quasi-static models of \citet{Niemeyer1984,Mansell2005,Riousset2007,krehbiel2008ued}, which disregard time evolution and therefore cannot determine current pulse shapes, electromagnetic wave emissions, or the time structure of channel development.
Our simulation, described in Section~\ref{sect:simulations}, is based directly on Maxwell's equations, explicitly includes time evolution of electric charge and current, and retains the dynamics of channel heating and charge migration away from the channel. We approximate the details of channel behavior to limit the model complexity and the number of free parameters. The remaining parameters are all physical properties of lightning channel behavior such as specific heat or the timescale for charge migration to the corona sheath. The dependence of these results on the parameters of the simulation is described in Section~\ref{sect:params}. We then use our simulation to predict possible electromagnetic emissions from stepwise channel extension in preliminary breakdown as described in Section~\ref{sect:results}. The simulation results are compared with observations in Section~\ref{sect:comparison}. The results of the comparison are used to suggest processes not properly captured in the simulation. We then suggest future studies and conclude in Section~\ref{sect:discussion}.
\section{Simulations}
\label{sect:simulations}
The simulation technique described in this paper, which we dub time-domain fractal lightning (TDFL) modelling, is an electrodynamic simulation of charge and current flow on a narrow branched conducting channel embedded in 3-dimensional space capable of reproducing fractal lightning geometry. (For an alternative approach to TDFL modelling as used to describe return strokes, see \citet{Liang2014}.) The simulation acts on the assumption that the electric field is the dominant driver of the electrodynamic behavior of lightning. This assumption is justified by the weakness of magnetic forces relative to electric forces, especially when charge imbalances are present as is the case for lightning. The simulation is described in detail as follows, but overall proceeds as a series of time-steps. In each step, the history of the channel is used to determine the electric field. The electric field is then used to determine the current evolution during the time-step. The current then determines how charge distributions evolve during the time-step. The charges and currents thus determined are then recorded, and the process is repeated to determine the time evolution of the system. This scheme is implemented as a method of moments solution to the electric field integral equation with the thin wire approximation and marching on in time, largely following and extending the method described in \citet{Miller1973}.
More specifically, in general electric fields can be calculated using the retarded-time electric field integral equation (EFIE, equation 6.55 in \citet{Jackson1999}):
\begin{equation}
\mathbf{E_\mathrm{t}}(\mathbf{x},t) = \frac{1}{4 \pi \epsilon_0} \int d^3 x' \left \{ \frac{\mathbf{\hat{R}}}{R^2}\left[ \rho(\mathbf{x'},t') \right]_{\mathrm{ret}} + \frac{\mathbf{\hat{R}}}{cR} \left[\frac{\partial \rho(\mathbf{x'},t')}{\partial t'} \right]_{\mathrm{ret}} - \frac{1}{c^2 R} \left[\frac{\partial \mathbf{J}(\mathbf{x'},t')}{\partial t'} \right]_{\mathrm{ret}} \right \}
\end{equation}
where $\mathbf{E_\mathrm{t}}$ is the total electric field, $\mathbf{x}$ is the observation point, $t$ is the observation time, $\mathbf{x'}$ is the source point, $d^3x'$ signifies integration over sources at points in (three dimensional) space, $\mathbf{R} = \mathbf{x}-\mathbf{x'}$ is a vector from source to observation location, $R = |\mathbf{R}|$, $\mathbf{\hat{R}} = \mathbf{R}/R$ is the corresponding unit vector, $t' = t - R/c$ is the (retarded) time at the source, $\epsilon_0$ is the permittivity of free space, $c$ is the speed of light, $\rho$ is charge density, $\mathbf{J}$ is current density, and $[...]_\mathbf{ret}$ emphasizes the evaluation at retarded time $t'$. The EFIE is a Green's function solution to the full set of Maxwell's equations, where the first term $\propto 1/R^2$ is the familiar static electric field while the last two terms derive from time derivatives of magnetic fields and contribute inductive and radiative effects. The use of a full time-domain solution to Maxwell's equations allows us to treat current and charge on the lightning channel as functions of time, in contrast to solution techniques based on the Poisson equation.
The EFIE must be integrated over all space, but since we wish to focus our attention on lightning, we treat charges not directly associated with the lightning channel as an external static applied electric field, $\mathbf{E_\mathrm{a}}$, which adds to the lightning electric field $\mathbf{E_\mathrm{l}}$ to give the total electric field $\mathbf{E_\mathrm{t}}$. $\mathbf{E_\mathrm{a}}$ is treated as an input to the simulation that encompasses the static effects of thunderstorm charge centers, screening charge layers, and net charges in the ground beneath the storm. Since in this work we consider channels and timescales that are short compared to the scale of variability of cloud charge density, $\mathbf{E_\mathrm{a}}$ is taken to be constant.
The electric field $\mathbf{E_\mathrm{l}}$ associated with the lightning channel is where we must apply the EFIE, observing that the lightning channel is effectively a long narrow structure. Partially integrating the EFIE over the cross sectional area of the channel converts the three dimensional volume integral into a one-dimensional integral along the channel:
\begin{equation}
\mathbf{E_\mathrm{l}}(\mathbf{x},t) = \frac{1}{4 \pi \epsilon_0} \int ds' \left \{ \frac{\mathbf{\hat{R}}}{R^2}\left[ \lambda(s',t') \right]_{\mathrm{ret}} + \frac{\mathbf{\hat{R}}}{cR} \left[\frac{\partial \lambda(s',t')}{\partial t'} \right]_{\mathrm{ret}} - \frac{1}{c^2 R} \left[\frac{\partial \mathbf{I}(s',t')}{\partial t'} \right]_{\mathrm{ret}} \right \}
\label{eqn:efie1d}
\end{equation}
where volume charge density $\rho$ has become linear charge density $\lambda$, vector current density $\mathbf{J}$ has become a vector total current $\mathbf{I}$ directed along the channel, and the three degrees of freedom of the vector $\mathbf{x'}$ are now represented as a scalar length coordinate $s'$ specifying position along the channel.
\begin{figure*}
\begin{center}
\includegraphics[width=6.0in]{seggeom.pdf}
\end{center}
\caption{(a) Channel discretization and representation geometry. The lightning channel is divided into straight current segments that connect charges, where charges are groups of straight segments and are surrounded by corona sheath segments with the same basic geometry as the associated charge segments. (b) Simulation geometry. The simulations in this paper largely consider a straight 1 km vertical channel placed 5 km above a perfectly conducting ground with receivers (``Rec'' in the figure) placed at ground level to record the vertical electric field.}
\label{fig:seggeom}
\end{figure*}
This one-dimensional integral version of the EFIE is unfortunately still not mathematically tractable, so we treat the EFIE numerically with the method of moments \citep{rao1999time} by dividing the channel into current and charge segments. Each current segment flows into a charge and out of a charge. Charges are represented as groups of straight segments partially overlapping the connected current segments as shown in Figure~\ref{fig:seggeom}a. Current and charge density are assumed to be uniform over the charge and current segments and thus piecewise uniform over the channel. This assumption means that the 1-d EFIE can be further simplified, separating the integral into a sum of many relatively simple sub-integrals, one over each charge or current segment. For example, consider the electrostatic term in the EFIE:
\begin{equation}
\frac{1}{4 \pi \epsilon_0} \int ds' \frac{\mathbf{\hat{R}}}{R^2}\left[ \lambda(s',t') \right]_{\mathrm{ret}} = \frac{1}{4\pi\epsilon_0} \sum_i \int_i ds' \frac{\mathbf{\hat{R}}}{R^2} \left[ \lambda_i(t') \right]_{\mathrm{ret}}
\end{equation}
where $i$ is an index for the set of charge segments, $\int_i$ represents integration over segment $i$, and $\lambda_i(t')$ is the charge density of segment $i$, which is uniform and thus only a function of $s'$ through the effect of $s'$ on the retarded time $t'$. We further simplify this by assuming the segment is short compared to the speed of light timescale of the processes to be captured by the simulation, so the effects of retarded time do not vary significantly over a segment. This means the retarded time charge density can be factored out of the integral:
\begin{equation}
\frac{1}{4\pi\epsilon_0} \sum_i \left[ \lambda_i(t') \right]_{\mathrm{ret}} \int_i ds' \frac{\mathbf{\hat{R}}}{R^2}
\end{equation}
leaving the integral solely to treat geometric effects. Applying this process to the $\partial\lambda/\partial t'$ and $\partial\mathbf{I}/\partial t'$ terms, the EFIE in full is:
\begin{equation}
\mathbf{E_\mathrm{l}}(\mathbf{x},t) = \frac{1}{4\pi\epsilon_0}\left\{
\sum_i\left(\left[ \lambda_i(t')\right]_\mathrm{ret} \int_i ds' \frac{\mathbf{\hat{R}}}{R^2}\right)
+ \sum_i\left(\left[ \frac{\partial \lambda_i(t')}{\partial t'}\right]_\mathrm{ret} \int_i ds' \frac{\mathbf{\hat{R}}}{cR}\right)
- \sum_i\left(\left[ \frac{\partial I_i(t')}{\partial t'}\right]_\mathrm{ret}\mathbf{\hat{I_i}} \int_i \frac{ds'}{c^2R}\right)
\right\}
\end{equation}
where $\mathbf{\hat{I_i}}$ is a unit vector capturing the direction of current segment $i$. Note that a system with $N$ current segments has $N+1$ charges, so the first two sums over $i$ run up to $N+1$ while the last sum only runs up to $N$. Assuming the lightning channel does not physically move significantly during a discharge, the geometric factors as calculated by the integrals above do not evolve with time. As such, they can be calculated once and treated simply as constants in the simulation, leaving the resulting equation simply a sum of geometric factors (constants) multiplied by charges and currents, a linear equation.
The time evolution of the system is also discretized: currents and charges are recorded at a set of times $t_j$ separated by a time-step $\delta t$ at locations represented by the center point of the segment in question, see Figure~\ref{fig:timing}. Time evolution of channel current is assumed to be piecewise linear (piecewise constant current time derivative), leading to quadratic time-variation in charge density. The interpolation scheme is implemented as a set of basis functions (e.g. triangle functions for linear interpolation) whereby the desired quantity at a desired time (one of the filled squares in Figure~\ref{fig:timing}) is determined as a sum of the values in the grid before and after the point in question (i.e. those points connected by a dotted line with the filled square in question) multiplied by the appropriate interpolation basis function. Such interpolation basis functions can capture the time derivative behavior as well, and the resulting scheme is purely linear, meaning the contribution to the electric field of each point in the space/time grid is simply proportional to the value at that grid point.
\begin{figure}
\includegraphics[width=3.5in]{interpScheme.pdf}
\caption{Time and space discretization, interpolation, and retarded time integration. The current and charge segment geometry is shown at the top, aligned with the space/time grid points where charge and current values are recorded shown below (filled green and black circles represent known history of charge and current respectively, open green and black circles represent unknown charge and current values, respectively, that are needed for time evolution). An example EFIE retarded time integral to calculate the electric field at the center of current segment 2 is shown in grey (labeled ``light cone''), as are the points where interpolated charge and current values are used (open squares). Slices through this space/time grid to demonstrate the interpolation scheme used are shown at the bottom.}
\label{fig:timing}
\end{figure}
The assumption that the lightning channel does not move significantly during a discharge also means the interpolation basis function evaluations necessary to apply the EFIE can also be effectively represented as a set of constants, fully reducing the EFIE to its form used here:
\begin{equation}
\mathbf{E_\mathrm{l}}(\mathbf{x},t_n) = \frac{1}{4\pi\epsilon_0}\left\{ \sum_i \left[ \left(\sum_j^n \lambda_i^j \alpha_i^j \mathbf{G}_{r^2}^i\right) + \left(\sum_j^n \lambda_i^j \beta_i^j \mathbf{G}_{rc}^i\right)\right] - \sum_i \left[\sum_j^n I_i^j \gamma_i^j \mathbf{G}_{rc^2}^i\right]\right\}
\end{equation}
Where the first and second sums over $i$ replace the integral over the charges and currents in the channel respectively, the sum over $j$ carries out the sum over multiple grid times necessary to account for time interpolation to the desired non-grid times needed in the retarded time integral, $\mathbf{G}_{r^2}^i$, $\mathbf{G}_{rc}^i$, and $\mathbf{G}_{rc^2}^i$ are the geometric factors resulting from the spatial integral over segment $i$ as seen by an observer at point $\mathbf{x}$ (the subscript represents the denominator of the corresponding term in equation \ref{eqn:efie1d}), while $\alpha_i^j$, $\beta_i^j$, and $\gamma_i^j$ are the interpolation basis function evaluations necessary to find the contribution from charge, charge time derivative, and current time derivative respectively for segment $i$ at time-step $j$ as seen by an observer at point $\mathbf{x}$ at time $t_n$. Note that due to the retarded time structure of the system, the $\alpha_i^j$, $\beta_i^j$, $\gamma_i^j$ factors are dependent on the position of the observation point so the above expression is specific to the value of $\mathbf{x}$, and also that most of the $\alpha_i^j$, $\beta_i^j$, $\gamma_i^j$ are zero, so in practice the computational complexity of this expression is that of a single sum instead of two nested sums.
The two additional ingredients needed to complete the simulation are Ohm's law and a method to determine the time evolution of channel current. Ohm's law, $\mathbf{J} = \sigma\mathbf{E}_t$, relates current density and total electric field given a conductivity $\sigma$. Taken over the cross sectional area of a narrow channel assuming current flows along the channel only and the electric field is approximately constant across the cross section, Ohm's law becomes $I = \sigma A \mathbf{E_\mathrm{t}}\cdot\mathbf{\hat{s}}$ giving the scalar current at the point in question in terms of the dot product of the total electric field, the cross sectional area $A$, and $\mathbf{\hat{s}}$, a unit vector giving the direction of the channel at the point in question. Note that the resistance per unit length $R_l = 1/(\sigma A)$.
This suggests a method to determine currents in a given segment, supposing the first $n-1$ time-steps of the simulation are complete. Apply the EFIE to determine the component of the electric field along the channel at the location of the given segment (segment $k$) for time-step $n$ and apply Ohm's law to determine the current. This overall gives a linear equation:
\begin{equation}
I_k^n R_l = \frac{1}{4\pi\epsilon_0} \left\{\sum_i \left[ \left(\sum_j^n \lambda_i^j \alpha_{ik}^j G_{r^2\,k}^i\right) + \left(\sum_j^n \lambda_i^j \beta_{ik}^j G_{rc\,k}^i\right)\right] - \sum_i \left[\sum_j^n I_i^j \gamma_{ik}^j G_{rc^2\,k}^i\right]\right\}
\label{eqn:efieReduced}
\end{equation}
Where the subscript $k$ indexes the segment of interest and the $G_{\cdots}^{\cdots}$ are now scalar contributions to the electric field component along segment $k$ instead of vectors giving the full electric field. Note first that most of the terms on the right hand side are contributions from the charge and current in the past to the electric field in the present. However, the sum over $j$ representing the sum over history runs all the way up to and includes time-step $n$, i.e. including influences from the recent past, nearby segments, and the contribution of a segment to itself. This implies that the unknown $I_k^n$ appears on both the left- and right-hand sides of the equation (as in Figure~\ref{fig:timing} the integration for $E_2$ required $I_2^n$ for interpolation), and that the equation for $I_k^n$ will also involve nearby unknown currents (as seen in Figure~\ref{fig:timing} the integration for $E_2$ explicitly required $I_1$ and $I_3$ for interpolation). Note finally that the contributions required from interpolations involving unknown charges can be expressed in terms of contributions from known past charges interpolated up to time-step $n$ by use of the unknown currents, adding further contributions from various unknown $I_i^n$. These complications pose no serious problems, and in a system with $N$ current segments ($N+1$ charge segments), repeating this process for all current segments provides a system of $N$ linear equations in $N$ unknown currents that can easily be rearranged into a form convenient for numerical solution.
One final subtlety is the influence of a segment on itself. Since the channel is effectively one-dimensional, the $1/R$ and $1/R^2$ terms in the geometric factors $G_{\cdots\,k}^k$ described above will diverge. We solve this problem by making the thin wire approximation, $R \rightarrow \sqrt{R^2 + r_0^2}$ in the denominators of terms of the EFIE, where $r_0$ is the effective radius of the channel \citep{Miller1973}. This approximation is valid if the typical value of $R$ is much greater than that of $r_0$, which it is even for such self-contributions if the segment length is much larger than the channel radius. Since the typical radius of a lightning channel is a few mm \citep{Rakov1998} and segments used to represent a lightning channel are at least several meters long this approximation is clearly justified. We further improve the calculation of geometric factors $G_{\cdots\,k}^i$ (i.e. those needed to calculate the electric field at segment $k$ due to segment $i$) by calculating the result of the integral over segment $i$ as observed by many points on segment $k$ and computing an average. In our implementation, this averaging is especially important for proper calculation of segment self-contributions (i.e. $G_{\cdots\,k}^k$) which are very important for correct time-evolution features such as current wave speed.
The system of linear equations that results can easily be solved on the computer by matrix techniques. The solution gives the currents at the next time-step. The charge values at the next time-step can be determined by applying charge conservation along the channel and integrating the current flow into and/or out of a given segment over the given time-step in question. The simulation keeps track of this by working internally with charges $Q_i$ instead of charge densities $\lambda_i$ ($Q_i = \lambda_i L_i$ where $L_i$ is the total length of charge segment $i$) and by defining a connectivity matrix $C_{il}$ defined as
\begin{equation}
C_{il} =
\left\{
\begin{array}{lr}
+1 & \text{if } I_l \text{ flows into charge } Q_i \\
-1 & \text{if } I_l \text{ flows out of charge } Q_i \\
0 & \text{otherwise}
\end{array}
\right.
\end{equation}
Though such a matrix is unnecessarily complex for the simple unbranched geometry considered in this work (here $C_{i\ i}=-1$ and $C_{i+1\ i}=+1$ for $i=1\ldots N$), more complicated network connectivity can easily be captured. It is also quite computationally convenient, for example giving the net charge flow into a segment as a matrix-vector product:
\begin{equation}
\frac{dQ_i}{dt} = \sum_l C_{il}I_l
\end{equation}
With this definition of the connectivity matrix, charge conservation and integration forward for over a time-step is straightforward. The piece-wise linear time interpolation for current integrated over a time-step simply becomes an average, giving
\begin{equation}
Q_i^n = Q_i^{n-1} + \frac{1}{2}\left(\sum_l C_{il}I_l^n + \frac{1}{2}\sum_l C_{il}I_l^{n-1}\right)\delta t
\end{equation}
where the first sum gives the net current flow into charge segment $i$ at time $t_n$ and the second gives the same quantity at time $t_{n-1}$. This expression gives time evolution for a simple case that will become more complicated when we add features to the model in the next section. Regardless, these new current and charge values complete a new state of the system to be added to the history. Repeating this procedure marches the system forward in time.
One further advantage of this system is that the time-step size is flexible. For periods during long duration simulations when no short-duration processes are happening, the time-step can be extended, trading time resolution for simulation speed. The system smoothly transitions from a full electrodynamic to a quasi-static simulation. As time-step size increases, more and more nearby unknowns appear in equation \ref{eqn:efieReduced}, and when the time-step is infinitely large, the system of linear equations that results is equivalent to the solving the Poisson equation over the lightning channel.
We have verified our simulation as described above against time-domain results from \citet{Poggio1973} and static results from \citet{jackson2000cdo} with very good agreement. While such time domain method of moments calculations often have high-frequency stability problems, we follow the time averaging scheme described in \citet{smith1990instabilities} to dampen out these high-frequency oscillations. Von Neumann stability analysis of the method shows good stability characteristics provided the time-step is not too short, and stability improves for longer time-steps though resolution is lost.
The simulation itself is written in C with sparse matrix operations from CXSparse \citep{Davis2006} and is written to run in parallel on multi-CPU computers with ScaLAPACK \citep{slug} and MPI.
\subsection{Additional features for leader extension simulation}
The simulation system described above cannot immediately duplicate the features seen in natural lightning step pulses. In particular, the corona sheath must be included in order to reproduce the quantity of charge transfered by a channel. Capturing the details of the radial distribution of charge would vastly increase the computational complexity of the model, so here we simply include the corona sheath as a secondary set of charges, located on top of the main channel charges that enter the EFIE with the same formalism as described above but with a larger effective radius $r_\mathrm{CS}$ in the thin wire approximation as used in computation of geometric factors $G_{\cdots}^{\cdots}$. It is important to note that this radius parameter is not the exact radius of a cylindrical sheath of charge, especially since this radius may become comparable to the segment length and thus leaving the thin wire approximation unjustified, but the corona sheath radius parameter does capture the behavior of a diffuse charge region surrounding the channel with effective size tunable by $r_\mathrm{CS}$.
A corona sheath charge segment is filled with charge from the channel charge segment it encloses, i.e. we assume charge flows only outward or inward from a given channel charge segment to its corona sheath, not longitudinally along the corona sheath. Higher charge density on the main channel charge segment would lead to more rapid transfer, so we further assume the charge migrates outward at a rate proportional to the charge stored on the channel. This effectively means we treat the corona discharge processes surrounding the channel as represented by a constant conductivity; this is not a good assumption but it is convenient: the linearity of the charge transfer process means it can be interpolated in time with the same sort of interpolation basis functions described above. Ignoring charge flow along the channel, the overall result of this approach would be an exponential decrease in channel charge, so we parametrize this process by the characteristic timescale $\tau_\mathrm{CS}$ for charge transfer to the corona sheath. Thus, the time derivative of the charge on corona sheath segment $i$ ($Q_{i\mathrm{CS}}$) is given by
\begin{equation}
\frac{dQ_{i\mathrm{CS}}}{dt} = \frac{Q_i}{\tau_\mathrm{CS}}
\end{equation}
where $Q_i$ is the charge on the corresponding channel segment. The charges in the equation above are functions of time, so time evolution of the system becomes more complicated than described above. Instead of simply integrating the current flow along the channel into or out of a channel charge segment to determine the change in the charge, we must solve a differential equation for the evolution from one timestep ($t_{n-1}$) to the next ($t_n$):
\begin{equation}
\frac{dQ_i}{dt} = -\frac{Q_i}{\tau_\mathrm{CS}} + \mathcal{I}_i^{n-1}\left[1-\frac{t-t_{n-1}}{\delta t}\right] - \mathcal{I}_i^n \left[\frac{t-t_{n-1}}{\delta t}\right]
\end{equation}
where $-\frac{Q_i}{\tau_\mathrm{CS}}$ gives the current flow outward to the corona sheath due to the charge on the segment in question, $\mathcal{I}_i^{n-1} = \sum_l C_{il}I_l^{n-1}$ and $\mathcal{I}_i^n = \sum_l C_{il}I_l^n$ are the net current flow into the charge segment at times $t_n$ and $t_{n-1}$, respectively, in terms of the connectivity matrix $C_{il}$ as discussed previously, and the terms in square brackets are the piecewise linear time interpolation basis functions discussed above appropriate for interpolation between $I_i^{n-1}$ and $I_i^n$ for times between $t_{n-1}$ and $t_n$. The equation is a first-order non-homogeneous ordinary differential equation that is straightforward to solve by variation of parameters. The result, valid for times between $t_{n-1}$ and $t_n$, is
\begin{equation}
Q_i(t) = \left(Q_i^{n-1} - \mathcal{I}_i^{n-1} \tau_\mathrm{CS} + \frac{\mathcal{I}_i^n - \mathcal{I}_i^{n-1}}{\delta t}\tau_\mathrm{CS}^2\right)e^{-\frac{t-t_{n-1}}{\tau_\mathrm{CS}}} + \mathcal{I}_i^{n-1} \tau_\mathrm{CS} + \frac{\mathcal{I}_i^n - \mathcal{I}_i^{n-1}}{\delta t}\left((t-t_{n-1})\tau_\mathrm{CS} - \tau_\mathrm{CS}^2\right)
\end{equation}
Evaluating this solution at $t_n = t_{n-1} + \delta t$ thus determines the evolution of the net charge carried on segment $i$ from time $t_{n-1}$ to time $t_n$. The evolution of charge on the corresponding corona sheath segment can then be computed easily by considering charge conservation on the channel segment in question:
\begin{equation}
Q_{i\mathrm{CS}}^n = Q_{i\mathrm{CS}}^{n-1} + \left(\frac{\mathcal{I}_i^{n-1} + \mathcal{I}_i^n}2\delta t - (Q_i^n - Q_i^{n-1})\right)
\end{equation}
where $\frac{\mathcal{I}_i^{n-1} + \mathcal{I}_i^n}2\delta t$ is the net charge flow along the channel onto the channel charge segment in question during the time-step in question and $Q_i^n - Q_i^{n-1}$ is the net change in charge on the channel charge segment in question. Any imbalance between these terms is due to charge flow onto the corona sheath.
Furthermore, in channel extension simulations, the channel itself must evolve with time. This evolution is determined by heating and cooling processes that also must be included in the simulation in order to reproduce the features of step pulses. Fundamentally, the heating process is Joule heating, with a power per length proportional to $I^2 R_l$, where $R_l$ is the resistance of the channel per unit length. Cooling is determined by a combination of radiative, conductive, and convective cooling. The fundamental physics of heating and cooling of a non-equilibrium plasma channel is very complex, and we have not attempted to capture its nuances here. For a more detailed consideration, see \citet{Liang2014}. Here we simply assume that the conductivity of the channel can be determined by an effective temperature, and that the effective temperature changes according to a effective heat capacity per unit length $C_l$. On timescales shorter than 20~$\mu$s as examined here, cooling is not a major factor \citep[Chapter 6]{Heckman1992}, so we simply have $C_l dT/dt = I^2 R_l$. In the segmented representation of the channel, a temperature is assigned to each current segment and evolved according to the heating at the end of each time-step.
Given a temperature, it remains to calculate the resistance per unit length of the channel, $R_l$. This is another complex topic that we can only qualitatively approximate. The conductivity calculation here is motivated by the Saha equation of ionization equilibrium and results from plasma conductivity studies. The Saha equation gives ratios of various ionization states in terms of in terms of their degeneracies and thermal energy effects, but if the temperature dependence is the only effect of importance, it becomes simply a proportionality $n_e^2 \propto T^{3/2}e^{-\epsilon/k_\mathrm{B} T}$ where $n_e$ is the electron number density, $T$ is the temperature, $k_\mathrm{B}$ is Boltzmann's constant, and $\epsilon \approx 14$~eV is the approximate ionization energy relevant for atomic oxygen or nitrogen. The plasma conductivity adapted to non-ideal plasma conditions as in \citet{Zollweg1987} is typically given in terms of a reduced conductivity, $\sigma^{*}$ as $\sigma \propto T^{3/2} \sigma^{*}$. The reduced conductivity $\sigma^{*}$ is roughly proportional to the square root of a non-ideality parameter, $\sigma^{*} \propto \gamma^{1/2}$ (see Figure~1, \citet{Zollweg1987}), where $\gamma \propto n_e^{1/3}/T$. Combining these proportionalities, we obtain $\sigma \propto T^{9/8} e^{-\epsilon/12 k_\mathrm{B}T}$ and a corresponding resistance per unit length $R_l \propto T^{-9/8} e^{\epsilon/12 k_\mathrm{B}T}$. Comparison to arc conductivity measurements in \citet{Schreiber1973} adjusted to our channel radius give $R_l(10^4 \mathrm{K}) \approx 2.5$ $\Omega$/m sets the proportionality constant. This calculation gives the resistance vs temperature shown in Figure~\ref{fig:rPerL}. This is of course at best an approximate treatment, and changes in the resistance as a function of temperature may make channels heat up more or less quickly which will have an effect on the electromagnetic radiation produced. Regardless, since channel heating is a relatively slow process compared to our time-step size, in each time-step, the current and charge are updated assuming constant temperature and conductivity. The temperature and conductivity are then updated before the next time-step.
\begin{figure}
\includegraphics[width=3.0in]{rPerL_vs_T.pdf}
\caption{The formulation of resistance per unit length of the channel as a function of temperature of the channel used in the simulations.}
\label{fig:rPerL}
\end{figure}
Though approximate, these treatments of corona sheath behavior and temperature-dependence of resistance capture the dominant behavior of the lightning channel.
\subsection{Parameters}
\label{sect:params}
Such a complex simulation naturally has parameters. These parameters break down into 3 main categories: physical constants, computational parameters, and initial conditions. The physical constants include not only true physical constants like the speed of light or the permittivity of free space, but also the physical properties of the channel that do not evolve with the simulation: channel heat capacity and effective radius for example. The initial conditions (e.g. the initial geometry of the channel or the length of the step in question) are discussed later in the context of specific simulations. A full list of the parameters and their values is given in Table~\ref{table:params0}.
\begin{table}
\caption{List of simulation parameters}
\label{table:params0}
\centering
\begin{tabular}{r|l}
\multicolumn{2}{l}{Physical constants} \\ \hline
Channel radius & 3 mm \\
Corona sheath radius & 4 m \\
Corona sheath timescale & 0.5 $\mu$s \\
Channel heat capacity & 2 J/K/m \\
& \\
\multicolumn{2}{l}{Computational parameters} \\ \hline
time-step size & 35 ns \\
Segment length & 5 m \\
& \\
\multicolumn{2}{l}{Initial conditions} \\ \hline
Applied electric field & $\sim 100$ kV/m \\
Channel length & $\sim 1$ km \\
Channel shape & straight \\
Channel orientation & vertical or angled \\
Channel position & above origin\\
Step length & $\sim 100$ m \\
\end{tabular}
\end{table}
The channel radius is taken to be $3$ mm, consistent with \citet{Rakov1998}, though this parameter does not significantly affect the results. The corona sheath radius (i.e. $r_\mathrm{CS}$ in the thin wire approximation) is taken to be $4$ m, consistent estimates from \citet[page 292]{cooray2004lf} based on the distance from the channel for electric fields to decrease below breakdown for typical linear charge densities inferred from stepped leader measurements. The corona sheath timescale depends on the time needed for charge to leave the channel, and is chosen to be $0.5$ $\mu$s as a compromise between the rapid motion of charge away from the channel out to 0.1 m distances on a timescale of 0.1 $\mu$s (i.e. a streamer speed of $\sim 1$ mm/ns \citep{briels2008pan}) and slower migration out to larger radii. The initial charge motion outward has a stronger effect on electric fields than motion to larger radii, so the corona timescale is shorter rather than longer. The heat capacity of the channel of $2$ J/K/m with a channel radius of 3 mm corresponds to a heat capacity of $\sim 50$~J/K/g for air at or below atmospheric pressure, consistent with experimental estimates \citep{D'Angola2007}.
The parameters specific to the simulation, the time-step size and channel segment length, are chosen based on the desired resolution. We seek to resolve processes on shorter than 10 $\mu$s scales, requiring time-steps shorter than 0.1 $\mu$s, so here we use 35~ns. 35~ns time-step interval, given our Von Neumann stability analysis, requires segments at most $\sim10$~m long, so here we use 5~m. We tested a variety of other spatial and temporal resolutions, and the values given above ensure good convergence, retain resolution, and are quite stable.
The simulation technique described above is applicable to many problems in lightning physics. The geometry of the channel is unconstrained, so branched channels with arbitrary shape and connectivity are allowed. The time-step is flexible, allowing efficient simulations both of large-scale channel development and short-duration charge motions. Inclusion of stochastic channel extension motivated by fractal geometry allows for simulations of the full lightning discharge. The radiative terms in the EFIE allow for prediction of electromagnetic emissions from lightning channels. Those advanced features aside, however, we start simple in this paper.
\section{Simulation results}
\label{sect:results}
As the focus of this paper is preliminary breakdown pulses, consider simulations of single steps. The simplest configuration that captures this phenomenon is a single isolated straight channel, with an uncharged non-conductive ``step'' portion at one end as shown in Figure~\ref{fig:seggeom}b. This initial channel is 1~km long, unbranched, and straight, with its bottom end at 5~km altitude. Since the discharge is so far from the ground, the effects of the charges induced in the ground by slowly-varying thunderstorm charges can be included as part of the applied field as described above.
For simulation of a step, since inter-step intervals are relatively long, we assume that the charge distribution along the channel has reached equilibrium prior to the step. We simulate this by initially allowing charge and current to flow on the main conductive channel for a time much longer than that needed for equilibrium to be reached without allowing the step to evolve in any way. Once the main channel has reached equilibrium, its resistance is set to 48~$\Omega$/m to represent an existing active channel and the resistance of the step portion of the channel is set to an initial non-infinite but very large value ($8 \times 10^4$ $\Omega$/m, corresponding to an initial temperature of 1400~K in our conductivity scheme). This artificial heating of the new step to the point where it can further heat itself by current flow hides the details of the near-channel physics that somehow leads to step-wise channel extension. The physics of this process is not well-known, though it presumably includes the effects of electric field-induced ionization, photoionization, corona, and streamer behavior ahead of the existing channel. Regardless of the details, once our crude approximation of the initial temperature increase has been applied, the entire system is allowed to evolve freely. The applied field and the field from charges accumulated on the main channel then drive small currents on the step which gradually heat the step until the positive feedback from rising temperatures and increasing conductivity results in a rapid current pulse from the main channel onto the newly-active step.
Given the large charge accumulation at and near the end of the main channel, the electric field is strongest at points on the step closest to the main channel. Thus, for our geometry, the current and resulting heating are strongest close to the main channel. If allowed to evolve without any additional requirements, the step heats starting closest to the main channel in a process akin to a dart leader. This is logical, but contrary to observations of stepped leaders that, as discussed above, seem to involve space stems and leap forward at velocities faster than those of dart leaders \citep[see speed and duration estimates in][sections 4.4.6 and 4.7.2]{Rakov2007}. Creating a space stem artificially entails careful tuning of the initial resistance over the newly-evolving step. As there is no clear justification for why or how we should accomplish this, we instead take a more blunt approach and enforce a uniform resistance per unit length of the step at each time-step. This is done by calculating the total heating and specific heat of the step channel and calculating the average effects on resistance. Though artificial, this smoothed resistance structure means our results are not tied to any particular ideas about pre-step channel structure. Though the resistance is enforced to be uniform, the current and charge density evolve without any smoothing.
\begin{figure}
\includegraphics[width=7in]{sampleResults_combined.pdf}
\caption{Sample results of a 100~m step simulation. Current (left) and charge (right) are shown vs time at various points (top) and as an image (bottom). The current flow shown at top left is through the channel-step junction. The charge shown at top right includes both channel and sheath charge and is plotted at three locations: the end of the old channel, the start of the step (just past the end of the channel), and the end of the step (the end of the new channel). The images at bottom both show the evolution of the bottom 600~m of the channel, i.e. the bottom half (500~m) of the old channel with the 100~m step at the bottom.}
\label{fig:sampleresults}
\end{figure}
Sample simulation results are shown in Figure~\ref{fig:sampleresults}. The first feature of note is that there are no high-speed propagating features since this process is dominated by resistance and heating that happen on a 10~$\mu$s timescale, much longer than the speed of light propagation time over the region shown. The contours on the current plot (Figure~\ref{fig:sampleresults} bottom left) do suggest some propagating feature moving along the channel, but this is basically a diffusion of charge, not a return-stroke like current pulse. Second, since current and charge evolve on the new step without any smoothing (only temperature and conductivity are smoothed), the current flow on the step first becomes significant near the end of the channel and is highest there. This also explains the kink in the ``step start'' charge density (Figure~\ref{fig:sampleresults} top right): initially this charge density increases before the step effectively turns on, but at around 35~$\mu$s when the current is large, much of it carries charge away from the beginning of the step, decreasing the magnitude of the local charge density before charge flow from elsewhere on the channel catches up. Third, the charge density associated with the end of the channel lingers for a relatively long time. This is expected, even given the relatively short $\tau_\mathrm{CS}$ used in the simulation, since $\tau_\mathrm{CS}$ is the timescale for charge transfer \emph{to} the corona sheath as driven by the focused charges on a given channel segment channel, while charge transfer \emph{from} the corona sheath is driven by the charge on the sheath itself, which exerts its effect only on portions of the channel away from the segment in question and thus is a relatively slow indirect effect. As the extension comes to equilibrium, the linear charge density is consistent with the $\lesssim 1$~mC/m inferred from measurements near lightning leaders \citep[e.g.][]{Winn2011}.
For comparison with data, the resulting currents and charges are used to calculate the electromagnetic fields observed by hypothetical receivers positioned on the ground at various positions near the channel as shown in Figure~\ref{fig:seggeom}b. While the ground is sufficiently far below the channel to neglect the effects of charges induced in the ground by lightning on the lightning itself, this is not true for observers on the ground. Here we simply treat the ground as a perfect conductor and apply the method of images, a reasonable first approach since ground conductivity for earth of $\sim 10^{-3}$ S/m \citep{itu1992} gives a relaxation time of $\sim 10$ ns, shorter than the processes we consider here. The electromagnetic wave radiated by the image superposed with the electromagnetic wave radiated by the channel itself result in valid perfect-conductor boundary conditions at ground level, simply cancelling out the horizontal electric fields at the location of the receiver and doubling the vertical electric field. This can easily be shown by consideration of the geometry of the image charges and currents and working out the vector geometry in the EFIE. Since our applied electric field is only intended as a driver of processes on the channel itself, it does not capture screening charge layers on the cloud or local to the receiver, so we only consider changes in electric field due to charge motions during the step, not the overall field.
\begin{figure}
\includegraphics[width=7in]{tdflAngleDist_combined.pdf}
\caption{Sample electric field change recordings at a variety of positions relative to the discharge for a variety of discharge orientations. The geometry is as shown in Figure~\ref{fig:seggeom}b, but with the top of channel tilted away from the vertical $z$ axis toward the $+x$ axis, pivoting about the junction between channel and step. Each row of plots in the figure corresponds to a different orientation of channel, with the angle marked at right in degrees deviation from vertical as the channel tilts. Each column in the plot shows predicted observations at ground level at locations displaced from the sub-step point in a given direction ($+x$, $+y$, or $-x$). Each curve in the plot shows the predicted field change for a receiver at a distance indicated by the color of the curve. For example, the black curve in the upper left plot shows the observations predicted at a point 2~km in the $+x$ direction from the sub-step point with a vertical channel.}
\label{fig:sampleresultsAngle}
\end{figure}
Samples of such electric field change records are show in Figure~\ref{fig:sampleresultsAngle}. Each panel shows the predicted observations at locations displaced a variety of distances from the sub-step point in a given direction for a discharge oriented at a given angle. The simulations shown use a 100~m step on the bottom of a 1~km channel in a 100~kV/m applied electric field directed along the channel. The signals received depend strongly on the direction of the channel and on the location of the observation point, but some trends are evident. In all panels, observers far from the sub-discharge point observer smaller static field changes, while the radiated pulse (amplitude $\propto 1/R$) remains evident out to long distances. For observers displaced in the $+x$ direction (the left column of plots), the upright channels result in a net transport of negative charge toward the observer (and thus producing an upward-directed static electric field change), while steeply tilted channels move negative charge on average away and positive charge toward the observer (and thus producing a downward-directed static electric field change). No such sign change is present in observation locations displaced along the $-x$ axis (right column), since regardless of angle, the step results in a net transport of negative charge closer to the observer, though such observers see a sign change in the \emph{radiated} signals as the channel tilts and thus hits the observer with signals radiated in different directions relative to the channel orientation. The observations displaced along the $+y$ axis (the middle column, observers for which the channel is neither angled toward nor away from the observer) fall generally in between the corresponding observations for observers displaced along the $+x$ and $-x$ axes. One exception to this general trend is for horizontal channels, where observers along the $+y$ axis are predicted to detect very small DC field changes due to a motion of negative charge on average toward the observer, together with a small transient non-radiated pulse contributed by the $\partial\rho/\partial t$ term in the EFIE.
Clearly, even a single stepwise channel extension event can produce a wide variety of static and radiated electric field changes. We hope at the very least that these results will be useful in qualitative interpretation of data, and with known geometry constraints or the plausible assumption of a vertical channel, such simulations can illuminate quantitative connections with individual pulses.
\subsection{Parameter dependence}
In order to complete our discussion of simulation results, we examine the dependence of the simulation results on the step properties and physical parameters of the model. Throughout our discussion we have given plausibility arguments and citations for parameter values, but these parameters are at least slightly tunable since the values are either not known precisely or appear only as ``effective'' values. Tunable parameters remove some of the predictive power of such a model, but tuning the parameters to match observations provides information about the allowable effective values of the parameters, making a connection between observations and more fundamental processes.
As a way to study the parameter dependence, consider the signals detected by a hypothetical observer a moderate distance from the sub-step point. Provided the channel is not too steeply angled, all observers more than a few km from the sub-step point agree on the shape of the radiated electric field pulse, so this is a useful diagnostic that is somewhat less dependent on the details of the geometry. Sample simulated current and radiated electric field waveforms are shown together with the effects of the most important parameters in Figures \ref{fig:paramsI} and \ref{fig:paramsEZ}. During the step, the step channel heats such that its resistance decreases to 30 -- 70 $\Omega/m$, consistent with $R_l$ of the main channel and thus with growth of the main channel. Overall, the electromagnetic signals produced (Figure \ref{fig:paramsEZ}) have 3 main identifiable features: amplitude, duration, and the relative height of negative and positive excursions (``asymmetry''). For simulations of single lightning leader steps, extensive numerical exploration shows that the most important parameters are the specific heat of the channel, the timescale over which charge migrates to the corona sheath, the step length, and the applied electric field strength which interacts with the channel length. As the specific heat of the channel increases, the duration of the pulse increases, its amplitude decreases, and the pulse becomes more asymmetric. As the corona timescale increases, the pulse amplitude decreases and the pulse becomes more asymmetric, leaving the pulse duration unaffected. Longer steps take longer to heat over their entire length and thus radiate longer duration pulses with similar amplitudes and increased symmetry. Finally, increasing the applied electric field strength is similar to decreasing the channel heat capacity: pulse duration decreases, amplitude increases, and asymmetry decreases. Unfortunately, applied electric field affects results similarly to the length of the pre-existing channel: longer channels lead to greater intensification of the electric field in the region of the step, and the equivalence between non-channel applied electric field and channel electric field makes long channels produce steps essentially identical to shorter channels in stronger applied electric fields.
\begin{figure}
\includegraphics[width=3.0in]{paramsI.pdf}
\caption{Parameter dependence of channel current at the base of the step, showing the effects on the current if the major parameters are increased or decreased as labeled (increase or decrease is by a factor of 2 except for applied field, which is increased and decreased by 25\%).}
\label{fig:paramsI}
\end{figure}
\begin{figure}
\includegraphics[width=3.0in]{paramsEZ.pdf}
\caption{Parameter dependence of received electric field showing the effects on the signal if the major parameters are increased or decreased as labeled (increase or decrease is by a factor of 2 except for applied field, which is increased and decreased by 25\%).}
\label{fig:paramsEZ}
\end{figure}
Physically, these pulse features and their parameter dependence shed light on the physical origin of the pulse features. First, note that the radiated electromagnetic wave comes largely from the $\partial J/\partial t$ term of the EFIE. Strong radiated electric fields thus correspond to rapidly-changing currents. The initial negative excursion comes from a rapid increase in upward current flowing onto the new step (the ``turn-on'' phase), while the smaller positive excursion that immediately follows comes from the decrease in current as the step gradually fills with change (the ``turn-off'' phase). The risetime of the pulse is directly connected to the rise-time of the current, and thus to heating of the channel. The duration of the initial negative excursion is simply the time required for the current to reach its maximum value as the channel heats, which is determined both by the energy necessary to heat the channel (heat capacity) and the amount of energy available (applied field). The maximum current is determined by the length of the step and also by the charge accumulation necessary to counteract the applied field and thus reflects the formation of the corona sheath: rapid sheath formation draws more charge away from the channel and thus leads to higher currents since charge on the corona sheath is less able to counteract an applied field on the channel than charge on the channel itself. Once the current reaches its maximum value, it decreases on a timescale determined by the formation of the corona sheath. With this physical framework in mind, the simulation results can be compared to observation.
\section{Comparison to observation}
\label{sect:comparison}
These simulation results can easily be compared to data collected with the Huntsville Alabama Marx Meter Array (HAMMA). HAMMA is a network of electric field change meters (Marx meters) located in the area surrounding the University of Alabama Huntsville. The electric field change meters have 100 ms time constant and are sampled at 1 MHz with GPS time synchronization. These meters provide high-resolution, high-dynamic-range measurements of electric field changes associated with both slow and fast processes in lightning discharge. GPS time accuracy allows the location of fast processes to be determined by time of arrival fitting. The Alabama Lightning Mapping Array (LMA) \citep{Goodman2005}, a VHF time of arrival lightning mapper, covers the same area.
In this paper, we focus on fast pulses measured during the initial growth of a lightning discharge on October 26, 2010 at 19:04:59 UT. This lightning discharge lasted more than 100 ms, and included multiple K-changes and return strokes, but here we focus on the preliminary breakdown pulses during the growth of the channel just after initiation and prior to the first K-change. A map of the discharge and the preliminary breakdown period in question is shown in Figure~\ref{fig:dataoverview}. HAMMA detector 5 is 4.4~km from the sub-median point, and sees a positive $\Delta E_z$ due to negative charge motion toward the detector, while the other detectors are far enough away that they see a net negative $\Delta E_z$ that can be understood in the context of the curvature of electric field lines of a dipole. For the rest of this paper, we will examine three representative detectors: detector 5 (4.4~km away, very close to the discharge), detector 2 (8.4~km away, moderate distance), and detector 4 (30.8~km, relatively large distance).
\begin{figure}
\includegraphics[width=6.0in]{map_overview.pdf}
\caption{A map of the discharge according to LMA and HAMMA time of arrival (left) and an overview of the HAMMA data for the preliminary breakdown phase (right). At left, LMA points during and after the preliminary breakdown period shown in the data overview are shown as black and grey circles, respectively, while HAMMA time of arrival locations during and after the preliminary breakdown period are shown as black and grey open squares, respectively. The median HAMMA time of arrival position for the preliminary breakdown period is shown as a large $+$, and has an altitude of $\sim 5$~km.}
\label{fig:dataoverview}
\end{figure}
The preliminary breakdown pulses considered in this paper are too short to be resolved on the relatively large timescale of Figure~\ref{fig:dataoverview}. Focusing on the first few milliseconds as shown in Figure~\ref{fig:datazoom}, pulses with a variety of features can be seen. Most pulses in detectors 2 and 4 show the same features as those in the more distant detectors in Figure~\ref{fig:sampleresultsAngle}: a relatively short intense downward excursion followed by a relatively weak and long upward excursion. Many of the pulses in detector 5 also show this pattern, but include a more clear stepwise increase in upward electric field, a feature present in our simulation results in Figure~\ref{fig:sampleresultsAngle} for relatively nearby detectors.
Focusing on the group of three pulses at 740--810~$\mu$s in detector 5, the first pulse shows a relatively small downward excursion associated with a relatively large DC change, while the last shows a relatively large downward excursion with very little DC change. Comparison to our simulation results suggest the first pulse was associated with channel extension directed somewhat but not directly toward detector 5 (see the $-x$, $22.5^\circ$, 4~km curve in Figure~\ref{fig:sampleresultsAngle}), while the third pulse was associated with extension directed more perpendicular to the line of sight from detector to channel (see the $+x$, $22.5^\circ$, 4~km curve). Detector 4, on the opposite side of the discharge as detector 5, is well-placed to test this hypothesis. If the first pulse of the trio was toward detector 5 (matching $-x$ curves), it should have been away from detector 4 (matching $+x$ curves), while if the third pulse of the trio was away from detector 5 (matching $+x$ curves), it should have been toward detector 4 (matching $-x$ curves). Unfortunately, the only visible difference between the $+x$ (away from, hypothetical first pulse) and $-x$ (toward, hypothetical third pulse) curves for detector 4 is amplitude, with the toward ($-x$) curve having a slightly lower amplitude. The third pulse in detector 4 does indeed have a slightly lower amplitude than the first, consistent with the predictions of the model, but the relative amplitude of the simulation results comes from a single simulation, while in the data we are comparing two distinct pulses. Detector 2, however, can address this uncertainty; located approximately perpendicular to the line connecting detector 5, the lightning channel, and detector 4, the symmetry of the situation suggests that whether channel extension is directed toward detector 5 or toward detector 4 should not affect the pulse observed by detector 2, so detector 2 can be used to judge the relative amplitude of pulses as emitted by the channel. Detector 2 sees approximately equal amplitudes, which indicates that the channel extension events responsible for the first and third pulses are of approximately equal intensity. This lends support to the comparison between a single simulation and two pulses seen in detector 4 as described above, and suggests that the amplitude difference between the first and third pulses seen in detectors 3 and 4 can be attributed to the different directions of channel extension relative to detector location. Our interpretation of channel directions and pulse intensities as seen by detectors 4 and 5 as motivated by the pulse shapes seen by detector 5 and supported by detector 2 is thus at least qualitatively self-consistent.
\begin{figure}
\includegraphics[width=6.0in]{hammaData12.pdf}
\caption{Zoomed views of the data from Figure~\ref{fig:dataoverview}, showing the first millisecond of data (left) and a further zoom into the second half of that first millisecond. Data from detectors 2, 4, and 5 are shown as representative.}
\label{fig:datazoom}
\end{figure}
Quantitative consistency requires direct comparison and manual iterative adjustment of simulation initial conditions. The results of such a process for the first and third pulses discussed above are shown in Figure~\ref{fig:simAndData}. The geometry of the simulations is exactly as in Figures~\ref{fig:seggeom}b and \ref{fig:sampleresultsAngle}, with the channel tip placed at 5~km altitude and positioned relative to the detectors as suggested by the median HAMMA source as shown in Figure~\ref{fig:dataoverview}. We slightly adjust the channel direction to attempt to fit the observations, changing the angle from $22.5^\circ$ to $20^\circ$ to better emphasize the initial downward excursion at detector 5. The quantitative agreement is much better for the third pulse than the for the first. This is unsurprising given the fine structure evident in the first pulse of the three; the first pulse is likely due to a more complex process than a simple single step extension, perhaps a superposition of two overlapping extension events as suggested by the two negative excursions visible in detector 2 data. The third pulse as seen in detector 2 is stronger than expected based on the simulation. This suggests the directionality of the channel extension is not as simple as described above. Adjusting the directionality such that detector 2 receives more of the radiated electric field can improve the match, as does moving the simulation channel closer to detector 2, but an automated fit would be required to improve the results significantly, the time required to run such simulations makes this difficult, and the match between a simple simulation and a complicated lightning channel is not expected to be perfect.
\begin{figure}
\includegraphics[width=6in]{trios.pdf}
\caption{Quantitative comparison between simulation results and observations for the pulses discussed geometrically in the text (i.e. those at $\sim 800$ $\mu$s in Figure~\ref{fig:datazoom}), showing at left and right the first and third pulses in the group of three, respectively.}
\label{fig:simAndData}
\end{figure}
The comparison in Figure~\ref{fig:simAndData} does point out some detailed qualitative deviations between simulation results and observations. First, the simulated DC field change that develops by around 25 $\mu$s after the pulse often disagrees with the observations. This suggests either the amount of charge transferred on the simulated step is wrong or there is some other charge transfer occurring elsewhere on the channel that confounds the data. Second, the positive excursion peaks later in the data than in the simulation. For radiated pulses, this indicates that the current begins decreasing more quickly in the simulation than in reality. The fact that the positive excursion is often higher in the data than in the simulation indicates that the maximum rate of decrease of current is higher in reality. Relative to simulation, therefore, in nature, the current that flows onto a new portion of channel increases in roughly the same way, but remains near its peak for longer before turning off more quickly. Such deviations are perhaps unsurprising given our enforced ignorance of the geometry of the heating of the step and our approximate treatment of the corona sheath, but broad qualitative agreement is promising, especially given that the data do not come from a straight channel that comes to equilibrium before extending in a single isolated step, but from active development throughout a dynamic branched channel.
\section{Summary and future work}
\label{sect:discussion}
This paper describes a simulation technique that captures the details of charge and current flow on an evolving lightning channel. The simulation includes approximate treatments of channel resistance evolution due to heating and the migration of charge outward from the channel to the corona sheath. Inclusion of these processes leads to a model capable of reproducing the detailed features of preliminary breakdown pulses as shown in Figure~\ref{fig:simAndData}, lending support to the interpretation of such pulses as from stepwise extensions of an existing channel.
Much work remains to be done, however, as seen both in the deviations between simulation and data and in the fact that we only consider a small portion of the overall evolution of the channel. In this, the mismatch between simulation and observation is encouraging; such mismatch means the results of the simulation are sensitive to the details of the processes at work in a preliminary breakdown pulse, so further study can shed light on such details. For example, the framework described here can be extended to include more detailed treatments of the plasma physics of the channel (see for example \citet{Liang2014}), and the resistance of the new step channel, here forced to be uniform, can be allowed to vary with better initial conditions, perhaps approximating a space leader process. The channel extension process can be simulated further by including more steps and variation of step properties with altitude. On longer timescales, the simulated channel can be allowed to extend and branch stochastically by implementing results from fractal lightning (e.g. \citet{Niemeyer1984,Riousset2007}, justifying our name for the technique as time-domain fractal lightning modeling), and preliminary results show excellent qualitative agreement with longer timescales of channel evolution (e.g. reproducing K-changes), to be described in a subsequent paper. Finally, though the simulation reproduces many features of lightning electric fields, there are still features that are difficult to explain, like the unusually large and symmetric pulse in Figure~\ref{fig:datazoom} near $375$ $\mu$s. Such a feature must represent a current pulse that turns on rapidly and turns off just as rapidly, suggesting that the effect of processes like the formation of the corona sheath are not as important for such pulses. Such speculation can easily be tested, and the future work described above is ongoing. It is our hope that such full electrodynamic simulations, motivated by physics, can help bridge the gap between plasma physics and lightning observation, helping both to constrain our understanding of the physics of the lightning channel and to interpret lightning observations.
\begin{acknowledgments}
The HAMMA data used in this paper are available from the authors on request.
The authors are very grateful for useful discussions with Thomas Gjesteland, Nikolai {\O}stgaard, and Forrest Foust, and especially Justin Barhite for assistance and sanity-checking the von Neumann stability analysis. Referee comments on a previous draft were also very helpful. This work was supported by Norwegian Research Council grant 197638/V30, US National Science Foundation grant ATM-0836326, DARPA NIMBUS grant HR0011-10-1-0058, and NASA grant NNM05AA22A.
\end{acknowledgments}
|
\section{Introduction}
\label{sec:intro}
The data from proton-proton (pp) collisions produced at the CERN LHC
provide an excellent environment to investigate properties of the top quark, in the context of its production and decay, with unprecedented precision.
Such measurements enable rigorous tests of the standard model (SM), and deviations from the SM predictions would indicate signs of possible new physics \cite{eqSaav3,np1,np2,np3}.
In particular, the $\PW$ boson helicity fractions in top quark decays are very sensitive to the Wtb vertex structure.
The $\PW$ boson helicity fractions are defined as the partial decay rate for a given helicity state divided by the total decay rate: $F_\mathrm{L,R,0}\equiv \Gamma_\mathrm{L,R,0}/\Gamma$, where $F_\mathrm{L}$, $F_\mathrm{R}$, and $F_\mathrm{0}$ are the left-handed, right-handed, and longitudinal helicity fractions, respectively.
The helicity fractions are expected to be $F_\mathrm{0}=0.687 \pm 0.005$, $F_\mathrm{L}=0.311 \pm 0.005$, and $F_\mathrm{R}=0.0017 \pm 0.0001$ at next-to-next-to-leading order (NNLO) in the SM,
including electroweak effects, for a top quark mass $m_{\PQt} =172.8 \pm 1.3\GeV$~\cite{Czarnecki:2010gb}.
Anomalous Wtb couplings, i.e. those that do not arise in the SM, would alter these values.
Experimentally, the $\PW$ boson helicity can be measured through the study of angular distributions of the top quark decay products.
The helicity angle $\theta^*$ is defined as the angle between the direction of either the down-type quark or the charged lepton arising from the $\PW$ boson decay and
the reversed direction of the top quark, both in the rest frame of the $\PW$ boson.
The distribution for the cosine of the helicity angle depends on the helicity fractions in the following way,
\begin{linenomath}
\begin{equation}
\ifthenelse{\boolean{cms@external}}
{
\begin{split}
\frac{1}{\Gamma}\frac{\rd\Gamma}{\rd\cos{\theta^*}} = & \phantom{+}\ \, \frac{3}{8}\left(1-\cos{\theta^*}\right)^2 F_\mathrm{L} \\
& + \frac{3}{4} (\sin{\theta^*} )^2 F_\mathrm{0} \\
& + \frac{3}{8}\left(1+\cos{\theta^*}\right)^2 F_\mathrm{R}.
\end{split}
}
{
\frac{1}{\Gamma}\frac{\rd\Gamma}{\rd\cos{\theta^*}} =
\frac{3}{8}\left(1-\cos{\theta^*}\right)^2 F_\mathrm{L} +
\frac{3}{4} (\sin{\theta^*} )^2 F_\mathrm{0} +
\frac{3}{8}\left(1+\cos{\theta^*}\right)^2 F_\mathrm{R}.
}
\label{eq:costhetastar}
\end{equation}
\end{linenomath}
This dependence is shown in Fig. \ref{fig:theory} for each contribution separately, normalised to unity, and for the SM expectation. Charged leptons (or down-type quarks) from left-handed W bosons are preferentially emitted in the opposite direction of the W boson,
and thus tend to have lower momentum and be closer to the b jet from the top quark decay,
as compared to charged leptons (or down-type quarks) from longitudinal or right-handed W bosons.
\begin{figure*}[h]
\centerline{ \includegraphics[width = 0.5\textwidth]{Figure_001.pdf} }
\caption{Predicted cos$\theta^*$ distributions for the different helicity fractions.
The distributions for the fractions $F_\mathrm{0},\ F_\mathrm{L}$, and $F_\mathrm{R}$
are shown as dashed, dotted, and dash-dotted lines, respectively, and the sum of the three contributions according to the SM predictions is displayed as a solid line. \label{fig:theory} }
\end{figure*}
The measurement of the $\PW$ boson helicity is sensitive to the presence of non-SM couplings between the $\PW$ boson, the top quark, and the bottom quark.
A general parametrisation of the Wtb vertex can be expressed as~\cite{eqSaav1,eqSaav3}
\begin{linenomath}
\begin{equation}
\ifthenelse{\boolean{cms@external}}
{
\begin{split}
\mathcal{L}_\mathrm{Wtb} = & - \frac{g}{\sqrt 2} \bar{b} \gamma^{\mu} ( V_\mathrm{L} P_\mathrm{L} + V_\mathrm{R} P_\mathrm{R} ) t W_\mu^- \\
& - \frac{g}{\sqrt 2} \bar {b} \frac{i \sigma^{\mu \nu} q_\nu}{M_\PW} \left( g_\mathrm{L} P_\mathrm{L} + g_\mathrm{R} P_\mathrm{R} \right) t\; W_\mu^- + \mathrm{h.c.} ,
\end{split}
}
{
\mathcal{L}_\mathrm{Wtb} = - \frac{g}{\sqrt 2} \bar{b} \gamma^{\mu} ( V_\mathrm{L} P_\mathrm{L} + V_\mathrm{R} P_\mathrm{R} ) t W_\mu^-
- \frac{g}{\sqrt 2} \bar {b} \frac{i \sigma^{\mu \nu} q_\nu}{M_\PW} \left( g_\mathrm{L} P_\mathrm{L} + g_\mathrm{R} P_\mathrm{R} \right) t\; W_\mu^- + \mathrm{h.c.} ,
}
\label{eq:Wtb0}
\end{equation}
\end{linenomath}
where $V_\mathrm{L},\ V_\mathrm{R},\ g_\mathrm{L},\ g_\mathrm{R}$ are vector and tensor couplings (complex constants), $q=p_{\PQt}-p_\PQb$, and $p_{\PQt}$ ($p_\PQb$) is the four-momentum of the top quark (b quark), $P_\mathrm{L}\ (P_\mathrm{R})$ is the left (right) projection operator, and h.c.\ denotes the Hermitian conjugate.
Hermiticity conditions on the possible dimension-six Lagrangian terms also impose $\mathrm{Im}(V_\mathrm{L})=0$~\cite{eqSaav2}.
In the SM and at tree level, \mbox{$V_\mathrm{L} = V_\mathrm{tb}$}, where $V_\mathrm{tb} \approx 1$ is the Cabibbo--Kobayashi--Maskawa matrix element connecting the top and the bottom quarks and
$V_\mathrm{R}=g_\mathrm{L}=g_\mathrm{R}=0$.
The relationships between the $\PW$ boson helicity fractions and the anomalous couplings including dependences on the b quark mass
are given in Ref.~\cite{SaavedraBernabeu}.
The helicity fractions of $\PW$ bosons in top quark decays were first measured at the Tevatron Collider \cite{d0new,cdfFull,TevComb}. They have been also measured at the LHC, using samples containing \ttbar~events obtained in pp collisions at 7\TeV, and having either one \cite{AtlasPaper,TOP-11-020} or two \cite{AtlasPaper} charged leptons in the final state.
The CMS Collaboration also reported measurements using event topologies that contain one single reconstructed top quark \cite{singtop}, in pp collisions at 8\TeV.
Limits on anomalous couplings have also been reported, derived from W boson helicity measurements \cite{AtlasPaper,TOP-11-020,singtop}, and from single top quark differential cross section
production measurements \cite{AnomATLAS}.
This Letter describes a measurement of the $\PW$ boson helicity fractions in \ttbar events involving one lepton and multiple jets, $\ttbar\to (\PW^+\PQb) \ (\PW^-\bar\PQb) \ \to \ (\ell^+\nu_\ell\PQb) \ (\PQq\PAQq^\prime\PAQb)$,
and its charge conjugate, where $\ell$ is an electron or a muon, including those from leptonic decays of a tau lepton.
Final states corresponding to such processes are referred to
as lepton+jets. The measurement relies on the analysis strategy described in Ref. \cite{TOP-11-020}.
The measurement is performed using pp collisions at centre-of-mass energy of 8\TeV, corresponding to an integrated luminosity of $19.8\fbinv$, collected during 2012 by the CMS detector.
\section{The CMS detector}
The CMS detector is a multipurpose apparatus of cylindrical design with respect to the proton beams.
The main features of the detector relevant for this analysis are briefly described here.
Charged particle trajectories are measured by a silicon pixel and strip tracker, covering the pseudorapidity range $|\eta| < 2.5$.
The inner tracker is immersed in a 3.8\unit{T} magnetic field provided by a superconducting solenoid of 6\unit{m} in diameter that also encompasses several calorimeters.
A lead tungstate crystal electromagnetic calorimeter (ECAL), and a brass and scintillator hadronic calorimeter surround the tracking volume and cover the region $|\eta| < 3$.
Quartz fibre and steel hadron forward calorimeters extend the coverage to $|\eta| \le 5$.
Muons are identified in gas ionisation detectors embedded in the steel return yoke of the magnet.
The data for this analysis are recorded using a two-level trigger system.
A more detailed description of the CMS detector, together with a definition of the coordinate system used and the relevant kinematic variables, can be found in Ref.~\cite{CMS_detector}.
\section{Data and simulated samples}
Signal events
corresponding to top quark pairs that decay to lepton+jets final states are expected to
contain one isolated lepton (electron or muon) together with at least four
jets, two of which originate from b quark fragmentation. Such events are
referred to separately
as $\Pe$+jets or $\mu$+jets, respectively, or when combined as $\ell$+jets.
Background events containing a single isolated lepton and four reconstructed jets arise mainly from processes that produce events containing a
single top quark, processes that produce multijet events in association with a $\PW$ boson that decays leptonically ($\PW$+jets), or Drell--Yan processes accompanied by multiple jets (DY+jets) when one of the leptons is misidentified as a jet or goes undetected.
Multijet processes can also mimic lepton+jets final states, if a
jet is reconstructed as an electromagnetic shower or, more unlikely, if a nonprompt muon from a hadron decay in flight fulfils all identification criteria of a prompt muon.
Simulated Monte Carlo (MC) samples, interfaced with \GEANTfour \cite{geant},
are used to account for detector resolution and acceptance effects, as well as to estimate the contribution from
background processes that have characteristics similar to lepton+jets
final states in \ttbar~decays.
A signal \ttbar sample, which also provides a reference for the SM (see Eq. (\ref{eq:Reweighiting:weightDef})), is simulated using \MADGRAPH v5.1.3.30 \cite{MADGRAPH} with matrix elements having up to three extra partons in the final state.
The parton distribution function (PDF) set {\sc CTEQ6L1}~\cite{cteq} is used when simulating
this reference \ttbar~sample.
The \MADGRAPH generator is interfaced with \PYTHIA v6.426 \cite{PYTHIA}, tune Z2* \cite{tuneZ2}, to simulate hadronisation and parton fragmentation, and also with \TAUOLA v27.121.5 \cite{tauola} to simulate $\tau$ lepton decays.
This SM reference \ttbar sample is simulated assuming
$m_{\PQt}=172.5\GeV$, which results in the following leading-order (LO) $\PW$ boson helicity fractions for that sample:
\begin{linenomath}
\begin{equation}
F^\mathrm{SM}_\mathrm{0}=0.6902,\ \ F^\mathrm{SM}_\mathrm{L}=0.3089,\ \ F^\mathrm{SM}_\mathrm{R}=0.0009.
\label{eq:refhel}
\end{equation}
\end{linenomath}
Single top quark events in the $s$, $t$, and tW
channels are generated using \POWHEG v1.0~\cite{POWHEG} and \PYTHIA interfaced with \TAUOLA, with the PDF set {\sc CTEQ6M}~\cite{cteq}.
Background $\PW$+jets and DY+jets processes are simulated using \MADGRAPH with the PDF set {\sc CTEQ6L1}, followed by \PYTHIA for fragmentation and hadronisation.
Finally, background multijet processes are simulated using the \PYTHIA event generator.
Corrections are applied to the simulated samples so that resolutions, energy scales, and efficiencies as functions of \pt and $\eta$ of jets
\cite{ref:jets} and leptons \cite{diff2} measured in data are well described.
The effect of multiple pp collisions occurring in the same bunch crossing (pileup) is also taken into account in the simulation.
The data samples selected for this measurement were recorded using inclusive single-lepton triggers, which require at least one isolated electron (muon) with $\pt>27\ (24)$\GeV, used to define the $\Pe$+jets ($\mu$+jets) data sample.
The decay products of candidate top quarks are reconstructed
using the CMS particle-flow (PF) algorithm, described in detail elsewhere~\cite{CMS-PAS-PFT-09-001,CMS-PAS-PFT-10-001}.
Individual charged particles identified as coming from pileup interactions are removed from the event.
Effects of neutral particles from pileup interactions are mitigated by applying corrections based on event properties.
Leptons are required to originate from the primary vertex of the event~\cite{ref:TRK-11-001}.
A lepton is determined to be isolated using
a variable computed as the
total transverse momentum of all particles (except the lepton itself) contained within a cone of radius 0.4,
centred on the lepton direction, relative to the transverse momentum of the lepton.
Electrons are identified by using a multivariate analysis (MVA)~\cite{ref:electrons} based on information from the inner tracker and the ECAL.
Events are selected for the $\Pe$+jets data sample if the identified electron has an MVA discriminant value greater than 0.9,
is determined to be isolated, has $\pt>30\GeV$, and $|\eta|<2.5$.
Muons are identified
by matching information from the inner tracker and the muon spectrometer~\cite{ref:muons}.
Events are selected for the $\mu$+jets data sample if they contain an isolated muon,
$\pt>26$\GeV, and $\abs{\eta} < 2.1$.
Events with at least one additional isolated electron or muon are vetoed to reject backgrounds from dileptonic decay modes of \ttbar and DY+jets processes.
Jets are reconstructed \cite{ref:jets} using the anti-$\kt$ clustering algorithm \cite{antikt_jet_algo}, with a distance parameter of 0.5.
The selected or vetoed leptons described above are not allowed to be clustered into jets, to avoid ambiguities.
The event selection requires at least four reconstructed jets having $|\eta|<2.4$,
of which the four most energetic jets are required to have $\pt$ higher than 55, 45, 35, and 20\GeV.
Events with additional jets are not vetoed. The transverse momentum imbalance of the event $\vec{p}_\mathrm{T}^\mathrm{miss}$ is determined by summing the
negative transverse momentum over
all reconstructed particles, excluding those charged
particles not associated with the primary vertex.
The transverse mass of the $\PW$ boson is defined as $M_\mathrm{T} = \sqrt{\smash[b]{ 2 p^\ell_\mathrm{T} p_\mathrm{T}^\mathrm{miss}(1-\cos(\Delta\phi))}}$, where $p^\ell_\mathrm{T}$ is the transverse momentum of the lepton, $p_\mathrm{T}^\mathrm{miss}$ is the magnitude of $\vec{p}_\mathrm{T}^\mathrm{miss}$, and $\Delta\phi$ is the angle in the $(x,y)$ plane between the direction of the lepton and $\vec{p}_\mathrm{ T}^\mathrm{miss}$.
To reduce the multijet background and suppress dilepton events from \ttbar~processes, events are required to have $30 < M_\mathrm{T} < 200\GeV$.
All backgrounds are further suppressed by requiring that at least two jets be identified as originating from b quarks. All jets
with $p_\mathrm{T}>20\GeV$ are considered as b quark candidates, including those that are not among the four most energetic.
The combined secondary vertex algorithm \cite{ref:btag,btagpas} tags
b quark jets with an efficiency of about 70\% and mistags jets originating from gluons, u, d, or s quarks with a probability of about 1\%, for the typical $p_\mathrm{T}$ ranges (30--100 GeV) probed in \texorpdfstring{\ttbar}{t-tbar} events.
Charm jets have a probability of $\approx$20\% of being tagged as b quark jets.
The residual multijet backgrounds, already strongly suppressed by the b tagging requirement described above, are estimated by normalising simulated event samples to yields in control data samples.
The control samples are defined by selection criteria which are similar to those for the signal, but which have no b tagging requirement, and have $ M_\mathrm{T} < 30\GeV$ for the $\mu$+jets channel or have an electron
MVA discriminant value smaller than 0.5 for the $\Pe$+jets channel.
The estimated amount of multijet events is $\approx$2\% of the $\Pe$+jets sample, and less than 1\% of the $\mu$+jets sample.
The contributions of all other residual backgrounds are determined using simulation.
\section{Reconstruction of the \texorpdfstring{\ttbar}{t-tbar} system and reweighting method}
The reconstruction of the \ttbar system, described in detail in Ref. \cite{TOP-11-020}, relies on testing the selected lepton,
the measured $\ptvecmiss$, and all selected jets for their compatibility with the top quark decay products from the leptonic
$(\PQt \to \PQb \PW\ \to \PQb \ell \nu)$ and hadronic $(\PQt \to \PQb \PW\to \PQb \PQq \PAQq^\prime)$ branches.
The unmeasured component of the neutrino momentum $p^\nu_z$ is determined by
requiring the \ttbar system to be consistent with the invariant masses of two top quarks and two $\PW$ bosons.
With these constraints,
\PQb jets are correctly assigned to the leptonic (hadronic) branch in about 74\% (71\%) of signal events.
After the assignment, a kinematic fit is performed, where the momenta of the measured
jets and lepton are allowed to vary within their resolutions to better comply with the mass constraints, leading to an improved determination of $p^\nu_z$ and a more accurate reconstruction of the $\ttbar$ system.
In about 5--7\% of the selected events, the fit fails to find a solution that is compatible with the constraints and such events are discarded.
The number of data events passing all selection criteria, including the fit convergence, is $71\,458$ in the $\Pe$+jets sample and 70 986 in the $\mu$+jets sample.
A study using simulations normalised to the most precise theoretical cross sections available to date \cite{ttbarNLO1,ttbarNLO2,singlet8Tev,fewz} indicates that the final sample composition is largely dominated by \ttbar~events, with about 82\% of events
from the $\ell$+jets decay mode, ${\approx}10\%$ from other decay modes (including $\tau$ leptons), and ${\approx}3.5\%$ of the events
from single top quark processes.
The remaining events come from backgrounds not containing top quarks in the final state.
The method~\cite{TOP-11-020} employed to measure the $\PW$ boson helicity fractions $(F_\mathrm{L},F_\mathrm{0},F_\mathrm{R})\equiv \vec{F}$ consists of maximising a binned Poisson likelihood function constructed using the number of observed events in data ${N_\text{data}}(i)$ and expected events from MC simulation ${ N_\mathrm{MC}}(i;\vec{F})$,
in each bin $i$ of the reconstructed $\cos\theta^*_\mathrm{rec}$ distribution,
\begin{linenomath}
\begin{equation}
\mathcal{L}(\vec{F}) = \prod_{i} \frac{N_\mathrm{MC}(i;\vec{F})^{\displaystyle~N_\text{data}(i)}}{\Bigl[ N_\text{data}(i)\Bigr]!}~\exp{[-N_\mathrm{MC}(i;\vec{F})]}.
\end{equation}
\end{linenomath}
\noindent
While the charged lepton is easily identified in the leptonic branch of \ttbar~decays, the down-type quark jet arising from the $\PW$ boson decay in the hadronic branch of \ttbar~decays can not be experimentally distinguished from the up-type quark jet.
Due to this ambiguity, only the absolute value $|\cos\theta^*_\text{rec}|$ can be reconstructed for the hadronic branch.
Hence, only the leptonic branch measurement of $\cos\theta^*_\text{rec}$ is used in this analysis.
The expected numbers of events from background processes, ${N_{\PW+\text{jets}}}(i)$, ${N_{\mathrm{DY}+\text{jets}}}(i)$, and ${N_\text{multijet}}(i)$ represent $\PW$ boson production in association with multiple jets, Drell--Yan production in association with multiple jets, and production of multiple jets, which do not depend on the $\PW$ boson helicity fractions.
For the processes containing top quarks, the number of expected events in a given bin $i$ is modified by reweighting each event in that bin by
a factor $w$, defined for each decaying branch as
\begin{equation}
\ifthenelse{\boolean{cms@external}}
{
\begin{split}
w_\text{lep/had/single-\PQt}(\cos\theta^{*}_\text{gen}; & \vec{F}) \equiv \\
& \left[\ \phantom{+} \frac{3}{8} F_\mathrm{L} (1-\cos\theta^{*}_\text{gen})^2 \right. \\
& \left. \ + \ \frac{3}{4} F_\mathrm{0} \sin^{2}\theta^{*}_\text{gen} \right. \\
& \left. \ + \ \frac{3}{8} F_\mathrm{R} (1+\cos\theta^{*}_\text{gen})^{2} \ \right] / \\
& \left[ \ \phantom{+} \frac{3}{8} F^\mathrm{SM}_\mathrm{L }(1-\cos\theta^{*}_\text{gen})^2 \right. \\
& \left. \ + \ \frac{3}{4} F^\mathrm{SM}_\mathrm{0 } \sin^{2}\theta^{*}_\text{gen} \right. \\
& \left. \ + \ \frac{3}{8} F^\mathrm{SM}_\mathrm{R } (1+\cos\theta^{*}_\text{gen})^{2} \ \right],
\end{split}
}
{
w_\text{lep/had/single-\PQt}(\cos\theta^{*}_\text{gen};\vec{F}) \equiv
\frac
{\displaystyle
\frac{3}{8} F_\mathrm{L} (1-\cos\theta^{*}_\text{gen})^2 +
\frac{3}{4} F_\mathrm{0} \sin^{2}\theta^{*}_\text{gen} +
\frac{3}{8} F_\mathrm{R} (1+\cos\theta^{*}_\text{gen})^{2}
}
{\displaystyle
\frac{3}{8} F^\mathrm{SM}_\mathrm{L }(1-\cos\theta^{*}_\text{gen})^2 +
\frac{3}{4} F^\mathrm{SM}_\mathrm{0 } \sin^{2}\theta^{*}_\text{gen} +
\frac{3}{8} F^\mathrm{SM}_\mathrm{R } (1+\cos\theta^{*}_\text{gen})^{2}
},
}
\label{eq:Reweighiting:weightDef}
\end{equation}
where $\theta^{*}_\text{gen}$ is the helicity angle (specified at matrix element level) of a particular decay branch,
and $F^\mathrm{SM}_\mathrm{L}, F^\mathrm{SM}_\mathrm{0}, F^\mathrm{SM}_\mathrm{R}$ are given in Eq.~(\ref{eq:refhel}).
Therefore, the number of expected events, as a function of the helicity fractions to be measured, is
\begin{equation}
\ifthenelse{\boolean{cms@external}}
{
\begin{split}
{N_\mathrm{MC}}(i;\vec{F}) = & \phantom{+} \ \, {N_{\ttbar}}(i;\vec{F}) \\
& + {N_\text{single-\PQt}}(i;\vec{F}) \\
& + {N_{\PW+\text{jets}}}(i) \\
& + {N_{\mathrm{DY}+\text{jets}}}(i) \\
& + {N_\text{multijet}}(i),
\end{split}
}{
{N_\mathrm{MC}}(i;\vec{F}) = {N_{\ttbar}}(i;\vec{F}) + N_\text{single-\PQt}(i;\vec{F}) + {N_\mathrm{\PW+jets}}(i)+{N_\mathrm{DY+jets}}(i)+{N_\mathrm{multijet}}(i),
}
\end{equation}
where
\ifthenelse{\boolean{cms@external}}
{
\begin{equation}
\begin{split}
N_{\ttbar} (i;\vec{F}) = \mathcal{ F}_{\ttbar} \left[ \sum_\text{\ttbar events in bin $i$} \right. & \left. \ \ \ \ w_\mathrm{lep}(\cos\theta^{*}_\text{gen};\vec{F}) \right. \\
& \left. \times \ w_\text{had}(\cos\theta^{*}_\text{gen};\vec{F}) \ \ \right] , \\
N_\mathrm{single\textnormal{-}t}(i;\vec{F}) = \sum_{\text{single-\PQt events in bin $i$}} & w_\text{single-\PQt}(\cos\theta^{*}_\text{gen};\vec{F})
\end{split}
\label{eq::Reweighting::WeightDef1}
\end{equation}
}
{
\begin{eqnarray}
{N_{\ttbar}}(i;\vec{F}) & = & \mathcal{ F}_{\ttbar}\left[\sum_\text{\ttbar events in bin $i$} w_\mathrm{lep}(\cos\theta^{*}_\text{gen};\vec{F}) \times w_\mathrm{had}(\cos\theta^{*}_\text{gen};\vec{F}) \right]
\label{eq::Reweighting::WeightDef1}, \\
N_\mathrm{single\textnormal{-}t}(i;\vec{F}) & = & \sum_\text{single-\PQt events in bin $i$} w_\text{single-\PQt}(\cos\theta^{*}_\text{gen};\vec{F})
\end{eqnarray}
}
represent the expected number of events fulfilling event selection criteria for processes involving top quark pair, and single top quark production, respectively.
The normalisation factor $\mathcal{F}_{\ttbar}$ for the \ttbar sample is a single free parameter in the fit across all bins.
The expected cross section for the simulated reference \ttbar sample is $252.9^{+13.3}_{-14.5}$\unit{pb}, calculated at NNLO and next-to-next-to-leading-log (NNLL) accuracy \cite{ttbarNLO1,ttbarNLO2}, and describes the data well.
The fitted values of $\mathcal{F}_{\ttbar}$ in both $\Pe$+jets and $\mu$+jets channels
are compatible with 1.00 within 3\%.
The overall normalisation factor for simulated single top quark events is not modified in the fit and the uncertainty in the assumed cross section is considered as a source of systematic uncertainty.
Finally, the unitarity constraint ($F_\mathrm{L}+F_\mathrm{0}+F_\mathrm{R}=1$) is imposed, so that one of the helicity fractions, namely $F_\mathrm{R}$, is
bound by the measurement of the other two.
The method was validated using pseudo-experiments, where the fitting procedure was performed on pseudo-data, mimicking altered helicity fractions.
Linearity tests show that the fitting procedure correctly retrieves
the helicity fractions of altered input values for $F_0\in[0.50,0.85]$
and $F_L\in[0.20,0.50]$. Likewise the corresponding
statistical uncertainties were verified using sets of statistically uncorrelated pseudo-data.
\begin{table*}[h]
\begin{center}
\topcaption{ Systematic uncertainties on the measurements of the $\PW$ boson helicity fractions
from lepton+jets events. The cases in which the statistical precision of the limited sample size was assigned as systematic uncertainties are denoted by
the symbol (*).
\label{tab:system}}
\tabcolsep=0.15cm
\ifthenelse{\boolean{cms@external}}{}{\resizebox{\textwidth}{!}}
{
\begin{tabular}{lcc|c c |c c } \cline{2-7}
& \multicolumn{2}{c|}{$\Pe$+jets} & \multicolumn{2}{c|}{$\mu$+jets} & \multicolumn{2}{c}{$\ell$+jets} \\ \cline{2-7}
& $\pm$ $\Delta F_\mathrm{0}$ & $\pm$ $\Delta F_\mathrm{L}$ & $\pm$ $\Delta F_\mathrm{0}$ & $\pm$ $\Delta F_\mathrm{L}$ & $\pm$ $\Delta F_\mathrm{0}$ & $\pm$ $\Delta F_\mathrm{L}$ \\ \hline
JES & 0.004 & 0.003 & 0.005 & 0.003 & 0.005 & 0.003 \\
JER & 0.001 & 0.002 & 0.004 & 0.003 & 0.003 & 0.003 \\
b tagging eff. & 0.001 & ${<}10^{-3}$ & 0.001 & ${<}10^{-3}$ & 0.001 & ${<}10^{-3}$ \\
Lepton eff. & 0.001 & 0.002 & 0.001 & 0.001 & 0.001 & 0.001 \\
Single top normal. & 0.002 & ${<}10^{-3}$ & 0.003 & 0.001 & 0.003 & 0.001 \\
$\PW$+jets bkg. & 0.008 & 0.001 & 0.007 & 0.001 & 0.007 & 0.001 \\
DY+jets bkg. & 0.002 & ${<}10^{-3}$ & 0.001 & ${<}10^{-3}$ & 0.001 & ${<}10^{-3}$ \\
Multijet bkg. & 0.023 & 0.007 & 0.007 & 0.003 & 0.008 & 0.001 \\
Pileup & 0.001 & 0.001 & ${<}10^{-3}$ & ${<}10^{-3}$ & 0.001 & ${<}10^{-3}$ \\
Top quark mass & 0.012 & 0.008 & 0.010 (*) & 0.008 (*) & 0.010 & 0.007 \\
\ttbar~scales & 0.011 & 0.008 (*) & 0.014 & 0.007 (*) & 0.012 & 0.007 \\
\ttbar~match. scale & 0.011 (*) & 0.007 (*) & 0.010 & 0.007 & 0.009 & 0.007 \\
\ttbar~MC and hadronisation & 0.015 & 0.009 & 0.005 & 0.003 & 0.006 & 0.004 \\
\ttbar \pt reweight & 0.011 & 0.010 & ${<}10^{-3}$ & 0.001 & ${<}10^{-3}$ & 0.002 \\
Limited MC size & 0.002 & 0.001 & 0.002 & 0.001 & 0.002 & 0.001 \\
PDF & 0.004 & 0.001 & 0.002 & 0.001 & 0.002 & 0.001 \\ \hline
Total & 0.037 & 0.020 & 0.024 & 0.014 & 0.023 & 0.014 \\ \hline
\end{tabular}
}
\end{center}
\end{table*}
\begin{figure*}[h]
\centerline{ \includegraphics[width = 0.5\textwidth]{Figure_002-a.pdf} \includegraphics[width = 0.5\textwidth]{Figure_002-b.pdf}}
\centerline{ \includegraphics[width = 0.5\textwidth]{Figure_002-c.pdf} \includegraphics[width = 0.5\textwidth]{Figure_002-d.pdf}}
\caption{
Distributions for the cosine of the helicity angle in the leptonic (upper row) and hadronic (lower row) branches, for the $\Pe$+jets (left) and $\mu$+jets (right) decay channels.
The combined $\ell$+jets post-fit measurements of the helicity fractions were used in the simulation of \ttbar~and single top quark events.
The data are displayed as solid points, simulated samples of \ttbar (signal) processes
and the contribution from background processes as histograms.
At the bottom of each plot, the ratio between MC simulation and data is displayed. The error bars correspond to the statistical uncertainties. \label{fig2}
}
\end{figure*}
\begin{figure*}[h]
\begin{center}
\centerline{ \includegraphics[scale=0.45]{Figure_003-a.pdf} \includegraphics[scale=0.45]{Figure_003-b.pdf} }
\caption{
Left: the measured $\PW$ boson helicity fractions in the $(F_\mathrm{0},F_\mathrm{L})$ plane.
The dashed and solid ellipses enclose the allowed two-dimensional 68\% and 95\% CL regions, for the combined $\ell$+jets measurement, taking into account the correlations on the total (including systematic) uncertainties.
The error bars give the one-dimensional 68\% CL interval for the separate $F_\mathrm{0}$ and $F_\mathrm{L}$ measurements, with the inner-tick (outer-tick) mark representing the statistical (total) uncertainty.
Right: the corresponding allowed regions for the real components of the anomalous couplings $g_\mathrm{L}$ and $g_\mathrm{R}$ at 68\% and 95\% CL, for $V_\mathrm{L}=1$ and $V_\mathrm{R}=0$.
A region near $\mathrm{Re}(g_\mathrm{L})=0$ and $\mathrm{Re}(g_\mathrm{R}) \gg 0$, allowed by the fit but excluded by the CMS single top quark production measurement, is omitted.
The SM predictions are shown as stars.
\label{fig:limanom}}
\end{center}
\end{figure*}
\section{Systematic uncertainties}
Systematic effects which could potentially bias the measurement of the $\PW$ boson helicity fractions have been investigated and their corresponding uncertainties determined, as presented in Table~\ref{tab:system}.
Residual corrections are applied in simulation to the jet energy scale (JES), to account for differences between data and simulation.
The momenta of the jets in simulation are also smeared so that the jet energy resolution (JER) in simulation agrees with that in data.
These corrections and smearings are propagated into $\ptvecmiss$ to correct its momentum scale.
The uncertainties~\cite{ref:jets} associated with the JES and JER corrections
are also propagated to $\ptvecmiss$, and the full analysis, including the \ttbar~reconstruction and the resulting measurements of the $\PW$ boson helicity fractions, is repeated.
Scale factors are used to correct the b tagging efficiency in simulation,
where those corrections are shifted by their estimated uncertainties, and the full analysis repeated.
Scale factors are also used to correct leptons for their identification, isolation and trigger efficiencies, which are varied within their uncertainties so as to maximise potential shape variations of the predicted \ensuremath{\cos \theta^*}\xspace~distributions.
To account for any possible bias of the $\PW$ boson helicity measurements due to uncertainties in the normalisation of simulated backgrounds, the assumed cross section for each sample is varied individually \cite{TOP-11-020}.
An uncertainty of 30\% is used for the normalisation of DY+jets, single top quark, and $\PW$ boson production in association with light-quark or gluon
jet production.
Since the modelling of the simulated heavy-flavour content of the
$\PW$+jets sample is known to be inaccurate, an uncertainty of ~$^{+100\%}_{-50\%}$ is assumed for simulated events involving a $\PW$ boson produced in association with b quark jets.
The impact of the
DY+jets normalisation uncertainty
in the analysis is small, since it corresponds to only a few percent of the sample composition.
The normalisation of the multijet background is estimated from control samples and results in an uncertainty of $^{+50\%}_{-50\%}$ in the $\Pe$+jets channel and $^{+40\%}_{-50\%}$ in the $\mu$+jets channel.
Shape uncertainties on the multijet background templates were investigated by comparing the distributions in several different control regions, both in MC and in data, and were found to be negligible, compared to the much larger normalisation uncertainties.
Several uncertainties from possible systematic effects related to theoretical modelling of the signal are estimated by replacing the default \ttbar samples
with alternative \ttbar samples and repeating the entire analysis.
Specifically, for the \MADGRAPH interfaced with PYTHIA event generation,
the default $m_{\PQt}$ value of 172.5\GeV is shifted up and down by 1\GeV;
the renormalisation and factorisation scales are varied down (up) by a factor of 0.5 (2); the kinematic scale used to match jets to partons (matching threshold) is varied down (up) by factor of 0.5 (2); finally, the parton shower and hadronisation modelling is studied in a \ttbar sample simulated with MC@NLO v3.41 \cite{mcnlo} using the PDF set CTEQ6M and interfaced with \HERWIG v6.520 \cite{herwig}.
Uncertainties in the helicity fractions arising from the limited size of the simulated \ttbar samples are
taken into account, both in the main analysis and in the determination of the modelling uncertainties.
In the former case, these effects are added as a separate source of uncertainty.
In the latter case, the systematic uncertainties in the $\PW$ boson helicity are assigned to be the larger of either (i) the statistical precision of the limited sample size or (ii) the systematic shift of the central value with respect to the reference \ttbar sample.
The shape of the \pt spectrum for top quarks, as measured by the differential cross section for top quark pairs \cite{diff1,diff2}, has been found to be softer than the predictions from \MADGRAPH simulation.
The effect of this mismodelling is estimated by reweighting the events in the simulated \ttbar sample, so that
the top quark \pt~at parton level in the MC describes the unfolded data distribution.
Further, the systematic effects due to the PDFs used to simulate the signal and background samples are estimated according to
the prescriptions described in \cite{pdf4lhc,pdf4lhc1}, using NNPDF21~\cite{NNPDF}
and MSTW2008lo68cl~\cite{pdfs} PDF sets as alternatives to those used at generation.
Finally, uncertainties related to the modelling of the pileup in simulated events are also taken into account.
The total systematic uncertainty is given by the sum in quadrature of all uncertainties described above.
\section{Results}
The measurements of the $\PW$ boson helicity fractions, using \ensuremath{\cos \theta^*}\xspace~from the leptonic branch of \ttbar~events that decay into $\Pe$+jets or $\mu$+jets, including the full combination of these two measurements, are shown in Table~\ref{tab:results}.
Within an individual channel, the helicity parameters $F_\mathrm{0}$ and $F_\mathrm{L}$ are fit simultaneously, but they are strongly anti-correlated due to the unitarity constraint $F_\mathrm{L}+F_\mathrm{ 0}+F_\mathrm{R}=1$, as indicated by the statistical correlation coefficient $\rho_\mathrm{0,L}$ given in the table.
The separate helicity measurements from the $\Pe$+jets and $\mu$+jets channels are combined into a single $\ell$+jets measurement using the BLUE method \cite{Lyons:1988rp,Valassi:2003mu}, taking into account all uncertainties and their possible correlations.
Uncertainties related to lepton efficiency, multijet background estimations, and statistical uncertainties are considered uncorrelated between the $\Pe$+jets and $\mu$+jets analyses, while all other uncertainties are assumed to be fully correlated.
The combined $\ell$+jets measurement of the helicity fractions is dominated by the $\mu$+jets channel, with
weights more than double those of the $\Pe$+jets channel.
The $\chi^2$ of the combination is 2.13 for 2 degrees of freedom, corresponding to a probability of 34.5\%. The measurement uncertainties are dominated by systematic effects that are correlated between both the $\Pe$+jets and $\mu$+jets channels.
The combined $F_\mathrm{0}$ and $F_\mathrm{L}$ values are anti-correlated with a statistical correlation coefficient $\rho_\mathrm{0,L}=-0.959$. The total correlation coefficient,
considering both statistical and systematic uncertainties, is found to be $-0.870$.
The measured helicity fractions presented in Table~\ref{tab:results} are consistent with the SM predictions given at NNLO accuracy~\cite{Czarnecki:2010gb}.
Figure~\ref{fig2} shows, separately for the $\Pe$+jets and $\mu$+jets channels, the distributions for the cosine of the helicity angles from the leptonic branch,
which are used in the helicity measurements, and the distributions of the corresponding absolute values from the hadronic branch, for comparison purposes.
The simulated samples involving top quarks used in the figure were produced using the measured values for the $\PW$ boson helicity fractions, as determined from the combined $\ell$+jets fit.
Left-handed W bosons tend to populate the region at $\cos\theta^{*}\approx -1$, where the charged lepton overlaps with the b quark.
However, the angular separation requirement between leptons and jets removes most of the events
near $\cos\theta^{*}= -1$.
Very few events are expected in the region preferred by right-handed bosons, near $\cos\theta^{*}= +1$.
However, due to resolution effects,
the reconstructed distribution does not fall as rapidly as
expected in that region, where the charged lepton and b quark have opposite directions.
For these reasons, the shape of the reconstructed $\cos\theta^{*}$ distribution differs from that expected in the SM (Fig. 1). These features are well reproduced by the simulation, and taken into account in the measurement.
\begin{table*}[h]
\begin{center}
\topcaption{Measurements of the $\PW$ boson helicity fractions from lepton+jets
final states in \ttbar~decays.
The helicity fractions $ F_\mathrm{0}$ and $ F_\mathrm{L}$ are measured simultaneously and are strongly anti-correlated, as indicated by a
correlation coefficient $\rho_\mathrm{0,L}$, because $F_\mathrm{R}$ is derived from the unitarity condition. \label{tab:results}}
\tabcolsep=0.15cm
\ifthenelse{\boolean{cms@external}}{}{\resizebox{\textwidth}{!}}
{
\begin{tabular}{lrrrc} \hline
Channel & $F_0$ $\pm$ (stat) $\pm$ (syst) & $F_L$ $\pm$ (stat) $\pm$ (syst) & $F_R$ $\pm$ (stat) $\pm$ (syst) & $\rho_{0,L}$
\\ \hline
$\Pe$+jets & 0.705 $\pm$0.013$\pm$0.037 & 0.304$\pm$0.009$\pm$0.020 & $-0.009\pm 0.005 \pm$0.021 & $-0.950$ \\
$\mu$+jets & 0.685 $\pm$0.013$\pm$0.024 & 0.328$\pm$0.009$\pm$0.014 & $-0.013\pm 0.005 \pm$0.017 & $-0.957$ \\
$\ell$+jets & 0.681 $\pm$0.012$\pm$0.023 & 0.323$\pm$0.008$\pm$0.014 & $-0.004\pm 0.005 \pm$0.014 & $-0.959$ \\ \hline
\end{tabular}
}
\end{center}
\end{table*}
Using these results, limits on anomalous couplings are obtained by fixing the two vector couplings in Eq. (\ref{eq:Wtb0}) to their SM values, $V_\mathrm{L}=1$ and $V_\mathrm{R}=0$, and choosing the tensor couplings, $\mathrm{Re}(g_\mathrm{L})$ and $\mathrm{Re}(g_\mathrm{R})$, as free parameters.
The combined $\ell$+jets measurement of the $\PW$ boson helicity fractions $F_\mathrm{0}$ and $F_\mathrm{L}$ is reinterpreted in terms of the tensor couplings, $\mathrm{Re}(g_\mathrm{L})$ and $\mathrm{Re}(g_\mathrm{ R})$, using the relationships between the $\PW$ boson helicity fractions given in Ref.~\cite{SaavedraBernabeu}.
The $\PW$ boson helicity measurements are displayed in the $(F_\mathrm{0},F_\mathrm{L})$ plane in Fig. \ref{fig:limanom} (left), together with their one-dimensional statistical (inner-tick mark) and total (outer-tick mark) error bars.
The full two-dimensional confidence level (CL) contours corresponding to 68\% (dashed line) and 95\% (solid line)
probabilities are also displayed for the combined measurement.
The SM prediction is shown as a star and lies within the 68\% CL contour.
The corresponding regions in the $(\mathrm{Re}(g_\mathrm{L}),\mathrm{Re}(g_\mathrm{R}))$ plane, allowed at 68\% (dark contour) and 95\% CL (light contour), are shown in
Fig. \ref{fig:limanom} (right),
together with the SM value.
They are derived from Fig. \ref{fig:limanom} (left), using the relationships between the $\PW$ boson helicity fractions and the anomalous
couplings given in Ref.~\cite{SaavedraBernabeu}.
A region near $\mathrm{Re}(g_\mathrm{L})=0$ and $\mathrm{Re}(g_\mathrm{R}) \gg 0$, allowed by the fit but excluded by the CMS single top quark production measurement \cite{tchann}, is not shown.
If the right-handed component $F_\mathrm{R}$ is bound to zero,
consistently with the SM within the precision of
the current measurement, the combined $\ell$+jets measurement amounts to
$F_\mathrm{0}=0.661\pm0.006\stat\pm 0.021\syst$.
In this case, $F_\mathrm{L}$ is obtained via the unitarity constraint and yields $F_\mathrm{L}=0.339\pm0.006\stat\pm 0.021\syst$.
\section{Summary}
A measurement of the $\PW$ boson helicity fractions in top quark pair events decaying in the $\Pe$+jets and $\mu$+jets channels
has been presented, using proton-proton collision data at \mbox{$\sqrt{s}= 8\TeV$,} and corresponding to an integrated luminosity of 19.8\fbinv.
The helicity fractions $F_\mathrm{0}$ and $F_\mathrm{L}$ are measured with a precision of better than 5\%,
yielding the most accurate experimental determination of the $\PW$ boson helicity fractions in \ttbar processes to date.
The measured $\PW$ boson helicity fractions are
$F_\mathrm{0}=0.681\pm0.012\stat\pm 0.023\syst$, $F_\mathrm{L}=0.323\pm 0.008\stat\pm 0.014\syst$,
and $F_\mathrm{R}=-0.004\pm 0.005\stat\pm 0.014\syst$,
with a correlation coefficient of $-0.87$ between $F_\mathrm{0}$ and $F_\mathrm{L}$,
and they are consistent with the expectations from the standard model.
\begin{acknowledgments}
\hyphenation{Bundes-ministerium Forschungs-gemeinschaft Forschungs-zentren}
We congratulate our colleagues in the CERN accelerator departments for the excellent performance of the LHC and thank the technical and administrative staffs at CERN and at other CMS institutes for their contributions to the success of the CMS effort. In addition, we gratefully acknowledge the computing centres and personnel of the Worldwide LHC Computing Grid for delivering so effectively the computing infrastructure essential to our analyses. Finally, we acknowledge the enduring support for the construction and operation of the LHC and the CMS detector provided by the following funding agencies: BMWFW and FWF (Austria); FNRS and FWO (Belgium); CNPq, CAPES, FAPERJ, and FAPESP (Brazil); MES (Bulgaria); CERN; CAS, MOST, and NSFC (China); COLCIENCIAS (Colombia); MSES and CSF (Croatia); RPF (Cyprus); MoER, ERC IUT and ERDF (Estonia); Academy of Finland, MEC, and HIP (Finland); CEA and CNRS/IN2P3 (France); BMBF, DFG, and HGF (Germany); GSRT (Greece); OTKA and NIH (Hungary); DAE and DST (India); IPM (Iran); SFI (Ireland); INFN (Italy); MSIP and NRF (Republic of Korea); LAS (Lithuania); MOE and UM (Malaysia); BUAP, CINVESTAV, CONACYT, LNS, SEP, and UASLP-FAI (Mexico); MBIE (New Zealand); PAEC (Pakistan); MSHE and NSC (Poland); FCT (Portugal); JINR (Dubna); MON, RosAtom, RAS and RFBR (Russia); MESTD (Serbia); SEIDI and CPAN (Spain); Swiss Funding Agencies (Switzerland); MST (Taipei); ThEPCenter, IPST, STAR and NSTDA (Thailand); T\"UBITAK and TAEK (Turkey); NASU and SFFR (Ukraine); STFC (United Kingdom); DOE and NSF (USA).
Individuals have received support from the Marie-Curie programme and the European Research Council and EPLANET (European Union); the Leventis Foundation; the Alfred P. Sloan Foundation; the Alexander von Humboldt Foundation; the Belgian Federal Science Policy Office; the Fonds pour la Formation \`a la Recherche dans l'Industrie et dans l'Agriculture (FRIA-Belgium); the Agentschap voor Innovatie door Wetenschap en Technologie (IWT-Belgium); the Ministry of Education, Youth and Sports (MEYS) of the Czech Republic; the Council of Science and Industrial Research, India; the HOMING PLUS programme of the Foundation for Polish Science, cofinanced from European Union, Regional Development Fund; the Mobility Plus programme of the Ministry of Science and Higher Education (Poland); the OPUS programme of the National Science Center (Poland); the Thalis and Aristeia programmes cofinanced by EU-ESF and the Greek NSRF; the National Priorities Research Program by Qatar National Research Fund; the Programa Clar\'in-COFUND del Principado de Asturias; the Rachadapisek Sompot Fund for Postdoctoral Fellowship, Chulalongkorn University (Thailand); the Chulalongkorn Academic into Its 2nd Century Project Advancement Project (Thailand); and the Welch Foundation, contract C-1845.
\end{acknowledgments}
\clearpage
|
\section{Introduction}
It was recently shown that the replica trick \cite{EEandQFT} can be used to measure
the R{\'e}nyi entropies for single subsystems of a lattice in a globally pure state
\cite{RenyiMelko}.
This allows the quantification of the entanglement across the cut but only if the state of
the entire system is pure. As such, it cannot tell us anything about the entanglement
between two subsystems, or what happens at non-zero temperatures. Moreover there have also
been developments in adapting this method for use in experimental systems
\cite{NatureTwins}, as well as methods \cite{twoPhoton1,twoQubit2}
based on the circuits for measuring the moments of the partially transposed density
matrix developed by the author in \cite{PT}. These expectation values have already been
studied as individual moments using Quantum Monte Carlo in \cite{Lauchli3}, and the reader
is referred to that paper as well as \cite{RenyiMelko,genericRenyi,pathIntegralRenyi}
for the details of sampling methods.
The problem of determining whether a state is bipartite entangled is known to be NP-hard
in the size of the subsystems\cite{GurvitsSTOC}, see also \cite{Gharibian,Ioannou}.
However, the problem of determining whether multiple copies of a state can be ``distilled''
\cite{firstDistillation} into a smaller number that are more entangled is believed to be
a lot more tractable as the subsystems involved become large. This question can be
addressed by evaluating an entanglement measure \cite{computEnt,negativity1} known as
the \emph{entanglement negativity}, $\mathcal{N}(\rho).$
This is the sum of the moduli of the negative eigenvalues of $\rho^{T_2},$ where
$\rho^{T_2}$ is the partially transposed density matrix. This can be written in terms of
the one-norm $||X||_1,$ which is defined to be the sum of the moduli of the eigenvalues of
$X,$ thus
\begin{equation}\label{linearNegativity}
\mathcal{N}(\rho)= \frac{1}{2}\left(||\rho^{T_2}||_1-1\right)
\end{equation}
or equivalently, as the log-negativity:
\begin{equation}\label{logNegativity}
\mathcal{E}(\rho)= \ln(||\rho^{T_2}||_1).
\end{equation}
The entanglement negativity is believed to have a specific operational meaning: given an
infinite number of copies of the state $\rho,$ what is the asymptotic yield of some
maximally entangled state that we extract by doing local measurements on all the copies
\cite{computEnt}?
The entanglement negativity gives an upper bound on the yield for bipartite
entanglement distillation protocols \cite{computEnt}, which is conjectured to be an equality.
The negativity is also a good choice of entanglement measure because it is defined for
all bipartite states, including mixed ones, as it is relatively simple to calculate;
no convex roofs over ensemble decompositions of the density matrix are required.
Thus a general method for evaluating the negativity would allow us to probe the entanglement
properties of finite temperature systems without modification.
Indeed, many other natural choices of entanglement and correlation measures
including the entanglement cost, relative entropy of entanglement and the quantum discord
are NP-hard/NP-complete \cite{discordNPcomplete} in the subsystem size.
(This reference also contains a more complete list of computationally hard measures; note
that the negativity is not on that list.)
The setting we have in mind is one in which we have two subsystems of interest, as opposed
to the single subsystem of interest that is usually considered in calculations of the
R{\'e}nyi entropy.
Rather than look at the behaviour of our entanglement measure only in the limit as the
subsystems of interest become large, instead we would also like to explore how the entanglement
between the two subsystems scales as the subsystems are moved apart.
This will allow us to detect the presence of long range entanglement and other corrections
to area laws which cannot be detected using two-point functions alone.
Some form of long-range entanglement is believed to be related to topological quantum order
\cite{XGWenBook1,KimLongRangeNec}, though whether or not the negativity captures the relevant
type of entanglement seems to depend on the system. Various authors have tried to use the
entanglement negativity as a measure of long-range entanglement in systems that are tractable
using analytic methods. Some topologically ordered systems do not exhibit long-range negativity
\cite{CastelnovoToric,LeeVidal},
but some other systems do
\cite{EdgeNegaChernSimons,N-2dHarmonicCorner}
and where present, there does seem to be a connection with topological quantum order.
Several other authors have already examined the entanglement negativity using a combination
of analytic and numerical methods; small systems can be handled directly using exact
diagonalization \cite{ultracold}.
Larger systems have been studied using a variety of methods. Some authors examine the moments
of $\rho^{T_2}$ only, or the quotients
\begin{equation}
R_n(y)=\frac{Tr((\rho_{AB}^{T_B})^n)}{Tr(\rho_{AB}^n))}
\end{equation}
introduced in \cite{NegativityQFT}; see also \cite{FiniteT-N-CFT}.
This approach was combined with Quantum Monte Carlo methods in \cite{AlbaNegativity,Lauchli3}.
More generally, several authors have aleady used the entanglement negativity in numerical
studies to probe the entanglement properties of specific phenomena
\cite{NsepBlocksCritChain,N-RandomSpinChains,N-universalityLMG,NegativityKondoChain,Nega2dFreeLattice}.
There have also been attempts to use such an analytical extrapolation, in
\cite{disjCFTnumExtrap}, though as yet these methods are uncontrolled.
If existence or non-existence proofs are desired,
such as when considering the ways in which entanglement can be shared between
multiple systems,
then the extrema are required instead of some kind of maximum entropy estimate.
(Moreover, it is known that maximum entropy methods must be used with care
when trying to estimate the distillable entanglement
\cite{fakeEntanglementH3,EntangUnknownStates,infoVsDistillEnt}.)
A note on terminology: the word ``negativity'' can have two inequivalent meanings in the
quantum information literature. One of them is the entanglement negativity defined above; the
other one refers to situations where quantities that are usually interpreted as probabilities
can nevertheless take on negative values. (These are sometimes called quasi-probabilities,
see for example \cite{Hakop,quasiprobQC}.)
For the rest of this paper, the word ``negativity'' should be interpreted as referring to the
entanglement negativity only.
\section{Outline of moment problems}
The negativity can be computed exactly for finite-dimensional systems either by exact
diagonalization of $\rho^{T_2}$ or by first obtaining a complete set of moments
$\mu_n=\text{Tr}((\rho^{T_2})^n)$
and then using these in the Newton-Girard formulae to find the eigenspectrum \cite{PT}.
(This reference explains the construction in terms of quantum circuits but those can be
re-interpreted as examples of the multi-copy replica trick in a natural way, by the same
argument as given in \cite{RenyiMelko}.)
However, in the context of large system sizes or experimental settings, even finding the
eigenspectrum has a prohibitive overhead, whether in terms of computational cost or the
difficulty of precision control of quantum systems.
A useful method can use at most a few, low-order moments of the partially-transposed
density matrix $\rho^{T_2},$ which can be written
\begin{align}
\mu_n &= \text{Tr}((\rho^{T_2})^{n}) =\sum_i |\lambda_i|^n, &\text{for} \; n=2m,\label{evenMmts} \\
\nu_n &= \text{Tr}((\rho^{T_2})^{n})
=\sum_{\lambda_i>0} \lambda_i^n -\sum_{\lambda_i<0} |\lambda_i|^n, &\text{for} \; n=2m-1
\end{align}
Approximate values for these moments can be obtained from a Quantum Monte Carlo routine
\cite{RenyiMelko,Lauchli3}, using the index contractions given in ref.~\cite{PT}.
(We use different names for the moment sets to avoid confusion between
$\nu_1=$Tr$(\rho^{T_2})=1$ and $\mu_1=||\rho^{T_2}||_1,$
which is the quantity we are trying to estimate!)
The moments of the partially transposed density matrix have two distinct analytic
continuations \cite{CCT-EN-FT}. The analytic continuation of the even-powered moments
$\widetilde{\mu}(n)$ can be obtained by allowing $n$ to take continuous values in
\eqref{evenMmts}, and it has the limiting behaviour
\begin{equation}\label{evenContinuation}
\lim_{n \rightarrow 1} \widetilde{\mu}(n) = ||\rho^{T_2}||_1.
\end{equation}
The odd powered moments $\nu_n$ have their own analytic continuation, which must tend to $1$
in the limit as $n\rightarrow 1$ since the partial transpose is a trace-preserving map.
We will not be using an analytic continuation for the odd moments in this paper, only the
odd-powered moments $\nu_n$ themselves.
It can be seen by repeated differentiation with respect to $n$ that the even analytic
continuation $\widetilde{\mu}(n)$ is downward convex.
Furthermore, if we take the natural logarithms of the even moments $\mu_{n}$ it can be seen
by repeated differentiation that $\ln(\widetilde{\mu}(n))$ is also downward convex.
Thus if we pick any two points on $\widetilde{\mu}(n)$ and join them with a two-parameter
exponential fitting curve
$g_1(n)=M\Lambda^{n},$
where $\Lambda$ is a coarse-graining of the true eigenvalues, then the curve $g_1(n)$ will
always be above $\widetilde{\mu}(n)$ when it is between the two fitting points and below
$\widetilde{\mu}(n)$ outside that interval.
If we choose our fitting points to be the values of $\mu_2$ and $\mu_4,$ this gives us a simple
lower bound for the negativity using only four replicas.
There are systems for which this lower bound would be exact \cite{TopoRenyiDensity}
but in general it will not be very accurate.
However it is possible to use the convexity of this problem to obtain tighter lower bounds using
more moments.
In this paper we will develop methods for bounding the set of admissable values for the
unknown first moment $||\rho^{T_2}||_1$ based on those for studying a class of problems
related to Laplace transform inversion, known as Hausdorff moment problems \cite{Hausdorff}.
The moments $\{c_k\}_{k=0}^{n}$ are defined in terms of an integral over a finite range
\cite{Akhiezer,ShohatTamarkin}
\begin{equation}\label{classicalForm}
c_k = \int_a^b \lambda^k \sigma(\lambda) d\lambda.
\end{equation}
where the distribution $\sigma(\lambda)$ must be a non-negative function.
If we only know a finite number of moments, we are dealing with an instance of the truncated
moment problem.
The precise statement of the truncated moment problem is: Given a set of moments
$\{c_k\}_{k=0}^n,$
find (one or more) distributions $\sigma(\lambda)$ consistent with the given moment values.
These integrals can be replaced by sums not only for finite dimensional systems, but also for extremal solutions of continuous systems, which will allow us to extremize integrals of the
form
\begin{equation}\label{MarkovProblem}
I = \int_a^b \Omega(\lambda)\sigma(\lambda)d\lambda,
\end{equation}
where we have only incomplete knowledge of the distribution $\sigma(\lambda).$
\subsection{Properties of the set of possible solutions}
It is known \cite{KreinNudel} that the extreme points of the set of solutions
(known as the \emph{canonical representations} of the moment system $\{c_k\}_{k=0}^{n}$)
are distributions with a minimal number of point masses. This set of representations have
their own extrema, known as the \emph{principal representations}.
We will refer to the positions of the weights as the ``roots'' of the representation, following
\cite{KarlinStudden}.
All other solutions lie in the convex hull of the boundary points \cite{KarlinShapleyOpus}.
As this is a continuously parametrized set infinite combinations are possible, such as those
obtained by maximum entropy methods.
The lower bounds found in this paper reflect the fact that the utility of a given supply of
entanglement to a set of agents is limited by their knowledge of how much entanglement they have.
If the parties to a distillation attempt are only able to perform operations on a small number
of copies of a state at any given time, they may be unable to detect the presence of
some entanglement and therefore be unable to distill it; there are states for which many
copies are required for any distillation to succeed \cite{WatrousManyCopies}.
It is also important to note that the positions of the point masses in the canonical
representations are not arbitrary; one cannot simply pick $\kappa$ roots and then tune the
weights at those positions to get a solution consistent with the moment constraints.
Instead, the roots of the canonical representations are continuous functions of each other:
move one, and the others adjust their positions to maintain consistency with the moment
constraints.
It can be shown (subject to some conditions on $\Omega(\lambda),$ which include continuity
and strict convexity over the range of integration \cite{KreinNudel,KarlinStudden})
that the set of distributions giving the extrema of the integral in \eqref{MarkovProblem}
will always include a pair of principal representations \cite{KarlinStudden} unless the
solution is unique, whereupon the two principal representations coincide.
The \emph{lower principal representation} has that name because it achieves the lower bound
for the simplest versions of the extremization problem in \eqref{MarkovProblem} and many
other versions of this problem.
(The other principal representation is known as the \emph{upper principal representation}
because it usually corresponds to the upper bounds in such problems.)
Some of the conditions on $\Omega(\lambda)$ refer to the existence conditions for the moment
problem to be soluble, so we will return to them after we have introduced the existence
conditions below.
It should be noted that it is possible to relax some of the conditions on $\Omega(\lambda)$
in certain special cases and still be able to find the extrema using the principal
representations, though this must be done with care and the methods tend not to generalize.
However, some of the more straightforward generalizations that can be made have the effect
of reversing the roles of the extremal representations, so that the lower principal
representation can correspond to the upper bound, and vice versa\cite{KarlinStudden}.
\subsection{Existence conditions}\label{Econds}
The precise form of the conditions for the Hausdorff moment problem to have a solution
depend on whether we have an odd or an even number of moments available
\cite{ShohatTamarkin,KreinNudel,Akhiezer}
because the extremal solutions are defined by an even number of parameters.
Thus if we know an even number of moments the problem will be well-constrained,
and if not it will be ill-constrained.
For the well-constrained case when we are given an even number of moments $\{c_k\}_0^{2m+1}$
a necessary and sufficient condition for a solution to exist is that the determinants of the
Hankel matrices
\begin{equation}\label{WellHankela}
H_1(k)= \sum_{i,j=0}^{m} (c_{i+j+1}-a c_{i+j})|i\rangle \langle j|
\end{equation}
and
\begin{equation}\label{WellHankelb}
H_2(k)= \sum_{i,j=0}^{m} (b c_{i+j}-c_{i+j+1})|i\rangle \langle j|
\end{equation}
must be greater than or equal to zero for all $k \leq m.$
If at least one of these inequalites is zero, the system is singularly positive and there is
only one solution.
If all the inequalities are strict, then the system is called \emph{strictly positive}.
If instead we have an odd number of moments $\{c_k\}_{k=0}^{2m}$ we have an ill-constrained
instance. Then the determinants of the set of Hankel matrices
\begin{equation}\label{BadHankel1}
H_1(k)= \sum_{i,j=0}^{m} c_{i+j}|i\rangle \langle j|
\end{equation}
must be non-negative for all $k\leq m,$ and the determinants of
\begin{equation}\label{BadHankelab}
H_2(k)= \sum_{i,j=0}^{m-1} ((a+b)c_{i+j+1}-abc_{i+j}-c_{i+j+2})|i\rangle \langle j|
\end{equation}
for $k=0,\ldots, m-1$ must also be greater than or equal to zero.
If we have a strictly positive instance, there will be infinitely many distributions that are
consistent with the set of known moments $\{c_k\}_{k=0}^{n}.$
If we have an even number $2\kappa$ of constraints, the canonical representations are
those which consist of no more than $\kappa+1$ roots, with the exception of one with
$\kappa$ roots, which will be the lower principal representation.
If there are $2\kappa+1$ moments, the canonical representations are the those with $\kappa+1$
roots.
For these systems the lower principal representation has $\kappa+1$ roots, and occurs when one
of the roots reaches the lower integration endpoint, $a.$
The other $\kappa$ roots occur within the interior of the integration range $[a,b].$
Thus the well-constrained lower principal representations tend to collect its roots in the
interior of the range of integration, whereas the ill-constrained version tends to push as
much spectral weight as it can towards the lower end of the range of integration.
In the event that the instance is singularly positive, there is a loss of independence in the
constraint equations and the unique solution will have fewer point masses, mimicking one with
fewer moments.
However, only instances with extremely simple spectra can be singularly positive when
constrained by only a small number of moments: the eigenvalues of such a spectrum can only
take at most $\kappa$ distinct values.
More typical systems, in which the eigenvalues can take many different values or even feature
quasi-continuous bands will have true spectra that are safely inside the interior of the set
of all possible distributions consistent with the known moments.
Thus if a given instance outright fails to satisfy the corresponding existence conditions, we
should interpret this as meaning that the data from at least the highest order moment involved
is too noisy to use. If this happens we can progressively omit the highest order moments until
we are left with a system of moments that admit a solution.
\subsection{Characteristic equations for the roots of the lower principal representations}\label{chareqns}
For $n=2m-1,$ the roots of the lower principal representation are the values of $\lambda$
which solve the equations \cite{KreinNudel}:
\begin{equation}\label{wellCharEqn}
\det
\begin{pmatrix} c_0 & c_1 & \ldots & c_{m-1} & 1 \\
c_1 & c_2 & \ldots & c_{m} & \lambda \\
\vdots & \vdots & \ddots & \vdots & \vdots \\
c_{m} & c_{m+1} & \ldots & c_{2m-1} & \lambda^{m}
\end{pmatrix}
=0.
\end{equation}
Likewise for $n=2m,$ the roots for the lower principal representation are the solutions of
\begin{equation}\label{illCharEqn}
(\lambda-a)\det
\begin{pmatrix} c_1-ac_0 & \ldots & c_{m}-ac_{m-1} & 1 \\
c_2-ac_1 & \ldots & c_{m+1}-ac_{m} & \lambda \\
\vdots & \ddots & \vdots & \vdots \\
c_{m+1}-ac_{m} & \ldots & c_{2m}-ac_{2m-1} & \lambda^{m}
\end{pmatrix}
=0.
\end{equation}
We will call the multiplicities of the eigenvalues $M_i.$ Then we can write
\begin{equation}\label{muSetupGen}
c_n = \sum_i M_i \lambda_i^n = \int_{a}^{b} \lambda^{n-1}\rho^{T_2}d\lambda,
\end{equation}
for integer values of $n.$ The right hand side is almost of the form of a ``classical'' moment
problem\cite{ShohatTamarkin}, in \eqref{classicalForm}.
Since $\rho^{T_2}$ is not positive semi-definite it cannot play the role of the distribution
$\sigma(\cdot),$ so we must choose an alternative. Note that we could define an alternative
measure in eqn.\eqref{muSetupGen} by relabelling $n$ by a ``shift'' of $\tau,$ i.e., defining
the distribution to be $\sigma_{\tau}(\lambda)=\lambda^{-\tau}(\rho^{T_2})^{1+\tau}.$
\subsection{Backwards extensions of moment systems}
It turns out that choosing $\tau=1$ and $\sigma(\lambda):=(\rho^{T_2})^2$ doesn't work because
choosing $\Omega(\lambda):=1/\lambda$ leads to a singularity at $\lambda=0$ when we try to
extremize the integral in \eqref{MarkovProblem}.
Instead we will have to shift the zero-position of the problem in the opposite direction
(i.e., choose $\tau=-1$) and define $\sigma(\lambda)=(\rho^{T_2})^{0},$ so that
\begin{equation}
\mu_0 = \int_{a}^{b} \sigma(\lambda)d\lambda.
\end{equation}
The unknown zeroth moment $\mu_0$ is the number of non-zero eigenvalues.
This means that the distribution $\sigma(\lambda)$ will be analogous to a state-counting
function, modulo the fact that the eigenvectors in question were produced by the partial
transpose operation, rather than being physical states.
However, we may not have any \emph{a priori} knowledge of the value of $\mu_0,$ apart from the
dimension of the Hilbert space in which our problem lives. Even if we have restricted ourselves
to some variational basis, we can't just use the dimension of that, as the partial transpose
map may take the density matrix to something outside the original set of states.
Therefore we will have to perform what is known as a \emph{backwards extension}
(in the literature on the moment problem)
to estimate $\mu_0,$ so we can then ``step forward'' to estimate $||\rho^{T_2}||_1$ by using
$\Omega(\lambda):=\lambda.$ The existence conditions are necessary and sufficient for the
moment problem to have at least one solution. We may therefore find a lower bound for $\mu_0$
by using those existence conditions.
The need for a backwards extension is another reason for restricting this analysis to use only
the lower principal representations, because the upper principal representations are generally
not well-defined for backwards extensions.
Finding expressions for them that appear to have the correct form is not difficult;
the hard part of the problem is showing that they are informative estimates, or even that they
correspond to the extrema at all.
The fundamental reason why the problem of finding upper principal representations for backwards
extensions is hard can be seen from the original intuition behind moment problems. Consider a rod
on a pivot that has an unknown mass distribution along its length.
Given only the moments of order two or higher, find the maximum possible mass of the rod.
This is generically insoluble if there could be a finite mass accumulation at the pivot, because
its presence cannot be detected by the other moments but it will still contribute to the total
mass of the rod.
Thus finding the upper principal representation of the mass distribution of the bar is impossible
unless we can find the right kind of additional information about the problem.
By contrast the lower principal representations require only a lower bound on $\mu_0,$ which can
be obtained using the existence conditions, above.
Once we have performed the backwards extension to lower bound $\mu_0,$ we can then proceed to
solve the characteristic equations for the lower principal representations. Finding the weights
for those roots is then a linear Gaussian elimination problem.
The resulting estimates for $||\rho^{T_2}||_1$ are indeed extremal, though it turns out that some
of them are actually upper bounds, in an example of the role-reversal phenomenon pointed out in
ref.\cite{KarlinStudden}.
The final step is to perform the extremization in \eqref{MarkovProblem}, which requires that
we revisit the conditions on $\Omega(\lambda)$ required for this method to work. This function
is continuous, but it isn't strictly convex (only convex) so we need to check the more
general conditions in \cite{KreinNudel,KarlinStudden}.
For our purposes, once we have chosen to back-step the problem, the conditions on
$\Omega(\lambda)=\lambda$ will amount to ensuring that the determinant
\begin{equation}\label{TsysCheck}
\det
\begin{pmatrix} 1 & \lambda_0^2 & \ldots & \lambda_0^n & \lambda_0 \\
1 & \lambda_1^2 & \ldots & \lambda_1^n & \lambda_1 \\
\vdots & \, & \ddots & \, & \vdots \\
1 & \lambda_{n+1}^2 & \ldots & \lambda_{n+1}^n & \lambda_{n+1}
\end{pmatrix}
\end{equation}
does not change sign for any pairwise distinct values of $\{\lambda_i\}_{i=0}^{n+1}$
in $[a,b].$
A cyclic permutation of the columns shows that this will revert to the standard existence
criterion for the power moment problem, albeit with the sign reversed for one of the cases.
(This will lead to a role-reversal between the upper and lower principal representations.)
\section{Obtaining the estimates}
The most general range of integration $[a,b]$ for this problem is between $a=-1/2$ and
$b=1,$ as shown in \cite{computEnt}.
However we will continue to write the integration limits as $a$ and $b,$ because there may
be situations in which the range of the eigenvalues is known to be smaller and incorporating
that information will yield tighter bounds. Thus we can write
\begin{equation}\label{muSetupOdd}
c_n = \sum_i M_i \lambda_i^n \int_{a}^{b} \lambda^{n-1}\rho^{T_2}d\lambda.
\end{equation}
In the context of an experiment when we might have only three replicas available, the only
non-trivial existence condition is that $\nu_3-\mu_2^2\geq 0,$ since
$\nu_1=$Tr$(\rho^{T_2})=1.$
Given that, we may proceed directly to obtaining a lower bounds for $\mu_0$ and hence for
$\mu_1.$
Therefore we will exhibit the initial safety checks for four and five replicas to show the
forms for even and odd numbers of replicas respectively. For the initial existence check
the problem is in the well-constrained form if the largest number of replicas used is even.
Thus for four replicas eqn.\eqref{WellHankela} becomes
\begin{equation}
\det\begin{pmatrix} (\mu_2-a) & (\nu_3-a\mu_2) \\
(\nu_3-a\nu_2) & (\mu_4-a\nu_3)
\end{pmatrix}
\geq 0,
\end{equation}
and eqn.\eqref{WellHankelb} becomes
\begin{equation}
\det\begin{pmatrix} (b-\mu_2) & (b\mu_2-\nu_3) \\
(b\mu_2-\nu_3) & (b\nu_3-\mu_4)
\end{pmatrix}
\geq 0.
\end{equation}
Conversely for five replicas we are initially dealing with the ill-constrained version of the
problem, so we must have that eqn.\eqref{BadHankel1} becomes
\begin{equation}
\det\begin{pmatrix} 1 & \mu_2 & \nu_3 \\
\mu_2 & \nu_3 & \mu_4 \\
\nu_3 & \mu_4 & \nu_5
\end{pmatrix}
\geq 0
\end{equation}
and eqn.\eqref{BadHankelab} is now
\begin{equation}
\det\begin{pmatrix} (a+b)\mu_2 -ab -\nu_3 & (a+b)\nu_3-ab\mu_2-\mu_4 \\
(a+b)\nu_3-ab\mu_2-\mu_4 & (a+b)\mu_4+ab\nu_3-\nu_5
\end{pmatrix}
\geq 0,
\end{equation}
where we have used the fact that $\nu_1=1.$
We now perform the backwards extension, which has the effect of interchanging the
well-constrained instances with the ill-constrained ones, as we have inserted an extra
moment-like parameter into the problem.
Rather than keep switching back and forth between the two types of solution, we will work
through first the well-constrained case, and then the ill-constrained one.
\subsection{The well-constrained lower bounds}
This is the form to use if we start with data from $3,5,7\ldots$ replicas.
We will use the three replica case as an example. The constraints for five replicas work the
same way, only the matrices will be correspondingly larger.
There are two backstep constraints for the three replica case. The first one is
\begin{equation}\label{lowerBnda}
\det\begin{pmatrix} (1-a\mu_0) & (\mu_2-a) \\
(\mu_2-a) & (\nu_3-a\mu_2)
\end{pmatrix}
\end{equation}
and the second:
\begin{equation}\label{lowerBndb}
\det\begin{pmatrix} (b\mu_0-1) & (b-\mu_2) \\
(b-\mu_2) & (b\mu_2-\nu_3)
\end{pmatrix}.
\end{equation}
The lower bound for $\mu_0$ is whichever one of these gives the higher value.
The lower bound for $\mu_1$ can then be found by substituting that value for $\mu_0$ in the
lower principal characteristic equations, and solving this to find the roots
$\lambda_1,\lambda_2$ for the lower principal representation.
We will follow the method in ref.~\cite{KreinNudel}. The relevant characteristic equation
is \eqref{wellCharEqn}.
\begin{equation}
\det\begin{pmatrix}
\mu_0 & 1 & 1 \\
1 & \mu_2 & \lambda \\
\mu_2 & \nu_3 & \lambda^2
\end{pmatrix}
=0.
\end{equation}
For larger systems these will need to be solved by numerical methods.
Now we substitute for $\mu_0$ with the lower bound we found above,
to obtain the characteristic equation for the lower bound,
which can be solved using standard methods.
We will label the lower root $\lambda_1,$ which is the only one that can be negative when
we have only three moments. If $\lambda_1 \geq 0$ then our lower bound for $\mu_1$ is zero.
If $\lambda_1$ is negative, we need to find its weight $M_1.$
We can do this by Gaussian elimination, starting from
\begin{equation}
c_j = \sum_i M_i \lambda_i^j
\end{equation}
where $c_{2k}=\mu_{2k}$ and $c_{2k+1}=\nu_{2k+1}.$ In this lowest-order case $j=2,3$ only.
We do not need to find the weight of the positive eigenvalue, $M_2,$ as the negativity is the
sum of the absolute values of the negative eigenvalues only. Thus we obtain a lower bound for
the negativity from three replicas:
\begin{equation}\label{N3lower}
\mathcal{N}_3 \geq \frac{\mu_2\lambda_2-\nu_3}{\lambda_1(\lambda_1-\lambda_2)},
\end{equation}
when $\lambda_1<0.$
For higher order systems where we have more moments to work with, the number of eigenvalues
that may be less than zero is unknown;
indeed it is possible for $\rho^{T_2}$ to have more negative eigenvalues than it does positive
ones \cite{Rana,NatLargeNPT}, and still satisfy Tr$(\rho^{T_2})=1.$ Nevertheless, the basic
principle remains that while we will need all of the eigenvalues to find the estimate, we will
only need to find the weights of the negative eigenvalues: the weights of the positive
eigenvalues can just be eliminated. Note that for strictly positive moment systems this
estimate will almost invariably be an infimum instead of a minimum.
This is because we have not invoked the fact that $\sigma(\lambda)$ is a counting function
anywhere in the calculation above, so the weights have not been constrained to only take
integer values.
Therefore unless $M_1$ and $M_2$ happen to take integer values by coincidence, this lower
bound cannot actually be obtained.
\subsection{The ill-constrained lower principal representation}
This is the form required if we start with data from $4,6,8,\ldots$ replicas.
We will work through the four replica case as an example. The backwards extension constraints
for $\mu_0$ are those in eqns.~\eqref{BadHankel1} and \eqref{BadHankelab}, starting from
$i=j=0,$ so eqn.~\eqref{BadHankelab} becomes
\begin{equation}
\det \begin{pmatrix}
(a+b) -ab\mu_0 -\mu_2 & (a+b)\mu_2 -ab -\nu_3 \\
(a+b)\mu_2 -ab -\nu_3 & (a+b)\nu_3 -ab \mu_2 -\mu_4
\end{pmatrix}
\geq 0.
\end{equation}
Once again, the lower bound is given by the larger solution for $\mu_0.$
The lower principal characteristic equation \eqref{illCharEqn} for four replicas
becomes
\begin{equation}
(\lambda-a)\det\begin{pmatrix} (1-a\mu_0) & (\mu_2-a) & 1 \\
(\mu_2-a) & (\nu_3-a\mu_2) & \lambda \\
(\nu_3-a\mu_2) & (\mu_4-a\nu_3) & \lambda^2
\end{pmatrix}
=0.
\end{equation}
This
will have three roots, one of which will always be at $a,$ and the larger interior
root $\lambda_2>0.$ Of the interior roots, only $\lambda_1$ may be of either sign.
This type of lower principal representation accumulates as much weight as possible at the
lower end of the range of integration.
Recall that the negativity is the sum of the absolute values of the negative eigenvalues only.
As such, a principal representation that tries to maximize the size and weights of those will
actually give us an upper bound for the negativity, so the principal representations will swap
roles in the ill-constrained cases.
Note that the estimate may be non-zero even if $\lambda_1>0$ because there must be a root
at $a.$ However this does not necessarily mean that our estimate must be greater than or
equal to $a,$ because we have not incorporated the constraint that the weights must take
integer values. Instead the requirement that there be a root at $a$ even though its weight
may be less than one indicates that our method will generally give a supremum rather than
a maximum.
We can only omit finding weights for eigenvalues that we know are positive.
Since we don't know the sign of $\lambda_1$ in general we will have to allow for both
possibilities.
As before, $M_1$ and $M_2$ will be the weights for the eigenvalue with the same label, and
we will label the weight of root at $a$ as $M_a.$ So our estimate can be written thus
\begin{equation}
\mathcal{N} = \frac{1}{2}((|\lambda_1|-\lambda_1)M_1+M_a).
\end{equation}
We can now find the counting weights $M_1$ and $M_a$ by Gaussian elimination, starting from
\begin{equation}
c_j = M_a a^j + \sum_i M_i \lambda_i^j
\end{equation}
where $c_{2k}=\mu_{2k}$ and $c_{2k+1}=\nu_{2k+1}$ as before. In this case $j=0,\ldots,4.$
A little algebra gives us the following:
\begin{equation}
M_a a=\frac{\lambda_1^2(\lambda_2-\mu_2)-\lambda_2\nu_3+\mu_4}{(a-\lambda_2)(a^2-\lambda_1^2)}
\end{equation}
and
\begin{equation}
M_1 \lambda_1=\frac{\lambda_2\nu_3-\mu_4-a^2(\lambda_2-\mu_2)}{(\lambda_2
-\lambda_1)(\lambda_1^2-a^2)}
\end{equation}
and therefore
\begin{multline}
\mathcal{N}_4 \leq
\frac{\lambda_1^2(\lambda_2 -\mu_2)-\lambda_2\nu_3+\mu_4}{(a-\lambda_2)(a^2-\lambda_1^2)} \\
+\left(\frac{|\lambda_1|-\lambda_1}{2\lambda_1}\right)
\frac{\lambda_2\nu_3-\mu_4-a^2(\lambda_2-\mu_2)}{(\lambda_2-\lambda_1)(\lambda_1^2-a^2)}.
\end{multline}
\section{Summary and open questions}\label{SummaryEtc}
In this paper we have shown how to obtain families of infima and suprema for the entanglement
negativity, given only a few, low order moments of the partially transposed density matrix.
Estimates for the entanglement negativity can be obtained even in the absence of information
about the number of non-zero eigenvalues in the system's partially-transposed density matrix.
A lower bound can be obtained with just three replicas, and an upper bound with four.
The methods in this paper can be extended further if more replicas are available,
though the existence conditions will need to be checked at each order because the method
becomes progressively more sensitive to statistical noise as the order of the approximation
increases.
The method presented in this paper is fully model-independent. If model-specific information
can be included in the calculation, then more structure can be resolved within the spectrum.
After the first version of this manuscript appeared, an analytic derivation of the
distribution of the eigenvalues of the partially transposed density matrix for one-dimensional
conformal field theories was presented in \cite{N-Spectrum}, including detailed numerical
results.
There are still some open questions regarding the interpretation of the bounds. In particular,
are the families of lower bounds in this paper entanglement monotones
\cite{VidalMonotones,simplifyMonoConds,DiffCondMonotones},
and can they be implemented as quantum circuits\cite{BrunPolynomial} to measure entanglement,
or even perform batch-size limited entanglement distillation? If an entanglement distillation
protocol is optimal, its yield will be an entanglement monotone\cite{VidalMonotones}.
However, the converse is not necessarily true: the entanglement of formation is a monotone but
does not correspond to the yield of any distillation protocol, because not all entanglement can
be distilled\cite{boundEntanglementH3}.
\begin{acknowledgments}
This work is supported in part by the M. Hildred Blewett Fellowship of the American
Physical Society, www.aps.org.
I would also like to thank the Department of Physics and Astronomy at the University of
Waterloo, The Institute for Quantum Computing for additional support,
and The Perimeter Institute for Theoretical Physics for hospitality.
I would also like to thank Roger Melko, Alioscia Hamma, William Donnelly, Todd Brun,
Marco Piani and Vern Paulsen for some interesting discussions,
and Yichen Huang for informative comments on the first version of this manuscript.
\end{acknowledgments}
|
\section{Introduction}
Despite the fact that the standard theory of gravity, general relativity (GR),
is already around one century old, there has always been interest
in finding viable modifications to it. In particular, the discovery of the late-time acceleration
of our Universe \cite{Riess:1998cb,Perlmutter:1998np} driven by some dark energy has
led to an additional motivation, as GR requires a technically unnatural
cosmological constant (CC) in order to be compatible with observations.
The list of problems with the standard theory goes much further (see for example Ref.~\cite{Bull:2015stt} for a recent review): GR
is not renormalizable and can only be regarded as an effective field
theory (EFT). Furthermore, the formation of structure at early times needs
an additional inflationary epoch, and even the requirement of some additional
dark matter might be the consequence of the inability of GR to properly describe
the evolution of the cosmic structure. It is however not only these
problems that make a search for modifications of GR attractive; there
is also the more fundamental question of which classes of theories are
allowed and consistent.
Under certain assumptions, Vermeil and Cartan independently proved
that Einstein equations are the only allowed field equations to describe gravity \cite{Vermeil1917,Cartan1922}.
In particular, if a rank-2 tensor $K$ is naturally constructed from
only a pseudo-Riemannian metric $g$, and is symmetric, divergence-free,
only second order in the derivatives of $g$, and linear in these derivatives,
then $K$ has to be a linear combination of the Einstein tensor and
the metric itself. Later, Lovelock showed that the requirements of symmetry
and linearity are redundant in four dimensions \cite{Lovelock1972}.
A generalization of this theorem has recently been suggested by Navarro
and Sancho \cite{NavarroSancho2007}, who replaced the assumptions
of the number of dimensions and absence of higher order derivatives
by a simpler requirement that $K$ is homogeneous, i.e., $K(\lambda^2 g) = \lambda^w K(g)$ $\forall g$, $\forall \lambda > 0$, and of weight $w > -2$.
Thinking about modifying GR can be translated into relaxing these
so-called Lovelock assumptions. Higher dimensional spacetimes, as well
as unnaturalness (understood in the mathematical sense, i.e., either breaking of locality or generally
non-$\mathcal{C}^{\infty}$), enable a richer phenomenology and do
not necessarily require a CC in order to fit current observations.
Additionally, a pseudo-Riemannian geometry is quite restrictive as
it both implies a vanishing torsion and enforces the connection to
be metric-compatible. Certainly the strongest assumption, however, is
the dependence on the metric only. Consequently, most theories of
modified gravity assume additional fields that can be either scalar,
vector, or tensor.
There is, however, one assumption that usually stays untouched: the
absence of higher order derivatives. An old theorem from Ostrogradski
states that non-degenerate Lagrangians that lead to third or higher
order derivatives in the equations of motion (EoM) always house an
additional ghost, i.e., a degree of freedom with the wrong kinetic sign.
But even degenerate Lagrangians producing third order derivatives
are affected by ghosts \cite{MotohashiSuyama2014}. The consequences
that come along with a ghost are usually believed to be fatal (see, e.g., Refs. \cite{Twain1875,Woodard2006}). Such a negative energy
mode could drive the classical theory into an instability. Even though
this might still be acceptable as long as the theory is in agreement
with observations, it indeed limits the number of viable theories
drastically. The real catastrophe appears, however, at the quantum level:
ghost fields can decay into ordinary matter fields by reaching
arbitrarily large negative energy states. And, even worse, this decay
will practically happen instantaneously (see Refs. \cite{Woodard2006,Sbisa2014}
for more details). Such a theory cannot describe a stable vacuum
and therefore has to be ruled out. The origin of the fast decay lies
in an integration over the entire phase space when computing scattering
amplitudes which diverge in the ultraviolet (UV) region. Therefore,
the only way to tame the ghost is to modify the integration in the
UV. In Ref. \cite{Carroll2003}, the authors suggested that new operators beyond the EFT would allow us to cut this integral and, therefore,
a theory with ghosts could theoretically be cured. In fact, it has been shown
that the vacuum in simple theories with two oscillators, of which one is a ghost, can indeed have a decay time that is larger than
the Hubble time \cite{Carroll2003,KaplanSundrum2005} (see also Refs. \cite{Dyda2012,Ramazanov2016} for discssions of ghosts in Chern-Simons and Ho\v{r}ava-Lifshitz theories, respectively). The energy
scale at which new physics might enter and break Lorentz invariance (LI) can be low enough
to slow down the vacuum decay sufficiently and circumvent any violation
of experimental constraints, but at the same time be high enough to be above the
cutoff of the EFT. In fact, a Lorentz breaking (LB) does not render the theory unappealing as long as it occurs above the EFT cutoff scale.
In this work, we discuss a theory of modified gravity that automatically
introduces a ghost instead of adding a simple ghost field by hand
to a well behaved theory. In fact, as will be shown, many properties of such a theory, like the decay
time of the vacuum, may be significantly different,
and therefore the conclusions from simpler toy models should not be adopted blindly. In order to modify GR suitably, we assume a massive graviton. Even though the
idea of studying massive gravity is very old \cite{FierzPauli1939},
ghost-free non-linear theories were discovered only recently \cite{HassanRosen2012c,Creminelli2005,deRhamGabadadze2010,HassanRosen2012b,HassanRosenSchmidt-May2012,deRhamGabadadzeTolley2011,HassanSchmidt-MayStrauss2012c}.
Here we use the so-called ghost-free de Rham-Gabadadze-Tolley (dRGT)
theory to construct a theory of a massive graviton with an
additional Boulware-Deser (BD) ghost \cite{BoulwareDeser1972}, which we then dub {\it haunted massive
gravity} (HMG). We first study the classical behavior of the theory for a Friedmann-Lema\^{i}tre-Robertson-Walker (FLRW) background
in order to identify the models that do not introduce
potentially dangerous instabilities already at this level. With HMG we find the first theory of a canonical non-linear massive gravity which possesses models that are free of any background pathologies, and allows for dynamical, even self-accelerating, FLRW solutions. We finally discuss the quantum stability of the viable models by computing the ghost and vacuum decay rates in HMG.
Although the theory that is discussed in this work has some nice features,
e.g., it can provide a solution to the dark energy problem, it is certainly not the most promising contender of GR. For example, the construction of the mass term is mainly based on keeping simplicity, and the strong coupling scale of the theory is very low, losing predictivity on small scales. This work, however, does not intend to present a new theory of modified gravity in order to address the problems with GR, but rather to examine the behavior of ghosts appearing in more realistic
theories of gravity. We find that the scattering processes that dominate in the ghost and vacuum decay rates
do not coincide with those that appear in theories with two coupled
canonical scalar fields where one of the fields is a ghost. While in such a
simple scenario the decay time has been found to scale only quadratically
with the cutoff at which Lorentz violation (LV) occurs \cite{Carroll2003}
(see also Refs. \cite{KaplanSundrum2005,Cline2003} for a discussion
on decay rates for other setups), we find a completely different scaling for HMG. Furthermore, we expect the
type of interaction that we find in this work as the most important one in HMG to
in fact determine the decay time also in many other theories of gravity
with a present ghost mode.
\section{HMG with $\text{d}$RGT limit}
Since we are interested in a theory of a massive graviton with an additional
BD ghost, we could simply study any non-linear theory that does not
coincide with the dRGT theory. However, we would like to keep a ghost-free
limit, and therefore, we start with dRGT and modify the tuning between
the coefficients of interaction terms.
The dRGT massive gravity can be written as \cite{HassanRosen2011a,deRhamGabadadzeTolley2011}
\begin{equation}
S_{\text{dRGT}}=M_{\text{P}}^{2}\int \text{d}^{4}x\sqrt{-g}\left[R+2m^{2}U(\mathbb{K})\right],
\end{equation}
where $U(\mathbb{K})$ is the mass term, which depends on the eigenvalues
of $\mathbb{K}\equiv\sqrt{g^{-1}f}-\mathbb{1}\equiv\mathbb{X}-\mathbb{1}$,
and is equivalent to $\sum_{n=0}^{3}\beta_{n}U_{n}$ with
$U_{n}\equiv e_{n}\left(\mathbb{X}\right)$, $e_{n}$ being the elementary symmetric polynomials of the eigenvalues
of $\mathbb{X}$. Additionally, $f$ is a non-dynamical symmetric rank-2 tensor field, and $\beta_{n}$ depend
on the two free dimensionless parameters of the theory, $\alpha_{3}$
and $\alpha_{4}$ \cite{HassanRosen2011a}:
\begin{align}
\beta_{0} & =6-4\alpha_{3}+\alpha_{4},\\
\beta_{1} & =-3+3\alpha_{3}-\alpha_{4},\\
\beta_{2} & =1-2\alpha_{3}+\alpha_{4},\\
\beta_{3} & =\alpha_{3}-\alpha_{4}.
\end{align}
Let us first discuss the simplest model, the so-called minimal model
\cite{HassanRosen2011a}, and choose $\alpha_{3}$ and $\alpha_{4}$
such that we can switch off the highest order interactions, i.e. $\beta_{2}=\beta_{3}=0$,
and obtain
\begin{align}
\alpha_{3}=\alpha_{4}=1\quad\Leftrightarrow\quad\beta_{0}=3,\;\beta_{1}=-1.
\end{align}
The action in this case becomes
\begin{align}
S_{\text{min}} & =M_{\text{P}}^{2}\int \text{d}^{4}x\sqrt{-g}\left[R+2m^{2}\left(3-\left[\mathbb{X}\right]\right)\right],
\end{align}
where $\mbox{\ensuremath{\left[\mathbb{X}\right]}}$ denotes the trace of $\mathbb{X}$.
If we change the prefactors of the mass term, we then
change either the CC or the graviton mass (or make it tachyonic). Thus, in
order to introduce a ghost, we should switch on higher order interactions.
One possibility would be to remove the CC (which would make the model very appealing especially
for cosmology) and allow for $\beta_{1}$
and $\beta_{2}$ to be non-vanishing. We then find $\beta_{0}=\beta_{3}=0$ together with
\begin{align}
\alpha_{3}=\alpha_{4}=2\quad\Leftrightarrow\quad\beta_{1}=1=-\beta_{2},
\end{align}
which results in the action
\begin{align}
S & =M_{\text{P}}^{2}\int \text{d}^{4}x\sqrt{-g}\left[R+2m^{2}\left(\beta_{1}\left[\mathbb{X}\right]+\frac{1}{2}\beta_{2}\left(\left[\mathbb{X}\right]^{2}-\left[\mathbb{X}^{2}\right]\right)\right)\right]\\
& =M_{\text{P}}^{2}\int \text{d}^{4}x\sqrt{-g}\left[R+2m^{2}\left(\left[\sqrt{g^{-1}f}\right]-\frac{1}{2}\left(\left[\sqrt{g^{-1}f}\right]^{2}-\left[g^{-1}f\right]\right)\right)\right].
\end{align}
Note that this choice, like all other combinations that satisfy $\alpha_{3}+\alpha_{4}>0$,
does not lead to a Higuchi ghost, at least around an FLRW background
for large $H^{2}$ \cite{Fasiello2012}. This is important because
we will use this theory as a ghost-free limit which should
ensure not only the absence of the BD ghost but also the presence of five healthy
graviton degrees of freedom.
We now modify the theory to introduce a ghost.
The simplest way would be to modify the prefactor in one of the interaction
terms which in the linear theory corresponds to a violation of the
Fierz-Pauli (FP) tuning. However, we do not expect this modification to enable dynamical
FLRW solutions for a flat reference metric since the combination of
the Bianchi identities and the conservation of energy-momentum tensor
will still be a constraint for the scale factor, as it has been shown for
dRGT in Ref. \cite{DAmico2011}. One way out could be to introduce a
metric-dependent (and, thus, lapse-dependent) prefactor like $\alpha \left[g^{-1}f\right]$, with $\alpha \in \mathbb R$,
and study the action
\begin{equation}
S=M_{\text{P}}^{2}\int \text{d}^{4}x\sqrt{-g}\left[R+2m^{2}\left(\left[\sqrt{g^{-1}f}\right]+\frac{1}{2}\left(\alpha\left[g^{-1}f\right]-1\right)\left(\left[\sqrt{g^{-1}f}\right]^{2}-\left[g^{-1}f\right]\right)\right)\right].
\end{equation}
Although this theory would certainly allow for dynamical FLRW backgrounds,
we found only unviable solutions for which the scale factor
would become imaginary or the lapse would cross zero, indicating an instability.
Therefore, we consider a slightly more complicated theory in which
both interaction terms are modified, and dub this theory haunted massive
gravity (HMG):
\begin{equation}
S_{\text{HMG}}=M_{\text{P}}^{2}\int \text{d}^{4}x\sqrt{-g}\left[R+2m^{2}\left(\left(1-\alpha_{1}\left(g,f\right)\right)\left[\sqrt{g^{-1}f}\right]-\frac{1}{2}\left(1-\alpha_{2}\left(g,f\right)\right)\left(\left[\sqrt{g^{-1}f}\right]^{2}-\left[g^{-1}f\right]\right)\right)\right],\label{eq:action_HMG}
\end{equation}
with
\begin{align}
\alpha_{1}\left(g,f\right) & \equiv\bar{\alpha}_{1}\mathbb{X}^{2}=\bar{\alpha}_{1}g^{\alpha\beta}f_{\beta\alpha},\\
\alpha_{2}\left(g,f\right) & \equiv\bar{\alpha}_{2}\mathbb{X}^{2}=\bar{\alpha}_{2}g^{\alpha\beta}f_{\beta\alpha},
\end{align}
where $\bar{\alpha}_{i}$ are two dimensionless parameters.
This theory has some interesting properties. Firstly, the limit $\bar{\alpha}_{i}\rightarrow0$
corresponds to the ghost-free dRGT theory, whereas any other values
should introduce a new ghost degree of freedom as it does not coincide with dRGT,
the unique non-linear ghost-free theory of a massive graviton. Secondly,
the additional dynamical factors will enable us to have dynamical FLRW solutions
by modifying the Bianchi constraint, and, finally, we expect the vacuum
to decay more slowly at late times since $\left[g^{-1}f\right]\propto a^{-2}$
for FLRW backgrounds.%
\footnote{This could have an interesting impact on the phenomenology at early
times and might lead to an enhanced creation of particles. The relevant
time period would, however, presumably lie above the cutoff scale of the
theory.%
}
\section{Background cosmology}
From now on, let us fix the reference metric to a flat Minkowski background, i.e.,
$f_{\mu\nu}=\eta_{\mu\nu}$. Since massive gravity with $f_{\mu\nu}=\eta_{\mu\nu}$
breaks diffeomorphism invariance, the lapse of $g_{\mu\nu}$ must not be chosen arbitrarily. For an FLRW background we therefore choose
\begin{equation}
\text{d} s^{2}=-N_{g}^{2}\text{d} t^{2}+a^{2}\text{d} \vec{x}^{2},
\end{equation}
with $N_{g}$ and $a$ denoting the lapse and the scale factor, respectively, and $t$ being cosmic time.
Varying the action (\ref{eq:action_HMG}) with respect to $g_{\mu\nu}$
yields
\begin{align}
-2m^{2}M_{\text{P}}^{2}\delta\left(\sqrt{-g}U_{1}\left(X\right)\right) & =-\sqrt{-g}\beta_{1}m^{2}M_{\text{P}}^{2}g^{\mu\alpha}Y_{(1)\,\alpha}^{\nu}\left(\sqrt{g^{-1}f}\right)\delta g_{\mu\nu},\\
-2m^{2}M_{\text{P}}^{2}\delta\left(\sqrt{-g}U_{2}\left(X\right)\right) & =\sqrt{-g}\beta_{2}m^{2}M_{\text{P}}^{2}g^{\mu\alpha}Y_{(2)\,\alpha}^{\nu}\left(\sqrt{g^{-1}f}\right)\delta g_{\mu\nu},
\end{align}
with
\begin{align}
Y_{(1)}\left(\mathbb{X}\right) & \equiv \mathbb{X}-\mathbb{1}\left[\mathbb{X}\right],\\
Y_{(2)}\left(\mathbb{X}\right) & \equiv \mathbb{X}^{2}-\mathbb{X}\left[\mathbb{X}\right]+\frac{1}{2}\mathbb{1}\left(\left[\mathbb{X}\right]^{2}-\left[\mathbb{X}^{2}\right]\right).
\end{align}
Furthermore, we need the variation of $\alpha_{i}$:
\begin{align}
\delta\alpha_{i}\left(g\right) & =-\bar{\alpha}_{i}g^{\alpha\mu}g^{\nu\beta}\eta_{\beta\alpha}\delta g_{\mu\nu}.
\end{align}
With this, the variation of the mass term yields
\begin{equation}
\sqrt{-g}M_{\text{P}}^{2}V^{\mu\nu}\equiv-m^{2}\sqrt{-g}M_{\text{P}}^{2}\left[\left(\alpha_{1}\left(g\right)-1\right)g^{\mu\alpha}Y_{(1)\,\alpha}^{\nu}+\left(\alpha_{2}\left(g\right)-1\right)g^{\mu\alpha}Y_{(2)\,\alpha}^{\nu}-2U_{1}\left(\mathbb{X}\right)\frac{\delta\alpha_{1}\left(g\right)}{\delta g_{\mu\nu}}-2U_{2}\left(\mathbb{X}\right)\frac{\delta\alpha_{2}\left(g\right)}{\delta g_{\mu\nu}}\right].
\end{equation}
Combining the Bianchi identities with the assumption of a conserved energy-momentum
tensor leads to the Bianchi constraint
\begin{equation}
\nabla_{\mu}V^{\mu\nu}=0,
\end{equation}
which implies
\begin{align}
\left(1+3a^{-2}N_{g}^{2}\right)\left[N_{g}'\left(a\left(a\bar{\alpha}_{1}+6\bar{\alpha}_{2}\right)+2N_{g}\left(2a\bar{\alpha}_{1}+\bar{\alpha}_{2}\right)\right)\right.\nonumber \\
\left.+HN_{g}\left(-6a\bar{\alpha}_{2}+N_{g}^{2}\left(4\bar{\alpha}_{1}-(a-2)a\right)-2N_{g}\left(a\bar{\alpha}_{1}+\bar{\alpha}_{2}\right)+9\bar{\alpha}_{1}a^{-1}N_{g}^{3}\right)\right] & =0.\label{eq:Bianchi_constraint}
\end{align}
Here, $H^2\equiv a'/a$ is the Hubble rate, and a prime denotes a derivative with respect to $t$. In the limit $\bar{\alpha}_{i}\rightarrow0$, Eq. (\ref{eq:Bianchi_constraint}) fixes the scale factor confirming the no-go theorem for
FLRW solutions in dRGT massive gravity. In our case, however, we are
able to switch on the dynamics since the Bianchi constraint now depends on both the scale factor and the lapse. This constraint together with the Friedmann equation
\begin{align}
3H^{2} & =\rho+V_{00}\\
& =\rho+\frac{m^{2}}{a^{4}N_{g}}\left[a^{3}\left(-\left(a\bar{\alpha}_{1}+6\bar{\alpha}_{2}\right)\right)-3a^{2}N_{g}\left(2a\bar{\alpha}_{1}+\bar{\alpha}_{2}\right)+3N_{g}^{3}\left(a\left((a-1)a-3\bar{\alpha}_{1}\right)+3\bar{\alpha}_{2}\right)\right]
\end{align}
and assuming a universe filled with dark matter only,
\begin{equation}
\rho\equiv\rho_{0}a^{-3},
\end{equation}
can be solved numerically. In the limit $a\ll 1$, the combination of the Bianchi constraint and the Friedmann equation provides
\begin{equation}
N_g = \pm \frac{1}{3}\sqrt{\frac{\bar \alpha_2}{\bar \alpha_1} a},
\end{equation}
and implies $H^2 \propto a^{-3}$. Therefore, we find a singularity for $a\rightarrow 0$. The time at which it occurs will be denoted by $t_c$, i.e., $N_g (t_c) = 0$.
Interestingly, for a given $\bar{\alpha}_{1}$ one can find a value for the parameter $\bar{\alpha}_{2}$ that maximizes the timescale of the background evolution by reaching $t_c \rightarrow 0$. One example
for such a model is
\begin{equation}
\bar{\alpha}_{1}=0.9\;\Rightarrow\;\bar{\alpha}_{2}\simeq0.1025.\label{eq:model4plot}
\end{equation}
By solving the background equations numerically, we can search for the
parameter region that leads to $t_c = 0$ and find
that it can be fitted very well linearly with
\begin{equation}
\bar{\alpha}_{2}\simeq\frac{1}{6}\bar{\alpha}_{1}-\frac{2}{45}.\label{eq:stability_curve}
\end{equation}
If we promote the maximization of the classical timescale to a constraint,
then this model will effectively lose one free parameter. Furthermore, Eq. (\ref{eq:stability_curve}) indicates that both $\bar \alpha_1$ and $\bar \alpha_2$ can be of $\mathcal O \left(1\right)$.
Surprisingly, after solving the background equations for the model (\ref{eq:model4plot}) numerically, we find that at late times the effective equation of state parameter
$w_{\text{eff}}<-1/3$ indicates a period of self-acceleration and we thus
have found a candidate for a model that could be able to provide a
solution to the dark energy problem. See Fig. \ref{fig:background_numerical_solution}
for a numerical solution of the background evolution, corresponding to model (\ref{eq:model4plot}).
\begin{figure}[t]
\includegraphics[width=0.45\textwidth]{img/background/a}$\quad$\includegraphics[width=0.45\textwidth]{img/background/H}\\
\includegraphics[width=0.45\textwidth]{img/background/N}$\quad$\includegraphics[width=0.45\textwidth]{img/background/weff}
\protect\caption{Numerical solution of the FLRW background evolution in HMG, corresponding to the model with
$\bar{\alpha}_{1}=0.9$ and $\bar{\alpha}_{2}\simeq0.1025$. The plots show the scale factor (upper left), the expansion rate
(upper right), the lapse (lower left), and the effective equation of state
(lower right). All the quantities are plotted versus the cosmic time, $t$, scaled in such a way that $a(t=1)=1$.}
\label{fig:background_numerical_solution}
\end{figure}
\section{Second order action for HMG around Minkowski}
In order to compute the vacuum decay, we have to first identify the ghost mode.
For this, we perturb the background, expand the action to second order,
and finally integrate out \footnote{Since we are interested in computing only tree-level diagrams in order to discuss the quantum behavior of the theory, the elimination of all auxiliary fields in the action by using their EoM is equivalent to properly integrating out these fields.} all auxiliary fields.
\subsection{Gravity sector}
We now choose to work with perturbations around a Minkowski background. Note that we expect a generalization to an FLRW background to merely modify the decay rate insignificantly. In fact, corrections from an FLRW background are proportional to $H/k$ and become negligible in the high-momentum limit, on which we will focus later. Furthermore, ignoring the cosmological
expansion and, therefore, a smaller volume at early times, should then correspond
to maximizing the decay rate and computing an upper bound for a more
realistic scenario.
In this case, the Bianchi constraint (\ref{eq:Bianchi_constraint})
enforces the lapse to be a constant; here we set $N_{g}=1$.
Furthermore, in order to excite the scalar BD ghost we consider the following scalar perturbations $\delta g$ around the background $\bar g\equiv \eta$:
\begin{equation}
\text{d} s_{\delta g}^{2}=2\left[-\Psi \text{d} t^{2}+B_{,i}\text{d} x^{i}\text{d} t+\left(\Phi\delta_{ij}+E_{,ij}\right)\text{d} x^{i}\text{d} x^{j}\right].\label{eq:pert_ansatz}
\end{equation}
The Einstein-Hilbert action at second order therefore reads
\begin{equation}
S_{\text{EH}}^{(2)}=4M_{\text{P}}^{2}\int \text{d}^{4}x\left(\Phi_{i}^{2}+2\Phi_{i}\Psi_{i}-2\Phi'\Delta E'-3\Phi'^{2}-2B_{i}\Phi_{i}'\right),\label{eq:action_EH}
\end{equation}
where we have used the notation
\begin{equation}
X_{i}Y_{i}\equiv\sum_{j}X_{,j}Y_{,j}\;,\qquad\Delta X\equiv\sum_{j}X_{,jj}.
\end{equation}
We now expand the mass term of action (\ref{eq:action_HMG}) to second order and use
\begin{align}
\sqrt{g^{-1}\eta} & \simeq\sqrt{\bar{g}^{-1}\eta}\left[1-\frac{1}{2}\bar{g}^{-1}\left(\delta g\right)+\frac{3}{8}\bar{g}^{-1}\left(\delta g\right)\bar{g}^{-1}\left(\delta g\right)\right]\\
& =\mathbb{1}-\frac{1}{2}\eta\left(\delta g\right)+\frac{3}{8}\eta\left(\delta g\right)\eta\left(\delta g\right),
\end{align}
to obtain
\begin{align}
\left(1-\bar{\alpha}_{1}\left[g^{-1}\eta\right]\right)\left[\sqrt{g^{-1}\eta}\right]\simeq & \left[\mathbb{1}-\frac{1}{2}\eta\left(\delta g\right)+\frac{3}{8}\eta\left(\delta g\right)\eta\left(\delta g\right)\right]\nonumber\\
& -\bar{\alpha}_{1}\left(\left[\mathbb{1}\right]-\left[\left(\delta g\right)\eta\right]+\left[\eta\left(\delta g\right)\eta\left(\delta g\right)\right]\right)\left(\left[\mathbb{1}\right]-\left[\frac{1}{2}\eta\left(\delta g\right)\right]+\left[\frac{3}{8}\eta\left(\delta g\right)\eta\left(\delta g\right)\right]\right)\\
\simeq & 4\left(1-4\bar{\alpha}_{1}\right)-\left(\frac{1}{2}-6\bar{\alpha}_{1}\right)\left[\eta\left(\delta g\right)\right]-\frac{1}{2}\bar{\alpha}_{1}\left[\eta\left(\delta g\right)\right]^{2}+\left(\frac{3}{8}-\frac{11}{2}\bar{\alpha}_{1}\right)\left[\eta\left(\delta g\right)\eta\left(\delta g\right)\right]
\end{align}
and
\begin{align}
\left(1-\bar{\alpha}_{2}\left[g^{-1}\eta\right]\right)\left(\left[\sqrt{g^{-1}\eta}\right]^{2}-\left[g^{-1}\eta\right]\right)\simeq & \left(1-\bar\alpha_2\left(4-\left[\left(\delta g\right)\eta\right]+\left[\eta\left(\delta g\right)\eta\left(\delta g\right)\right]\right)\right) \times \nonumber\\
& \times \left(\left[\mathbb{1}-\frac{1}{2}\eta\left(\delta g\right)+\frac{3}{8}\eta\left(\delta g\right)\eta\left(\delta g\right)\right]^2-4+\left[\left(\delta g\right)\eta\right] - \left[ \eta \left(\delta g\right) \eta \left(\delta g\right) \right] \right)\\
\simeq & 12\left(1-4\bar{\alpha}_{2}\right)-3\left(1-8 \bar{\alpha}_{2}\right)\left[\left(\delta g\right)\eta\right]+\left(\frac{1}{4}-4\bar{\alpha}_{2}\right)\left[\eta\left(\delta g\right)\right]^{2}\nonumber\\
&+2\left(1-10\bar{\alpha}_{2}\right)\left[\eta\left(\delta g\right)\eta\left(\delta g\right) \right].
\end{align}
Finally, with
\begin{align}
\sqrt{-g} & \simeq\frac{1}{4}\sqrt{-\bar{g}}\left(4+2\left(\delta g\right)_{\;\mu}^{\mu}-\left(\delta g\right)_{\mu\nu}\left(\delta g\right)^{\mu\nu}+\frac{1}{2}\left(\left(\delta g\right)_{\;\mu}^{\mu}\right)^{2}\right)\\
& =\frac{1}{4}\left(4+2\left[\eta\left(\delta g\right)\right]-\left[\eta\left(\delta g\right)\eta\left(\delta g\right)\right]+\frac{1}{2}\left[\eta\left(\delta g\right)\right]^{2}\right)
\end{align}
we find the second order action of the mass term as
\begin{align}
S_{\text{mass}}^{(2)}=M_{\text{P}}^{2}m^{2}\int \text{d} ^{4}x & \left(\frac{1}{4}+\bar\alpha_1-2\bar\alpha_2\right)\left[\eta\left(\delta g\right)\right]^{2}-\left(\frac{1}{4}+3\bar{\alpha}_{1}-8\bar{\alpha}_{2}\right)\left[\eta\left(\delta g\right)\eta\left(\delta g\right)\right].\label{eq:pot_2nd}
\end{align}
For the ansatz (\ref{eq:pert_ansatz}), this becomes
\begin{align}
S_{\text{mass}}^{(2)}= & \frac{1}{2} M_{\text{P}}^{2}m^{2}\int \text{d} ^{4}x\left[-c_{1}B\Delta B+c_{2}\left(\Psi^{2}+\left(\Delta E\right)^{2}\right)+8c_{3}\Phi\Delta E+12c_{3}\Phi^{2}+4c_{4}\Psi\left(\Delta E+3\Phi\right)\right],\label{eq:action_pot}
\end{align}
where we have defined the parameters $c_{i}$ as
\begin{align}
c_{1} & \equiv1+12\bar{\alpha}_{1}-32\bar{\alpha}_{2},\label{eq:def_c1}\\
c_{2} & \equiv-16\bar{\alpha}_{1}+12\bar{\alpha}_{2},\label{eq:def_c2}\\
c_{3} & \equiv1+4\bar{\alpha}_{2},\label{eq:def_c3}\\
c_{4} & \equiv1+4\bar{\alpha}_{1}-8\bar{\alpha}_{2}\label{eq:def_c4}.
\end{align}
Therefore, the (minimal) ghost-free massive gravity corresponds to
the limit $c_1,c_3,c_4\rightarrow1$ and $c_2\rightarrow 0$. Note that we should not necessarily
expect a smooth limit due to the change in the number of degrees of freedom in HMG compared to dRGT.
\subsection{Full action including matter}
For simplicity, we assume the matter sector to contain only a single, minimally-coupled, scalar field
$\varphi$ with mass $m_{\varphi}$, i.e.,
\begin{equation}
S_{\text{matter}}=-\int \text{d} ^{4}x\sqrt{-g}\left(\partial^{\mu}\varphi\partial_{\mu}\varphi+m_{\varphi}^{2}\varphi^{2}\right),\label{eq:action_matter}
\end{equation}
and consider the weak-field limit where only the gravitational sector is expanded to second order and matter is kept unperturbed. The kinetic and mass terms of $\varphi$ are
then given by its coupling to the background metric.
By combining the actions (\ref{eq:action_EH}), (\ref{eq:action_pot}), and (\ref{eq:action_matter}),
we obtain the final leading order action of HMG, containing
a Boulware-Deser ghost and a matter field, around Minkowski (modulo total
derivatives),
\begin{align}
S_{\text{HMG}}^{(2)}=\int \text{d} ^{4}x & \left[4M_{\text{P}}^{2}\left(\Phi_{i}^{2}-2\Delta\Phi\Psi-2\Phi'\Delta E'-3\Phi'^{2}-2B_{i}\Phi_{i}'\right)\right.\nonumber \\
& \left.+\frac{1}{2}m^{2}M_{\text{P}}^{2}\left(c_{1}B_{i}^{2}+c_{2}\left(\Psi^{2}+\left(\Delta E\right)^{2}\right)+8c_{3}\Phi\Delta E+12c_{3}\Phi^{2}+4c_{4}\Psi\left(\Delta E+3\Phi\right)\right)\right.\nonumber \\
& \left.-\left(1+B_{i}^{2}-\left(\Delta E\right)^{2}+3\Phi^{2}+6\Phi\Psi-\Psi^{2}+2\Delta E\left(\Phi+\Psi\right)\right)X_{\varphi}\right],\label{eq:action_HMG_2nd_order}
\end{align}
where we have defined
\begin{equation}
X_{\varphi}\equiv-\varphi'^{2}+\varphi_{i}^{2}+m_{\varphi}^{2}\varphi^{2}.\label{eq:def_X_varphi}
\end{equation}
In total, all the scalar potentials $\Phi$, $\Psi$, $B$, $E$, and the matter field
$\varphi$ should describe at most three propagating scalar degrees of freedom: one
helicity-0 mode of the graviton, one scalar field from the matter sector,
and one additional Boulware-Deser ghost. All these scalar degrees of freedom are,
however, not always excited around all backgrounds. Since we are interested
in the interaction of the ghost with the matter field, we need to
ensure that the BD ghost is indeed a propagating mode in Eq. (\ref{eq:action_HMG_2nd_order}).
To see this, we first integrate out all the auxiliary fields by using
their EoM
\begin{equation}
\frac{\partial\mathcal{L}}{\partial X}-\partial_{t}\left(\frac{\partial\mathcal{L}}{\partial X'}\right)-\partial_{i}\left(\frac{\partial\mathcal{L}}{\partial X_{i}}\right)+\partial_{t}^{2}\left(\frac{\partial\mathcal{L}}{\partial X''}\right)+\partial_{i}^{2}\left(\frac{\partial\mathcal{L}}{\partial\left(\partial_{i}^{2}X\right)}\right)=0.
\end{equation}
For $X\in\left\{ \Psi,\, B,\,\Delta E\right\} $ this leads to
\begin{align}
\Psi & =\frac{8M_\text{P}^2\Delta\Phi-\left(\Delta E+3\Phi\right)\left(2c_{4}m^{2}M_{\text{P}}^{2}+X_{\varphi}\right)}{c_{2}m^{2}M_{\text{P}}^{2}-X_{\varphi}},\label{eq:EoM_auxvars_Psi}\\
B_{i} & =\frac{8M_\text{P}^2\Phi_{i}'}{c_{1}m^{2}M_{\text{P}}^{2}+X_{\varphi}},\label{eq:EoM_auxvars_Bi}\\
\Delta E & =-\frac{\Phi\left(4c_{3}m^{2}M_{\text{P}}^{2}+X_{\varphi}\right)+\Psi\left(2c_{4}m^{2}M_{\text{P}}^{2}+X_{\varphi}\right) + 8 M_\text{P}^2 \Phi''}{c_{2}m^{2}M_{\text{P}}^{2}-X_{\varphi}}.\label{eq:EoM_auxvars_DelE}
\end{align}
Solving this set of equations for $\Psi$, $B$, and $\Delta E$ as functions of $\Phi$, $X_\varphi$, and their derivatives, yields
\begin{align}
\Psi & =\frac{8M_\text{P}^2 \Delta\Phi\left(c_{2}m^{2}M_{\text{P}}^2-X_{\varphi}\right)-\left(2 c_4 m^2 M_\text{P}^2 + X_{\varphi}\right)\left[ \Phi \left( \left(3c_2-4c_3\right)m^2 M_\text{P}^2 - 4X_\varphi \right)-8 M_\text{P}^2 \Phi'' \right]}{\left(c_{2}^2-4c_{4}^2\right)m^{4}M_\text{P}^4-2\left(c_{2}+2c_{4}\right)m^{2}M_{\text{P}}^{2}X_{\varphi}},\label{eq:sol_Psi}\\
B_{i} & =\frac{8M_\text{P}^2\Phi_{i}'}{c_{1}m^{2}M_{\text{P}}^{2}+X_{\varphi}},\label{eq:sol_B}\\
\Delta E & =\frac{8\Delta\Phi\left(2c_{4}m^{2}M_{\text{P}}^{2}+X_{\varphi}\right) +\Phi\left[4\left(c_{2}c_{3}-3c_{4}^{2}\right)m^{4}M_{\text{P}}^{4} +\left(c_{2}-4c_{3}-12c_{4}\right)m^{2}M_{\text{P}}^{2}X_{\varphi}-4X_{\varphi}^{2}\right] +8M_\text{P}^2 \Phi'' \left( c_2 m^2 M_\text{P}^2 - X_\varphi \right)}{-\left(c_{2}^2-4c_{4}^2\right)m^{4}M_\text{P}^4+2\left(c_{2}+2c_{4}\right)m^{2}M_{\text{P}}^{2}X_{\varphi}}.\label{eq:sol_DeltaE}
\end{align}
Note that the determinant of the mixing matrix for $\Psi$ and $\Delta E$ in Eqs. (\ref{eq:EoM_auxvars_Psi}) and (\ref{eq:EoM_auxvars_DelE}) is proportional to the factors $m^2$ and $c_2+2c_4=2-8\bar\alpha_1 +32\bar\alpha_2$. If one of these terms vanishes, the mixing matrix becomes singular, which is equivalent to it having a zero eigenvalue. These eigenvalues correspond to the kinetic terms of the diagonalized degrees of freedom (given by the eigenvectors). Therefore, the singular situation corresponds to a combination of the auxiliary fields losing its kinetic term, and leads to a strong coupling and a breakdown of perturbativity.
Finally, the full second order HMG action
\begin{equation}
S_{\text{HMG}}^{(2)}=S_{\text{EH}}^{(2)}+S_{\text{mass}}^{(2)}+S_{\text{matter}}
\end{equation}
can be written as
\begin{equation}\label{action_HMG_ghost_matter}
S_{\text{HMG}}^{(2)}\left[\Phi,\varphi\right]=S_{\text{m}}^{(2)}\left[\Phi,\varphi\right]+S_{\text{kin}}^{(2)}\left[\Phi,\varphi\right]+S_{\text{int}}^{(2)}\left[\Phi,\varphi\right].
\end{equation}
The action now depends only on two remaining interacting massive scalar fields $\varphi$ and $\Phi$
described by the mass term
\begin{equation}
S_{\text{m}}^{(2)}=-\int \text{d} ^{4}x\left(-6c_{3}m^{2}M_\text{P}^2 \Phi^{2}+m_{\varphi}^{2}\varphi^{2}\right),
\end{equation}
and rather complicated kinetic and interaction terms, $S_{\text{kin}}^{(2)}$ and $S_{\text{int}}^{(2)}$, respectively. Because the action (\ref{eq:action_HMG_2nd_order}) contains a term that is proportional to $\Phi' \Delta E'$, we find, after integrating out $\Delta E$, terms that include $\left(\Phi''\right)^2$. The occurrence of fourth-order derivatives in $\Phi$ signals that the theory is inevitably plagued by an Ostrogradsky ghost. In total, we expect a composition of three scalar degrees of freedom consisting of a helicity-0 mode from the graviton, a matter field, and a ghost. In order to analyze the interactions between the ghost and the other degrees of freedom, we need to decouple all of them.
\subsection{Decoupling of the ghost}
\subsubsection{Decoupling in vacuum}
Before analyzing the UV limit, i.e., $X_\varphi \gg m^2 M_\text{P}^2$, we study the simpler case first in which the matter field is absent, i.e., $X_\varphi = 0$. The action can then be written as
\begin{equation}
S_{\text{HMG}}^{(2)}=\int \text{d}^4 x \left[ C_1 \left(\Delta \Phi\right)^2 +C_2 \Phi \Delta \Phi + C_3 \Phi'' \Delta \Phi + C_4 \Phi \Phi'' + C_5 \left(\Phi''\right)^2 + C_6 \Phi^2 \right],\label{eq:action_Phi}
\end{equation}
where
\begin{align}
C_1 &\equiv -\frac{32 c_2 M_\text{P}^2}{m^2 \left(c_2^2-4 c_4^2\right)}, &\quad C_2 &\equiv -\frac{4 M_\text{P}^2 \left(c_2^2-12 c_2 c_4+4 c_4 (4 c_3-c_4)\right)}{c_2^2-4 c_4^2},\\
C_3 &\equiv -\frac{32 M_\text{P}^2 \left(4 c_4 (c_1-c_4)+c_2^2\right)}{c_1 m^2 \left(c_2^2-4 c_4^2\right)}, &\quad C_4 &\equiv \frac{4 M_\text{P}^2 \left(3 c_2^2-8 c_2 c_3+12 c_4^2\right)}{c_2^2-4 c_4^2},\\
C_5 &\equiv -\frac{32 c_2 M_\text{P}^2}{m^2 \left(c_2^2-4 c_4^2\right)}, &\quad C_6 &\equiv \frac{2 m^2 M_\text{P}^2 (3 c_2-4 c_3) \left(c_2 c_3-3 c_4^2\right)}{c_2^2-4 c_4^2}.
\end{align}
In order to make the additional scalar degree of freedom manifest we can try to find an equivalent action that descibes two fields with at most second derivatives instead of one field having fourth order derivatives. A special case where the interaction term is just $\left(\square \Phi\right)^2$ has already been presented in Ref.~\cite{Creminelli2005}.
For this, we introduce an auxiliary field $\chi$ together with seven unknown constants $D_i$, and consider a general action of two scalars $\Phi$ and $\chi$ that contains at most second-order derivatives,
\begin{equation}\label{eq:action_Phi_chi}
S'=\int \text{d}^4 x \left[ D_1 \Phi \Phi'' + D_2 \Phi \Delta \Phi + D_3 \chi \Phi'' + D_4 \chi \Delta \Phi + D_5 \chi^2 + D_6 \chi \Phi + D_7 \Phi^2\right].
\end{equation}
The coupling to $\chi$ is constructed such that this auxiliary field can easily be integrated out by using its equation of motion,
\begin{equation}
\chi = -\frac{1}{2 D_5}\left( D_6 \Phi + D_3 \Phi'' + D_4 \Delta \Phi \right).
\end{equation}
We would then obtain an action that looks similar to the action (\ref{eq:action_Phi}) with which we started, except for the coefficients that will now depend on the constants $D_i$. Since we are interested in finding an equivalent action with two degrees of freedom, we equate these coefficients and solve for the unknown constants $D_i$. Interestingly, a solution does exist only if
\begin{equation}
C_1 =\frac{C_3^2}{4C_5}\;\;\Leftrightarrow\;\; C_1 \left(\Delta \Phi\right)^2 + C_3 \Phi'' \Delta \Phi + C_5 \left(\Phi''\right)^2 = \left(\sqrt C_1 \Delta \Phi + \sqrt C_5 \Phi''\right)^2,
\end{equation}
which, as one can easily check, is also satisfied for HMG.\footnote{This condition enforces the theory to be covariant. Since we have started with a covariant theory and then performed a time-space splitting, it is indeed expected that this condition is satisfied. For theories that violate this constraint due to terms that break covariance, this does, however, not imply that there is no ghost but rather that our ansatz is not sufficient. This could indicate that the theory does not propagate only one but more degrees of freedom.} After fixing the redundancy due to a free rescaling of the actions by choosing $D_5 = - m^2 M_\text{P}^2$, \footnote{This choice does also ensure real coefficients for parameter values that we will focus on later. Otherwise, if $c_2^2 - 4 c_4^2 $ is negative, the coefficient $D_5$ should be positive such that $\sqrt C_5$ is real.} we find
\begin{align}
D_1 &=C_4 \pm \sqrt C_5 D_6, \;\; & D_2 &= C_2 \pm \frac{C_3 D_6}{2\sqrt C_5}, \;\;& D_3 &=\mp 2 \sqrt C_5,\label{eq:sol_Di_1}\\
D_4 &= \mp \frac{C_3}{\sqrt C_5}, \;\; & D_7 &=C_6 - \frac{1}{4}D_6^2.\label{eq:sol_Di_2}
\end{align}
We can now introduce two scalar fields $\pi$ and $\ghost$ (to be pronounced ``phi spectre''), described by a superposition of $\Phi$ and $\chi$, which can finally be decoupled with the transformations
\begin{align}
\Phi \longrightarrow A_1 \pi - A_2 \ghost\quad\text{and}\quad\chi \longrightarrow \ghost,
\end{align}
where $A_1$ and $A_2$ are free coefficients. They can be used to diagonalize the mass terms,
\begin{equation}\label{eq:action_matter_decoupled_vacuum}
S'_m=\int \text{d}^4 x \left[ D_5 \ghost^2 + D_6 \ghost \left(A_1 \pi -A_2 \ghost\right) + D_7 \left( A_1 \pi - A_2 \ghost \right)^2 \right],
\end{equation}
to obtain the physical degrees of freedom which can be achieved by setting
\begin{align}
A_1 D_6 = 2 A_1 A_2 D_7.
\end{align}
For the choice $A_1 = 1$ and $D_6 = m^2 M_\text{P}^2$, and using the solution corresponding to the upper signs in Eqs. (\ref{eq:sol_Di_1}) and (\ref{eq:sol_Di_2}), we see that if $c_2^2 - 4 c_4^2 > 0$ (the parameter region that we are interested in) then the prefactor in the kinetic terms for $\ghost$ is negative and, therefore, describes a (BD) ghost, whereas the one for $\pi$ is positive and, thus, corresponds to the healthy helicity-0 mode. From Eq. (\ref{eq:action_matter_decoupled_vacuum}) we can read off the diagonalized mass terms for both scalars and find
\begin{align}
m^2_{\ghost} M_\text{P}^2 &= \frac{4D_5 D_7 - D_6 ^2}{4 D_7} = \frac{-4m^2 M_\text{P}^2 D_7 - m^4 M_\text{P}^4}{4 D_7},\\
m^2_{\pi} M_\text{P}^2 &= D_7.
\end{align}
Since $D_7$ is positive if Eq. (\ref{eq:stability_curve}) is satisfied, this indicates that the ghost is indeed a tachyon.
\subsubsection{Decoupling in the presence of matter}
So far, we have been able to decouple the ghost and the helicity-0 in the absence of an additional matter field. Due to integrating out all auxiliary fields in the full action (\ref{eq:action_HMG_2nd_order}), the coupling between matter and both the ghost and the helicity-0 mode is, however, not trivial and requires a proper decoupling of all present degrees of freedom. Fortunately, the procedure is conceptually similar to what has been done in the vacuum case. Furthermore, we can simplify the calculations by considering the small scale limit $X_\varphi \gg m^2 M_\text{P}^2$. Because the analysis nevertheless becomes a bit lenghtier, we present some itermediate steps in Appendix \ref{app_dec_three_dof}.
In the presence of a matter field $\varphi$ and assuming small scales, the action can be decomposed as
\begin{align}\label{eq:action_Phi_phi}
S_{\text{HMG}}^{(2)}=\int \text{d}^4 x &\left[ \left(\Phi''\right)^2 \left(C_1 \Phi^2 \varphi^2 + C_2 \Phi \varphi + C_3 \right) + \left(\varphi''\right)^2 \left(C_4 \Phi^2 \varphi^2 + C_5 \Phi \varphi + C_6 \right) \right.\\
&\left. + \left(\Delta \Phi\right)^2 \left(C_7 \Phi^2 \varphi^2 + C_8 \Phi \varphi + C_9 \right) + \left(\Delta \varphi\right)^2 \left(C_{10} \Phi^2 \varphi^2 + C_{11} \Phi \varphi + C_{12} \right) +C_{13}\Delta \Phi \Delta \varphi \Phi \varphi \right.\nonumber\\
&\left. + \Phi'' \left(C_{14}\Delta \varphi \Phi \varphi+ C_{15}\varphi'' \Phi \varphi + C_{16}\Delta \Phi \varphi^2 \Phi^2 +C_{17} \Delta \Phi \varphi \Phi +C_{18} \Delta \Phi + C_{19}\Delta \varphi \varphi^2 \Phi^2 +C_{20} \Delta \Phi \varphi \Phi^2\right) \right.\nonumber\\
&\left. +\Phi'' \left( C_{21}\Delta \varphi \varphi^2 \Phi^2 + C_{22}\Delta \varphi + C_{23}\varphi'' \varphi^2 \Phi^2 + C_{24}\varphi'' \right) \right.\nonumber\\
&\left.+ \Delta \Phi \left( C_{25}\varphi'' \varphi^2 \Phi^2 + C_{26}\varphi'' + C_{27}\Delta \varphi \varphi^2 \Phi^2 + C_{28}\Delta \varphi \right) + C_{29}\Delta \varphi \varphi'' \Phi \varphi + C_{30}\Delta \varphi \varphi'' \right.\nonumber\\
&\left. +\Phi''\left( C_{31}\Phi \varphi^2 +C_{32} \Phi \right) + \varphi'' \left( C_{33}\Phi^2 \varphi^3 + C_{34}\Phi^2 \varphi + C_{35}\varphi \right) + \varphi \Phi \left( C_{36}\Phi \varphi^2 + C_{37}\Phi \right) \right.\nonumber\\
&\left. + \Delta \varphi \left( C_{38}\Phi^2 \varphi^3 +C_{39} \Phi^2 \varphi +C_{40} \varphi \right) + C_{41}\Phi^2 \varphi^4 + C_{42}\Phi^2 \varphi^2 +C_{43} \varphi^2 + C_{44}\Phi^2 \right].\nonumber
\end{align}
Note that for HMG, some of the constants $C_i$ do indeed vanish. All of them are explicitly listed in Eq. (\ref{coeff_Ci_Phi_phi}). Again, we start with an ansatz for an action that explicitly describes three degrees of freedom with at most second-order derivatives,
\begin{align}\label{eq:action_Phi_phi_chi}
S_{\text{HMG}}'^{(2)}=\int \text{d}^4 x &\left[ \Phi'' \left( D_{1}\Phi \varphi^2 + D_{2}\Phi \right) + \varphi'' \left( D_{3}\varphi^3 \Phi^2 + D_{4}\varphi \Phi^2 + D_{5}\varphi \right) \right.\\
&\left. + \Delta \Phi \left( D_{6}\Phi \varphi^2 + D_{7}\Phi \right) + \Delta \varphi \left( D_{8}\varphi^3 \Phi^2 + D_{9}\varphi \Phi^2 + D_{10}\varphi \right) \right.\nonumber\\
&\left. + D_{11}\Phi^2 \varphi^4 + D_{12}\Phi^2 \varphi^2 +D_{13} \varphi^2 +D_{14} \Phi^2 \right.\nonumber\\
&\left. + \chi \left( D_{15}\Phi'' + D_{16}\Phi'' \varphi \Phi + D_{17}\Delta \Phi + D_{18}\Delta \Phi \varphi \Phi +D_{19} \varphi'' + D_{20}\varphi'' \varphi \Phi +D_{21} \Delta \varphi +D_{22} \Delta \varphi \varphi \Phi \right) + D_{23}\chi^2 \right].\nonumber
\end{align}
After solving for all coefficients $D_i$, applying the field transformations
\begin{align}\label{eq:field_trafo_Phi_phi_chi}
\Phi &\longrightarrow A_1 \pi + \ghost + M_\text{P}^{-1} \xi,\\
\varphi &\longrightarrow M_\text{P} \pi - M_\text{P} \ghost + A_2 \xi,\\
\chi &\longrightarrow A_3 \pi - \ghost + M_\text{P}^{-1} \xi,
\end{align}
and setting $D_{23} = -m^2 M_\text{P}^2$, we find by checking the relative signs of all kinetic terms that $\ghost$ is the ghost mode with the mass
\begin{align}\label{eq:mass_ghost}
m^2_{\ghost} M_\text{P}^2 = C_{44} - m_\varphi^2 M_\text{P}^2 - m^2 M_\text{P}^2,
\end{align}
and $\pi$ and $\xi$ describe the helicity-0 mode and the matter field, respectively, with masses
\begin{align}\label{eq:mass_helicity0-and-matter}
m^2_{\pi} M_\text{P}^2 &= C_{43} M_\text{P}^2 + \frac{\left(M_\text{P}^2 \left(m^2-C_{43}\right)+C_{44}\right)^2}{4C_{44}}-\frac{\left(M_\text{P}^2 \left(C_{43}+m^2\right)+C_{44}\right)^2}{4m^2 M_\text{P}^2}\\
&= m^2_{\ghost} M_\text{P}^2 \left( -1 + \frac{1}{4} m^2_{\ghost} \left( C_{44}^{-1}M_\text{P}^2 - m^{-2}\right) \right),\\
m^2_{\xi} &= M_\text{P}^{-2}\left( C_{44} - m^2 M_\text{P}^2 \right) + \frac{1}{C_{43} M_\text{P}^4} \left( C_{44} + m^2 M_\text{P}^2 \right)^2\\
&=\frac{m^2_{\ghost} m^2 M_\text{P}^2 - C_{44} \left( m^2_{\ghost} + 4m^2 \right)}{C_{44} - \left( m^2_{\ghost} + m^2\right) M_\text{P}^2} .
\end{align}
We observe that all masses are mainly determined by the coefficient $C_{44}$. In our favored parameter region that satisfies Eq. (\ref{eq:stability_curve}) and $\alpha_1 = \mathcal O (1)$ we find $C_{44} \gg m^2 M_\text{P}^2$. Hence, if $m_\varphi^2$ is small (which we will also assume later for the analysis of the vacuum decay) then Eq. (\ref{eq:mass_ghost}) indicates a positive $m^2_{\ghost}$ but tachyonic scalars $\pi$ and $\xi$. This is not surprising as we have already seen the existence of a tachyon in the vacuum case. With an additional coupling to a new, even non-tachyonic, scalar field the tachyonic instability can leak into all other mass terms. However, this does not render the theory more dangerous and rather tells us that the decay processes in our theory of massive gravity with an additional scalar field can be described by the equivalent setting of one ghost and two tachyonic fields.
\subsection{Strong coupling scale of the theory\label{sub:Cutoff-of-the-theory}}
The constraint that removes the BD ghost in a non-linear theory of
a massive graviton automatically removes all interactions that are
suppressed by scales $\Lambda<\Lambda_{3}$ with
\begin{equation}
\Lambda_{\lambda}\equiv\left(M_{\text{P}}m^{\lambda-1}\right)^{1/\lambda}.
\end{equation}
All other non-linear theories that reduce to FP at the linear level contain
terms suppressed by $\Lambda_{5}$. However, this does not necessarily
hold for theories that do not reduce to the FP theory at the linear level.
We can find the cutoff scale of HMG by using the expansion of the mass term (\ref{eq:pot_2nd})
and introducing the St\"{u}ckelberg fields
\begin{equation}
\delta g_{\mu\nu}\longrightarrow h_{\mu\nu}+\partial_{\mu}A_{\nu}+\partial_{\nu}A_{\mu},
\end{equation}
and, subsequently,
\begin{equation}
A_{\mu}\longrightarrow A_{\mu}+\partial_{\mu}\phi.
\end{equation}
This decomposition into the three helicity modes allows us to read
off the energy scales with which all single interactions are suppressed
(see, e.g., Refs. \cite{Hinterbichler2011,deRham2014}). For this we
need to canonically normalize all the modes through the rescaling
\begin{align}
h_{\mu\nu}\longrightarrow & \frac{2}{M_{\text{P}}}h_{\mu\nu},\\
A_{\mu}\longrightarrow & \frac{2}{mM_{\text{P}}}A_{\mu},\\
\phi\longrightarrow & \frac{2}{m^{2}M_{\text{P}}}\phi.
\end{align}
One now finds that the interactions in HMG that are suppressed by
the smallest scale are of the type
\begin{equation}
\propto\frac{\bar{\alpha}_{i}}{M_{\text{P}}m^{4}}\left(\partial\partial\phi\right)^{3}=\frac{\bar{\alpha}_{i}}{\Lambda_{5}^{5}}\left(\partial\partial\phi\right)^{3},
\end{equation}
and correspond to the cutoff scale $\Lambda_{5}$.
\section{Quantum instability}
\subsection{Most dominant interaction terms}
The quantum stability depends on the scattering
between the ghost $\ghost$ and the matter field $\xi$. In order to compute the scattering amplitude
we move to Fourier space and introduce $k_{\ghostsmall}$ and $k_{\xi}$
for the momenta of $\ghost$ and $\xi$, respectively. The
final action of the interaction between these two fields is rather
complicated. In general, the interaction terms contain derivatives
of both fields that describe the so-called derivative interactions, which,
thus, have momentum-dependent vertices. It is exactly this type of interaction
that might be dangerous since the scattering amplitude requires an integration
over the entire phase space of the initial and final states of the
fields. Therefore, all derivative interactions lead to UV divergent
terms $\propto k^{\alpha}$ with $\alpha\in\mathbb{R}^{+}$. Even
though such derivative interactions exist in the Standard Model (SM),
this problem is usually solved by introducing counter terms which
regularize the divergent parts. In our case, we require a Lorentz
violation to cut the integral over the phase space.
If the integral of the phase space is cut at some energy level due to some new Lorentz breaking operators, then
the decay rate might not necessarily be dominated by the UV behavior
anymore. As seen in Sect. \ref{sub:Cutoff-of-the-theory}, the cutoff
of the EFT is much below the Planck scale. Depending on the mass of
the graviton, terms with a lower number of derivatives could then
become dominant. At the end of this section we will, however, find
that these types of interactions are indeed less important.
From now on, we will need to only focus on the interactions with the highest
number of derivatives where we are allowed to assume that both $k_{\ghostsmall}$
and $k_{\xi}$ are of the same order since the cutoff scales above which
Lorentz invariance is broken are equivalent for both momenta. Even though one can directly see from the action (\ref{eq:action_Phi_phi_chi}) that there are many different types of derivative interations, most of them are suppressed by powers of $M_\text{P}^{-1}$ or $m_\varphi$. We find the Lagrangian corresponding to the most dangerous process to be
\begin{equation}\label{eq:dominant_interaction1}
\mathcal{L}^{\text{dom}}\simeq \frac{\left(c_2^2+4 c_2 (4 c_3+3 c_4+1)-8 \left(2 c_3^2+2 c_3 c_4+c_4 (2 c_4-1)\right)\right)^4 m^6}{32 m_\varphi^6 M_\text{P}^2 (c_2+2 c_4)^5} \; \ghost^2\; \xi^3 \;\partial_\mu \partial^\mu \xi.
\end{equation}
For the analysis of the vacuum decay it will be useful to apply the transformation $\ghost \longrightarrow M_\text{P}^{-1} \ghost$ to obtain the same dimensions for both $\ghost$ and $\xi$. Finally, the interaction in Fourier-space becomes
\begin{equation}\label{eq:dominant_interaction}
\mathcal{L}^{\text{dom}}\simeq \frac{\left(c_2^2+4 c_2 (4 c_3+3 c_4+1)-8 \left(2 c_3^2+2 c_3 c_4+c_4 (2 c_4-1)\right)\right)^4 m^6}{32 m_\varphi^6 M_\text{P}^4 (c_2+2 c_4)^5} k_\xi^2 \; \ghost^2\; \xi^4.
\end{equation}
Note that this interaction arises from a matter sector, which, as in
many other theories of modified gravity, couples minimally to gravity.
Thus, this type of derivative interaction is not only a property of HMG
but rather occurs in a much broader class of theories, even beyond
massive gravity. Since we are studying the Lagrangian on-shell, the
exact term describing the most dominant interaction is, of course,
still model-dependent. Especially the occurrence of derivatives in
the potential term of the theory might lead to different results. However, we expect
that the qualitative results for HMG will still be valid for a huge
class of theories of modified gravity that introduce a ghost and
have a matter sector minimally coupled to gravity.
\subsection{Ghost decay}
\begin{figure}
\includegraphics[trim={0 0 0 0},clip,width=0.35\textwidth]{diagrams/ghost_decay-1}$\qquad$
\includegraphics[trim={7cm 0 0 0},clip,width=0.35\textwidth]{diagrams/vac_decay-1}
\protect\protect\caption{\label{fig:FeynmanDiagrams} Left: Feynman diagram for the most dominant decay of a Boulware-Deser ghost particle (left dashed leg) into another ghost and four minimally coupled matter particles (solid legs). Right: Feynman diagram for the most dominant vacuum decay into ghost and matter paricles.}
\end{figure}
The total decay rate of the ghost particle is the sum of the decay rates from all possible decay channels.%
\footnote{Since the ghost is a boson, due to spin-statistics its production rate will be enhanced by a factor $1+n_{\vec p}$ depending on the occupation number of the final state. However, we are interested in the case in which the phase-space density of ghosts is negligible.}
As already discussed, the dominant contribution to the total decay rate comes from the process shown in Fig. \ref{fig:FeynmanDiagrams} (left) and, therefore, the rate can be very well approximated by
\begin{equation}\label{eq:ghost_decay_rate}
\Gamma_{\ghostsmall}=\frac{1}{2m_{\ghostsmall}}\int\prod_{f}\frac{\text{d} ^{3}p_{f}}{\left(2\pi\right)^{3}2E_{f}}\left|\mathcal{M}\right|^{2}\left(2\pi\right)^{4}\delta^{\left(4\right)}\left(p_{\ghostsmall}-\sum_{f}p_{f}\right).
\end{equation}
Here, $\mathcal{M}$ is the scattering amplitude, $m_{\ghostsmall}$ and $p_{\ghostsmall}$ are the mass and four-momentum of the ghost particle,
respectively,\footnote{Even though a (non-tachyonic) ghost is usually recognized as a field with negative
kinetic energy, its mass term does also carry an additional minus
sign \cite{Sbisa2014}.%
} and $E_{f}$ is the energy of a particle appearing in a
final decay product. The dominance of high momenta in the decay rate further justifies the high energy limit leading to Eq. (\ref{eq:dominant_interaction1}).
It is important to note that we have different dispersion relations for a ghost and
a standard field. While for
a ghost we have $E_{\ghostsmall}=-\sqrt{m_{\ghostsmall}^{2}+\vec{p}_{\ghostsmall}^{2}}$, the
dispersion relation for standard matter fields is $E_{\text{sm}}=\sqrt{m_{\text{sm}}^{2}+\vec{p}_{\text{sm}}^{2}}$.
Here $\vec{p}_{\ghostsmall}$ and $\vec{p}_{\text{sm}}$ are the spatial momenta
for the ghost and matter fields, respectively.
Since Eq. (\ref{eq:ghost_decay_rate}) contains an integral over the entire phase space of all decay products, the decay rate is usually expected to be infinitely large. As mentioned earlier, a LB allows us to cut the integral at $\Lambda_{\text{LB}}$, which is, in fact, the energy scale that determines the decay time.
We now need to find the $\mathcal{M}$ matrix that corresponds to
the derivative interaction between $\ghost$ and $\xi$ as
described by Eq. ($\ref{eq:dominant_interaction1}$). The derivatives yield two
powers from the vertex of the interaction. Furthermore, we multiply
the vertex by a factor of $3!$ as we can freely swap all lines
that correspond to $\xi$. Thus, the scattering amplitude from
the Feynman diagram shown in Fig. \ref{fig:FeynmanDiagrams} (left)
becomes
\begin{align}\label{eq:mmatrix2}
\mathcal{M} =3!A\left(ip_{3}\right)\left(ip_{3}\right) =-3!A\eta^{\mu\nu}\left(p_{3}\right)_{\mu}\left(p_{3}\right)_{\nu},
\end{align}
where we have introduced
\begin{equation}
A\equiv\frac{\left(c_2^2+4 c_2 (4 c_3+3 c_4+1)-8 \left(2 c_3^2+2 c_3 c_4+c_4 (2 c_4-1)\right)\right)^4 m^6}{32 m_\varphi^6 M_\text{P}^4 (c_2+2 c_4)^5}.
\end{equation}
In order to find an upper bound on the decay rate or, equivalently, a lower
bound on the decay time, we consider the worst-case scenario
in which the matter field is almost massless.
Even though this will generally lead to higher decay rates, it is
still a good approximation as the decay will be dominant at energies
near the LI-violating cutoff scale. Assuming isotropy in the
decay process, i.e., $\text{d}^{3}p_{f}=4\pi p_{f}^{2}\text{d} p_{f}$, fixing the
angles between different vectors, and using the momentum conservation,
we finally obtain the differential decay rate,
\begin{equation}
\text{d} \Gamma_{\ghostsmall}\simeq-\frac{18 A^{2}p_{2}p_{3}p_{4}p_{5}p_{6} m_{\xi}^4(2\pi)^4\delta^{\left(4\right)}\left(p_{\ghostsmall}-\sum_{f=1}^{5}p_{f}\right)}{\left(2\pi\right)^{10}m_{\ghostsmall}}.\label{eq:decayrate_fin}
\end{equation}
We are now able to perform the phase-space integral in Eq. (\ref{eq:ghost_decay_rate})
up to the cutoff scale $\Lambda_{\text{LB}}$, at which Lorentz breaking occurs, and obtain
\begin{equation}
\Gamma_{\ghostsmall}\simeq\frac{3{A}^{2}m_{\xi}^4\Lambda_{\text{LB}}^{6}}{2\left(2\pi\right)^{10}m_{\ghostsmall}} + \mathcal O \left( \Lambda_{\text{LB}}^5 \right).\label{eq:decay_rate_result}
\end{equation}
Note that $A$ contains the scale with which the tree-level interaction term (\ref{eq:dominant_interaction}) is suppressed. If one would consider contributions from loops then their vertices that are suppressed by $\Lambda_5$ might lower the scale with which the decay rate is suppressed down to $\Lambda_5$.\footnote{We thank Claudia de Rham for discussions on this aspect.}
As mentioned before, the decay rate (\ref{eq:decay_rate_result})
corresponds only to the scattering process that dominates in the
UV. The validity of this assumption is not obvious for low cutoff scales.
From Eqs. (\ref{eq:sol_Psi}) and (\ref{eq:sol_B}) we find that interactions
with less derivatives of $\varphi$ introduce additional factors
of $m^{2}M_{\text{P}}^{2}$ in $A$. From a power counting we find that
the corresponding decay rate $\tilde{\Gamma}_{\ghostsmall}$ behaves like
\begin{equation}
\tilde{\Gamma}_{\ghostsmall} \propto m^{4}M_{\text{P}}^{4}\Lambda_{\text{LB}}^{-8}\Gamma_{\ghostsmall}.
\end{equation}
Therefore, for all cutoff scales that satisfy $\Lambda_{\text{LB}}\apprge\sqrt{mM_{\text{P}}}\simeq10^{-2}\,\text{eV}$
(for $m=\mathcal{O}\left(H_{0}\right)$) we do not expect higher decay
rates. As we will see in the next section, this condition is always satisfied
for cutoff scales $\Lambda^{(\text{max})}_{\text{LB}}$ with which the
decay would happen on a timescale of the Hubble time.
\subsection{Vacuum decay}
Besides the decay of a ghost, the vacuum itself can also decay into two ghosts and additional matter particles. The Feynman diagram for the most dominant vacuum decay is shown in Fig. \ref{fig:FeynmanDiagrams} (right). The main contribution to the decay rate of the vacuum comes from the same vertex that we found for the ghost decay and, thus, we find
\begin{equation}\label{eq:vac_decay_rate}
\Gamma_{\text{vac}}=\int\prod_{f_{\ghostsmall}}\frac{\text{d} ^{3}p_{f_{\ghostsmall}}}{\left(2\pi\right)^{3}2E_{f_{\ghostsmall}}} \prod_{f}\frac{\text{d} ^{3}p_{f}}{\left(2\pi\right)^{3}2E_{f}}\left|\mathcal{M}\right|^{2}\left(2\pi\right)^{4}\delta^{\left(4\right)}\left(\sum_{f_{\ghostsmall}} p_{f_{\ghostsmall}} + \sum_{f}p_{f}\right).
\end{equation}
After choosing the rest-frame of the ghost particle $f_1$, the $\mathcal{M}$ matrix is, up to a symmetry factor $2!$, similar to Eq. (\ref{eq:mmatrix2}) and, thus, Eq. (\ref{eq:vac_decay_rate}) reduces to
\begin{equation}
\Gamma_{\text{vac}} = 2\Gamma_{\ghostsmall}.
\end{equation}
The total decay rate of the interaction described by Eq. (\ref{eq:dominant_interaction}) is, however, not simply the sum of all decay rates $\Gamma_i$. The vacuum $|0\rangle$ is defined as the state without any excitations, which is not a stable state if the theory contains ghost fields. The particle production rate from the vacuum decay is, therefore, only a good approximation for an initial vacuum state and might become less trustable as the vacuum decays. For this reason and since the decay of the vacuum is not more dangerous than the decay of the ghost, we now focus on the ghost decay only.
\subsection{Numerical calculations}
The decay rate possesses some model dependencies.
A priori the graviton mass scale $m$ is a free parameter. However,
if HMG is to be regarded as a theory of modified gravity that is supposed to solve the
dark energy problem by providing self-accelerating solutions, then
$m$ should not be chosen arbitrarily. The mass parameter $m$ determines
the scale at which modifications to GR become important and is therefore expected
to be $\sim H_{0}$.
There is however an additional model dependency. By using the fit (\ref{eq:stability_curve}), we get
\begin{equation}
A\propto 1/\left(c_{2}+2c_{4}\right)^5\simeq\left(\frac{45}{26-120\bar{\alpha}_{1}}\right)^5,
\end{equation}
which, by tuning
$\bar{\alpha}_{1}$, might diverge, leading to an infinite decay rate. As discussed previously, this limit corresponds to a strong coupling of the matter and ghost mode, and thus, the perturbative approach breaks down.
Since the classical background should also be unstable if the
vacuum decays at tree level, we do not expect to find stable classical
backgrounds for
$\bar{\alpha}_{1}\simeq 13/60$.
For a cross-check, we determine the parameter region that maximizes the timescale of the classical instability to develop. As shown in Fig. \ref{fig:stability_cutoff} (left panel),
we find that the results of the background analysis indeed agree
with this constraint.
\begin{figure}
\includegraphics[width=0.45\textwidth]{img/alpha1_vs_alpha2}$\qquad$
\includegraphics[width=0.45\textwidth]{img/cutoff}
\protect\protect\caption{\label{fig:stability_cutoff} Left: Constraints on the parameter space of HMG. The blue solid line indicates
the region in which Eq. (\ref{eq:stability_curve}) is satisfied, i.e., the timescale of the instability at the classical level to develop is maximized. Models on
the red dashed line satisfy $c_{2}+2c_{4}=0$, indicating the strong coupling regime. Right: Numerical results for the upper bound on the Lorentz-breaking cutoff scale, $\Lambda^{(\text{max})}_{\text{LB}}$
(red dashed line) corresponding to a decay time of the order of
the Hubble time $H_{0}^{-1}$. As indicated by the lowest
and second-lowest dashed lines, denoting $\Lambda_{5}$ and $\Lambda_{3}$,
respectively, the LB cutoff scale can be much larger than the strong coupling scale of the
EFT.}
\end{figure}
The viability of the theory depends on the decay time of the most
dominant scattering process, and requires $\Gamma^{-1}\apprge H_{0}^{-1}$,
which sets an upper bound on the scale $\Lambda_{\text{LB}}$. For a graviton mass $m=\mathcal{O}\left(H_{0}\right)$ and $\bar{\alpha}_{1}=\mathcal{O}\left(1\right)$, where we approximately find $m_\xi\simeq m_{\ghostsmall}\simeq m$, we can estimate
the order of magnitude for the upper bound $\Lambda^{(\text{max})}_{\text{LB}}$,
\begin{equation}
\Lambda_{\text{LB}}\lesssim\Lambda^{(\text{max})}_{\text{LB}} \equiv\left( \frac{ 2 \left(2\pi\right)^{10} m m_{\ghostsmall}}{ 3 A^2 m_\xi^4}\right)^{1/6} = \mathcal O \left( \left( \frac{m_\varphi^{12} M_\text{P}^8}{m^{14}} \right)^{1/6} \right).\label{eq:lambda_upper_bound}
\end{equation}
This gives us an upper limit that is much above the cutoff scale of the theory.
For more accurate numerical results see Fig. \ref{fig:stability_cutoff} (right panel). In the limit $m_\varphi \rightarrow 0$, the amplitude $A$ diverges and indicates an infinitely large decay rate. However, in this limit the Lagrangian of the interaction (\ref{eq:dominant_interaction}) would enter a strongly coupled regime and our perturbative ansatz would not be trustable anymore. Nevertheless, even considering extremely small masses $m_\varphi \simeq m=\mathcal O (H_0)$ would lead to $\Lambda^{(\text{max})}_{\text{LB}} > M_\text{P}$.
Even though this LB cutoff scale $\Lambda^{(\text{max})}_{\text{LB}}$ is much above the strong coupling scale
that was found in Sect. \ref{sub:Cutoff-of-the-theory},
all of our results are still trustable and should be taken seriously. In
fact, the decay products can reach energies near $\Lambda^{(\text{max})}_{\text{LB}}$.
In addition, we should expect the decay processes
to occur even above $\Lambda_{\text{EFT}}$ (which is $\Lambda_5$ for our massive gravity theory). We should however note that it could indeed
be possible that energies above $\Lambda_{\text{EFT}}$ would lead to new
interactions that dominate and result in much larger decay rates, depending on the underlying new physics at energies above the EFT cutoff scale.
\subsection{Comparison to observations}
To date, no high-energy physics experiments have found any signals for the violation of Lorentz invariance, which may seem to indicate that Lorentz-violating operators, if exist, play a role only at very high energies, perhaps even above the Planck scale. Even though our results are compatible with this conclusion, a LB at much smaller energy scales but above $\Lambda_3$ would nevertheless be allowed.
Even though the arguments above require some speculation about the
UV-completed physics, there is a more profound reason why it is
not surprising to find no LV at higher energies. As recently pointed
out in Ref. \cite{Afshordi2015}, most of the operators that break
LI lead to a strong coupling already above energies of $\mathcal{O}\left(\text{meV}\right)$. It has been conjectured that the strongly coupled degree of freedom (in our case the one that leads to a LV) effectively decouples
from the high-energy theory and can therefore not be observed
yet \cite{Afshordi2015}. Similar problems appear in QCD (confinement) and massive gravity
(Vainshtein screening).
Fortunately, there are possibilities to indirectly detect a breaking
of LI that stabilizes the vacuum decay. For the decay products it is most likely to have
energies of order $\Lambda_\text{LB}$, even if $\Lambda_\text{LB}\gg\Lambda_{\text{EFT}}$.
As long as they do not scatter at these energy levels, they can still
consistently be described by our EFT. A direct observation of these
decay products could then hint towards a breaking of LI. If one assumes a LB above $\sim 1 \text{MeV}$ then one could search for observable effects such as peaks in the gamma-ray background, along the lines of the studies in Ref~\cite{Cline2003}. However, the background flux is not well constrained yet for all (especially higher) energies.
\section{Summary and conclusions}
In this work, we have discussed the influence of a ghost on the viability
of an EFT by considering the violation of Lorentz invariance above certain energy scales in a particular theory of modified gravity describing
a massive graviton with an additional Boulware-Deser ghost, which we called haunted massive gravity (HMG). Even though
we do not believe that our HMG model is able to play a
major role in the class of theories of modified gravity attempting
to explain, e.g., the late-time acceleration of our Universe, we do expect that
its quantum properties can be mapped onto a huge class of other theories
of gravity that also introduce an Ostrogradski ghost.
In contrast to simple toy models with a canonical scalar field interacting
with a ghost, we have found a decay rate that does not scale as $\Lambda_{\text{LB}}^{2}$, where $\Lambda_{\text{LB}}$ denotes the energy scale above which Lorentz invariance is broken, or $\Lambda_{\text{LB}}^{8}$ if one assumes the simplest interaction with a graviton; the decay rate scales, instead, as $\Lambda_{\text{LB}}^{6}$. The origin of this difference
lies in the different dominating scattering processes involved. If the
ghost mass is of the order of the Hubble parameter $H_{0}$, which is expected
for theories of modified gravity that provide solutions
to the dark energy problem, then the upper bound on the cutoff scale at
which LB has to occur is allowed to be extremely high and could even be above the Planck scale.
Finally, with HMG we have found an example of a massive gravity theory which
allows for dynamical, and even self-accelerating, FLRW solutions
with a flat reference metric, contrary to the ghost-free dRGT theory. Furthermore, we obtained a parameter region in which both free parameters of the theory are of $\mathcal O \left( 1 \right)$ and maximizes the timescale on which the classical instability is suppressed to obtain a viable cosmological solution. This is indeed surprising as one might expect that a ghost that is present at the background level (which is required in order to obtain dynamical FLRW solutions) will automatically destabilize the theory. We have however studied only the background solutions, and one should therefore note that it is very likely that the cosmological perturbations
would be classically unstable, although it is not obvious with which timescale this instability is suppressed. Furthermore, it might also be possible that
quantum loops would render the theory unviable due to interactions that could
theoretically be much more dangerous than the tree-level interactions which
we have studied in this work; we leave the investigation of these questions for future work.
In general, ghosts are potentially dangerous and can rule
out a theory if the quantum behavior is not under control. However,
if one accepts the possibility of Lorentz-violating physics above
the cutoff of the theory, then all these theories should be studied
carefully and might be acceptable and well behaved.
\begin{acknowledgments}
We are grateful to Arthur Hebecker for initial discussions. We would also like to thank Claudia de Rham, Florian F\"{u}hrer, Johannes Noller, Sabir Ramazanov, Javier Rubio, Angnis Schmidt-May, Adam R. Solomon, and Mikael von Strauss for useful comments and suggestions. We are also grateful to the anonymous referee for important and helpful comments on a previous version of the manuscript. Y.A., L.A., F.K., and H.N. acknowledge support from DFG through the TRR33 project ``The
Dark Universe.'' F.K. is also supported by the Landesgraduiertenf\"{o}rderung
(LGFG) through the Graduate College ``Astrophysics of Fundamental
Probes of Gravity.'' H.N. also acknowledges financial support from DAAD through the program ``Forschungsstipendium f\"{u}r Doktoranden und Nachwuchswissenschaftler.''
Significant parts of the tensor algebra benefited from the package
xAct for Mathematica \cite{xAct}.
\end{acknowledgments}
\newpage
|
\section{Introduction}
Elkies, Kuperberg, Larsen and Propp in their paper \cite{diamond} introduced a new class of object which they called Aztec Diamonds. The Aztec Diamond of order $n$ (denoted by $\ad(n)$) is the union
of all unit squares inside the contour $\abs{x}+\abs{y}=n+1$ (see Figure \ref{fig:diamond} for an Aztec Diamond of order $3$). A domino is the union of any two unit squares sharing an edge, and a domino tiling of a region is a covering of the region by dominoes so that there are no gaps or overlaps. The authors in \cite{diamond} and \cite{diamond2}
considered the problem of counting the number of domino tiling the Aztec Diamond with dominoes and presented four different proofs of the following result.
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{diamond3.pdf}
\caption{$\ad(3)$, Aztec Diamond of order $3$}
\label{fig:diamond}
\end{figure}
\begin{theorem}[Elkies--Kuperberg--Larsen--Propp, \cite{diamond, diamond2}]\label{adm}
The number of domino tilings of an Aztec Diamond of order $n$ is $2^{n(n+1)/2}$.
\end{theorem}
This work subsequently inspired lot of follow ups, including the natural extension of the Aztec Diamond to the Aztec rectangle (see Figure \ref{fig:check}). We denote by $\mathcal{AR}_{a,b}$
the Aztec rectangle which has $a$ unit squares on the southwestern side and $b$ unit squares on the northwestern side. In the remainder of this paper, we assume $b\geq a$ unless
otherwise mentioned. For $a<b$, $\mathcal{AR}_{a,b}$ does not have any tiling by dominoes. The non-tileability of the region $\mathcal{AR}_{a,b}$ becomes evident if we look at the checkerboard representation of $\mathcal{AR}_{a,b}$ (see Figure \ref{fig:check}).
However, if we remove $b-a$ unit squares from the southeastern side then we have a simple product formula found by Helfgott and Gessel \cite{gessel}.
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{check.pdf}
\caption{Checkerboard representation of an Aztec Rectangle with $a=4, b=10$}
\label{fig:check}
\end{figure}
\begin{theorem}[Helfgott--Gessel, \cite{gessel}]\label{ar}
Let $a<b$ be positive integers and $1\leq s_1<s_2<\cdots <s_a\leq b$. Then the number of domino tilings of $\mathcal{AR}_{a,b}$ where all unit squares from the southeastern side are
removed except for those in positions $s_1, s_2, \ldots, s_a$ is \[2^{a(a+1)/2}\prod_{1\leq i<j\leq a}\frac{s_j-s_i}{j-i}.\]
\end{theorem}
Tri Lai \cite{lai} has recently generalized Theorem \ref{ar} to find a generating function, following the work of Elkies, Kuperberg, Larsen and Propp \cite{diamond, diamond2}. Motivated by the recent work of Ciucu and Fischer \cite{ilse}, here we look at the problem of tiling an Aztec rectangle
with dominoes if arbitrary unit squares are removed along the boundary of the Aztec rectangle.
This paper is structured as follows: in Section \ref{s2} we state our main results, in Section \ref{cond-sec} we introduce our main tool in the proofs and present a
slight generalization of it, in Section \ref{s3} we look at tilings of some special cases which are used in our main results. Finally, in Section \ref{s4} we
prove the results described in Section \ref{s2}. The main ingredients in most of our proofs will be the method of condensation developed by Kuo \cite{kuo} and its subsequent generalization by
Ciucu \cite{ciucu}.
\section{Statements of Main Results}\label{s2}
In order to create a region that can be tiled by dominoes we have to remove $k$ more white squares than black squares along the boundary of $\mathcal{AR}_{a,b}$. There are $2b$ white squares and $2a$ black squares on the boundary of $\mathcal{AR}_{a,b}$. We choose
$n+k$ of the white squares that share an edge with the boundary and denote them by $\beta_1, \beta_2, \ldots, \beta_{n+k}$ (we will refer to them as defects of type $\beta$). We choose any
$n$ squares from the black squares which share an edge with the boundary and denote them by $\alpha_1, \alpha_2, \ldots, \alpha_n$ (we refer to them as defects of type $\alpha$). We consider
regions of the type $\mathcal{AR}_{a,b}\setminus \{\beta_1, \ldots, \beta_{n+k}, \alpha_1, \ldots, \alpha_n\}$, which are more general than the type considered in \cite{gessel}.
It is also known that domino tilings of a region can be identified with perfect matchings of its planar dual graph, so for any region $R$ on the square lattice we denote by $\m (R)$ the number of domino tilings
of $R$. We now state the main results of this paper below. The first result is concerned with the case when the defects are confined to three of the four sides of the Aztec
rectangle (defects do not occur on one of the sides with shorter length), and provides a Pfaffian expression for the number of tilings of such a region, with each entry in the Pfaffian being given by a simple product or by a sum or product of quotients of factorials and powers of $2$. The second result gives a nested Pfaffian expression for the general case when we do not restrict the occurence of defects
on any boundary side. The third result deals with the case of an Aztec Diamond with arbitrary defects on the boundary and gives a Pfaffian expression for the number of tilings of such a
region, with each entry in the Pfaffian being given by a simple sum of quotients of factorials and powers of $2$.
We define the region $\mathcal{AR}_{a,b}^k$ to be the region obtained from $\mathcal{AR}_{a.b}$ by adding a string of $k$ unit squares along the boundary of the southeastern side as shown in Figure
\ref{fig:mt1}. We denote this string of $k$ unit squares by $\gamma_1, \gamma_2, \ldots, \gamma_k$ and refer to them as defects of type $\gamma$.
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{rectangle-k.pdf}
\caption{$\mathcal{AR}_{a,b}^k$ with $a=4,b=9,k=5$}
\label{fig:mt1}
\end{figure}
\begin{theorem}\label{mt1}
Assume that one of the two sides on which defects of type $\alpha$ can occur does not actually have any defects on it. Without loss of generality, we assume this to be the southwestern side.
Let $\delta_1, \ldots, \delta_{2n+2k}$ be the elements of the set $\{\beta_1, \ldots, \beta_{n+k}\}\cup \{\alpha_1, \ldots, \alpha_{n}\}\cup \{\gamma_1, \ldots, \gamma_{k}\}$ listed in a cyclic order.
Then we have
\begin{equation}\label{em1}
\m (\mathcal{AR}_{a,b}\setminus \{\beta_1, \ldots, \beta_{n+k}, \alpha_1, \ldots, \alpha_n\})=\frac{1}{[\m (\mathcal{AR}_{a,b}^k)]^{n-k+1}}\pf[(\m (\mathcal{AR}_{a,b}^k\setminus\{\delta_i, \delta_j\}))_{1\leq i<j\leq 2n+2k}],
\end{equation}
\noindent where all the terms on the right hand side are given by explicit formulas:
\begin{enumerate}
\item $\m (\mathcal{AR}_{a,b}^k)$ is given by Theorem \ref{adm},
\item $\m (\mathcal{AR}_{a,b}^k\setminus\{\beta_i, \alpha_j\})$ is given by Proposition \ref{ad_i_j} if $\beta_i$ is on the south-eastern side and not above a $\gamma$ defect; otherwise it is $0$,
\item $\m (\mathcal{AR}_{a,b}^k\setminus\{\beta_i, \gamma_j\})$ is given by Theorem \ref{adm} if $\beta_i$ is above a $\gamma$ defect; it is given by Proposition \ref{ar_k-1_i} if the $\beta$ defect is in the northwestern side at a distance of more than $k-1$ from the western corner; it is given by Propositions \ref{ar_k_i} if
the $\beta$ dent is on the southeastern side; otherwise it is $0$,
\item $\m (\mathcal{AR}_{a,b}^k\setminus\{\beta_i, \beta_j\})=\m (\mathcal{AR}_{a,b}^k\setminus\{\alpha_i, \alpha_j\})=\m (\mathcal{AR}_{a,b}^k\setminus\{\alpha_i, \gamma_j\})=\m (\mathcal{AR}_{a,b}^k\setminus\{\gamma_i, \gamma_j\})=0$.
\end{enumerate}
\end{theorem}
\begin{theorem}\label{mt2}
Let $\beta_1, \ldots, \beta_{n+k}$ be arbitrary defects of type $\beta$ and $\alpha_1, \ldots, \alpha_n$ be arbitrary defects of type $\alpha$ along the boundary of $\mathcal{AR}_{a,b}$. Then $\m ({\mathcal{AR}_{a,b}}\setminus\{\beta_1, \ldots, \beta_{n+k}, \alpha_1, \ldots, \alpha_n\})$ is equal to the Pfaffian of a $2n\times 2n$
matrix whose entries are Pfaffians of $(2k+2)\times (2k+2)$ matrices of the type in the statement of Theorem \ref{mt1}.
\end{theorem}
In the special case when the number of defects of both types are the same, that is when $k=0$ we get an Aztec Diamond with arbitrary defects on the boundary and the number of tilings can be given by a Pfaffian
where the entries of the Pfaffian are explicit, as stated in the theorem below.
\begin{theorem}\label{mt3}
Let $\beta_1, \ldots, \beta_{n}$ be arbitrary defects of type $\beta$ and $\alpha_1, \ldots, \alpha_n$ be arbitrary defects of type $\alpha$ along the boundary of $\ad(a)$, and let $\delta_1, \ldots, \delta_{2n}$
be a cyclic listing of the elements of the set $\{\beta_1, \ldots, \beta_n\}\cup \{\alpha_1, \ldots, \alpha_n\}$. Then
\begin{equation}\label{emt3}
\m (\ad(a)\setminus\{\beta_1, \ldots, \beta_n, \alpha_1, \ldots, \alpha_n\})=\frac{1}{[\m (\ad(a))]^{n-1}}\pf [(\m (\ad(a)\setminus\{\delta_i, \delta_j\}))_{1\leq i<j\leq 2n}],
\end{equation}
\noindent where the values of $\m (\ad(a)\setminus\{\delta_i, \delta_j\}))$ are given explicitly as follows:
\begin{enumerate}
\item $\m (\ad(a)\setminus\{\beta_i, \alpha_j\}))$ is given by Proposition \ref{ad_i_j},
\item $\m (\ad(a)\setminus\{\beta_i, \beta_j\}))=\m (\ad(a)\setminus\{\alpha_i, \alpha_j\}))=0$.
\end{enumerate}
\end{theorem}
\section{A result on Graphical Condensation}\label{cond-sec}
The proofs of our main results are based on Ciucu's generalization \cite{ciucu} of Kuo's graphical condensation \cite{kuo} which we state below. The aim of this section is also to present
our small generalization of Ciucu's result.
Let $G$ be a weighted graph, where the weights are associated with each edge of $G$, and let $\m (G)$ denote the sum of the weights of the perfect matchings of $G$, where the weight of a perfect matching is taken to be the product
of the weights of its constituent edges. We are interested in graphs with edge weights all equaling $1$, which corresponds to tilings of the region in our special case.
\begin{theorem}[Ciucu, \cite{ciucu}]\label{condensation}
Let $G$ be a planar graph with the vertices $a_1, a_2, \ldots, a_{2k}$ appearing in that cyclic order on a face of $G$. Consider the skew-symmetric matrix $A=(a_{ij})_{1\leq i,j\leq 2k}$ with entries given by
\begin{equation}\label{ciucu1}
a_{ij} := \m (G\setminus \{a_i, a_j\}), \text{if } i<j.
\end{equation}
Then we have that
\begin{equation}\label{ciucu2}
\m (G\setminus \{a_1, a_2, \ldots, a_{2k}\})=\frac{\pf(A)}{[\m (G)]^{k-1}}.
\end{equation}
\end{theorem}
Although Theorem \ref{condensation} is enough for our purposes, we state and prove a slightly more general version of the theorem below. It turns out that our result is a common generalization for
the condensation results in \cite{kuo} as well as Theorem \ref{condensation} which follows immediately from Theorem \ref{condensation-2} below if we consider $a_1, \ldots, a_{2k}\in \vv(G)$. We also mention
that Corollary \ref{cond-cor} of Theorem \ref{condensation-2}, does not follow from Theorem \ref{condensation}.
To state and prove our result, we will need to make some notations and concepts clear. We consider the symmetric difference on the vertices and edges of a graph. Let $H$ be a planar graph and $G$ be an
induced subgraph of $H$ and let $W\subseteq \vv(H)$. Then we define $G+W$ as follows: $G+W$ is the induced subgraph of $H$ with vertex set $\vv(G+W)=\vv(G)\Delta W$, where $\Delta$
denotes the symmetric difference of sets. Now we are in a position to state our result below.
\begin{theorem}\label{condensation-2}
Let $H$ be a planar graph and let $G$ be an induced subgraph of $H$ with the vertices $a_1, a_2, \ldots, a_{2k}$ appearing in that cyclic order on a face of $H$. Consider the skew-symmetric matrix $A=(a_{ij})_{1\leq i,j\leq 2k}$ with entries given by
\begin{equation}\label{ciucu1-2}
a_{ij} := \m (G+ \{a_i, a_j\}), \text{if } i<j.
\end{equation}
Then we have that
\begin{equation}\label{cond4}
\m (G+ \{a_1, a_2, \ldots, a_{2k}\})=\frac{\pf(A)}{[\m (G)]^{k-1}}.
\end{equation}
\end{theorem}
\begin{corollary}\cite[Theorem 2.4]{kuo}\label{cond-cor}
Let $G=(V_1, V_2, E)$ be a bipartite planar graph with $\abs{V_1}=\abs{V_2}+1$; and let $w, x, y$ and $z$ be vertices of $G$ that appear in cyclic order on a face of $G$. If
$w, x, y \in V_1$ and $z\in V_2$ then
\begin{align}
\m (G-\{w\})\m (G-\{x,y,z\})+\m (G-\{y\})\m (G-\{w, x,z\})=\m (G-\{x\})\m (G-\{w, y,z\}) & \\
+\m (G-\{z\})\m (G-\{w, x,y\}). & \nonumber
\end{align}
\end{corollary}
\begin{proof}
Take $n=2$, $a_1=w, a_2=x, a_3=y, a_4=z$ and $G=H\setminus\{a_1\}$ in Theorem \ref{condensation-2}.
\end{proof}
The proof of Theorem \ref{condensation} follows from the use of some auxillary results. In the vein of those results, we need the following proposition to complete our proof of
Theorem \ref{condensation-2}.
\begin{proposition}\label{ck3}
Let $H$ be a planar graph and $G$ be an induced subgraph of $H$ with the vertices $a_1, \ldots, a_{2k}$ appearing in that cyclic order among the vertices of some face of $H$.
Then
\begin{align}\label{prope1}
\m(G)\m(G+\{a_1, \ldots, a_{2k}\})+\sum_{l=2}^{k}\m(G+ \{a_1, a_{2l-1}\})\m(G+ \overline{\{a_1, a_{2l-1}\}}) & \nonumber \\
= \sum_{l=1}^{k}\m(G+ \{a_1, a_{2l}\})\m(G+ \overline{\{a_1, a_{2l}\}}), &
\end{align}
\noindent where $\overline{\{a_i, a_j\}}$ stands for the complement of $\{a_i, a_j\}$ in the set $\{a_1, \ldots, a_{2k}\}$.
\end{proposition}
Our proof follows closely that of the proof of an analogous proposition given by Ciucu \cite{ciucu}.
\begin{proof}
We recast equation \eqref{prope1} in terms of disjoint unions of cartesian products as follows
\begin{align}\label{prope2}
\mathcal{M}(G)\times \mathcal{M}(G+\{a_1, \ldots, a_{2k}\})\cup \mathcal{M}(G+ \{a_1, a_{3}\})\times \mathcal{M}(G+ \overline{\{a_1, a_{3}\}})\cup \ldots & \nonumber \\
\cup \mathcal{M}(G+ \{a_1, a_{2k-1}\})\times \mathcal{M}(G+ \overline{\{a_1, a_{2k-1}\}})&
\end{align}
\noindent and
\begin{align}\label{prope3}
\mathcal{M}(G+ \{a_1, a_{2}\})\times \mathcal{M}(G+ \overline{\{a_1, a_{2}\}})\cup\mathcal{M}(G+ \{a_1, a_{4}\})\times \mathcal{M}(G+ \overline{\{a_1, a_{4}\}}) \cup \ldots & \nonumber \\
\cup \mathcal{M}(G+ \{a_1, a_{2k}\})\times \mathcal{M}(G+ \overline{\{a_1, a_{2k}\}})\cup&
\end{align}
\noindent where $\mathcal{M}(F)$ denotes the set of perfect matchings of the graph $F$. For each element $(\mu, \nu)$ of \eqref{prope2} or \eqref{prope3}, we think of the edges of $\mu$ as being marked by solid lines and
that of $\nu$ as being marked by dotted lines, on the same copy of the graph $H$. If there are any edges common to both then we mark them with both solid and dotted lines.
We now define the weight of $(\mu, \nu)$ to be the product of the weight of $\mu$ and the weight of $\nu$. Thus, the total weight of the elements in the set \eqref{prope2} is same as the left
hand side of equation \eqref{prope1} and the total weight of the elements in the set \eqref{prope3} equals the right hand side of equation \eqref{prope1}. To prove our result, we
have to construct a weight-preserving bijection between the sets \eqref{prope2} and \eqref{prope3}.
Let $(\mu, \nu)$ be an element in \eqref{prope2}. Then we have two possibilities as discussed in the following. If $(\mu, \nu)\in \mathcal{M}(G)\times \mathcal{M}(G+\{a_1, \ldots, a_{2k}\})$ we note that when considering the edges of $\mu$ and $\nu$ together on the
same copy of $H$, each of the vertices $a_1, \ldots, a_{2k}$ is incident to precisely one edge (either solid or dotted depending on the graph $G$ and the vertices $a_i$'s), while all the other
vertices of $H$ are incident to one solid and one dotted edge. Thus $\mu \cup \nu$ is the disjoint union of paths connecting the $a_i$'s to one another in pairs, and cycles covering
the remaining vertices of $H$. We now consider the path containing $a_1$ and change a solid edge to a dotted edge and a dotted edge to a solid edge. Let this pair of matchings be $(\mu^\prime, \nu^\prime)$.
The path we have obtained must connect $a_1$ to one of the even-indexed vertices, if it connected $a_1$ to some odd-indexed vertex $a_{2i+1}$ then it would isolate the $2i-1$ vertices $a_2, a_3,
\ldots, a_{2i}$ from the other vertices and hence we do not get disjoint paths connecting them. Also, we note that the end edges of this path will be either dotted or solid
depending on our graph $G$ and the vertices $a_i$'s. So $(\mu^\prime, \nu^\prime)$ is an element of \eqref{prope3}.
If $(\mu, \nu)\in \mathcal{M}(G+\{a_1, a_3\})\times \mathcal{M}(G+ \overline{\{a_1, a_{3}\}})$, then we map it to a pair of matchings $(\mu^\prime, \nu^\prime)$ obtained by reversing the solid and dotted edges
along the path in $\mu\cup \nu$ containing $a_3$. With a similar reasoning like above, this path must connect $a_3$ to one of the even-indexed vertices and a similar argument will show
that indeed $(\mu^\prime, \nu^\prime)$ is an element of \eqref{prope3}. If $(\mu, \nu)\in \mathcal{M}(G+\{a_1, a_{2i+1}\})\times \mathcal{M}(G+ \overline{\{a_1, a_{2i+1}\}})$ with $i>1$, we have the same construction with
$a_3$ replaced by $a_{2i+1}$.
The map $(\mu, \nu)\mapsto (\mu^\prime, \nu^\prime)$ is invertible because given an element in $(\mu^\prime, \nu^\prime)$ of \eqref{prope3}, the pair $(\mu, \nu)$ that is mapped to it is obtained by shifting
along the path in $\mu^\prime \cup \nu^\prime$ that contains the vertex $a_{2i}$, such that $(\mu^\prime, \nu^\prime)\in \mathcal{M}(G+\{a_1, a_{2i}\})\times \mathcal{M}(G+\overline{\{a_1, a_{2i}\}})$. The map we
have defined is weight-preserving and this proves the proposition.
\end{proof}
Now we can prove Theorem \ref{condensation-2}, which is essentially the same proof as that of Theorem \ref{condensation}, but now uses our more general Proposition \ref{ck3}.
\begin{proof}[Proof of Theorem \ref{condensation-2}]
We prove the statement by induction on $k$. For $k=1$ it follows from the fact that \[ \pf \left( \begin{array}{cc}
0 & a \\
-a & 0 \end{array} \right) =a.\]
For the induction step, we assume that the statement holds for $k-1$ with $k\geq 2$. Let $A$ be the matrix
\[\left( \begin{array}{ccccc}
0 & \m(G+\{a_1, a_2\}) & \m(G+\{a_1, a_3\}) & \cdots & \m(G+\{a_1, a_{2k}\}) \\
-\m(G+\{a_1, a_2\}) & 0 & \m(G+\{a_2, a_3\}) & \cdots & \m(G+\{a_2, a_{2k}\})\\
-\m(G+\{a_1, a_3\}) & -\m(G+\{a_2, a_3\}) & 0 & \cdots & \m(G+\{a_3, a_{2k}\}) \\
\vdots & \vdots & \vdots & & \vdots \\
-\m(G+\{a_1, a_{2k}\}) & -\m(G+\{a_2, a_{2k}\}) & -\m(G+\{a_3, a_{2k}\}) & \cdots & 0\end{array} \right).\]
\noindent By a well-known property of Pfaffians, we have
\begin{equation}\label{cond1}
\pf(A)=\sum_{i=2}^{2k}(-1)^i\m(G+\{a_1, a_i\})\pf (A_{1i}).
\end{equation}
Now, the induction hypothesis applied to the graph $G$ and the $2k-2$ vertices in $\overline{\{a_i, a_j\}}$ gives us
\begin{equation}\label{cond2}
[\m(G)]^{k-2}\m(G+\overline{\{a_1, a_i\}})=\pf(A_{1i}),
\end{equation}
\noindent where $A_{1i}$ is same as in equation \eqref{cond1}. So using equations \eqref{cond1} and \eqref{cond2} we get
\begin{equation}\label{cond3}
\pf(A)=[\m(g)]^{k-2}\sum_{i=2}{2k}(-1)^i\m(G+\{a_1, a_i\})\m(G+\overline{\{a_1, a_i\}}).
\end{equation}
\noindent Now using Propositition \ref{ck3}, we see that the above sum is $\m(G)\m(G+\{a_1, \ldots, a_{2k}\})$ and hence equation \eqref{cond3} implies \eqref{cond4}.
\end{proof}
\section{Some family of regions with defects}\label{s3}
In this section, we find the number of tilings by dominoes of certain regions which appear in the statement of Theorem \ref{mt1} and Theorem \ref{mt3}. We define the binomial coefficients that appear in this section as follows
\begin{equation*}
\binom{c}{d} := \begin{cases} \dfrac{c(c-1)\cdots(c-d+1)}{d!}, &\text{if } d\geq0\\
0, &\text{otherwise} \end{cases}.
\end{equation*}
\noindent Our formulas also involve hypergeometric series. We recall that the hypergeometric series of parameters $a_1, \ldots, a_r$ and $b_1, \ldots, b_s$ is defined as
\[_rF_s\left[\hyper{a_1, \ldots, a_r}{b_1, \ldots, b_s}\,;z\right]=\sum_{k=0}^{\infty}\frac{(a_1)_k\cdots (a_r)_k}{(b_1)_k\cdots (b_s)_k}\frac{z^k}{k!}.\]
We also fix a notation for the remainder of this paper as follows, if we remove the squares labelled $2,4,7$ from the south-eastern boundary of $\mathcal{AR}_{4,7}$, we denote it by $\mathcal{AR}_{4,7}(2,4,7)$. In the derivation of the results in this section, the following two corollaries of Theorem \ref{ar} will be used.
\begin{corollary}\label{cor1}
The number of tilings of $\mathcal{AR}_{a,a+1}(i)$ is given by \[2^{a(a+1)/2}\binom{a}{i-1}.\]
\end{corollary}
\begin{corollary}\label{cor2}
The number of tilings of $\mathcal{AR}_{a,b}(2,\ldots, b-a+1)$ is given by \[2^{a(a+1)/2} \binom{b-1}{a-1}.\]
\end{corollary}
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop31.pdf}
\caption{Aztec rectangle with $k-1$ squares added on the southeastern side and a defect on the $j$-th position shaded in black; here $a=4,b=10,k=6,j=8$}
\label{fig:ar_k_i}
\end{figure}
\begin{proposition}\label{ar_k_i}
Let $1\leq a\leq b$ be positive integers with $k=b-a>0$, then the number of domino tilings of $\mathcal{AR}_{a,b}(j)$ with $k-1$ squares added to the southeastern side
starting at the second position (and not at the bottom) as shown in the Figure \ref{fig:ar_k_i} is given by
\begin{equation}\label{k-1,j}
2^{a(a+1)/2}\binom{a+k-1}{j-1}\binom{j-2}{k-1}~_3F_2\left[\hyper{1,1-j,1-k}{2-j, 1-a-k}\,;1\right].
\end{equation}
\end{proposition}
\begin{proof}
Let us denote the region in Figure \ref{fig:ar_k_i} by $\ar^{k-1,j}_{a,b}$ and we work with the planar dual graph of the region $\ar^{k-1,j}_{a,b}$ and count the number of matchings of
that graph. We first notice that the first added square in any tiling of the region in Figure \ref{fig:ar_k_i} by dominoes has two possibilities marked in grey in the Figure
\ref{fig:ar_k_i1}. This observation
allows us to write the number of tilings of $\ar^{k-1,j}_{a,b}$ in terms of the following recursion
\begin{equation}\label{ep31}
\m (\ar^{k-1,j}_{a,b})=\m (\ar_{a,b-1}^{k-2,j-1})+\m (\mathcal{AR}_{a,b}(2,3,\ldots, k, j)).
\end{equation}
\noindent which can be verified from Figure \ref{fig:grey}.
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop311.pdf}
\caption{$\ar^{k-1,j}_{a,b}$ with the possible choices for the first added square in a tiling; here $a=4, b=10, k=6, j=8$}
\label{fig:ar_k_i1}
\end{figure}
\noindent Repeatedly using equation \eqref{ep31} $k-1$ times on succesive iterations, we shall finally obtain
\begin{equation}\label{ep31-2}
\m (\ar^{k-1,j}_{a,b})=\sum_{l=0}^{k-2}\m (\mathcal{AR}_{a,b-l}(2,3,\ldots,k-l,j-l))+\m (\mathcal{AR}_{a,a+1}(j-k+1)).
\end{equation}
\begin{figure}[!htb]
\minipage{0.50\textwidth}
\includegraphics[width=\linewidth, center]{grey-1.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[width=\linewidth, center]{grey-2.pdf}
\endminipage
\caption{Choices for the tilings of $\ar^{k-1,j}_{a,b}$ with forced dominoes; here $a=4,b=10,k=6,j=8$}
\label{fig:grey}
\end{figure}
Now, plugging in the values of the quantities in the right hand side of equation \eqref{ep31-2} from Theorem \ref{ar} and Corollary \ref{cor1} we shall obtain equation \eqref{k-1,j}.
\end{proof}
One of the main ingredients in our proofs of the remaining results in this section are the following results of Kuo \cite{kuo}.
\begin{theorem}\cite[Theorem 2.3]{kuo}\label{kk1}
Let $G=(V_1, V_2, E)$ be a plane bipartite graph in which $\abs{V_1}=\abs{V_2}$. Let $w, x, y$ and $z$ be vertices of $G$ that appear in cyclic order on a face of $G$. If $w,x\in V_1$ and $y, z\in V_2$ then
$$\m (G-\{w,z\})\m (G-\{x,y\})=\m (G)\m (G-\{w,x,y,z\})+\m (G-\{w,y\})\m (G-\{x,z\}).$$
\end{theorem}
\begin{theorem}\cite[Theorem 2.5]{kuo}\label{kk}
Let $G=(V_1, V_2, E)$ be a plane bipartite graph in which $\abs{V_1}=\abs{V_2}+2$. Let the vertices $w,x,y$ and $z$ appear in that cyclic order on a face of $G$. Let $w,x,y,z\in V_1$, then
$$\m (G-\{w,y\})\m (G-\{x,z\})=\m (G-\{w,x\})\m (G-\{y,z\})+\m (G-\{w,z\})\m (G-\{x,y\}).$$
\end{theorem}
The following proposition does not appear explicitely in the statement of Theorem \ref{mt1}, but it is used in deriving Proposition \ref{ar_k-1_i}.
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop32.pdf}
\caption{An $a\times (a+2)$ Aztec rectangle with some labelled squares; here $a=5$}
\label{fig:ar_i_j}
\end{figure}
\begin{proposition}\label{ar_i_j}
Let $1\leq a$ be a positive integer, then the number of tilings of $\mathcal{AR}_{a, a+2}$ with a defect at the $i$-th position on the southeastern side counted from the south corner and a defect on the $j$-th position on the northwestern side
counted from the west corner is given by
\begin{equation}\label{arij}
2^{a(a+1)/2}\left[\binom{a}{i-2}\binom{a}{j-1}+\binom{a}{i-1}\binom{a}{j-2}\right].
\end{equation}
\end{proposition}
\begin{proof}
\begin{figure}[!htb]
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop32-ac.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop32-bd.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop32-ab.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop32-cd.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop32-ad.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop32-bc.pdf}
\endminipage
\caption{Some forced dominoes in the proof of Proposition \ref{ar_i_j} where the vertices we remove are labelled}
\label{fig:kuo-1}
\end{figure}
If $j=1$ or $j=a+2$, then the region we want to tile reduces to the type in Theorem \ref{ar} and it is easy to see that the expression \eqref{arij} is satisfied in these cases. By symmetry, this
also takes care of the cases $i=1$ and $i=a+2$.
In the rest of the proof, we now assume that $1<i,j<a+2$ and let us denote the region we are interested in by $\ooo(a)_{i,j}$. We now use Theorem \ref{kk} with the vertices as indicated in Figure \ref{fig:ar_i_j} to obtain the following identity
(Figure \ref{fig:kuo-1}).
\begin{align}\label{ep32}
\m (\ad(a))\m (\ooo(a)_{i,j}) =& \m (\mathcal{AR}_{a,a+1}(i-1))\m (\mathcal{AR}_{a,a+1}(j))\\ \nonumber
&+ \m (\mathcal{AR}_{a,a+1}(j-1))\m (\mathcal{AR}_{a,a+1}(i)).
\end{align}
\noindent Now, using Theorem \ref{adm} and Corollary \ref{cor1} in equation \eqref{ep32} we get \eqref{arij}.
\end{proof}
\begin{remark}\label{rem1}
Ciucu and Fischer \cite{ilse}, have a similar result for the number of lozenge tiling of a hexagon with dents on opposite sides (Proposition 4 in their paper). They also make use of
Kuo's condensation result, Theorem \ref{kk1} and obtain the following identity
\begin{align*}
\opp(a,b,c)_{i,j}&\opp(a-2,b,c)_{i-1, j-1} \\
=& \opp(a-1,b,c)_{i-1,j-1}\opp(a-1,b,c)_{i,j}\\
&- \opp(a-1,b-1,c+1)_{i,j-1}\opp(a-1,b+1,c-1)_{i-1,j}
\end{align*}
\noindent where $\opp(a,b,c)_{i,j}$ denotes the number of lozenge tilings of a hexagon $H_{a,b,c}$ with opposite side lengths $a,b,c$ and with two dents in position $i$ and $j$ on
opposite sides of length $a$, where $a,b,c,i,j$ are positive integers with $1\leq i,j \leq a$.
In their use of Kuo's result, they take the graph $G$ to be $\opp(a,b,c)_{i,j}$, but
if we take the graph $G$ to be $H_{a,b,c}$ and use Theorem \ref{kk1} with an appropriate choice of labels, we get the following identity
\begin{align*}
\opp(a,b,c)_{i,j}\hex(a-1,b,c) =& \hex(a,b,c)\opp(a-1,b,c)_{i,j} \\
&+ \hex(a,c-1,b+1,a-1,c,b)_i\hex(a,c,b,a-1,c+1,b-1)_{a-j+1}
\end{align*}
\noindent where $\hex(a,b,c)$ denotes the number of lozenge tilings of the hexagon with opposite sides of length $a,b,c$ and $\hex(m,n,o,p,q,r)_k$ denotes the number of
lozenge tilings of a hexagon with side lengths $m,n,o,p,q,r$ with a dent at position $k$ on the side of length $m$. Then, Proposition 4 of Ciucu and Fischer \cite{ilse} follows more easily without the need for contigous relations of hypergeometric series that they use in their paper.
\end{remark}
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop34.pdf}
\caption{An $a\times b$ Aztec rectangle with defects marked in black; here $a=4, b=9. k=5, i=5$}
\label{fig:ar_k-1_i}
\end{figure}
\begin{proposition}\label{ar_k-1_i}
Let $1\leq a, i \leq b$ be positive integers with $k=b-a>0$, then the number of domino tilings of $\mathcal{AR}_{a,b}(2, 3, \ldots, k)$ with a defect on the northwestern side in the $i$-th position counted from the west corner as shown in the Figure \ref{fig:ar_k-1_i} is given by
\[2^{a(a+1)/2}\binom{a+k-2}{k-1}\binom{a}{a-i+k}~_3F_2\left[\hyper{1, -k-1, i-a-k}{i-k+1, 2-a-k}\,;-1\right].\]
\end{proposition}
\begin{proof}
Our proof will be by induction on $b=a+k$. The base case of induction will follow if we verify the
result for $a=2, k=1$ in which case $b=3$. We also need to check the result for $i=1$ and $i=b$. If $i=1$ we have many forced dominoes and we get the region shown in Figure \ref{fig:prop34-2}, which is $\ad(a)$. Again, if $i=b$, then also we get a region of the type in Theorem \ref{ar}.
In both of these cases the number of domino tilings of these regions satisfy the formula mentioned in the statement. To check our base case it is now enough to verify the formula for $a=2, k=1, i=2$ as the other cases of $i=1$
and $i=3$ are already taken care of. In this case, we see that the region we obtain is of the type as described in Corollary \ref{cor1} and this satisfies the statement of our result.
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop34-2.pdf}
\caption{Forced tilings for $i=1$ in Proposition \ref{ar_k-1_i}}
\label{fig:prop34-2}
\end{figure}
From now on, we assume $b> 3$ and $1<i<b$. We denote the region of the type shown in Figure \ref{fig:ar_k-1_i} by $\mathcal{AR}_{a,b,k-1}^{i}$. We use Theorem \ref{kk} here, with the vertices $w,x,y$ and $z$
marked as shown in Figure \ref{fig:prop34-1}, where we add a series of unit squares to the northeastern side to make it into an $a\times (b+1)$ Aztec rectangle. Note that the square in the
$i$-th position to be removed is included in this region and is labelled by $z$. The identity we now obtain is the following (see Figure \ref{fig:kuo-2} for forcings)
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop34-1.pdf}
\caption{Labelled $a\times (b+1)$ Aztec rectangle; here $a=4, b=9$}
\label{fig:prop34-1}
\end{figure}
\begin{equation}\label{ep34}
\m (\ad(a))\m (\mathcal{AR}_{a,b+1,k}^{i}) = \m (\ad(a))\m (\mathcal{AR}_{a,b,k-1}^{i}) + Y\cdot \m (\mathcal{AR}_{a,b}(2,3,\ldots, k, k+1))
\end{equation}
\noindent where
\begin{equation}\label{ep34-1}
Y := \begin{cases} 0, &\text{if } i\leq k\\
\m (\mathcal{AR}_{a,a+1}(a+k+2-i), &\text{if } i\geq k+1\end{cases}.
\end{equation}
\begin{figure}[!htbp]
\minipage{0.50\textwidth}
\includegraphics[scale=0.5, center]{prop34-ac.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.5, center]{prop34-bd.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.5, center]{prop34-ab.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.5, center]{prop34-cd.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.5, center]{prop34-ad.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.5, center]{prop34-bc.pdf}
\endminipage
\caption{Forced dominoes in the proof of Proposition \ref{ar_k-1_i} where the vertices we remove are labelled}
\label{fig:kuo-2}
\end{figure}
Using equation \eqref{ep34-1} in equation \eqref{ep34}, we can simplify the relation further to the following
\begin{equation}\label{ep34-m}
\m (\mathcal{AR}_{a,b+1,k}^{i})=\m (\mathcal{AR}_{a,b,k-1}^{i})+Z\cdot \m (\mathcal{AR}_{a,b}(2,3,\ldots, k+1))
\end{equation}
\noindent where
\begin{equation}\label{ep34-m-1}
Z := \begin{cases} 0, &\text{if } i\leq k\\
\dfrac{\m (\mathcal{AR}_{a,a+1}(a+k+2-i)}{\m (\ad(a))}, &\text{if } i\geq k+1\end{cases}.
\end{equation}
It now remains to show that the expression in the statement satisfies equation \eqref{ep34-1}. This is now a straightforward application of the induction hypothesis and some
algebraic manipulation.
\end{proof}
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop-ad-i-j-n.pdf}
\caption{Aztec Diamond with defects on adjacent sides; here $a=6$, $i=4$, $j=4$}
\label{fig:ad-i-j-n}
\end{figure}
\begin{proposition}\label{ad_i_j}
Let $a, i, j$ be positive integers such that $1\leq i, j\leq a$, then the number of domino tilings of $\ad(a)$ with one defect on the southeastern side at the $i$-th position counted from the south corner and one defect on the northeastern side on the $j$-th position counted from the north corner
as shown in Figure \ref{fig:ad-i-j-n} is given by
\[ 2^{a(a-1)/2}\binom{a-1}{i-1}\binom{a-1}{j-1}~_3F_2\left[\hyper{1, 1-i, 1-j}{1-a, 1-a}\,;2\right].\]
\end{proposition}
\begin{figure}[!htb]
\centering
\includegraphics[scale=.7]{prop35.pdf}
\caption{Aztec Diamond with some labelled squares; here $a=6$}
\label{fig:ad_i_j}
\end{figure}
\begin{proof}
We use induction with respect to $a$. The base case of induction is $a=2$. We would also need to check for $i=1,j=1,i=a$ and $j=a$ separately.
If $a=2$, then the only possibilities are $i=1$ or $i=a$ and $j=1$ or $j=a$, so we do not have to consider this case, once we consider the other mentioned cases.
We now note that when either $i$ or $j$ is $1$ or $a$, some dominoes are forced in any tiling
and hence we are reduced to an Aztec rectangle of size $(a-1)\times a$. It is easy to see that our formula is correct for this.
In the rest of the proof we assume $a\geq 3$ and $1<i,j<a$. Let us now denote the region we are interested in this proposition as $\ad_a(i,j)$. Using the dual graph of this region and applying
Theorem \ref{kk1} with the vertices as labelled in Figure \ref{fig:ad_i_j} we obtain the following identity (see Figure \ref{fig:kuo-3} for details),
\begin{align}\label{ep35}
\m (\ad_a(i,j))\m (\ad(a-1)) =& \m (\ad(a))\m (\ad_{a-1}(i-1,j-1))\\ \nonumber
&+ \m (\mathcal{AR}_{a-1,a}(j))\m (\mathcal{AR}_{a-1,a}(i)).
\end{align}
\begin{figure}[!htb]
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop36-abcd.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, right]{prop36-ac.pdf}
\endminipage\hfill
\vspace{4mm}
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, left]{prop36-bd.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{prop36-bc.pdf}
\endminipage
\caption{Forced dominoes in the proof of Proposition \ref{ad_i_j} where the vertices we remove are labelled}
\label{fig:kuo-3}
\end{figure}
\noindent Simplifying equation \eqref{ep35}, we get the following
\begin{equation}\label{ll}
\m (\ad_a(i,j))=2^a\m (\ad_{a-1}(i-1,j-1))+2^{a(a-1)/2}\binom{a-1}{j-1}\binom{a-1}{i-1}
\end{equation}
\noindent where we used Theorem \ref{adm} and Corollary \ref{cor1}.
Now, using our inductive hypothesis on equation \eqref{ll} we see that we get the expression in the proposition.
\end{proof}
\begin{remark}
Ciucu and Fischer \cite{ilse}, have a similar result for the number of lozenge tiling of a hexagon with dents on adjacent sides (Proposition 3 in their paper). They make use of the
following result of Kuo \cite{kuo}.
\begin{theorem}[Theorem 2.1]\cite{kuo}\label{kj}
Let $G=(V_1, V_2, E)$ be a plane bipartite graph with $\abs{V_1}=\abs{V_2}$ and $w,x,y,z$ be vertices of $G$ that appear in cyclic order on a face of $G$. If $w,y \in V_1$
and $x,z\in V_2$ then
\[ \m(G)\m(G-\{w,x,y,z\})=\m(G-\{w,x\})\m(G-\{y,z\})+\m(G-\{w,z\})\m(G-\{x,y\}).\]
\end{theorem}
\noindent They obtain the following identity
\begin{align*}
\adj(a,b,c)_{j,k}&\adj(a-1,b,c-1)_{j,k}\\
=& \adj(a,b,c-1)_{j,k}\adj(a-1,b,c)_{j,k}\\
&+ \adj(a-1,b+1,c-1)_{j,k}\adj(a,b-1,c)_{j,k}
\end{align*}
\noindent where $\adj(a,b,c)_{j,k}$ denotes the number of lozenge tilings of a hexagon $H_{a,b,c}$ with opposite side lengths $a,b,c$ with two dents on adjacent sides of length
$a$ and $c$ in positions $j$ and $k$ respectively, where $a,b,c,j,k$ are non-negative integers with $1\leq j\leq a$ and $1\leq k\leq c$.
In their use of Theorem \ref{kj}, they
take the graph $G$ to be $\adj(a,b,c)_{j,k}$, but if we take the graph $G$ to be $H_{a,b,c}$ and use Theorem \ref{kk1} with an appropriate choice of labels we obtain the following
identity
\begin{align*}
\hex(a-1,b,c)\adj(a,b,c)_{j,k} &= \hex(a,b,c)\adj(a-1,b,c)_{j,k} \\
&+ \hex(c,a-1,b+1, c-1, a,b)_k\hex(b-1, c+1, a-1, b,c,a)_j
\end{align*}
\noindent with the same notations as in Remark \ref{rem1}. Then, Proposition 3 of Ciucu and Fischer \cite{ilse} follows more easily without the need for contigous relations of hypergeometric series that they use in their paper.
\end{remark}
\section{Proofs of the main results}\label{s4}
\begin{proof}[Proof of Theorem \ref{mt1}]
\begin{figure}[!htb]
\centering
\includegraphics[scale=.6]{rect-aztec.pdf}
\caption{Removing the forced dominoes from $\mathcal{AR}_{a,b}^k$; here $a=5, b=10$, $k=5$}
\label{fig:mt11}
\end{figure}
We shall apply the formula in Theorem \ref{condensation} to the planar dual graph of our region $\mathcal{AR}_{a,b}^k$, and the vertices $\delta_1, \ldots, \delta_{2n+2k}$. Then the left hand side
of equation \eqref{ciucu2} becomes the left hand side of equation \eqref{em1}, and the right hand side of equation \eqref{ciucu2} becomes the right hand side of \eqref{em1}. We
just need to verify that the quantities expressed in equation \eqref{em1} are indeed given by the formulas described in the statement of Theorem \ref{mt1}.
The first statement follows immediately by noting that the added squares on the south eastern side of $\mathcal{AR}_{a,b}^k$ forces some domino tilings. After removing this forced dominoes we are left
with an Aztec Diamond of order $a$ as shown in Figure \ref{fig:mt11}, whose number of tilings is given by Theorem \ref{adm}.
\begin{figure}[!htb]
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{alpha-beta-no-tile.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{mt1-2-1.pdf}
\endminipage
\caption{Choices of $\beta$-defects that lead to no tiling of $\mathcal{AR}_{a,b}^k$}
\label{fig:mt12}
\end{figure}
\begin{figure}[!htb]
\centering
\includegraphics[scale=.6]{mt1-2.pdf}
\caption{Choice of $\beta$-defect, not sharing an edge with some $\gamma_l$}
\label{fig:mt13}
\end{figure}
The possibilities in the second statement are as follows. If an $\beta$ square shares an edge with some $\gamma_l$, then the region cannot be covered by any domino as illustrated in the right image of Figure \ref{fig:mt12}.
Again, if $\beta_i$ is on the northwestern side at a distance of atmost $k$ from the western corner, then the strips of forced dominoes along the sourthwestern side interfere with the $\beta_i$ and hence there
cannot be any tiling in this case as illustrated in the left image of Figure \ref{fig:mt12}. If neither of these situation is the case, then due to the squares $\gamma_1, \ldots, \gamma_k$ on the southeastern side, there are forced dominoes as shown in
Figure \ref{fig:mt13} and then $\beta_i$ and $\alpha_j$ are defects on an Aztec Diamond on adjacent sides and then the second statement follows from Proposition \ref{ad_i_j}.
\begin{figure}[!htb]
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{al-ga-no.pdf}
\endminipage\hfill
\minipage{0.50\textwidth}
\includegraphics[scale=0.6, center]{al-ga-no-opp.pdf}
\endminipage
\caption{Choices of $\beta$ and $\gamma$-defects that lead to no tiling of $\mathcal{AR}_{a,b}^k$}
\label{fig:al-ga-no}
\end{figure}
To prove the validity of the third statement, we notice that if an $\beta$ and $\gamma$ defect share an edge then, there are two possibilities, either the $\beta$ defect is above the $\gamma$ defect in which
case we have some forced dominoes as shown in the left of Figure \ref{fig:al-ga-yes} and we are reduced to finding the number of domino tilings of an Aztec Diamond; or the
$\beta$-defect is to the left of a $\gamma$-dent, in which case, we get no tilings as shown in the left of Figure \ref{fig:al-ga-no} as the forced dominoes interfere in this case.
If $\beta_i$ and $\gamma_j$ share no edge in common, then we get no tiling if the $\beta$-defect is on the northwestern side at a distance of atmost $k-1$ from the western corner as illustrated
in the right of Figure \ref{fig:al-ga-no}. If the $\beta$-defect is in the northwestern side at a distance more than $k-1$ from the western corner then the situation is as shown in
the right of Figure \ref{fig:al-ga-yes} and is described in Proposition \ref{ar_k-1_i}. If the $\beta$-defect is in the southeastern side then the situation is as shown in the middle of Figure \ref{fig:al-ga-yes}
and is described in Proposition \ref{ar_k_i}.
\begin{figure}[!htb]
\minipage{0.33\textwidth}
\includegraphics[scale=0.45, left]{al-ga-aztec.pdf}
\endminipage\hfill
\minipage{0.33\textwidth}
\includegraphics[scale=0.45, center]{al-ga-yes-same.pdf}
\endminipage\hfill
\minipage{0.33\textwidth}
\includegraphics[scale=0.45, right]{al-ga-yes-opp.pdf}
\endminipage
\caption{Choices of $\beta$ and $\gamma$-defects that lead to tiling of $\mathcal{AR}_{a,b}^k$}
\label{fig:al-ga-yes}
\end{figure}
The fourth statement follows immediately from the checkerboard drawing (see Figure \ref{fig:check}) of an Aztec rectangle and the condition that a tiling by dominoes exists for such a board if and only if
the number of white and black squares are the same. In all other cases, the number of tilings is $0$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{mt2}]
Let $\mathcal{AR}$ be the region obtained from $\mathcal{AR}_{a,b}^k$ by removing $k$ of the squares $\beta_1, \ldots, \beta_{n+k}$. We now apply Theorem \ref{condensation} to the planar dual graph of $\mathcal{AR}$, with the
removed squares choosen to be the vertices corresponding to the $n$ $\beta_i$'s inside $\mathcal{AR}$ and to $\alpha_1, \ldots, \alpha_n$. The left hand side of equation \eqref{ciucu2} is now the
required number of tilings and the right hand side of equation \eqref{ciucu2} is the Pfaffian of a $2n\times 2n$ matrix with entries of the form $\m (\mathcal{AR} \setminus \{\beta_i, \alpha_j\})$,
where $\beta_i$ is not one of the unit squares that we removed from $\mathcal{AR}_{a,b}^k$ to get $\mathcal{AR}$.
We now notice that $\m (\mathcal{AR} \setminus \{\beta_i, \alpha_j\})$ is an Aztec rectangle with all its defects confined to three of the sides. So, we can apply Theorem \ref{mt1} and it gives us
an expression for $\m (\mathcal{AR} \setminus \{\beta_i, \alpha_j\})$ as the Pfaffian of a $(2k+2)\times (2k+2)$ matrix of the type described in the statement of Theorem \ref{mt1}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{mt3}]
We shall now apply Theorem \ref{condensation} to the planar dual graph of $\ad(a)$ with removed squares choosen to correspond to $\beta_1, \ldots, \beta_n, \alpha_1, \ldots, \alpha_n$. The right hand
side of equation \eqref{ciucu2} is precisely the right hand side of equation \eqref{emt3}. If $\delta_i$ and $\delta_j$ are of the same type then $\ad(a)\setminus \{\delta_i, \delta_j\}$ does
not have any tiling as the number of black and white squares in the checkerboard setting of an Aztec Diamond will not be the same (see Figure \ref{fig:check}). Finally, the proof
is complete once we note that $\ad(a)\setminus \{\beta_i, \alpha_j\}$ is an Aztec Diamond with two defects removed from adjacent sides for any choice of $\beta_i$ and $\alpha_j$ and is given by
Proposition \ref{ad_i_j}.
\end{proof}
\bibliographystyle{amsplain}
|
\section*{References}}
\DeclareMathOperator*{\argmin}{arg\,min}
\newtheorem{definition}{Definition}
\newtheorem{theorem}{Theorem}
\title{Understanding Convolutional Neural Networks}
\author{
Jayanth Koushik \\
Language Technologies Institute \\
Carnegie Mellon University \\
Pittsburgh, PA 15213 \\
\texttt{<EMAIL>} \\
}
\begin{document}
\maketitle
\begin{abstract}
Convoulutional Neural Networks (CNNs) exhibit extraordinary performance
on a variety of machine learning tasks. However, their mathematical
properties and behavior are quite poorly understood. There is some
work, in the form of a framework, for analyzing the operations that they
perform. The goal of this project is to present key results from this
theory, and provide intuition for why CNNs work.
\end{abstract}
\section{Introduction}
\subsection{The supervised learning problem}
We begin by formalizing the supervised learning problem which CNNs are
designed to solve. We will consider both regression and classification, but
restrict the label (dependent variable) to be univariate.
Let $X\in\mathcal{X}\subset\mathbb{R}^d$ and $Y\in\mathcal{Y}\subset\mathbb{R}$
be two random variables. We typically have $Y = f(X)$ for some unknown $f$.
Given a sample $\{(x_i,y_i)\}_{i=1,\dots,n}$
drawn from the joint distribution of $X$ and $Y$, the goal of supervised
learning is to learn a mapping $\hat{f}:\mathcal{X}\to\mathcal{Y}$ which
minimizes the expected loss, as defined by a suitable loss function
$L:\mathcal{Y}\times\mathcal{Y}\to\mathbb{R}$.
However, minimizing over the set of all functions from $\mathcal{X}$ to
$\mathcal{Y}$ is ill-posed, so we restrict the space of hypotheses to some set
$\mathcal{F}$, and define
\begin{align}
\label{eq:supl}
\hat{f} = \argmin_{f\in\mathcal{F}}\mathrm{E}[L(Y,f(X))]
\end{align}
\subsection{Linearization}
A common strategy for learning classifiers, and the one employed by kernel
methods, is to linearize the variations in $f$ with a feature representation.
A feature representation is any transformation of the input variable $X$; a
change of variable. Let this transformation be given by $\Phi(X)$. Note that
the transformed variable need not have a lower dimension than $X$. We would
like to construct a feature representation such that $f$ is linearly separable
in the transformed space i.e.
\begin{align}
f(X) = \langle\Phi(X),w\rangle
\end{align}
for regression, or
\begin{align}
f(X) = \mathit{sign}(\langle\Phi(X),w\rangle)
\end{align}
for binary classification\footnote{Multi-class classification problems can
be considered as multiple binary classification problems.}. Classification
algorithms like Support Vector Machines (SVM)~\cite{cortes1995support} use a
fixed feature representation that may, for instance, be defined by a kernel.
\subsection{Symmetries}
The transformation induced by kernel methods do not always linearize $f$
especially in the case of natural image classification. To find suitable
feature transformations for natural images, we must consider their invariance
properties. Natural images show a wide range of invariances e.g. to pose,
lighting, scale. To learn good feature representations, we must suppress these
intra-class variations, while at the same time maintaining inter-class
variations. This notion is formalized with the concept of symmetries as
defined next.
\begin{definition}[Global Symmetry]
Let $g$ be an operator from $\mathcal{X}$ to $\mathcal{X}$. $g$ is a
global symmetry of $f$ if $f(g.x) = f(x)\ \forall x \in \mathcal{X}$.
\end{definition}
\begin{definition}[Local Symmetry]
Let $G$ be a group of operators from $\mathcal{X}$ to $\mathcal{X}$ with norm
$|.|$. $G$ is a
group of local symmetries of $f$ if for each $x \in \mathcal{X}$, there exists
some $C_x > 0$ such that $f(g.x) = f(x)$ for all $g \in G$ such that
$|g| < C_x$.
\end{definition}
Global symmetries rarely exist in real images, so we can try to construct
features that linearize $f$ along local symmetries. The symmetries we will
consider are translations and diffeomorphisms, which are discussed next.
\subsection{Translations and Diffeomorphisms}
Given a signal $x$, we can interpolate its dimensions and define $x(u)$ for all
$u \in \mathbb{R}^n$ ($n = 2$ for images). A translation is an operator $g$
given by $g.x(u) = x(u - g)$. A diffeomorphism is a deformation; small
diffeomorphisms can be written as $g.x(u) = x(u - g(u))$.
We seek feature transformations $\Phi$ which linearize the action of local
translations and diffeomorphisms. This can be expressed in terms of a
Lipschitz continuity condition.
\begin{align}
\label{eq:lips}
\|\Phi(g.x) - \Phi(x)\| \le C|g|\| x\|
\end{align}
\subsection{Convolutional Neural Networks}
\label{subsec:cnn}
Convolutional Neural Networks (CNNs), introduced by~\citet{le1990handwritten}
are a class of biologically inspired neural networks which solve equation
$\eqref{eq:supl}$ by passing $X$ through a series of convolutional
filters and simple non-linearities. They have shown remarkable results in
a wide variety of machine learning problems~\cite{lecun2015deep}. Figure
$\ref{figure:cnn}$ shows a typical CNN architecture.
A convolutional neural network has a hierarchical architecture. Starting
from the input signal $x$, each subsequent layer $x_j$ is computed as
\begin{align}
x_j = \rho W_j x_{j-1}
\end{align}
Here $W_j$ is a linear operator and $\rho$ is a non-linearity. Typically,
in a CNN, $W_j$ is a convolution, and $\rho$ is a rectifier $\max(x, 0)$ or
sigmoid $\nicefrac{1}{1+\exp(-x)}$. It is easier to think of the operator
$W_j$ as a stack of convolutional filters. So the layers are filter maps and
each layer can be written as a sum of convolutions of the previous layer.
\begin{align}
x_j(u, k_j) = \rho\Big(\sum_k (x_{j-1}(.,k)\ast W_{j,k_j}(.,k))(u)\Big)
\end{align}
Here $\ast$ is the discrete convolution operator:
\begin{align}
(f \ast g)(x) = \sum_{u=-\infty}^{\infty} f(u)g(x - u)
\end{align}
The optimization problem defined by a convolutional neural network is
highly non-convex. So typically, the weights $W_j$ are learned by stochastic
gradient descent, using the backpropagation algorithm to compute gradients.
\begin{figure}[!t]
\includegraphics[width=\textwidth]{cnn}
\caption{Architecture of a Convolutional Neural Network (from
\citeauthor{lecun1998gradient}~\cite{lecun1998gradient})}
\label{figure:cnn}
\end{figure}
\subsection{A mathematical framework for CNNs}
\citet{mallat2016understanding} introduced a mathematical framework for
analyzing the properties of convolutional networks. The theory is based on
extensive prior work on wavelet scattering (see for example~
\cite{bruna2013invariant,anden2014deep}) and illustrates that
to compute invariants, we must separate variations of $X$ at different
scales with a wavelet transform. The theory is a first step towards
understanding general classes of CNNs, and this paper presents its key
concepts.
\section{The need for wavelets}
Although the framework based on wavelet transforms is quite successful
in analyzing the operations of CNNs, the motivation or need for wavelets
is not immediately obvious. So we will first consider the more general
problem of signal processing, and study the need for wavelet transforms.
In what follows, we will consider a function $f(t)$ where $t\in\mathbb{R}$
can be considered as representing time, which makes $f$ a time varying
function like an audio signal. The concepts, however, extend quite naturally
to images as well, when we change $t$ to a two dimensional vector.
Given such a signal, we are often interested in studying its variations
across time. With the image metaphor, this corresponds to studying
the variations in different parts of the image. We will consider a progression
of tools for analyzing such variations. Most of the following material is from
the book by \citeauthor{gerald1994friendly}~\cite{gerald1994friendly}.
\subsection{Fourier transform}
The Fourier transform of $f$ is defined as
\begin{align}
\label{eq:fourier}
\hat{f}(\omega) \equiv \int_{-\infty}^{\infty}f(t)e^{-2\pi i\omega t}dt
\end{align}
The Fourier transform is a powerful tool which decomposes $f$ into the
frequencies that make it up. However, it should be quite clear from equation
$\eqref{eq:fourier}$ that it is useless for the task we are interested in.
Since the integral is from $-\infty$ to $\infty$, $\hat{f}$ is an average
over all time and does not have any local information.
\subsection{Windowed Fourier transform}
To avoid the loss of information that comes from integrating over all time,
we might use a weight function that localizes $f$ in time. Without going
into specifics, let us consider some function $g$ supported on
$[-T, 0]$ and define the windowed Fourier transform (WFT) as
\begin{align}
\label{eq:wft}
\tilde{f}(\omega,t) \equiv \int_{-\infty}^{\infty}f(u)g(u-t)
e^{-2\pi i\omega u}du
\end{align}
It should be intuitively clear that the WFT can capture local variations
in a time window of width $T$. Further, it can be shown that the WFT also
provides accurate information about $f$ in a frequency band of some width
$\Omega$. So does the WFT solve our problem? Unfortunately not; and this is
a consequence of Theorem $\ref{thm:uncertain}$ which is stated very
informally next.
\begin{theorem}[Uncertainty Principle]
\label{thm:uncertain}
\footnote{Contrary to popular belief, the Uncertainty Principle is a
mathematical, not physical property.}
Let $f$ be a function which is small outside a time-interval of length $T$,
and let its Fourier transform be small outside a frequency-band of width
$\Omega$. There exists a positive constant $c$ such that
$$ \Omega T \ge c $$
\end{theorem}
Because of the Uncertainty Principle, $T$ and $\Omega$ cannot both be
small. Roughly speaking, this implies that the WFT cannot capture small
variations in a small time window (or in the case of images, a small patch).
\subsection{Continuous wavelet transform}
The WFT fails because it introduces scale (the width of the window) into
the analysis. The continuous wavelet transform involves scale too, but it
considers all possible scalings and avoids the problem faced by the WFT.
Again, we begin with a window function $\psi$ (supported on $[-T,0]$),
this time called a mother wavelet. For some fixed $p\ge 0$, we define
\begin{align}
\label{eq:wavelet}
\psi_s(u) \equiv |s|^{-p}\psi\Big(\frac{u}{s}\Big)
\end{align}
The scale $s$ is allowed to be any non-zero real number. With this family
of wavelets, we define the continuous wavelet transform (CWT) as
\begin{align}
\label{eq:cwt}
\tilde{f}(s, t) \equiv (f \ast \psi_s)(t)
\end{align}
where $\ast$ is the continuous convolution operator:
\begin{align}
\label{eq:conv}
(p \ast q)(x) \equiv \int_{-\infty}^\infty p(u)q(x - u)du
\end{align}
The continuous wavelet transform captures variations in $f$ at a particular
scale. It provides the foundation for the operation of CNNs, as will be
explored next.
\section{Scale separation with wavelets}
Having motivated the need for a wavelet transform, we will now construct
a feature representation using the wavelet transform. Note that convolutional
neural network are covariant to translations because they use convolutions
for linear operators. So we will focus on transformations that linearize
diffeomorphisms.
\begin{theorem}
\label{thm:scale}
Let $\phi_J(u) = 2^{-nJ}\phi(2^{-J}u)$ be an averaging kernel with
$\int\phi(u)du = 1$. Here $n$ is the dimension of the index in $\mathcal{X}$,
for example, $n = 2$ for images. Let $\{\psi_{k}\}_{k=1}^K$ be a set of $K$
wavelets with zero average: $\int\psi_k(u)du = 0$, and from them define
$\psi_{j,k}(u) \equiv 2^{-jn}\psi_k(2^{-j}u)$. Let $\Phi_J$ be a feature
transformation defined as
\begin{align*}
\Phi_Jx(u,j,k) = |x \ast \psi_{j,k}| \ast \phi_J(u)
\end{align*}
Then $\Phi_J$ is locally invariant to translations at scale $2^J$, and
Lipschitz continuous to the actions of diffemorphisms as defined by equation
$\eqref{eq:lips}$ under the following diffeomorphism norm.
\begin{align}
\label{eqn:diffnorm}
|g| = 2^{-J}\sup_{u\in\mathbb{R}^n}|g(u)| +
\sup_{u\in\mathbb{R}^n}|\nabla g(u)|
\end{align}
\end{theorem}
Theorem $\ref{thm:scale}$ shows that $\Phi_J$ satisfies the regularity
conditions which we seek. However, it leads to a loss of information due
to the averaging with $\phi_J$. The lost information is recovered by
a hierarchy of wavelet decompositions as discussed next.
\section{Scattering Transform}
\begin{figure}[!t]
\includegraphics[width=\textwidth]{scatter}
\caption{Architecture of the scattering transform (from
\citeauthor{estrach2012scattering}~\cite{estrach2012scattering})}
\label{figure:scatter}
\end{figure}
Convolutional Neural Networks transform their input with a series of linear
operators and point-wise non-linearities. To study their properties, we first
consider a simpler feature transformation, the scattering transform introduced
by \citeauthor{mallat2012group}~\cite{mallat2012group}. As was discussed in
section $\ref{subsec:cnn}$, CNNs compute multiple convolutions across channels
in each layer; So as a simplification, we consider the transformation obtained
by convolving a single channel:
\begin{align}
x_j(u, k_j) = \rho\Big((x_{j-1}(., k_{j-1}) \ast W_{j, h})(u)\Big)
\end{align}
Here $k_j = (k_{j-1}, h)$ and $h$ controls the hierarchical structure of the
transformation. Specifically, we can recursively expand the above equation
to write
\begin{align}
x_J(u, k_J) = \rho(\rho(\dots\rho(x\ast W_{1,h_1})\ast\dots)\ast W_{J,h_J})
\end{align}
This produces a hierarchical transformation with a tree structure rather than
a full network. It is possible to show that the above transformation has
an equivalent representation through wavelet filters i.e. there exists a
sequence $p \equiv (\lambda_{1},\dots,\lambda_{m})$ such that
\begin{align}
\label{eqn:scatter}
x_J(u, k_J) = S_J[p]x(u) \equiv (U[p]x \ast \phi_J)(u) \equiv
(\rho(\rho(\dots\rho(x\ast \psi_{\lambda_1})\ast\dots)\ast
\psi_{\lambda_m})\ast\phi_J)(u)
\end{align}
where the $\psi_{\lambda_i}$s are suitably choses wavelet filters and $\phi_J$
is the averaging filter defined in Theorem $\ref{thm:scale}$. This is the
wavelet scattering transfom; its structure is similar to that of a
convolutional neural network as shown in figure $\ref{figure:scatter}$, but its
filters are defined by fixed wavelet functions instead of being learned from
the data. Further, we have the following theorem about the scattering
transform.
\begin{theorem}
\label{thm:scatterdiff}
Let $S_J[p]$ be the scattering transform as defined by equation
$\eqref{eqn:scatter}$. Then there exists $C > 0$ such that for all
diffeomorphisms $g$, and all $L^2(\mathbb{R}^n)$ signals $x$,
\begin{align}
\|S_J[p]g.x - S_J[p]x\| \le Cm|g|\|x\|
\end{align}
with the diffeomorphism norm $|g|$ given by equation
$\eqref{eqn:diffnorm}$.
\end{theorem}
Theorem $\ref{thm:scatterdiff}$ shows that the scattering transform is
Lipschitz continuous to the action of diffemorphisms. So the action of small
deformations is linearized over scattering coefficients. Further, because of
its structure, it is naturally locally invariant to translations. It has
several other desirable properties~\cite{estrach2012scattering}, and can be
used to achieve state of the art classification errors on the MNIST digits
dataset~\cite{bruna2013invariant}.
\section{General Convolutional Neural Network Architectures}
The scattering transform described in the previous section provides a simple
view of a general convolutional neural netowrk. While it provides
intuition behind the working of CNNs, the transformation suffers from high
variance and loss of information because we only consider single channel
convolutions. To analyze the properties of general CNN architectures, we
must allow for channel combinations. \citeauthor{mallat2016understanding}
\cite{mallat2016understanding} extends previously introduced tools to develop
a mathematical framework for this analysis. The theory is, however, out of the
scope of this paper. At a high level, the extension is achieved by replacing
the requirement of contractions and invariants to translations by contractions
along \emph{adaptive} groups of local symmetries. Further, the wavelets are
replaced by adapted filter weights similar to deep learning models.
\section{Conclusion}
In this paper, we tried to analyze the properties of convolutional neural
networks. A simplified model, the scattering transform was introduced as
a first step towards understanding CNN operations. We saw that
the feature transformation is built on top of wavelet transforms which
separate variations at different scales using a wavelet transform. The
analysis of general CNN architectures was not considered in this paper, but
even this analysis is only a first step towards a full mathematical
understanding of convolutional neural networks.
\bibliographystyle{plainnat}
|
\section{Introduction}
\label{sec:intro}
A magician enters the room with a 32-card deck. He invites five volunteers to the stage and claims he will read their minds. Another volunteer is asked to cut the deck a few times and pass the top five cards to the volunteers, one for each. ``Now I need you to think about your card and I will tell what it is,'' the magician says. Silence. ``Please concentrate! Think harder.'' A long pause. ``Okay, the weather is not good today. It is interfering with the brainwaves between us. I need you to work with me a bit,'' the magician begs. ``Could the people with red cards move one step closer to me?'' Another long pause. ``Hmm, you have the six of clubs. You have the five of spades...'' Sure enough, he gets them all!
This is Diaconis' mind-reading trick~\cite{Diaconis--Graham2011,goresky2012algebraic}. The magic makes use of a binary de Bruijn sequence of order 5~\cite{deBruijn1946}, which is a length-32 circulant binary sequence such that every length-5 binary string occurs as a contiguous subsequence exactly once. The magician enters the room with the 32 cards prearranged such that their color (black/red) corresponds to the de Bruijn sequence. Cutting the deck only shifts the sequence cyclically. By the property of de Bruijn sequence, knowing the colors reveals the location (or \textit{phase}) of the 5 contiguous cards inside the deck, hence uniquely determines their identities. More generally, this trick can be performed with $k$ volunteers and a deck of size $n=2^k$, by using a de Bruijn sequence of order $k$, which is a binary sequence such that every length-$k$ binary string occurs as a contiguous subsequence exactly once~\cite{deBruijn1946}.
Suppose now that some of the volunteers are not collaborative and may lie when asked about their card color. Can the magician still guess the cards correctly? In other words, can one design a length-$n$ sequence such that the set of all length-$k$ contiguous subsequences forms a good error-correcting code? Besides its appeal as a card trick, such a sequence can also be useful e.g. for phase detection in positioning systems. Imagine
that a satellite sends the length-$n$ sequence periodically. A user hearing a noisy chunk of the sequence would like to figure out the location of his chunk within the original sequence, so as to measure the transmission delay and compute his distance to the satellite. Fixing the sequence length $n$ (which results in a given ambiguity of the distance estimation), it is clearly desirable to minimize $k$, as this results in the fastest positioning. Clearly, $k$ cannot be smaller than $\log{n}$, and this lower bound can be achieved in case there is no noise, by using a de Bruijn sequence of order $k$. As we shall see, in the noisy case $k=O(\log n)$ is also sufficient, and we will in fact be interested in characterizing the exact constant $\frac{\log{n}}{k}$, which will be referred to as \textit{rate}.
In reality, positioning systems typically employ multiple satellites, each transmitting its own length-$n_i$ sequence. Sequences get combined through a multiple access channel (MAC) when reaching the user. Upon hearing a length-$k$ chunk of the combined sequence, the user wishes to measure his distance to all of the satellites by locating the chunk within each one of the sequences. We note that existing techniques (such as GPS~\cite{wiki:gps}) typically employ sequences (e.g. Gold codes~\cite{Gold1967}) that possess good autocorrelation and cross-correlation properties, and use $k = N\cdot n_1 = \cdots = N\cdot n_L$, for some repetition factor $N\geq 1$. From our perspective, these systems hence operate at zero rates. In fact, when the repetition factor $N>1$, this does not precisely fall under our setup; we further remark on this in Example~\ref{ex:GPS}. In what follows, we focus on fast positioning at non-zero rates. We are interested in characterizing the optimal trade-offs among
$
\bigl(\frac{\log{n_1}}{k},\cdots,\frac{\log{n_L}}{k}\bigr)
$
that ensure successful detection, as well as in constructing sequences that achieve the optimal trade-offs.
In what follows, we refer to the first problem, which only involves a single-sequence design, as \emph{point-to-point phase detection}. We refer to the second problem as \emph{multiple access phase detection}. Different noise models are considered: the adversarial noise and the probabilistic noise. For the probabilistic noise, different error criteria are discussed: the vanishing error criterion and the zero error criterion. These models are defined formally in the sequel. We also compare the phase detection problems to their natural channel coding counterparts.
\subsection{Point-to-Point Phase Detection}
\label{sec:intro-p2p}
In Sections~\ref{sec:advs},~\ref{sec:prob},~and~\ref{sec:prob-zero}, we consider point-to-point phase detection.
An $(n,k)$ \emph{point-to-point phase detection scheme} consists of
\begin{itemize}
\item a sequence $x^n \triangleq (x_1, x_2,\ldots,x_{n}) \in \mathcal{X}^n$, and
\item a detector ${\hat{m}} \colon \mathcal{Y}^k \to [n]\cup \mathrm{e}$, where $[n]\triangleq\{1,2,\ldots,n\}$ and $\mathrm{e}$ is an error symbol.
\end{itemize}
We assume that the detector observes a noisy version $y^k$ of the sequence $x_m^{m+k-1}$, and attempts to correctly identify the \emph{phase} $m$. Clearly, any reliable scheme would require $k \ge \log_{|\mathcal{X}|}n$. Thus, it is natural to define the \emph{efficiency} of a scheme as the excess multiplicative factor it uses over the minimal possible, i.e., $k/\log_{|\mathcal{X}|}{n}$. However, for comparison to channel coding, it would be more convenient to work with the inverse of this quantity and take logarithms in base 2, namely work with the \emph{rate}
\[
R \triangleq \frac{\log_2{n}}{k}.
\]
We note that any phase detection scheme induces a \textit{codebook}\footnote{The codebook is treated as a multiset, namely there might be repetitions in its elements.} $\mathcal{C} = \{x_{m}^{m+k-1}\colon m\in[n]\}\subseteq \mathcal{X}^k$ of rate $R$. Here and throughout indices are taken cyclically, modulo the set $[n]$. Also, we assume throughout that $k\leq n$.
We discuss three distinct models: the adversarial noise model in Section~\ref{sec:advs}, the probabilistic noise with vanishing error in Section~\ref{sec:prob}, and the probabilistic noise with zero error in Section~\ref{sec:prob-zero}. For convenience, let the function $\phi(m; x^n)$ return the length-$k$ contiguous subsequence of $x^n$ starting at phase $m$, i.e., $\phi(m; x^n) = x_{m}^{m+k-1}$. We will typically omit the dependence on the sequence $x^n$, and simple write $\phi(m; x^n) = \phi(m)$.
For the adversarial noise model, we assume that $\mathcal{X} = \mathcal{Y} = \{0,1\}$ and the observation sequence $y^k$ is obtained from $\phi(m)$ by flipping at most $pk$ bits, where $m$ is the correct phase, and $p$ is fixed and given. We define the \emph{minimum distance} of a scheme as the minimum Hamming distance of its induced codebook. A rate $R$ is said to be \emph{achievable} in this setting if, for a divergent sequence of $k$'s, there exist $(n,k)$ schemes with $\frac{\log{n}}{k} \ge R$, such that $m$ can be recovered from $y^k$ without error. Namely, we require the scheme to have a minimum distance $d > 2pk$. The \emph{capacity} of adversarial phase detection $C_\text{ad}(p)$ is defined as the supremum over all achievable rates\footnote{Here we define capacity asymptotically. Note that similarly to adversarial channel coding, it is not guaranteed short sequences with rate above the capacity do not exist.}.
Several works have addressed this noise model in the literature. The trade-off between the rate and the minimum distance of the code was studied in~\cite{Kumar--Wei1992, Hagita--Matsumoto--Natsu--Ohtsuka2008}. Kumar and Wei provided a lower bound on $d$ in the regime of $d \le \sqrt{k}$ for \emph{$m$-sequences}, which are generated by linear feedback shift registers~\cite{Kumar--Wei1992}. Some explicit sequence constructions were also provided in~\cite{Krishnamachari--Yedavalli2007, Malzbender--Reth--Ordentlich2012, Jorissen--Maesen--Doshi--Bekaert2014, Horvath--Herout--Szentrandrasi--Zacharias2013,Berkowitz--Kopparty2015}.
By a concatenation of an optimal binary channel code with the Reed--Solomon code, Berkowitz and Kopparty have recently constructed a phase detection scheme with nonzero rate and nonzero relative distance~\cite{Berkowitz--Kopparty2015}. For generalization to two dimensional phase detection, see~\cite{MacWilliams--Sloane1976, Etzion1988, Paterson1994, Bruckstein--Etzion--Giryes--Gordon--Holt--Shuldiner2012}.
In Section~\ref{sec:advs}, we focus on the tradeoff between the rate and the minimum distance in the asymptotic limit. We note that a codebook induced by any phase detection scheme can be used as a channel code in the standard binary adversarial channel model~\cite{Hamming1950}. The capacity of the latter setup is unknown. Clearly however, any upper bound for that capacity, such as the MRRW upper bound~\cite{McEliece--Rodemich--Rumsey--Welch1977}, also serves as an upper bound for $C_\text{ad}(p)$. The best known binary adversarial channel coding lower bound is given by Gilbert and Varshamov~\cite{Gilbert1952,Varshamov1957}. Applying the Lov\'{a}sz local lemma~\cite{Lovasz--Erdos1975}, we show
in Section~\ref{sec:advs-limit} that this rate is also achievable for adversarial phase detection. In Section~\ref{sec:advs-code}, we characterize the family of \emph{linear} phase detection schemes and study their performance.
For the probabilistic noise model with vanishing error criterion, we assume that the phase is uniformly distributed, i.e., $M \sim \mathrm{Unif}[n]$. We further assume that the noisy observation $y^k$ is obtained from $\phi(m)$ via a discrete memoryless channel $p(y|x)$. The probability of error is defined as
\[
P_e^{(k)} = \P\{M \neq \hat{m}(Y^k)\}.
\]
A rate $R$ is said to be \emph{achievable} if, for a divergent sequence of $k$'s, there exist $(n,k)$ schemes with $\frac{\log{n}}{k} \ge R$ and $\lim_{k\to \infty}P_e^{(k)} = 0$. The \emph{vanishing error capacity} of probabilistic phase detection $C_\text{ve}$ is defined as the supremum over all achievable rates.
As before, the codebook induced by any phase detection scheme is also a channel code. Thus, the Shannon capacity of the channel $p(y|x)$ is an upper bound for $C_\text{ve}$. In Section~\ref{sec:prob-limit}, we show that in fact $C_\text{ve}$ equals the Shannon capacity. Moreover, we present in Section~\ref{sec:prob-code} a concatenated construction with $O(k\log k)$ complexity that achieves the capacity of probabilistic phase detection. As a consequence, this construction also establishes the equivalence between channel coding and phase detection for this noise model.
For the probabilistic noise model with zero error criterion, we again assume that the noisy observation $y^k$ is obtained from $\phi(m)$ via a discrete memoryless channel $p(y|x)$. A rate $R$ is said to be \emph{achievable} if, for a divergent sequence of $k$'s, there exist $(n,k)$ schemes with $\frac{\log{n}}{k} \ge R$ such that the phase $m$ can be recovered with \emph{zero error} for any $m \in [n]$. Similar to Shannon's zero error channel coding~\cite{Shannon1956}, achievable rates can be
equivalently defined on the \emph{confusion graph} $G = (\mathcal{X},E)$ associated with the channel $p(y|x)$. Here the vertex set is $\mathcal{X}$ and two distinct vertices are connected $(u,v) \in E$ if they may result in the same output, i.e., there exists a $y \in \mathcal{Y}$ such that $p_{Y|X}(y|u) > 0$ and $p_{Y|X}(y|v)>0$. Let $G^k = (\mathcal{X}^k,E_k)$ be the $k$-fold \emph{strong product} of $G$, where two distinct vertices are connected $(u^k, v^k) \in E_k$ if for all $i \in [k]$, either $u_i = v_i$ or $(u_i, v_i)
\in E$. Then, a rate $R$ is achievable if and only if, for a divergent sequence of $k$'s, there exist $(n,k)$ schemes with $\frac{\log{n}}{k} \ge R$ such that $(\phi(m), \phi(m')) \notin E_k$ for any two distinct phases $m, m'\in [n]$, or in other words, the induced codebook forms an independent set of $G^k$. The \emph{zero error capacity} $C_\text{ze}(G)$ is defined as the supremum over all achievable rates.
We note the distinction between phase detection and channel coding under the zero error criterion. For zero error channel coding (in contrast to vanishing error and adversarial channel coding) if a rate $R$ is achievable at some length $k$, it is also achievable for all multiples of $k$ (by concatenation) and thus for a divergent sequence of $k$'s. However, this argument cannot be applied to the phase detection setting, since concatenating the codewords of two induced codebooks may not necessarily result in a new codebook that can be chained up into a single sequence. Nevertheless, and despite the fact that the zero error channel capacity is generally unknown, we show in Section~\ref{sec:prob-zero} that the zero error capacity for phase detection coincides with its channel coding counterpart.
\subsection{Multiple Access Phase Detection}
In Sections~\ref{sec:mac-prob}~and~\ref{sec:mac-advs}, we consider multiple access phase detection. We only discuss the two-user case for simplicity. But all the results extend to more users.
An $(n_1,n_2,k)$ \emph{multiple access phase detection scheme} consists of
\begin{itemize}
\item two sequences $x_1^{n_1} = (x_{11}, x_{12},\ldots,x_{1,n_1}) \in \mathcal{X}_1^n$ and $x_{2}^{n_2} = (x_{21},x_{22},\ldots,x_{2,n_2}) \in \mathcal{X}_2^n$, and
\item a detector that declares two phase estimates ${\hat{m}}_1\colon \mathcal{Y}^k \to [n_1]\cup \{\mathrm{e}\}$ and ${\hat{m}}_2\colon \mathcal{Y}^k \to [n_2]\cup \{\mathrm{e}\}$.
\end{itemize}
We assume that the detector observes $y^k$, which is the output of a discrete memoryless multiple access channel $(\mathcal{X}_1\times \mathcal{X}_2, p(y|x_1,x_2), \mathcal{Y})$ with the two inputs $\phi_1(m_1) = \phi_1(m_1; x_1^{n_1})\triangleq (x_{1,m_1},x_{1,m_1+1},\ldots,x_{1,m_1+k-1})$ and $\phi_2(m_2) = \phi_2(m_2; x_2^{n_2}) \triangleq (x_{2,m_2},x_{2,m_2+1},\ldots,x_{2,m_2+k-1})$, and attempts to correctly identify the \emph{phases} $(m_1,m_2)$. Similar to the point-to-point case, we define the \emph{rates} of the two sequences as
\[
R_1 \triangleq \frac{ \log_2{n_1}}{k} \quad \text{ and } \quad R_2 \triangleq \frac{\log_2 n_2}{k}.
\]
We note that every multiple access phase detection scheme induces two (multiset) \textit{codebooks}
\begin{equation}
\label{eqn:c1}
\mathcal{C}_1 = \{\phi_1(m_1)\colon m_1\in[n_1]\}\subseteq \mathcal{X}_1^k
\end{equation}
and
\begin{equation}
\label{eqn:c2}
\mathcal{C}_2 = \{\phi_2(m_2)\colon m_2\in[n_2]\}\subseteq \mathcal{X}_2^k
\end{equation}
of rates $R_1$ and $R_2$ respectively.
We discuss two different error criteria: the vanishing error criterion in Section~\ref{sec:mac-prob} and the zero error criterion in Section~\ref{sec:mac-advs}.
Under the vanishing error criterion, we assume that the phase pair $(M_1,M_2)$ is uniformly distributed over $[n_1] \times [n_2]$. The probability of error is defined as
\[
P_e^{(k)} = \P\{(M_1,M_2) \neq ({\hat{m}}_1(Y^k),{\hat{m}}_2(Y^k))\}.
\]
A rate pair $(R_1,R_2)$ is said to be \emph{achievable} if, for a divergent sequence of $k$'s, there exist $(n_1,n_2,k)$ schemes with $\frac{\log{n_1}}{k} \ge R_1$, $\frac{\log{n_2}}{k} \ge R_2$, and $\lim_{k\to \infty}P_e^{(k)} = 0$. The \emph{vanishing error capacity region} $\mathscr{C}_\text{ve}$ is defined as the closure of the set of achievable rate pairs.
In Section~\ref{sec:mac-prob-limit}, we establish the vanishing error capacity region of multiple access phase detection. This region turns out to be \emph{strictly} included, in general, in the capacity region of its channel coding counterpart. This is in contrast to all models in the point-to-point case, in which phase detection either achieves the same best known rate or shares the same capacity as its channel coding counterpart. Due to the lack of synchronization between sequences, a phase detection scheme achieves at best the usual MAC capacity region without the time-sharing random variable. In Section~\ref{sec:mac-prob-code}, we provide a low-complexity ($O(k\log{k})$) sequence construction that achieves any rate pair in the capacity region.
Under the zero error criterion, a rate pair $(R_1,R_2)$ is said to be \emph{achievable} if, for a divergent sequence of $k$'s, there exist $(n_1,n_2,k)$ schemes with $\frac{\log{n_1}}{k} \ge R_1$ and $\frac{\log{n_2}}{k} \ge R_2$ such that $(m_1,m_2)$ can be recovered from $y^k$ with \emph{zero error} for any pair $(m_1,m_2)\in [n_1]\times [n_2]$. The \emph{zero error capacity region} $\mathscr{C}_\text{ze}$ is defined as the closure of the set of achievable rate pairs. We note that the problem of zero error phase detection in MACs is generally very difficult, as it is at least as hard as the zero error MAC coding problem, which in turn is open even in the simplest cases, e.g., the binary adder channel~\cite{Lindstrom1969,Kasami--Lin1978, Mattas--Ostergard2005, Urbanke--Li1998, Ordentlich--Shayevitz2016,Austrin--Kaski--Koivisto--Nederlof2016}. Nevertheless, in Section~\ref{sec:separation}, we demonstrate the distinction between the phase detection and the channel coding problems, by showing a separation between their capacity regions.
In Sections~\ref{sec:mac-advs-limit} and~\ref{sec:mac-advs-code}, we restrict our attention to a simple channel model, the modulo-2 addition channel with $\mathcal{X}_1 = \mathcal{X}_2 = \mathcal{Y} = \{0,1\}$ and $Y = X_1 \oplus X_2$. For this channel, a rate pair $(R_1,R_2)$ is achievable if every element in the sumset
\begin{equation}
\label{eqn:csum}
\mathcal{C}_\text{sum}\triangleq \{\phi_1(m_1)\oplus \phi_2(m_2)\colon m_1 \in [n_1], m_2 \in [n_2]\}
\end{equation}
can be uniquely expressed as an element in the induced codebook $\mathcal{C}_1$ plus an element in the induced codebook $\mathcal{C}_2$. Note that $\mathcal{C}_\text{sum}$ is defined as a regular set with distinct elements (rather than a multiset). Hence, any $\mathcal{C}_1$ and $\mathcal{C}_2$ induced by a valid scheme must also have distinct elements.
Clearly, the zero-error channel coding capacity region $\{(R_1,R_2)\colon R_1 + R_2 \le 1\}$ is an outer bound for that of phase detection. In Section~\ref{sec:mac-advs-limit}, we establish the achievability of this region by a random construction that exploits properties of linear codes, in a way that resembles Wyner's linear Slepian--Wolf codes~\cite{Wyner1974}. We further provide in Section~\ref{sec:mac-advs-code} an explicit sequence construction that achieves this region, by exploiting
properties from finite field theory. As an consequence, the induced code from our phase detection sequences can be used for channel coding and achieve any rate pair in the zero-error capacity region, \emph{without using time sharing}\footnote{For other channel codes that achieve this region without using time sharing, see for example~\cite{Cohen--Litsyn--Vardy--Zemor1996, Ostergard--Kaikkonen1996}. For a channel code that achieves the rate pair $(1/2,1/2)$ with the \emph{same} codebook,
see~\cite{Lindstrom1969,Poltyrev--Snyders1995} for a construction utilizing the parity check matrix of a BCH code.}.
\section{Point-to-Point: Adversarial Noise}
In this section we discuss the adversarial noise model. We first examine whether the
adversarial phase detection schemes achieve the best known rate for adversarial channel coding, namely the Gilbert--Varshamov (GV) bound~\cite{Gilbert1952,Varshamov1957}.
\label{sec:advs}
\subsection{Fundamental Limit}
\label{sec:advs-limit}
\begin{theorem}
\label{thm:advs-nkd-bound}
An $(n,k)$ point-to-point phase detection scheme with minimum distance $d$ exists if
\begin{equation}
\label{eqn:advs-nkd-bound}
n \le \frac{2^k}{16k \sum_{i=0}^d \binom{k}{i}}.
\end{equation}
\end{theorem}
\begin{corollary}
\label{thm:advs-limit}
The capacity for adversarial phase detection is lower bounded by
\[
C_\text{ad}(p) \geq 1-h(2p),
\]
where $h(\cdot)$ is the binary entropy function.
\end{corollary}
We show the existence of a good sequence using the probabilistic method. We note that while several different proofs of the GV bound exist~\cite{Gilbert1952,Varshamov1957,Tolhuizen1997}, none of them seem to directly extend to our setting. This is simply due to the fact that there is a dependence between the codewords in the induced codebook. To alleviate this technical difficulty, we need the following well-known lemma.
\begin{lemma}[Lov\'{a}sz Local Lemma~\cite{Lovasz--Erdos1975}]
\label{lem:lll}
Let $A_1,\ldots,A_N$ be a set of ``bad'' events with $\P(A_j) \le q < 1$, where each event
$A_j$ is mutually independent of all but at most $L$ of the other events. If $4qL\le 1$,
then
\[
\P\left\{\cap_{j=1}^N A_j^c\right\}>0.
\]
\end{lemma}
\begin{IEEEproof}[Proof of Theorem~\ref{thm:advs-nkd-bound}]
We generate the phase detection sequence $X^n$ i.i.d.$\sim \mathrm{Bern}(1/2)$ and apply minimum distance detection. Let $\{A_j\}$ be the collection of events where the Hamming distance between a pair of codewords $\mathrm{wt}\bigl(\phi(m_1)\oplus \phi(m_2)\bigr) \le d$ where $m_1<m_2$. We have
\begin{align*}
\P(A_j) &\stackrel{(a)}{=} \P\{\mathrm{wt}(Z^k) \le d, Z^k \text{ i.i.d.}\sim\mathrm{Bern}(1/2)\}\\
&= \sum_{i = 0}^{d} \binom{k}{i}\frac{1}{2^k},
\end{align*}
where ($a$) follows since for any two distinct phases $m_1 \ne m_2$, the sum of the two codewords $\phi(m_1)\oplus \phi(m_2)$ is i.i.d.$\sim \mathrm{Bern}(1/2)$ even if they are overlapping subsequences of $X^n$. Now each $A_j$ is mutually independent of all other
events, except for a set of at most $4kn$ events. This is because the random variable $\phi(m_1)\oplus \phi(m_2)$ is mutually independent of all $X_i$'s with $i \in [n]\setminus \{m_1-k+1,m_1-k+2,\ldots,m_1+k-1\} \setminus \{m_2-k+1,m_2-k+2,\ldots,m_2+k-1\}$, which excludes at most $4kn$ events. Applying Lemma~\ref{lem:lll}, the phase detection sequence $X^n$
has minimum distance greater than $d$ with positive probability
\[
\P\left\{\cap_{j=1}^{\frac12 n(n-1)}A_j^c\right\}>0
\]
if
\begin{equation}
\label{eqn:nkd-2}
16kn\sum_{i = 0}^{d} \binom{k}{i}\frac{1}{2^k} \le 1,
\end{equation}
or equivalently the condition in~\eqref{eqn:advs-nkd-bound}. This completes the proof of Theorem~\ref{thm:advs-nkd-bound}.
\end{IEEEproof}
\begin{IEEEproof}[Proof of Corollary~\ref{thm:advs-limit}]
Set $d = 2pk$ in~\eqref{eqn:advs-nkd-bound}. Applying the Hamming ball volume approximation
\[
\sum_{i = 0}^{2pk} \binom{k}{i}\leq 2^{kh(2p)}
\]
and plugging $R = \frac{\log{n}}{k}$ in~\eqref{eqn:nkd-2}, we have
\[
R \le 1-h(2p) - \frac{\log (16k)}{k}.
\]
Letting $k\to \infty$, it follows that a rate $R$ is achievable if $R < 1-h(2p)$.
\end{IEEEproof}
\begin{remark}
In the standard channel coding setup, a random codebook attains the GV bound with high probability. In contrast, the probability of randomly drawing a good scheme for our setup is exponentially small. This is most obvious in the noiseless case ($p=0$), where it is well known that the fraction occupied by de Bruijn sequences among all sequences vanishes exponentially fast~\cite{deBruijn1946}.
\end{remark}
\subsection{Linear Phase Detection Schemes}
\label{sec:advs-code}
Theorem~\ref{thm:advs-nkd-bound} and Corollary~\ref{thm:advs-limit} showed the existence of a good adversarial phase detection scheme. Now, we discuss explicit constructions of such schemes. First, we ask whether phase detection schemes are ``equivalent'' to error-correcting codes in a certain sense. Clearly, any adversarial phase detection scheme induces a codebook that can be used as an error-correcting code for the corresponding adversarial channel coding problem. The converse direction seems more challenging. Given an error-correcting code, is it possible to ``chain up'' all or a sizable fraction of its codewords to create a sequence, and use the decoding rule as the detector? If so, what structure should such a code possess? In the following, we answer these questions for the class of \emph{linear} error-correcting codes.
First, we note that in order to induce any error-correcting code with minimum distance $d>1$, the phase detection sequence $x^n$ should not contain $0^k$ as a contiguous subsequence, for otherwise a shift by one from that position would create a codeword that is at distance $1$ from $0^k$. Following that, an $(n,k)$ phase detection scheme is said to be \emph{linear} if $\mathcal{C}\cup \{0^k\}$, namely its induced codebook together with the zero codeword, forms a linear code. Let $r$ be the dimension of this linear code. Then, the length of the linear phase detection sequence is $n = 2^r-1$.
\begin{theorem}
\label{thm:lfsr}
A phase detection scheme with $n = 2^r-1$ is linear if and only if
it is generated by a linear feedback shift register (LFSR) with a primitive characteristic
polynomial $a(z) = \sum_{i=0}^{r-1} a_i z^i + z^r$ over $GF(2)$, i.e.,
\begin{equation}
\label{eqn:lfsr}
x_{r+j} = \sum_{i=0}^{r-1}a_i x_{i+j}, \quad j \in [n].
\end{equation}
\end{theorem}
\begin{corollary}
\label{cor:linear}
The non-zero codewords of a linear code of dimension $r$ can be
chained up to a sequence of length $2^r-1$ if and only if any $r$ contiguous columns
of the generator matrix
$
G_{r\times k} = [{\bf g}_1,{\bf g}_2,\ldots,{\bf g}_k]
$
are linearly independent, and
\begin{equation}
\label{eqn:genMtx}
{\bf g}_{r+j} = \sum_{i=0}^{r-1}a_i {\bf g}_{i+j},\quad j\in[k-r],
\end{equation}
where $a_i$'s are the coefficients of a primitive polynomial $a(z) = \sum_{i=0}^{r-1}a_i
z^i + z^r$ over $GF(2)$.
\end{corollary}
\begin{IEEEproof}[Proof of Theorem~\ref{thm:lfsr}]
To prove sufficiency, suppose that $x^n$ is generated by an LFSR with a primitive characteristic polynomial in~\eqref{eqn:lfsr} and a nonzero initial state vector $(x_1,x_2,\dots,x_r)$. Then, every length-$r$ string except $0^r$
occurs exactly once in $x^n$ (see~\cite[Theorem 8.33]{Lidl--Niederreiter2008}). It follows that for any distinct codewords $\phi(m_1) =
c^k$ and $\phi(m_2) = d^k$, there exists $\phi(m_3) = e^k$ such that $c_j + d_j =
e_j$ for $j \in [r]$. For $r < j \le k$, $c_{j} + d_{j} = e_{j}$ follows since the
sequence is generated by an LFSR of degree $r$.
For necessity, let $x^n$ be a sequence associated with a linear phase detection scheme. We show that the first
$r$ columns ${\bf g}_1,\ldots,{\bf g}_r$ of the generator matrix $G_{r\times k}= [{\bf g}_1,{\bf g}_2,\ldots,{\bf g}_k]$ must be linearly independent.
Assuming that contrary, there exist $f_1,\ldots,f_r \in \{0,1\}$ not all zero such that
\begin{equation}
\label{eqn:rank}
\sum_{i=1}^r f_i {\bf g}_i = {\bf 0}.
\end{equation}
Let $[x_1,\ldots,x_k] = [u_1,\ldots,u_r]G_{r\times k}$. Multiplying
both sides of~\eqref{eqn:rank} by $[u_1,\ldots,u_r]$, we have $\sum_{i=1}^r f_i x_i = 0$. Applying this to
every codeword in $\mathcal{C}$, and recalling that the codewords are all contiguous subsequences of $x^n$, we have
\[
\sum_{i=1}^r f_i x_{i+j} = 0, \quad j\in[n].
\]
Let $i_0 = \max\{i\in[r]\colon f_i=1\}$. If $i_0 = 1$, then $x^n$ has to be $0^n$,
in contradiction. For $i_0 > 1$, we have
\[
x_{j+i_0} = \sum_{i=1}^{i_0-1} f_i x_{i+j},\quad j\in [n],
\]
which implies $x^n$ is generated by an LFSR of degree $i_0-1 < r$. But this
contradicts the fact that $x^n$ is of length $2^r-1$ and all codewords $\phi(m), m\in[n]$,
are distinct.
Now, since the first $r$ columns of $G_{r\times k}$ are linearly independent, there
exist $a_0, \ldots, a_{r-1}$ such that
\[
{\bf g}_{r+1} = \sum_{i=0}^{r-1} a_i {\bf g}_{i+1}.
\]
From this it follows that~\eqref{eqn:lfsr} holds and $x^n$ is generated by an LFSR.
Finally, an LFSR sequence is of maximum length if and only if the characteristic polynomial
is primitive.
\end{IEEEproof}
\begin{IEEEproof}[Proof of Corollary~\ref{cor:linear}]
The sufficiency follows since for a linear code, the relation~\eqref{eqn:genMtx} implies~\eqref{eqn:lfsr}. The necessity follows the same way as the necessity in Theorem~\ref{thm:lfsr}.
\end{IEEEproof}
\begin{remark}
As an application of Theorem~\ref{thm:lfsr}, we can design a card trick for adversarial crowds. Picking the primitive polynomial $a(z) = z^5 + z^4 + z^2 + z + 1$ and $k = 9$, we get a sequence of length $n =31$ and minimum distance $d =3$. Ordering cards according to this sequence, the magician can now correct one lie out of 9 contiguous color reads.
\end{remark}
\begin{remark}
When the characteristic polynomial of the LFSR is irreducible but not primitive, the sequence it generates has length $t$, which equals the order of the characteristic polynomial. Depending on the initial state $x^r$, the LFSR generates one out of $s=\frac{2^r-1}{t}$ \emph{disjoint} sequences $x^t{(1)}, \ldots, x^t{(s)}$. The length-$k$ contiguous subsequences of each sequence $\mathcal{C}^{(i)} = \{x_{m}^{m+k-1}(i)\colon m \in [t]\}$ together with the zero codeword form a linear code
$
\cup_{i=1}^s\mathcal{C}^{(i)} \cup \{0^k\}.
$
Conversely for a linear code, if the first $r$ columns of its generator matrix are linearly independent and~\eqref{eqn:genMtx} holds with $a_i$'s being the coefficients of an irreducible but not primitive polynomial of order $t$, then its nonzero codewords can be partitioned into $s$ equal size subsets, each of which can be chained up to a phase detection sequence.
\end{remark}
We now provide two results on the performance of linear phase detection schemes. In Theorem~\ref{thm:concentration}, we cite a known result from~\cite[Theorem 8.85]{Lidl--Niederreiter2008} on asymptotic relative distance, which improves upon~\cite[Theorem~1]{Kumar--Wei1992}. Then, inspired by a linear programing bound for LDPC codes~\cite{Ben-Haim--Litsyn2006}, we provide in Theorem~\ref{thm:ldpc} an upper bound on the sequence length of a linear phase detection scheme of a given minimum distance, using the linear programing method originated by Delsarte~\cite{Delsarte1973}.
\begin{theorem}[Theorem 8.85~\cite{Lidl--Niederreiter2008}]
\label{thm:concentration}
For every $(n,k)$ linear phase detection scheme, for every $m \in [n]$,
\[
\left|\emph{wt}(x_m^{m+k-1})-\frac{k}{2}\right| \le \sqrt{n} \left(\frac{\log{n}}{\pi}+1\right).
\]
In particular, for $(n,k)$ such that $\lim_{k\to \infty} \frac{\sqrt{n}\log{n}}{k}= 0$, the relative distance of the induced code converges to
\begin{equation}
\label{eqn:relative-d}
\lim_{k \to \infty} \frac{\;d\;}{k} = \frac 12.
\end{equation}
\end{theorem}
\begin{remark}
We note a similar result in~\cite[Theorem~1]{Kumar--Wei1992}, which claims~\eqref{eqn:relative-d} for every $0 < \mu \le 1$ and $k = \mu n$. Theorem~\ref{thm:concentration} improves upon~\cite{Kumar--Wei1992} by allowing $k$ to be sublinear in $n$.
\end{remark}
For the next result, we need the following definitions. For $t \in [k]$ and $z \in \mathbb{R}$, let
\[
K_t(z) = \sum_{j=0}^t (-1)^j \binom{z}{j} \binom{l-z}{t-j}
\]
be the Krawtchouk polynomial~\cite[Ch.~5. \S~2]{MacWilliams--Sloane1977a}~\cite{MacWilliams--Sloane1977b}, where the binomial coefficient for $z \in \mathbb{R}$ is defined as
$
\binom{z}{i} = \frac{z(z-1)(z-2)\cdots (z-i+1)}{i\, !}.
$
For large $k$, the exponent of $K_t(z)$ can be approximated as~\cite[Equation (40)]{Ben-Haim--Litsyn2006}
\[
\frac{1}{k}\log{K_{\lfloor pk\rfloor}(\lfloor\lambda k \rfloor)} = h(p) + \text{Int}(p,\lambda) + o(1),
\]
where
\begin{align}
&\text{Int}(p,\lambda) \nonumber\\
&= \int_0^{\lambda} \log\left(\frac{1-2p + \sqrt{(1-2p)^2 - 4(1-y)y}}{2(1-y)}\right)dy. \label{eqn:int}
\end{align}
\begin{theorem}
\label{thm:ldpc}
Every $(n,k)$ linear phase detection scheme with length $n =2^r-1$ and minimum distance $d = 2t + 1$ must satisfy
\[
2^r \cdot \frac{K_{t}^2(ic)\binom{(k-r)/c^2}{i}c^{2i}}{\binom{k}{t}} \le 2^k
\]
for every $i\in [k]$ such that $ic^2 \le k - r$. Here $c$ is the number of nonzero coefficients of the characteristic polynomial $a(z) = \sum_{j=0}^r a_j z^j$.
\end{theorem}
\begin{remark}
Compared to Delsarte's linear programing bound for channel codes~\cite{Delsarte1973},
the bound in Theorem~\ref{thm:ldpc} can sometimes be better. For example, when $r = 20,t = 5$, and $c = 3$, the linear programing bound yields $k \ge 41$, while Theorem~\ref{thm:ldpc} requires $k \ge 42$. We note, however, that with further optimization for these specific parameters, the best known channel coding upper bound is $k \ge 43$~\cite{Grassl2007}.
\end{remark}
\begin{remark}
For low-complexity LFSR implementation, it may be desirable to choose a characteristic polynomial with low coefficient weight. According to a conjecture in finite field theory~\cite{Mullen--Shparlinski1996, Golomb2007}, there are infinitely many primitive polynomials with coefficient weight $c = 3$. For this class of primitive polynomials, Theorem~\ref{thm:ldpc} implies that when the adversarial channel can flip at most a fraction $p$ of the inputs, the rate of the linear phase detection scheme must satisfy
\begin{align}
&\max_{0\le \mu \le \frac{1-R}{9}}\left\{2\mu \log{3} + h\left(\tfrac{9\mu}{1-R}\right) \tfrac{(1-R)}{9} +2\, \text{Int}(p,3\mu) \right\} \nonumber\\
&\le 1 - h(p) - R, \label{eqn:new-ub}
\end{align}
where $\text{Int}(p,3\mu)$ is given in~\eqref{eqn:int}.
This bound can sometimes be better than the second MRRW bound~\cite{McEliece--Rodemich--Rumsey--Welch1977}, which is the best known asymptotic upper bound for binary channel codes. For example, when $p = 0.05$, the second MRRW bound requires $R \le 0.6927$. However, $p=0.05$ and $R = 0.6927$ violate condition~\eqref{eqn:new-ub} when $\mu = 0.03073$.
\end{remark}
\begin{IEEEproof}[Proof of Theorem~\ref{thm:ldpc}]
Following the same line of reasoning as in Section II-C (29)--(36) and (48)--(49) of~\cite{Ben-Haim--Litsyn2006}, we have for every $\alpha\in [k]$,
\begin{equation}
\label{eqn:krawtchouk}
2^r \cdot \frac{K_{t}^2(\alpha) B_{\alpha}}{\binom{k}{t}} \le 2^k,
\end{equation}
where $B_{\alpha}$ is the number of codewords of weight $\alpha$ in the dual code of the linear code induced by the phase detection scheme. Now we show that when the coefficient weight of the characteristic polynomial is $c$, for every $i \in [k]$ such that $ic^2 \le k-r$, we can lower bound
\begin{equation}
\label{eqn:weight}
B_{ic} \ge \binom{(k-r)/c^2}{i} c^{2i}.
\end{equation}
To that end, note that our $(k-r)\times k$ parity check matrix, which is also the generator matrix of the dual code, can be written in the following form
$$
\left[
\begin{array}{ccccccccc}
1 & a_1 & \cdots & a_{r-1} & 1 & 0 & 0 &\cdots & 0\\
0 & 1 & a_1 & \cdots & a_{r-1} & 1 & 0 & \cdots & 0\\
\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots\\
0 & 0 &\cdots &0 & 1 & a_1 & \cdots & a_{r-1} & 1 \\
\end{array}
\right]_{.}
$$
A weight $ic$ codeword of the dual code could come from the sum of $i$ rows of $H$ whose nonzero elements (the $1$'s) are in disjoint columns. We lower bound the number of such codewords. First, we select an arbitrary row from the $(k-r)$ rows. Since each row of $H$ has weight $c$, the locations of the $1$'s in the chosen row overlap that of at most $c^2$ rows (including itself). Then a second row is chosen from the $(k-r-c^2)$ remaining non-overlapping rows. We continue in this manner until we obtain $i$ rows (we will not exhaust all rows provided that $ic^2 \le k-r$). Hence, the number of choices is lower bounded by
\begin{align*}
&\frac{1}{i\,!}(k-r)(k-r-c^2)\cdots(k-r-(i-1)c^2)\nonumber\\
&={(k-r)/c^2\choose i}c^{2i},
\end{align*}
which establishes~\eqref{eqn:weight}. Plugging~\eqref{eqn:weight} into~\eqref{eqn:krawtchouk} with $\alpha = ic$ completes the proof.
\end{IEEEproof}
\section{Point-to-Point: Probabilistic Noise,\\ Vanishing Error}
\label{sec:prob}
In this section we discuss the probabilistic noise model with a vanishing error criterion. We first show that the capacity in this case coincides with the Shannon capacity of the observation channel. We then proceed to describe a low-complexity coding construction, based on a concatenation of a channel code and a de Bruijn sequence, that approaches this fundamental limit.
\subsection{Fundamental Limit}
\label{sec:prob-limit}
\begin{theorem}
\label{thm:prob-limit}
The vanishing error capacity for probabilistic phase detection over a channel $p(y|x)$ is
\[
C_\text{ve} = \max_{p(x)}I(X;Y).
\]
\end{theorem}
Before we proceed to the proof, we need a technical lemma. We denote the typical set of length-$k$ vectors corresponding to $(X,Y)$ by
\begin{align*}
&\mathcal{T}_\epsilon^{(k)}(X,Y)\\
&:= \Big\{(x^k,y^k)\colon \left|\frac{\#\{i\colon (x_i,y_i) = (x,y)\}}{n} - p(x,y)\right|\\
&\hspace{2em}\le \epsilon p(x,y)\text{ for all } x \in \mathcal{X}, y \in \mathcal{Y}\Big\}.
\end{align*}
\begin{lemma}[Lemma 24.2~\cite{El-Gamal--Kim2011}]
\label{lem:iid}
Let $(X,Y)\sim p(x,y) \neq p(x)p(y)$ and $(X^n,Y^n)\sim \prod_{i=1}^n p_{X,Y}(x_i,y_i)$.
If $\epsilon > 0$ is sufficiently small, then there exists $\gamma (\epsilon)>0$ that depends only on
$p(x,y)$ such that
\begin{equation}
\label{eqn:overlap}
\P\{(X_{m}^{m+k-1},Y^k)\in \mathcal{T}_\epsilon^{(k)}(X,Y) \} \le 2^{-k \gamma(\epsilon)}
\end{equation}
for every $m>1$. Moreover, for non-overlapping sequences, i.e., for $k+1 \le m \le n-k
+1$, \begin{equation}
\label{eqn:indep}
\P\{(X_{m}^{m+k-1},Y^k)\in \mathcal{T}_\epsilon^{(k)}(X,Y)\} \le 2^{-k(I(X;Y)-\delta(\epsilon))},
\end{equation}
where $\delta(\epsilon)$ tends to zero as $\epsilon \to 0$.
\end{lemma}
\begin{IEEEproof}[Proof of Theorem~\ref{thm:prob-limit}]
Clearly, any phase detection sequence is also a channel code. Thus, the above rate cannot be exceeded. We proceed to prove the achievability. Recall $\phi(m) = x_{m}^{m+k-1}$.
\smallskip
{\it Phase detection sequence generation.} We generate the sequence $X^n$ i.i.d.$\sim
p(x)$.
\smallskip
{\it Detection.} Upon receiving $y^k$, the detector declares ${\hat{m}}$ is the phase estimate if
it is the unique phase such that $(\phi({\hat{m}}),y^k) \in \mathcal{T}_\epsilon^{(k)}(X,Y)$; otherwise---if there is
none or more than one---it declares an error.
\smallskip
{\it Analysis of the probability of error.} Without loss of generality, we assume the phase $M=1$. The detector makes
an error only if one or more of the following events occurs:
\begin{align*}
\mathcal{E}_1 &= \{(\phi(1),Y^k) \notin \mathcal{T}_\epsilon^{(k)}(X,Y)\},\\
\mathcal{E}_2 &= \{(\phi(m),Y^k) \in \mathcal{T}_\epsilon^{(k)}(X,Y) \text{ for some } m \ne 1\}.
\end{align*}
By the law of large numbers, $\P(\mathcal{E}_1)$ tends to zero as $k \to \infty$. For the
second term, we have
\begin{align*}
&\P(\mathcal{E}_2) \\
&\le \left(\sum_{m=2}^k + \sum_{m=n-k+2}^n\right)
\P\{(\phi(m),Y^k) \in \mathcal{T}_\epsilon^{(k)}(X,Y)\} \\
&\hspace{2em} + \sum_{m=k+1}^{n-k+1} \P\{(\phi(m),Y^k) \in \mathcal{T}_\epsilon^{(k)}(X,Y)\}\\
&\stackrel{(a)}{=} 2(k-1)2^{-k\gamma(\epsilon)} + (2^{kR}-2k+1)2^{-k(I(X;Y)-\delta(\epsilon))},
\end{align*}
which tends to zero as $k \to \infty$ if $R < I(X;Y)-\delta(\epsilon)$. Here the first and the
second terms in ($a$) follow from~\eqref{eqn:overlap}~and~\eqref{eqn:indep} respectively.
Letting $\epsilon \to 0$ completes the proof.
\end{IEEEproof}
\begin{example}\label{ex:GPS}
Consider the case of GPS signaling. For GPS, the binary (BPSK) symbol duration is about 1$\mu$sec, and the length of the underlying Gold code sequence is $n=1023$. Consider a typical observation time of $1$ second, which corresponds to a repetition factor $N\approx 1000$ and $k\approx 1e6$ binary observations. A correlator receiver can thus increase the SNR by about $60\textrm{dB}$ by coherently integrating over this sequence (assuming symbol timing has been recovered). Due to the good autocorrelation structure of the Gold code, an SNR of $30\textrm{dB}$ is typically sufficient in order to distinguish the correct phase (out of the $1023$ possibilities, and typically also over several Doppler hypotheses), with a small enough error probability. Namely, one can operate at an SNR of $-30\textrm{dB}$, and provide positioning with uncertainty of $1023\,\mu$sec; multiplied by the speed of light, this yields a positioning modulo $\approx 30,000$ km, which is sufficient as it is of the same order of the distance to the satellites.
Let us now show that one can significantly improve sensitivity using a more general phase detection sequence. Using the same observation period of $1$ second, let us assume a much lower SNR of $-44\textrm{dB}$. Using the Gaussian capacity formula and Theorem~\ref{thm:prob-limit}, we have that
\begin{align*}
\frac{\log{n}}{k} \approx \frac{1}{2}\log_2(1+SNR)
\end{align*}
can be asymptotically achieved. Using our $k=1e6$ and solving for $n$, we get that the largest $n$ that can be supported is $n\approx 4e8$. Since this large $n$ is also (much) larger than $k$, we can in principle design a phase detection sequence with roughly these parameters that attains a low error probability. This will reliably find our distance to the satellite with an uncertainty of about $120$ billion km, a huge overkill, but saves $14\textrm{dB}$ in the SNR relative to the competing GPS solution operating with the same observation time. To make the comparison more precise, one should look more carefully at many important details such as the exact error probability performance, the effect of multiple Doppler hypotheses, complexity of detection, and accounting for multiple satellites. Most of these issues are beyond the scope of this paper. In the next subsection and in Secion~\ref{sec:mac-advs} we discuss the issues of complexity and multiple sequences.
\end{example}
\subsection{A Low-Complexity Construction}
\label{sec:prob-code}
Now we present a sequence construction with low-complexity detection that
achieves the capacity asymptotically. The construction consists of three main ingredients:
\begin{enumerate}
\item a de Bruijn sequence with an efficient decoding algorithm~\cite{Tuliani2001},
\item a capacity achieving low-complexity code, e.g. a polar code~\cite{Arikan2009}, that protects the de Bruijn
sequence against noise, and
\item an i.i.d. synchronization sequence, which is known at the detector {\it a priori}, that
allows the detector to find the block boundary.
\end{enumerate}
The details are as follows.
\begin{figure*}[t!]
\centering
\def2.1{2.1}
\input{prob-code2.eps_tex}
\caption{Construction for probabilistic phase detection.}
\label{fig:prob-code}
\end{figure*}
\smallskip
{\it Phase detection sequence design.}
\iffalse
Let $a(z) = \sum_{i=0}^{r-1} a_i z^i +
z^{r}$ be a primitive polynomial over $GF(2)$. A sequence of length $2^{r}-1$
sequence is generated through the linear feedback shift register with connection function
$a(z)$, i.e., $$u_{j+r} = \sum_{i=0}^{r-1}a_i u_{j+i}, \;\;j =
1,2,\ldots,2^r-r-1,$$
with the initial $u^{r}$ being not all zeros (for example, $u_1 = 1, u_2 =
\cdots = u_{r} = 0$). This generates a maximal length circulant sequence with the
property that any $r$ contiguous bits uniquely determines its location in the sequence.
\fi
We design a de Bruijn sequence $u^{2^r}$ of order $r$ according the method in~\cite{Tuliani2001}. To encode it to a phase detection sequence $x^n$, we let $r = sl$, where $s$ and $l$ are integers. The de Bruijn sequence is chopped up into length-$s$ chunks, each of which is encoded into a length-$t$ codeword using a channel code of rate $R_\mathrm{ch} = s/t$. Then, a synchronization sequence $b^{3\tau}$ is generated i.i.d.$\sim p(x)$, where the parameter $\tau$ is a linear function of $t$, i.e., $\tau = c_1 t + c_2$ for some constants $c_1>0$ and $c_2$. Below we use $\tau = t$, but $\tau \neq t$ will prove useful later in Section~\ref{sec:mac-prob-code}. This sequence $b^{3\tau}$ is inserted every $l$ blocks. The middle chunk of the synchronization sequence $b_{\tau+1}^{2\tau}$ is given to the detector. The chunks $b^{\tau}$ and $b_{2\tau+1}^{3\tau}$ play the role of ``guarding bits'' between codewords and the middle chunk $b_{\tau+1}^{2\tau}$. Their purpose is to simplify the analysis of the error probability event associated with the synchronization detection (later denoted $\mathcal{E}_1$), as will become clear in the sequel. This is illustrated in Figure~\ref{fig:prob-code}.
\smallskip
{\it Detection.} We choose the length of the detection window to be
\begin{equation}
\label{eqn:dec-window}
k = lt+3\tau+\max\{t,\tau\}.
\end{equation}
The extra $\max\{t,\tau\}$ symbols are the margin to ensure there are $l$ complete channel code blocks and a complete synchronization sequence in the received sequence. Upon receiving $y^k$, the detector first finds an ${\hat{w}}_1 \in \{0\} \cup [k-\tau]$ such that $(b_{\tau+1}^{2\tau},y_{{\hat{w}}_1 +1}^{{\hat{w}}_1+\tau}) \in \mathcal{T}_\epsilon^{(\tau)}(X,Y)$. If there are more than one, it chooses the smallest index. It declares an error if there is none. This determines the block boundary of the channel code blocks, i.e., a complete block starts from index $({\hat{w}}_1-\tau \mod t) + 1$ in $y^k$. By design, there are at least $l$ complete channel code blocks in $y^k$ (the dashed-line parts in $y^k$ of Figure~\ref{fig:prob-code}). The detector then applies the channel decoder to recover $l$ blocks of messages. This corresponds to $ls = r$ contiguous bits in the de Bruijn sequence (the dashed-line parts in the $u$ sequence of Figure~\ref{fig:prob-code}), which uniquely determines the location of these bits $u_{s{\hat{w}}_2 + 1}^{s{\hat{w}}_2+r}$ via the de Bruijn decoder of~\cite{Tuliani2001}. The phase estimate is then declared as
\[
{\hat{m}} = {\hat{w}}_2 t + \left\lceil\frac{{\hat{w}}_2}{l}\right\rceil 3\tau + 1 - ({\hat{w}}_1 -\tau \mod
t).
\]
\smallskip
{\it Analysis of the probability of error.} For clarity of notation, we set $\tau = t$ in the following analysis. Similar analysis can be done for other linear functions of $t$.
Let $W_1$ be the actual index of the noisy version of $B_{\tau}$ in $Y^k$. The detector makes an error only if at least one of the following events occurs:
\begin{align*}
\mathcal{E}_1 &= \{W_1 \ne {\hat{W}}_1\},\\
\mathcal{E}_2 &= \{\text{an error in channel decoding}\}.
\end{align*}
Given $\mathcal{E}_1^c \cap \mathcal{E}_2^c$, the de Bruijn decoder can figure out the phase of the decoded $r$ bits with zero error. Since we are using a good channel code, we have $\P(\mathcal{E}_2 \cap \mathcal{E}_1^c)\to 0$ as $t\to\infty$. To bound $\P(\mathcal{E}_1)$, assume for convenience and without loss of generality that $W_1=t-1$. We have
{\allowdisplaybreaks
\begin{align*}
\P(\mathcal{E}_1)&= \P\{(B_{t+1}^{2t},Y_{w_1+1}^{w_1+t}) \in \mathcal{T}_\epsilon^{(t)} \text{ for some } w_1 \ne t-1\}\\
&\le \sum_{w_1 = 0}^{t-2} \P\{(B_{t+1}^{2t},Y_{w_1+1}^{w_1+t}) \in \mathcal{T}_\epsilon^{(t)}\} \\
&\hspace{2em} + \sum_{w_1 = t}^{2t-2} \P\{(B_{t+1}^{2t},Y_{w_1+1}^{w_1+t}) \in \mathcal{T}_\epsilon^{(t)}\}\\
&\hspace{2em} + \sum_{w_1 = 2t-1}^{k-t} \P\{(B_{t+1}^{2t},Y_{w_1+1}^{w_1+t})
\in \mathcal{T}_\epsilon^{(t)}\}\\
&\stackrel{(a)}{\le} (t-1) 2^{-t \gamma(\epsilon)} + (t-1)2^{-t \gamma(\epsilon)}\\
&\hspace{2em}+ ((l+1)t+2)2^{-t(I(X;Y)-\delta(\epsilon))},
\end{align*}}%
which, for fixed $l$, tends to zero as $t \to \infty$.
Here, the first term in ($a$) follows from Lemma~\ref{lem:iid} and the fact that $B_{t+1}^{2t}$
and its preceding guarding block $B^{t}$ are i.i.d.$\sim p(x)$. The second term follows since $B_{t+1}^{2t}$
and its succeeding guarding block $B_{2t+1}^{3t}$ are i.i.d.$\sim p(x)$. The third term follows by virtue of the packing lemma~\cite[Lemma~3.1]{El-Gamal--Kim2011}, since any length-$t$ chunk from two channel code blocks is independent of $B_{t+1}^{2t}$. Note the role of the ``guarding bits'' here is to make sure $Y_{w_1+1}^{w_1+t}$ never overlaps with both $B_{t+1}^{2t}$ and a codeword, as we cannot generally assume too much about the statistics of a specific codeword. Therefore, the probability of error averaged over all possible realizations of $B^{3t}$ tends to zero as $t \to \infty$. It follows that a good deterministic sequence $b^{3t}$ exists (in fact, most choices are good).
\smallskip
{\it Rate.}
By design, the rate of the sequence is
\begin{align}
R &= \frac{\log n}{k}\nonumber\\
&= \frac{\log\left[2^{r}\frac{lt+3\tau}{ls}\right]}{lt+3\tau+\max\{t,\tau\}} \label{eqn:p2p-rate}\\
&\stackrel{(a)}{=} \frac{\log\left[2^{r}\frac{(l+3)t}{ls}\right]}{(l+4)t} \nonumber\\
&= R_\mathrm{code}\left(1-\frac{4}{l+4}\right) \nonumber\\
&+ \frac{R_\mathrm{code}}{s(l+4)}\log\left[\frac{1}{R_\mathrm{code}}\left(1+\frac
3l\right)\right], \nonumber
\end{align}
which, for fixed $R_\mathrm{code}$ and $l$, tends to $R_\mathrm{code}\left(1-\frac{4}{l+4}\right)$ as $s \to \infty$. Choosing a large $l$ and a capacity achieving code for the underlying channel $p(y|x)$ ensures the rate of the
phase detection sequence can be as close to $C_\mathrm{prob}$ as desired. Note that in step ($a$), we set $\tau = t$. But one can verify that the rate approaches capacity $C_\mathrm{prob}$ for other choices of $\tau$.
\smallskip
{\it Complexity.} Finding the block boundary is $O(k)$ complexity. Recalling that $r$ is linear in $k$ and using the method
of~\cite{Tuliani2001}, decoding the de Bruijn sequence is $O(k\log k)$ complexity. There exist capacity achieving channel codes
with $O(k \log k)$ decoding complexity, e.g., polar codes~\cite{Arikan2009}.
Therefore, the overall detection complexity is $O(k \log k)$.
\begin{remark
\label{rmk:p2p-code}
For future reference, we refer to the above construction an \emph{$(R_\text{ch},l,t,\tau)$ point-to-point phase detection sequence}. Once these four parameters are given, $s = tR_\text{ch}, r = ls$, and both $k$ and $R$ can be expressed as in~\eqref{eqn:dec-window} and~\eqref{eqn:p2p-rate}. As shown above, an $(R_\text{ch},l,t,\tau)$ point-to-point phase detection sequence has detection complexity $O(k \log{k})$. Moreover, for $\tau = c_1t + c_2$ with some constants $c_1 > 0$ and $c_2$, the achievable rate of the sequence satisfies
\[
\lim_{l\to \infty}\lim_{t \to \infty} R(R_\text{ch},l,t,\tau) = R_\text{ch}.
\]
This construction will also prove useful in Section~\ref{sec:mac-prob-code}.
\end{remark}
\begin{remark}
\label{rmk:eqvt-avg}
It appears plausible that the synchronization sequence could be discarded, and that the codeword boundary could be determined as part of the detection process. This coding scheme, in a sense, shows the equivalence between error-correcting codes and phase detection schemes for the probabilistic setting.
\end{remark}
\begin{remark}
\label{rmk:eqvt-max}
Our analysis for the point-to-point phase detection problem in the probabilistic noise model assumed a uniformly distributed phase, which in channel coding terms corresponds to an average error probability criterion. In channel coding, the capacity under a more stringent maximal error probability criterion remains the same; this is easily shown by throwing away the worse half of a good average error probability codebook. In the sequence phase detection problem however, it is not immediately clear whether the capacity remains the same, as throwing bad codewords can significantly shorten the sequence. However, using our specific construction above and using a maximal error capacity achieving channel code (which may increase the detection complexity), we can show that the resulting phase detection sequence is capacity achieving under maximal error probability criterion.
\end{remark}
\section{Point-to-Point: Probabilistic Noise,\\ Zero Error}
\label{sec:prob-zero}
In this section, we consider zero error phase detection. Let $\alpha(G)$ denote the \emph{independence number} of a graph $G$, i.e., the cardinality of a maximum independent set of $G$.
Then, the \emph{Shannon capacity of a graph} $G$ can be defined as~\cite{Shannon1956}
\[
C(G) \triangleq \sup_{k} \frac{\log{\alpha(G^k)}}{k} = \lim_{k \to \infty} \frac{\log{\alpha(G^k)}}{k},
\]
where $G^k$ is the $k$-fold strong product of $G$ (see definition in Section~\ref{sec:intro-p2p}). It is well known that $C(G)$ is the zero error capacity of any channel $p(y|x)$ with confusion graph $G$. An explicit expression for $C(G)$ is unknown. Nevertheless, the following theorem shows that $C(G)$ is also the fundamental limit in the zero error phase detection setting.
\begin{theorem}
\label{thm:prob-zero-limit}
The zero error capacity for phase detection in a channel with confusion graph $G$ coincides with the Shannon capacity of this graph, i.e.,
\[
C_\text{ze}(G) = C(G).
\]
\end{theorem}
\begin{IEEEproof}
Again, the induced codebook of every phase detection scheme is also a good channel code for the same confusion graph, and thus $C_\text{ze}(G) \le C(G)$. For the other direction, we show that every channel code of rate $R$ can be used to construct a phase detection scheme with the same rate in the asymptotic limit.
To this end, we first note that the rate $\log{\alpha(G)}$ can be readily achieved. This can be done by employing a one-shot zero error channel code of the same rate (which exists by definition), and using it to construct a de Bruijn sequence of alphabet size $\alpha(G)$ and order $k$ (cf. the existence of de Buijn sequences of any alphabet size and any order~\cite{Martin1934}). When $C(G) > \log{\alpha(G)}$, which means that the graph capacity can be achieved only by block coding over the product graph, then concatenating a length-$k$ zero error channel code according to a de Bruijn sequence of alphabet size $\alpha(G^k)$ (a naive extension of the one-shot approach above) does not immediately work (see also Remark \ref{rmk:last-one}). This is because the phase detector may not always know where a complete codeword starts or ends, which may result in detection errors. In what follows, we design a novel zero error synchronization sequence that enables the detector to determine the block boundary without error, and with vanishing loss in rate.
{\it Augmented codebook and synchronization sequence.} For any $G$ with $C(G)>0$, there exist two distinct vertices $\beta,\gamma \in \mathcal{X}$ such that $(\beta,\gamma) \notin E$. Let $\mathcal{C}^{(t)}$ be a zero error channel code of length $t$ and rate $R_\text{ch} = \frac{1}{t}\log{|\mathcal{C}^{(t)}|}$. We create an augmented codebook by sandwiching each codeword between two guarding $\gamma$'s, i.e.,
\[
\tilde{\mathcal{C}}^{(t+2)} = \{(\gamma,c^t,\gamma)\colon c^t \in \mathcal{C}^{(t)}\}.
\]
The sequence
\[
\beta^{t+2} = (\underbrace{\beta,\ldots,\beta}_{t+2})
\]
will be used as the synchronization sequence.
{\it Phase detection sequence design.} We take a de Bruijn sequence with alphabet size $|\mathcal{C}^{(t)}|$ and order $r$. We associate each symbol in the de Bruijn alphabet with a different codeword in the augmented codebook $\tilde{\mathcal{C}}^{(t+2)}$ (note that $|\tilde{\mathcal{C}}^{(t+2)}| = |\mathcal{C}^{(t)}|$ by design). Then, similar to the concatenated structure in Figure~\ref{fig:prob-code}, we concatenate (in a sequential manner) the augmented codewords according to the de Bruijn sequence. Between every $r$ consecutive blocks of augmented codewords, we insert a synchronization block $\beta^{t+2}$. This way, the de Bruijn sequence of length $|\mathcal{C}^{(t)}|^r$ is mapped to a phase detection sequence $x^n$ of length $n = \frac{(r+1)(t+2)}{r}|\mathcal{C}^{(t)}|^r$.
{\it Detection.} We choose the length of the detection window to be
\[
k = (r+2)(t+2).
\]
This ensures that the window will contain $r$ complete codeword blocks and one complete synchronization block. For each $w_1 \in\{0\} \cup [k-t-2]$, define
\[
\mathcal{S}_y(w_1) \triangleq \left\{u^{t+2} \in \mathcal{X}^{t+2}\colon \prod_{i=1}^{t+2}p_{Y|X}(y_{w_1+i}|u_i) >0\right\}
\]
as the set of input sequences that may result in the output sequence $y_{w_1+1}^{w_1+t+2}$. The detector finds a ${\hat{w}}_1 \in \{0\} \cup [k-t-2]$ such that $\beta^{t+2} \in \mathcal{S}_y(w_1)$. If there are more than one, it chooses the smallest index. It declares an error if there is none. If ${\hat{w}}_1$ is found, then the first complete block starts from index $({\hat{w}}_1 \mod t+2)+1$ of $y^k$. Knowing the block boundary, the detector can then decode the $r$ codewords from $\mathcal{C}^{(t)}$. This corresponds to $r$ contiguous symbols in the de Bruijn sequence, which uniquely determine the starting position ${\hat{w}}_2+1$ in the de Bruijn sequence. Then, the phase estimate is declared as
\[
{\hat{m}} = \left({\hat{w}}_2 + \left\lceil \frac{{\hat{w}}_2}{r}\right\rceil\right)(t+2) + 1 - ({\hat{w}}_1 \mod t+2).
\]
{\it Error Analysis.} The crucial part of the error analysis is to show the synchronization sequence $\beta^{t+2}$ can be detected with zero error. Once the block boundary is found, the $r$ codewords can be decoded with zero error, and the location of the corresponding $r$ symbols in the de Bruijn sequence can also be found with zero error.
To see the scheme ensures zero error detection of $\beta^{t+2}$, we show that for all $m \in [n]$ such that $x_{m}^{m+t+1} \ne \beta^{t+2}$, $(x_{m}^{m+t+1},\beta^{t+2}) \notin E_{t+2}$, where $E_{t+2}$ is the edge set of $G^{t+2}$. There are two cases. When $x_{m}^{m+t+1}$ is a complete block, we have $x_m = x_{m+t+1} = \gamma$ and hence $(x_{m}^{m+t+1}, \beta^{t+2}) \notin E_{t+2}$ since $(\beta,\gamma) \notin E$. When $x_{m}^{m+t+1}$ consists of two (partial) blocks, we know at least one block must be a codeword block. Thus, considering the last symbol of the first block and the first symbol of the second block, we know at least one is $\gamma$. This implies $(x_{m}^{m+t+1},\beta^{t+2}) \notin E_{t+2}$. In summary, $\beta^{t+2}$ is never confusable with other $x_{m}^{m+t+1}$ at the detector.
{\it Rate.} By design, the rate of the scheme is
\begin{align*}
R &= \frac{\log\left(\frac{(r+1)(t+2)}{r}|\mathcal{C}^{(t)}|^r\right)}{(r+2)(t+2)}\\
&= \frac{r\log |\mathcal{C}^{(t)}|}{(r+2)(t+2)} + \frac{\log\left(\frac{(r+1)(t+2)}{r}\right)}{(r+2)(t+2)}\\
&= \frac{r\,t}{(r+2)(t+2)} R_\text{ch} + \frac{\log\left(\frac{(r+1)(t+2)}{r}\right)}{(r+2)(t+2)},
\end{align*}
which tends to $R_\text{ch}$ as $r \to \infty$ and $t \to \infty$. Therefore, by choosing a zero error capacity achieving channel code, the phase detection scheme achieves $C(G)$.
\end{IEEEproof}
\begin{remark}\label{rmk:last-one}
Suppose $C(G)$ is achieved by a finite block code of length $s>1$ (e.g., for the pentagon graph \cite{Lovasz1979}). In this case, generating a phase detection sequence using a de Bruijn sequence with codewords of the capacity achieving code as symbols, cannot work. To see this, recall that the induced codebook associated with any zero error phase detection sequence forms a zero error channel code of the same rate. However, a simple calculation shows that this rate is equal to $\frac{\log{n}}{\log{(n\slash s)}}\cdot C(G)$, which exceeds the capacity $C(G)$ for $s>1$.
\end{remark}
\section{Multiple Access: Probabilistic Noise,\\ Vanishing Error}
\label{sec:mac-prob}
So far in all the models we have discussed, phase detection either achieves the best known achievable rate of its channel coding counterpart, or shares the same capacity as that of channel coding. In this section, we encounter the first model, the multiple access phase detection with vanishing error, whose capacity region is strictly included in that of its channel coding counterpart.
\subsection{Fundamental Limit}
\label{sec:mac-prob-limit}
\begin{theorem}
\label{thm:prob}
The vanishing error capacity region $\mathscr{C}_\text{ve}$ for phase detection over the channel $p(y|x_1,x_2)$ is the set of all rate pairs $(R_1,R_2)$ such that
\begin{equation}\label{eqn:mac-ve-limit}
\begin{split}
R_1 &\le I(X_1;Y|X_2),\\
R_2 &\le I(X_2;Y|X_1),\\
R_1+R_2 &\le I(X_1,X_2;Y)
\end{split}%
\end{equation}
for some $p(x_1)p(x_2)$.
\end{theorem}
\begin{remark}
\label{rmk:gap}
We note that this region is not convex in general. Compared to the usual MAC capacity region, which is the convex hull of $\mathscr{C}_\text{ve}$,
this region can be a strict subset (see, for example, the \emph{push-to-talk} MAC with binary inputs and output, given by $p(0|0,0) = p(1|0,1) = p(1|1,0) = 1$ and $p(0|1,1) = 1/2$~\cite[Problem 3.2.6]{Csiszar--Korner1981}).
\end{remark}
\begin{IEEEproof}[Proof of Theorem~\ref{thm:prob}] We prove the achievability through random sequence generation and joint typicality detection.
\smallskip
{\it Sequence generation.} Fix a pmf $p(x_1)p(x_2)$. Let $n_1 = 2^{kR_1}$ and $n_2 = 2^{kR_2}$. We generate the two sequences $X_1^{n_1}$ i.i.d.$~\sim p(x_1)$ and $X_2^{n_2}$ i.i.d.$~\sim p(x_2)$.
\smallskip
{\it Detection.} Upon receiving $y^{k}$, the detector declares the phase estimate $({\hat{m}}_1, {\hat{m}}_2) \in [n_1]\times [n_2]$ if it is the unique pair such that $(\phi_1({\hat{m}}_1),\phi_2({\hat{m}}_2),y^k) \in \mathcal{T}_\epsilon^{(k)}(X_1,X_2,Y)$; if there is none or more than one, it declares an error.
\smallskip
{\it Analysis of the probability of error.}
Without loss of generality, we assume that the correct phase pair is $(M_1,M_2) = (1,M_2)$. the detector makes an error only if one or more of the following events occur:
\begin{align*}
\mathcal{E}_1 &= \{(\phi_1(1),\phi_2(M_2),Y^k) \notin \mathcal{T}_\epsilon^{(k)}\},\\
\mathcal{E}_2 &= \{(\phi_1(m_1),\phi_2(M_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \text{ for some } m_1 \neq 1\},\\
\mathcal{E}_3 &= \{(\phi_1(1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \text{ for some } m_2 \neq M_2\},\\
\mathcal{E}_4 &= \{(\phi_1(m_1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \\
&\hspace{8em}\text{ for some } m_1 \neq 1 \text{ and } m_2 \neq M_2\}.
\end{align*}
By the law of large number, $\P(\mathcal{E}_1)$ tends to zero as $k \to \infty$. For $\mathcal{E}_2$, we have
\begin{align*}
&\P(\mathcal{E}_2) \le \sum_{m_1 = k+1}^{n_1-k+1} \P\{(\phi_1(m_1),\phi_2(M_2),Y^k) \in \mathcal{T}_\epsilon^{(k)}\}\\
& + \sum_{m_1 = 2}^{k} \P\{(\phi_1(m_1),\phi_2(M_2),Y^k) \in \mathcal{T}_\epsilon^{(k)}\}\\
& + \sum_{m_1 = n_1-k+2}^{n_1} \P\{(\phi_1(m_1),\phi_2(M_2),Y^k) \in \mathcal{T}_\epsilon^{(k)}\}\\
&\stackrel{(a)}{\le} (2^{kR_1}-2k+1)2^{-kI(X_1;Y|X_2)} + 2(k-1)2^{-k\gamma_1(\epsilon)},
\end{align*}
which tends to zero as $k \to \infty$ if $R_1 < I(X_1;Y|X_2) -\delta(\epsilon)$.
Here the first term in ($a$) follows since $\phi_1(m_1)$ in that range does not overlap with the right chunk $\phi_1(1)$ and the probability can be bounded by the packing lemma~\cite[Lemma~3.1]{El-Gamal--Kim2011}. The second term in ($a$) corresponds to the overlapping chunks and the probability is bounded by Lemma~\ref{lem:iid} with $X_{m}^{m+k-1} \leftarrow \phi_1(m_1), Y^k \leftarrow (\phi_2(M_2),Y^k)$.
We can similarly show that $\P(\mathcal{E}_3)$ tends to zero as $k \to \infty$ if $R_2 < I(X_2;Y|X_1) - \delta(\epsilon)$. For $\mathcal{E}_4$, there are four different cases:
\begin{itemize}
\item There are $(2^{kR_1}-2k+1)(2^{kR_2}-2k+1)$ pairs $(m_1,m_2)$ such that neither $\phi_1(m_1)$ nor $\phi_2(m_2)$ overlaps with the right chunks. By the packing lemma, we can bound
\begin{align*}
&\P\{(\phi_1(m_1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \}\le 2^{-kI(X_1,X_2;Y)}.
\end{align*}
\item There are $2(k-1)(2^{kR_2}-2k+1)$ pairs $(m_1,m_2)$ such that $\phi_1(m_1)$ overlaps with the right chunk while $\phi_2(m_2)$ does not. Applying Lemma~\ref{lem:iid} first and then the packing lemma (note the independence between $\phi_1(m_1)$ and $\phi_2(m_2)$ for any $m_1$ and $m_2$), we have
\begin{align*}
&\P\{(\phi_1(m_1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \}\\
&=\P\{(\phi_1(m_1),Y^k )\in \mathcal{T}_\epsilon^{(k)}\}\\
&\hspace{.6em}\cdot\P\{(\phi_1(m_1),\phi_2(m_2), Y^k) \in \mathcal{T}_\epsilon^{(k)}\big| \\
&\hspace{13em}(\phi_1(m_1),Y^k)\in \mathcal{T}_\epsilon^{(k)}\}\\
&\le 2^{-k(\gamma_2(\epsilon)+I(X_2;Y|X_1))}.
\end{align*}
\item There are $2(k-1)(2^{kR_1}-2k+1)$ pairs $(m_1,m_2)$ such that $\phi_2(m_2)$ overlaps with the right chunk while $\phi_1(m_1)$ does not. Similarly, we have
\begin{align*}
&\P\{(\phi_1(m_1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \}\\
&\le 2^{-k(\gamma_3(\epsilon)+I(X_1;Y|X_2))}.
\end{align*}
\item The rest $4(k-1)^2$ pairs are such that both $\phi_1(m_1)$ and $\phi_2(m_2)$ overlap with the right chunks. We note that the event $\{(\phi_1(m_1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)}(X_1,X_2,Y)\}$ implies $\{(\phi_1(m_1),Y^k) \in \mathcal{T}_\epsilon^{(k)}(X_1;Y)\}$. Thus, we can bound the error as
\begin{align*}
&\P\{(\phi_1(m_1),\phi_2(m_2),Y^k) \in \mathcal{T}_\epsilon^{(k)} \}\\
&\le \P\{(\phi_1(m_1),Y^k) \in \mathcal{T}_\epsilon^{(k)} \}\\
&\le 2^{-k\gamma_2(\epsilon)}.
\end{align*}
\end{itemize}
Combining all four cases, we have $\P(\mathcal{E}_4)$ tends to zeros as $k \to \infty$ if $R_1+R_2 < I(X_1,X_2;Y) - \delta(\epsilon)$, $R_1 < I(X_1;Y|X_2) -\delta(\epsilon)$, and $R_2 < I(X_2;Y|X_1)-\delta(\epsilon)$. Letting $\epsilon \to 0$ completes the proof of the achievability.
\smallskip
For the converse, we wish to show for any $(2^{kR_1},2^{kR_2},k)$ multiple access phase detection scheme with vanishing probability of error $\lim_{k \to \infty}P_e^{(k)} = 0$, the rate pair $(R_1,R_2) \in \mathscr{C}_\text{ve}$. Given the two sequences $x_1^{n_1}$ and $x_2^{n_2}$, the joint distribution of $(M_1,M_2,Y^k)$ is
\[
\frac{1}{2^{k(R_1+R_2)}}\prod_{i=0}^{k-1} p_{Y|X_1,X_2}(y_{1+i}|x_{1,m_1+i},x_{2,m_2+i}).
\]
By Fano's inequality, we have $H(M_1,M_2|Y^k)\le k(R_1+R_2)P_e^{(k)} + 1 \le k\epsilon_k$,
where $\epsilon_k$ tends to zero as $k \to \infty$. We bound the sum rate as follows
\begin{align*}
&k(R_1+R_2)\\
&= H(M_1,M_2)\\
&\stackrel{(a)}{\le} I(M_1,M_2;Y^k) + k\epsilon_k\\
&= \sum_{i=0}^{k-1} I(M_1,M_2;Y_{1+i}|Y^i) + k\epsilon_k\\
&\stackrel{(b)}{\le} \sum_{i=0}^{k-1} I(M_1, M_2,Y^i, x_{1,M_1+i},x_{2,M_2+i};Y_{1+i}) + k\epsilon_k\\
&\stackrel{(c)}{=} \sum_{i=0}^{k-1} I(x_{1,M_1+i},x_{2,M_2+i};Y_{1+i}) + k\epsilon_k,
\end{align*}
where ($a$) follows from Fano's inequality and ($c$) follows since $(M_1,M_2,Y^{i}) \to (x_{1,M_1+i},x_{2,M_2+i}) \to Y_{1+i}$ form a Markov chain due to the memorylessness of the channel. Note here in both ($b$) and ($c$), $x_{j,M_j+i}$ is a function of $M_j$ that takes value $x_{j,m_j+i}$ when $M_j = m_j$ for $j = 1,2$. Now we bound the individual rate as follows
\begin{align*}
kR_1 &= H(M_1|M_2)\\
&\stackrel{(c)}{\le} I(M_1;Y^k|M_2) + k\epsilon_k\\
&= \sum_{i=0}^{k-1} I(M_1;Y_{1+i}|M_2,Y^i) + k\epsilon_k\\
&\le \sum_{i=0}^{k-1} I(Y^i,M_1,M_2,x_{1,M_1+i};Y_{1+i}|x_{2,M_2+i}) + k\epsilon_k\\
&= \sum_{i=0}^{k-1} I(x_{1,M_1+i};Y_{1+i}|x_{2,M_2+i}) +k\epsilon_k,
\end{align*}
where ($c$) follows since $H(M_1|Y^k,M_2) \le H(M_1,M_2|Y^k) \le k\epsilon_k$. Now flipping the role of $1$ and $2$, we have
\[
kR_2 \le \sum_{i=0}^{k-1} I(x_{2,M_2+i};Y_{1+i}|x_{1,M_1+i}) +k\epsilon_k.
\]
Now we introduce a time-sharing random variable $Q \sim \mathrm{Unif}[k]$, which is independent of $(M_1,M_2,Y^k)$. We can write
\begin{align*}
R_1+R_2 &\le I(x_{1,M_1+Q},x_{2,M_2+Q};Y_{1+Q}|Q) + \epsilon_k,\\
R_1 &\le I(x_{1,M_1+Q};Y_{1+Q}|x_{2,M_2+Q},Q) + \epsilon_k,\\
R_2 &\le I(x_{2,M_2+Q};Y_{1+Q}|x_{1,M_1+Q},Q) + \epsilon_k.
\end{align*}
Note that $\P\{Y_{1+Q}=y|x_{1,M_1+Q} = x_1,x_{2,M_2+Q} = x_2\} = p(y|x_1,x_2)$, which is distributed according to the channel conditional pmf. Hence, we identify $X_1 = x_{1,M_1+Q}, X_2 = x_{2,M_2+Q}$, and $Y = Y_{1+Q}$ to obtain
\begin{align*}
R_1+R_2 &\le I(X_1,X_2;Y|Q) + \epsilon_k\\
&\le I(Q,X_1,X_2;Y) + \epsilon_k\\
&\stackrel{(d)}{=} I(X_1,X_2;Y) + \epsilon_k,
\end{align*}
where $(d)$ follows since $Q \to (X_1,X_2) \to Y$ form a Markov chain. We similarly obtain
\begin{align*}
R_1 &\le I(X_1;Y|X_2) + \epsilon_k,\\
R_2 &\le I(X_2;Y|X_1) + \epsilon_k.
\end{align*}
Note that since $M_1$ and $M_2$ are independent and uniform over $[n_1]$ and $[n_2]$, $M_1 + Q$ and $M_2+Q$ are independent, and so are $x_{1,M_1+Q}$ and $x_{2,M_2+Q}$. Therefore, we can restrict the inputs to independent distribution $p(x_1)p(x_2)$.
Letting $k \to \infty$ completes the proof of the converse.
\end{IEEEproof}
\begin{remark}
We note the connection between the above converse proof, and that of the totally asynchronous MAC~\cite{Poltyrev1983, Hui--Humblet1985}. Unlike channel coding in the usual (synchronous) MAC setting, where the two inputs can be correlated through the time-sharing random variable $Q$, the two inputs $x_{1,M_1+Q}$ and $x_{2,M_2+Q}$ in the phase detection setting are independent even with the time-sharing random variable. Therefore, while the input pmf for the channel coding problem is $p(q)p(x_1|q)p(x_2|q)$, it is $p(x_1)p(x_2)$ in the phase detection setting. This essentially results in the strict gap between the capacity regions of the two problems (see also Remark~\ref{rmk:gap}).
\end{remark}
\begin{remark}
One can similarly show that the vanishing error capacity region for phase detection in the $L$-user MAC $p(y|x_1,\ldots,x_L)$ is the set of rate tuples $(R_1,\ldots,R_L)$ such that
\begin{equation}
\label{eqn:lmac-prob}
\sum_{i \in \mathcal{J}} R_i \le I(X_\mathcal{J};Y|X_{\mathcal{J}^c}) \quad \text{for every } \mathcal{J} \subseteq [L]
\end{equation}
for some $\prod_{i=1}^L p(x_i)$. Here $X_\mathcal{J} = \{X_i\colon i\in\mathcal{J}\}$.
\end{remark}
\subsection{A Low-Complexity Construction}
\label{sec:mac-prob-code}
In this section, we build on the point-to-point phase detection sequence construction in Section~\ref{sec:prob-code} and provide an $O(k\log{k})$ complexity sequence construction that achieves any rate pairs $(R_1,R_2) \in \mathscr{C}_\text{ve}$. The construction consists of several ingredients:
\begin{enumerate}
\item the vanishing error capacity achieving phase detection sequence for a point-to-point channel $p(y|x)$, as given in Section~\ref{sec:prob-code},
\item the rate-splitting~\cite{Grant--Rimoldi--Urbanke--Whiting2001} technique, which is a point-to-point channel coding technique for achieving arbitrary rate pairs in the MAC region without time sharing, and
\item a novel symbol-by-symbol mapping that enables rate-splitting in the phase detection setting.
\end{enumerate}
Details are as follows.
\smallskip
{\it Rate splitting.} In the random coding scheme, we simultaneously detect the phases $m_1$ and $m_2$ by checking typicality of all possible pairs via brute force. In practice, it is unclear whether simultaneous detection can be implemented at low complexity. In our design, we circumvent this difficulty by employing the rate splitting technique of~\cite{Grant--Rimoldi--Urbanke--Whiting2001}, which transforms the MAC coding problem into three point-to-point channel coding problems. Fix a pmf $p(u)p(v)p(x_2)$ and a function $x_1(u,v)$. We target the rate pair
\begin{align}
\label{eqn:split-rate}
\begin{split}
R_1 &= I(U;Y) + I(V;Y|X_2,U),\\
R_2 &= I(X_2;Y|U).
\end{split}
\end{align}
It is known~\cite{Grant--Rimoldi--Urbanke--Whiting2001} that for any rate point $(I_1,I_2) \in \mathscr{C}_\text{ve}$, there exists a pmf $p(u)p(v)p(x_2)$ and a function $x_1(u,v)$ such that $I_1 = R_1$ and $I_2 = R_2$.
\smallskip
{\it Sequence construction.} We design three vanishing error phase detection sequences for three point-to-point channels $U \to Y$, $X_2 \to (Y,U)$, and $V \to (Y,U,X_2)$ respectively, each according to the construction in Section~\ref{sec:prob-code}. Specifically, $u^{n_u}$, $x_2^{n_2}$, and $v^{n_v}$ are $(I(U;Y),l,t,t_u)$, $(I(X_2;Y|U),l,t,t_2)$, and $(I(V;Y|U,X_2),l,t,t_v)$ point-to-point phase detection sequences, respectively (see Remark~\ref{rmk:p2p-code} for the definition of an $(R_\text{ch},l,t,\tau)$ point-to-point phase detection sequence).
Given $u^{n_u}$ and $v^{n_v}$, we form an $x_1$ sequence of length $n_u n_v$ through the symbol-by-symbol mapping
\begin{equation}
\label{eqn:mapping}
x_{1,m_1} = x_1(u_{m_u},v_{m_v}) \text{ for } m_1 \in [n_un_v],
\end{equation}
where
\begin{align*}
m_u &= m_1 \pmod{n_u},\\
m_v &= m_1\pmod{n_v}.
\end{align*}
Note that when $n_u$ and $n_v$ are relatively prime, each phase $m_1 \in [n_un_v]$ corresponds to a \emph{distinct} phase pair $(m_u,m_v)$. Moreover, the way the $u$ and the $v$ sequences are ordered ensures that any length-$k$ chunk of the $x_1$ sequence is formed from a length-$k$ chunk of the $u$ sequence and a length-$k$ chunk of the $v$ sequence. Such an $x_1^{n_un_v}$ sequence simulates the channel output when the two phase detection sequences $u^{n_u}$ and $v^{n_v}$ go through a deterministic MAC $x_1(u,v)$. Finally, recall that any $t_u$ (and $t_v$) that is a linear function of $t$ results in the same asymptotic rate of the phase detection sequence (cf. Section~\ref{sec:prob-code}). Hence, by adjusting the parameters $t_u$ and $t_v$, it is always possible to make $n_u$ and $n_v$ relatively prime.
\smallskip
{\it Detection.} The way the $u,x_2,v$ sequences are designed allows multiple access phase detection through successive point-to-point phase detection in the channels $U \to Y$, $X_2 \to (Y,U)$, and $V \to (Y,U,X_2)$. We choose the length of the detection window to be
\[
k = lt + 3\max\{t_u,t_v,t_2\} + \max\{t,t_u,t_v,t_2\}
\]
and successively detect the phases in the order ${\hat{m}}_u \to {\hat{m}}_{2} \to {\hat{m}}_v$. The phase of the $x_1$ sequence is declared to be the unique ${\hat{m}}_1 \in [n_un_v]$ such that ${\hat{m}}_1 \pmod{n_u} = {\hat{m}}_u$ and ${\hat{m}}_1 \pmod{n_v} = {\hat{m}}_v$.
\smallskip
{\it Analysis of the probability of error.} By the analysis in the point-to-point case, the probability of error for detecting each sequence $\P(\mathcal{E}_j)$, $j = 1,2,3$, tends to zero as $t \to \infty$. By successive cancellation, the total probability of error $\P(\mathcal{E}) \le \P(\mathcal{E}_1) + \P(\mathcal{E}_2) + \P(\mathcal{E}_3)$, which tends to zero as $t \to \infty$.
\smallskip
{\it Rate.} Letting $t \to \infty$ and then $l \to \infty$, the rates of the $u$, $x_2$, and $v$ phase detection sequences approach, respectively,
\begin{align*}
R_u &= I(U;Y),\\
R_2 &= I(X_2;Y|U),\\
R_v &= I(V;Y|U,X_2).
\end{align*}
Moreover, we have
\begin{align*}
R_1 &= \frac{\log{n_1}}{k} \\
&= \frac{\log{n_un_v}}{k}\\
&= \frac{\log{n_u}+\log{n_v}}{k}\\
&= R_u + R_v \\
&= I(U;Y) + I(V;Y|U,X_2),
\end{align*}
which, together with $R_2$, is exactly our target rate pair~\eqref{eqn:split-rate}.
\smallskip
{\it Complexity.} Each of the three point-to-point phase detection sequence has detection complexity $O(k\log{k})$. Therefore, the total complexity of the multiple access detection complexity is also $O(k\log{k})$.
\begin{remark}
The original symbol-by-symbol mapping in the channel coding setting, which maps
\[
x_{1i} = x_1(u_i,v_i),
\]
does not provide the desired relation $R_1 = R_u + R_v$ in the sequence setting. This is because knowing the phase $m_u$ simultaneously reveals the phase $m_v$. In contrast, the mapping in~\eqref{eqn:mapping} ensures that for each phase $m_u \in [n_u]$, all possible phases $m_v \in [n_v]$ appear in the $x_1$ sequence. This creates the independence between the two phases $M_u$ and $M_v$.
\end{remark}
\begin{remark}
Rate splitting can be generalized to $L$-user MACs. More precisely, one can split $X_j$, $j \in[L-1]$, into two auxiliary layers $U_j$ and $V_j$, and keep $X_L$ unsplit. It is shown~\cite{Grant--Rimoldi--Urbanke--Whiting2001} that there exists a successive decoding order that achieves any rate tuple in the $L$-user MAC region~\eqref{eqn:lmac-prob}. Together with the symbol-by-symbol mappings $x_j(u_j,v_j)$ applied as in~\eqref{eqn:mapping}, we can design a rate-optimal low-complexity phase detection scheme for an $L$-user MAC.
\end{remark}
\section{Multiple Access: Probabilistic Noise,\\ Zero Error}
\label{sec:mac-advs}
In this section, we consider zero error phase detection in multiple access channels. We first demonstrate a strict separation between the channel coding setting and the phase detection setting in Section~\ref{sec:separation}. Then, we restrict our attention to zero error phase detection in the modulo-2 addition MAC in Sections~\ref{sec:mac-advs-limit} and~\ref{sec:mac-advs-code}. We note that for channel coding in the modulo-2 addition MAC, any rate pair in the zero-error capacity region $\{(R_1,R_2)\colon R_1 + R_2 \le 1\}$ can be achieved by time sharing between two rate-one codes. However, time sharing is not applicable in the phase detection scenario. Thus, our sequence design requires different ideas.
\subsection{Separation Between Phase Detection and Channel Coding}
\label{sec:separation}
Let us consider again the push-to-talk MAC (see definition in Remark~\ref{rmk:gap}). The zero error capacity region for channel coding is the set of rate pairs $(R_1,R_2)$ such that
\begin{equation}
\label{eqn:push-to-talk}
R_1 + R_2 \le 1.
\end{equation}
To see this, first note that the two corner points $(0,1)$ and $(1,0)$ can be achieved with zero error using any channel code of rate 1, and other points are achievable by time sharing. Moreover, since the output alphabet is binary, the rate region~\eqref{eqn:push-to-talk} is also an outer bound.
For zero error phase detection, a simple outer bound of $\mathscr{C}_\text{ze}$ is its vanishing error counterpart $\mathscr{C}_\text{ve}$, which is shown to be the rate region~\eqref{eqn:mac-ve-limit} in Theorem~\ref{thm:prob}. For any rate pair $(R_1,R_2)$ in the rate region~\eqref{eqn:mac-ve-limit},
\begin{align*}
R_1 + R_2 &\le I(X_1,X_2;Y)\\
& = H(Y) - H(Y|X_1,X_2)\\
&\stackrel{(a)}{\le} 1 - p_{X_1}(1)p_{X_2}(1)\\
&\stackrel{(b)}{\le} 1.
\end{align*}
For equalities in both ($a$) and ($b$) to hold, we must have $p_{X_1}(1) = 0, p_{X_2}(1) = 1/2$ or $p_{X_1}(1) = 1/2, p_{X_2}(1) = 0$, which correspond to the two corner points $(0,1)$ and $(1,0)$ respectively. Any other input pmf $p(x_1)p(x_2)$ results in a sum rate strictly less than 1. Therefore, other than the two corner points, the rate pair $(R_1,R_2)$ along the line $R_1 + R_2 =1$ is not achievable in the phase detection setting, which establishes the separation.
\subsection{Fundamental Limit for Modulo-2 Addition MAC}
\label{sec:mac-advs-limit}
\begin{theorem}
\label{thm:mac-advs}
The zero error capacity region $\mathscr{C}_\text{ze}$ for multiple access phase detection over the channel $Y = X_1 \oplus X_2$ is the set of rate pairs $(R_1,R_2)$ such that
\[
R_1 + R_2 \le 1.
\]
\end{theorem}
\begin{IEEEproof}
As argued above, this rate region is an outer bound since any codebooks induced by the zero error phase detection scheme can be used for the zero error channel coding problem. In what follows, we prove the achievability of this region using properties of linear codes, in a way that resembles Wyner's linear code for the Slapian--Wolf problem~\cite{Wyner1974}.
We choose the first sequence $x_1^{n_1}$ to be a linear sequence generated by LFSR with a primitive characteristic polynomial $a(z) = \sum_{i=0}^{r-1} a_i z^i + z^r$ over $GF(2)$ (cf. Section~\ref{sec:advs-code}). Then, for $r \le k \le n_1$, the induced codebook together with the all-zero codeword $\mathcal{C}_1 \cup \{0^k\}$ form a linear code. Let $H_{(k-r)\times k}$ be a parity check matrix of this linear code. This allows us to define $2^{k-r}$ \emph{cosets}
\[
\mathcal{C}(s^{k-r}) = \{a^k\colon H a^k = s^{k-r}\} \subseteq \{0,1\}^k.
\]
Clearly, the linear code belongs to the zero coset $\mathcal{C}_1\cup \{0^k\} = \mathcal{C}(0^{k-r})$.
Now, suppose there exist a sequence $x_2^{n_2}$ such that each length-$k$ chunk $\phi_2(m_2)$ of the sequence belongs to a different non-zero coset $\mathcal{C}(s^{k-r})$ with $s^{k-r} \neq 0^{k-r}$. Then, the phase pair $(m_1,m_2)$ can be recovered with zero-error from their sum through successive cancellation detection as follows. We take $H(\phi_1(m_1) \oplus \phi_2(m_2)) = 0^{k-r} \oplus H\phi_2(m_2) \triangleq s^{k-r}$. By design, there is only one chunk of $x_2^{n_2}$ that belongs to the coset $\mathcal{C}(s^{k-r})$. This uniquely determines the phase $m_2$. Once $\phi_2(m_2)$ is recovered, we know $\phi_1(m_1) = y^k \oplus \phi_2(m_2)$. Then by design, $m_1$ can be uniquely determined by its first $r$ bits.
We now proceed to show the existence of such a sequence $x_2^{n_2}$, using Lov\'{a}sz local lemma (Lemma~\ref{lem:lll}). We generate $X_2^{n_2}$ i.i.d. uniform. Let the ``bad'' events be $A_j = \{H \phi_2(m_2) = H\phi_2(m_2') \text{ for some } m_2 \neq m_2'\}$. Since $\phi_2(m_2)\oplus \phi_2(m_2')$ is i.i.d. uniform whether or not the two chunks overlap, the probability that the sum falls in the null space of $H$ is
\[
\P(A_j) = \P\{H(\phi_2(m_2)\oplus \phi_2(m_2')) = 0^{k-r}\} = \frac{1}{2^{k-r}}.
\]
Now each $A_j$ is mutually independent of all other events, except for a set of at most $4kn_2$ events. This is because the random variable $\phi_2(m_2)\oplus \phi_2(m_2')$ is mutually independent of all $X_{2i}$'s with $i \in [n_2]\setminus \{m_2-k+1,m_2-k+2,\ldots,m_2+k-1\} \setminus \{m_2'-k+1,m_2'-k+2,\ldots,m_2'+k-1\}$, which excludes at most $4kn_2$ events. Applying Lemma~\ref{lem:lll}, the sequence $X_2^{n_2}$ exists with positive probability
\[
\P\left\{\cap_{j=1}^{\frac12 n_2(n_2-1)}A_j^c\right\}>0
\]
if
\[
16kn_22^{-(k-r)} \le 1,
\]
or equivalently
\[
\frac{\log{n_2}}{k} + \frac{r}{k} \le 1-\frac{\log{(16k)}}{k}.
\]
By the definition of the rates, $R_2 = \frac{\log{n_2}}{k}$ and $R_1 = \frac{\log{n_1}}{k} = \frac{\log(2^r-1)}{k}\approx \frac{r}{k}$.
Letting $k\to \infty$, we conclude that a good sequence $x_2^{n_2}$ exists if $R_1 + R_2 < 1$.
\end{IEEEproof}
\subsection{Sequence Construction for Modulo-2 Addition MAC}
\label{sec:mac-advs-code}
In this Section, we show that not only does the sequence $x_2^{n_2}$ from the previous section exists, but it can also be a linear sequence. Moreover, we provide an explicit sequence construction that achieves any rate pair $(R_1,R_2) \in \mathscr{C}_\text{ze}$.
{\it Sequence Construction.}
Let $a_1(z),a_2(z) \in \mathbb{F}_2(z)$ be two distinct primitive polynomials of degree $r_1$ and $r_2$ respectively. Let $x_1^{n_1}$ and $x_2^{n_2}$ be the two linear sequences generated by the $a_1(z)$ and $a_2(z)$ respectively (cf. Section~\ref{sec:advs-code}). Letting $k \ge r_1 + r_2$, the rates of the two sequences are
\[
R_1 = \frac{\log(2^{r_1}-1)}{k} \approx \frac{r_1}{k}
\]
and
\[
R_2 = \frac{\log(2^{r_2}-1)}{k} \approx \frac{r_2}{k}.
\]
\smallskip
{\it Analysis of Detectability.} We need to show that every element in $\mathcal{C}_\text{sum}$ can be uniquely expressed as an element in $\mathcal{C}_1$ plus an element in $\mathcal{C}_2$ (see definitions of $\mathcal{C}_1,\mathcal{C}_2$, and $\mathcal{C}_\text{sum}$ in~\eqref{eqn:c1}, \eqref{eqn:c2} and~\eqref{eqn:csum} respectively). Let us first recall some definitions and facts from the LFSR theory~\cite{Lidl--Niederreiter2008}.
A one-sided infinite binary sequence ${\bf x} = \{x_i\}_{i \in \mathbb{N}}$ is said to be an \emph{LFSR sequence} if it satisfies the recursion
$ x_{r+j} = \sum_{i=0}^{r-1} a_i x_{i+j}$, for all $j \in \mathbb{N}$.
The polynomial $a(z) = z^r + \sum_{i=0}^{r-1}a_iz^i$ is a \emph{characteristic polynomial} of $\bf x$. The first $r$ bits $(x_1,\ldots,x_r)$ is the \emph{initial state} of ${\bf x}$.
Let $\mathcal{S}(a(z)) = \{{\bf x}\colon a(z)$ is a characteristic polynomial of ${\bf x}\}$. Then $\mathcal{S}(a(z))$ contains $2^r$ sequences, each corresponding to an $r$-bit initial state. One can check that $\mathcal{S}(a(z))$ is an $r$ dimensional vector space over $\mathbb{F}_2$. Define $\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))$ $ = \{{\bf x} \oplus {\bf y}\colon {\bf x} \in \mathcal{S}(a_1(z)), {\bf y} \in \mathcal{S}(a_2(z))\}$. When $a_1(z)$ and $a_2(z)$ are relatively prime, we have~\cite[Theorems 8.54, 8.55]{Lidl--Niederreiter2008}
\begin{align}
\mathcal{S}(a_1(z)) \cap \mathcal{S}(a_2(z)) &= \{\bf 0\},\label{eqn:intersectoin}\\
\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))& = \mathcal{S}(a_1(z)a_2(z)).\label{eqn:sum}
\end{align}
This is exactly the case for our construction, since $a_1(z)$ and $a_2(z)$ are distinct primitive polynomials and hence relatively prime.
Now suppose that there exists a nonzero $c^k \in \mathcal{C}_1 \cap \mathcal{C}_2$. One can find an ${\bf x} \in \mathcal{S}(a_1(z))$ such that $x^k = c^k$, since the first $r_1$ bits of ${\bf x}$ can be arbitrary and the rest $k-r_1$ bits are generated by the same polynomial $a_1(z)$. Similarly there is a ${\bf y} \in \mathcal{S}(a_2(z))$ such that $y^k = c^k$. Note that ${\bf x} \oplus {\bf y} \in \mathcal{S}(a_1(z)a_2(z))$ by~\eqref{eqn:sum} and that $a_1(z)a_2(z)$ is a polynomial of degree $r_1+r_2 \le k$. The first $k$ bits $x^k \oplus y^k = 0^k$
fully determines the whole sequence, hence ${\bf x} \oplus {\bf y} = {\bf 0}$. It follows that ${\bf x} = {\bf y}$ and ${\bf x},{\bf y} \in \mathcal{S}(a_1(z)) \cap \mathcal{S}(a_2(z))$, which contradicts~\eqref{eqn:intersectoin} since ${\bf x}$ and ${\bf y}$ are nonzero sequences starting with $c^k \neq 0^k$. This proves that $\mathcal{C}_1 \cap \mathcal{C}_2 = \emptyset$. Notice that $\mathcal{C}_i \cup \{0^k\}$ is a linear code, and thus a vector space over $\mathbb{F}_2$ (cf. Theorem~\ref{thm:lfsr}). For two vector
spaces $A$ and $B$, we know $\dim(A)+\dim(B)=\dim(A\cap B)+\dim(A+B)$.
Therefore every element in $\mathcal{C}_\text{sum}$ can be uniquely
expressed as $c_1^k \oplus c_2^k$, where $c_i^k \in \mathcal{C}_i, i= 1,2$.
\iffalse
Now that $\mathcal{C}_1 \cap \mathcal{C}_2 = \emptyset$, every element in $\mathcal{C}_\text{sum}$ can be uniquely expressed as $c_1^k \oplus c_2^k$, where $c_i^k \in \mathcal{C}_i, i= 1,2$. Suppose not, i.e., there exist $d_i^k \in \mathcal{C}_i, d_i^k \neq c_i^k$ and $c_1^k \oplus c_2^k = d_1^k \oplus d_2^k$. Then $c_1^k \oplus d_1^k = c_2^k \oplus d_2^k$. Note that $c_i^k \oplus d_i^k \in \mathcal{C}_i$ since $\mathcal{C}_i \cup \{0^k\}$ is a linear code~\cite[Theorem 2]{Wang--Hu--Shayevitz2016a}. It follows that $c_i^k \oplus d_i^k \in \mathcal{C}_1 \cap \mathcal{C}_2$, in contradiction.
Since $\mathcal{S}(a_1(z))$ and $\mathcal{S}(a_2(z))$ are vector spaces, it follows from~\eqref{eqn:intersectoin} that the sumset $\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))$ equals the \emph{direct sum} of $\mathcal{S}(a_1(z))$ and $\mathcal{S}(a_2(z))$, i.e., every element in $\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))$ is uniquely expressible by an element in $\mathcal{S}(a_1(z))$ plus an element in $\mathcal{S}(a_2(z))$.
Now we relate the induced codebooks $\mathcal{C}_1, \mathcal{C}_2$ and the sumset $\mathcal{C}_\text{sum}$ of our phase detection sequences to $\mathcal{S}(a_1(z))$, $\mathcal{S}(a_2(z))$, and $\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))$. Since $a_i(z)$ is primitive, the first $r_i$ bits in its induced codebook contain all nonzero binary sequences of length $r_i$. The rest $k-r_i$ bits are generated the same way as those in $\mathcal{S}(a_i(z))$. Therefore, $\mathcal{C}_i \cup \{0^k\}$ equals the length-$k$ truncation of $\mathcal{S}(a_i(z))$, denoted as $[\mathcal{S}(a_i(z))]_k$. For the sumset $\mathcal{C}_\text{sum}$, there are more sequences missing compared to $[\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))]_k$. But one can check that all missing sequences are of the form $\phi_1(m_1) \oplus 0^k$, $0^k \oplus \phi_2(m_2)$, or $0^k \oplus 0^k$.
Finally, we are left to show $[\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))]_k$ still equals the direct sum of $[\mathcal{S}(a_1(z))]_k$ and $[\mathcal{S}(a_2(z))]_k$. To that end, note that $\text{dim}(\mathcal{S}(a_1(z))+\mathcal{S}(a_2(z))) = \text{dim}(\mathcal{S}(a_1(z))) + \text{dim}(\mathcal{S}(a_2(z))) - \text{dim}(\mathcal{S}(a_1(z)) \cap \mathcal{S}(a_2(z))) = r_1 + r_2$. Moreover, it follows from~\eqref{eqn:sum} that the first $r_1+r_2$ bits of the sequences in $\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))$ must be all distinct, since the sequences are generated by a degree $r_1+r_2$ characteristic polynomial $a_1(z)a_2(z)$. Therefore, for any $k \ge r_1 + r_2$, $\dim([\mathcal{S}(a_1(z)) + \mathcal{S}(a_2(z))]_k) = r_1 + r_2$ and the direct sum statement holds. This implies every element in $\mathcal{C}_\text{sum}$ can also be uniquely expressed as an element in $\mathcal{C}_1$ and an element in $\mathcal{C}_2$. In particular, when $k = r_1 + r_2$, we achieve the optimal rate trade-off.
\fi
\begin{remark}
Here the crucial property is that $a_1(z)$ and $a_2(z)$ are relatively prime. In order to generalize to more than two users, one can choose $L$ distinct primitive polynomials $a_1(z),\ldots,a_L(z)$. This ensures they are relative prime. Thus~\eqref{eqn:intersectoin}~and~\eqref{eqn:sum} generalize as
{\allowdisplaybreaks
\begin{align*}
\mathcal{S}(a_1(z)) \cap \cdots \cap \mathcal{S}(a_L(z)) &= \{\bf 0\},\\
\mathcal{S}(a_1(z)) + \cdots + \mathcal{S}(a_L(z))& = \mathcal{S}(a_1(z)\cdots a_L(z)).
\end{align*}}%
Constructing the phase detection sequences from these polynomials and following a similar analysis, one can show that the zero error capacity region for phase detection in the $L$-user modulo-2 addition MAC is the set of rate tuples $(R_1,\ldots,R_L)$ such that $\sum_{i=1}^L R_i \le 1$.
\end{remark}
\section{Future Research}
There are several remaining questions. In adversarial point-to-point channel coding, it is easy to show that the GV bound can be attained using linear codes. Is it also true that a linear phase detection scheme can achieve that bound? Furthermore, since sequence design is more difficult than codebook design, can we obtain upper bounds on $C_\text{ad}(p)$ that are tighter than the ones obtained for the adversarial channel coding setup?
In the zero error point-to-point setup, we have shown how to achieve $C(G)$ in the limit of long sequences. However, when $C(G)=\alpha(G)$, this can be achieved in finite length. Suppose that $C(G)>\alpha(G)$ and is achieved by a finite length channel block code. Does there exists a finite length phase detection sequence of rate exactly $C(G)$?
\section*{Acknowledgement}
We would like to thank Tuvi Etzion for the pointer to many relevant works and for helpful comments on an earlier version of the results. We are grateful to Ronny Roth for interesting discussions. We are also thankful to the Associate Editor and anonymous reviewers, who read the manuscript thoroughly and provided many helpful comments.
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Given a finite set $\{p_1,\ldots,p_k\}$ of prime numbers, let $p_{k+1}$ be
a prime factor of $1+p_1\cdots p_k$. Then, as shown by Euclid, $p_{k+1}$
is necessarily distinct from $p_1,\ldots,p_k$. Iterating this procedure,
we thus obtain an infinite sequence of distinct primes. For instance,
beginning with $k=0$ (with the convention that the empty product is
$1$) and choosing $p_{k+1}$ as small as possible at each step, one
obtains the \emph{Euclid--Mullin sequence} (\texttt{A000945} in the
OEIS \cite{a000945}). More generally, following Clark \cite{clark},
we call any sequence resulting from this construction a
\emph{Euclid sequence with seed $\{p_1,\ldots,p_k\}$}.
One of the central questions in this area was posed by Mullin
\cite{mullin} in 1963: Does the Euclid--Mullin sequence contain every
prime number? Despite a compelling heuristic argument of Shanks \cite{shanks}
that the answer is yes, even the broader question of whether there
is \emph{any} Euclid sequence containing every prime number remains
open. (On the other hand, there are Euclid sequences that
provably do not contain every prime. For instance, starting from $k=0$
and choosing $p_{k+1}$ as large as possible at each step, one obtains the
\emph{second Euclid--Mullin sequence}, which is known to omit infinitely
many primes \cite{booker,pt}.)
In \cite{bi} it was shown that, for any given seed
$\{p_1,\ldots,p_k\}$, the possible Euclid sequences have a natural
directed graph structure. Although one can prove many interesting properties
of the family of graphs obtained by varying the seed, proving much
about any particular graph remains an elusive goal.
In this note, following a suggestion of Trevor Wooley, we consider a
generalization of Euclid's construction, in the hope that it will be more
amenable to proof. Precisely, if $\{p_1,\ldots,p_k\}$ is a set of prime
numbers, then for any $I\subseteq\{1,\ldots,k\}$, the number $N_I=\prod_{i\in
I}p_i+\prod_{i\in\{1,\ldots,k\}\setminus I}p_i$ is coprime to $p_1\cdots
p_k$ and has at least one prime factor. Iteratively choosing a set $I$ and
a prime $p_{k+1}\mid N_I$, we obtain an infinite sequence $p_1,p_2,\ldots$
of distinct prime numbers, as in Euclid's proof. (Note that Euclid's
construction is the special case in which $I=\emptyset$ at each step.)
We call a sequence resulting from this more general construction
a \emph{generalized Euclid sequence with seed $\{p_1,\ldots,p_k\}$}.
Our result is that the construction is provably
general enough to obtain every prime.
\begin{theorem}\label{t:main}
For any finite set $P$ of prime numbers, there is a generalized Euclid
sequence with seed $P$ containing every prime.
\end{theorem}
One particular generalized Euclid sequence was defined by Chua
(\texttt{A167604} in the OEIS \cite{a167604}), starting with $k=0$
and choosing $p_{k+1}$ as small as possible at each step. A natural
question, analogous to Mullin's, is whether Chua's sequence itself
contains every prime. This seems very likely, but difficult to
prove, since there is an obstruction that prevents the terms from
always appearing in numerical order. Precisely, if $n=p_1\cdots p_k$
is the product of the first $k$ terms of Chua's sequence, then the
next term $p_{k+1}$ is the smallest prime factor of $\prod_{d\mid
n}(d+n/d)$; thus, $d^2+n\equiv0\pmod*{p_{k+1}}$ for some $d$, so that
$\left(\frac{-n}{p_{k+1}}\right)=1$. (Alekseyev has conjectured that
$p_{k+1}$ is always the smallest prime satisfying this constraint;
see \cite{a167604}.) Given the well-known difficulty of proving good
bounds for the gaps between sign changes of a quadratic character,
we cannot rule out the possibility that Chua's sequence is very thin.
We conclude the introduction by mentioning another variant of Euclid's
construction, due to Pomerance \cite[\S1.1.3]{cp}: given a set of
primes $\{p_1,\ldots,p_k\}$, let $p_{k+1}$ be a prime that is not one
of $p_1,\ldots,p_k$ and divides a number of the form $d+1$ for $d\mid
p_1\cdots p_k$. Then, starting from $k=0$ and choosing $p_{k+1}$ as
small as possible at each step, one obtains a sequence containing every
prime, and in fact $p_k$ is the $k$th smallest prime for $k\ge5$. While
our variant is arguably truer in spirit to Euclid's proof (since it is
guaranteed to produce only new primes at each step), Pomerance's variant
has the distinct advantage of exhibiting a specific sequence containing
every prime.
\subsection*{Acknowledgements}
I thank Trevor Wooley and Carl Pomerance for supportive comments.
\section{Proof of Theorem~\ref{t:main}}
Given a prime number $q$, let $S_q\subseteq(\mathbb{Z}/q\mathbb{Z})^\times$ be the set
of residue classes attained by the squarefree, $(q-1)$-smooth,
positive integers, i.e.\
$$
S_q=\left\{d+q\mathbb{Z}:d\in\mathbb{Z}_{>0},\;d\mid\prod_{p<q}p\right\}.
$$
One of the main ingredients in the proof of Theorem~\ref{t:main}
is that $S_q$ is large, so that if $q$ is the smallest prime not yet
attained in $p_1,\ldots,p_k$, then there is a significant chance that
$q$ is a prime factor of $d+n/d$ for some $d\mid n=p_1\cdots p_k$. From
computation for small $q$, it seems likely that $S_q=(\mathbb{Z}/q\mathbb{Z})^\times$
for all $q\notin\{5,7\}$. We are not aware of a proof of this, but it
turns out that the following weaker approximation is sufficient for
our purposes:
\begin{lemma}\label{l:Sq}
For any prime $q$, $\#S_q>\frac12(q-1)$.
\end{lemma}
\begin{proof}
For squarefree positive integers $d\le q-1$, the residue classes $d+q\mathbb{Z}$
are distinct and contained in $S_q$. By \cite{rogers}, the number of
such $d$ is at least $\frac{53}{88}(q-1)>\frac12(q-1)$.
\end{proof}
In addition, we need one further input from algebraic geometry:
\begin{lemma}\label{l:hyp}
Let $q$ be an odd prime number and $a\in(\mathbb{Z}/q\mathbb{Z})^\times$.
\begin{itemize}
\item[(i)]If $q\ne5$ or $q=5$ and $a\ne3+5\mathbb{Z}$ then
there exists $x\in(\mathbb{Z}/q\mathbb{Z})^\times$ such that
$\left(\frac{x+a/x}{q}\right)\ne1$.
\item[(ii)]If $q\notin\{7,13\}$ then
there exists $x\in(\mathbb{Z}/q\mathbb{Z})^\times$ such that
$\left(\frac{x^6+a}{q}\right)\ne1$.
\end{itemize}
\end{lemma}
\begin{proof}
We consider the sum
$$
\sum_{x\in(\mathbb{Z}/q\mathbb{Z})^\times}\left(\frac{x+a/x}{q}\right)
=\sum_{x\in(\mathbb{Z}/q\mathbb{Z})^\times}\left(\frac{x(x^2+a)}{q}\right).
$$
For $q\ge3$, $x(x^2+a)$ has no repeated roots modulo $q$, so that
$$
\{(x,y)\in(\mathbb{Z}/q\mathbb{Z})^2:y^2=x(x^2+a)\}
$$
are the affine points of an elliptic curve. The curve has one point
at infinity, so by the Hasse bound, we have
$$
1+\sum_{x\in\mathbb{Z}/q\mathbb{Z}}\left[1+\left(\frac{x(x^2+a)}{q}\right)\right]
\le q+1+2\sqrt{q},
$$
whence
$$
\sum_{x\in(\mathbb{Z}/q\mathbb{Z})^\times}\left(\frac{x(x^2+a)}{q}\right)\le 2\sqrt{q}.
$$
This last estimate is less than $q-1$ provided that $q\ge 7$, and we check
the claim for $q\in\{3,5\}$ directly.
Similarly, for $q\ge5$, $x^6+a$ has no repeated roots modulo $q$, so that
$$
\{(x,y)\in(\mathbb{Z}/q\mathbb{Z})^2:y^2=x^6+a\}
$$
are the affine points of a genus $2$ curve. The curve has two points
at infinity, so by the Weil bound, we have
$$
2+\sum_{x\in\mathbb{Z}/q\mathbb{Z}}\left[1+\left(\frac{x^6+a}{q}\right)\right]
\le q+1+4\sqrt{q},
$$
whence
$$
\sum_{x\in(\mathbb{Z}/q\mathbb{Z})^\times}\left(\frac{x^6+a}{q}\right)\le
4\sqrt{q}-1-\left(\frac{a}{q}\right)\le 4\sqrt{q}.
$$
This last estimate is less than $q-1$ provided that $q\ge 19$, and we check
the claim for $q\in\{3,5,11,17\}$ directly.
\end{proof}
Theorem~\ref{t:main} follows by induction from the following proposition.
\begin{proposition}
Let $P$ be a finite set of prime numbers and $q$ the smallest prime
not contained in $P$. Then there is a generalized Euclid sequence with
seed $P$ that contains $q$.
\end{proposition}
\begin{proof}
Suppose that $P=\{p_1,\ldots,p_k\}$, and put $n=p_1\cdots p_k$.
If $q=2$ then $n+1$ is even, so we may choose $2$ as the next term,
$p_{k+1}$. Hence we may assume that $q$ is odd.
Put
$$
S=\{d+q\mathbb{Z}:d\in\mathbb{Z}_{>0},\;d\mid n\}\subseteq(\mathbb{Z}/q\mathbb{Z})^\times,
$$
and note that $S\supseteq S_q$. Suppose first that
$S=(\mathbb{Z}/q\mathbb{Z})^\times$. If $\left(\frac{-n}{q}\right)=1$ then it
follows that there is a $d\mid n$ such that $d+n/d\equiv0\pmod*{q}$,
so we can choose $q$ as the next term. On the other hand, if
$\left(\frac{-n}{q}\right)=-1$ then by Lemma~\ref{l:hyp}(i) we may choose
$d\mid n$ such that $\left(\frac{d+n/d}{q}\right)=-1$, provided that
$q\ne5$ or $n\not\equiv3\pmod*{5}$. For this choice of $d$ there must be a
prime $p\mid(d+n/d)$ such that $\left(\frac{p}{q}\right)=-1$. Choosing
this $p$ as the next term, we replace $n$ by $n'=pn$, so that
$\left(\frac{-n'}{p}\right)=1$, and we may then follow this by
$q$, as above. For $q=5$ and $n\equiv3\pmod*{5}$ we choose $d=1$;
since $n+1\equiv-1\pmod*{5}$ there is a prime $p\mid(n+1)$ with
$p\not\equiv1\pmod*{5}$, and replacing $n$ by $pn$ gives a different
residue with which we can carry out the proof above.
Suppose now that $S\ne(\mathbb{Z}/q\mathbb{Z})^\times$. We seek to enlarge $S$ by
continuing the sequence, i.e.\ we choose $p=p_{k+1}$ from
$$
T=\{p:p\text{ prime and }p\mid(d+n/d)\text{ for some }d\mid n\},
$$
and replace $P$ by $P\cup\{p\}$, $n$ by $pn$ and $S$ by $S\cup pS$.
We are free to repeat this procedure until either $q\in T$ (in which
case we may choose $q$ as the next term) or $S$ stabilizes, so that
$pS\subseteq S$ for every choice of $p\in T$. If that is the case then
it is easy to see that for every $s\in S$, $S$ contains the coset $sG$,
where $G\le(\mathbb{Z}/q\mathbb{Z})^\times$ is the subgroup generated by $\{p+q\mathbb{Z}:p\in
T\}$. Thus, $S=\bigcup_{s\in S}sG$ is a union of cosets; in particular,
$\#G$ divides $\#S$.
Next, let $H$ be a subgroup of $(\mathbb{Z}/q\mathbb{Z})^\times$ of index at least $4$.
For any $h\in H$, the number of $d\in(\mathbb{Z}/q\mathbb{Z})^\times$ such that
$d+n/d=h$ is at most $2$. Hence,
$$
\#\{d\in(\mathbb{Z}/q\mathbb{Z})^\times:d+n/d\in H\}\le 2\#H\le\tfrac12(q-1).
$$
By Lemma~\ref{l:Sq},
it follows that there exists $d\mid n$ such that $(d+n/d)+q\mathbb{Z}\notin H$.
In turn this implies that $p+q\mathbb{Z}\notin H$ for some $p\in T$.
For any $r\mid(q-1)$, consider the subgroup
$$
H_r=\{h\in(\mathbb{Z}/q\mathbb{Z})^\times:h^{\frac{q-1}{r}}=1\}
=\{x^r:x\in(\mathbb{Z}/q\mathbb{Z})^\times\}.
$$
Let $q-1=\prod_{i=1}^mr_i^{e_i}$ be the prime factorization of $q-1$.
For each $r_i\ge5$ we apply the above argument with
$$
H=H_{r_i}=\{h\in(\mathbb{Z}/q\mathbb{Z})^\times:
r_i^{e_i}\text{ does not divide the order of }h\}
$$
to see that $G$ has order divisible by $r_i^{e_i}$. For $r_i\in\{2,3\}$
the index of $H_{r_i}$ is too small to apply the argument, but we may
still apply it to $H_{r_i^2}$ (when $r_i^2\mid(q-1)$) to see that $G$
has order divisible by $r_i^{e_i-1}$. Thus we find that the index of $G$
in $(\mathbb{Z}/q\mathbb{Z})^\times$ divides $6$.
If $q\not\equiv1\pmod*{3}$ then $G$ has index at most $2$, so that
$\frac12(q-1)\mid\#G\mid\#S$; by Lemma~\ref{l:Sq} it follows that
$S=(\mathbb{Z}/q\mathbb{Z})^\times$, as desired.
If $q\equiv1\pmod*{3}$ then we apply the
above argument with $H=H_6$ to see that there exists $p\in T$ such
that $p^{\frac{q-1}{6}}\not\equiv1\pmod*{q}$. Since
$p^{\frac{q-1}{6}}=p^{\frac{q-1}{2}}/p^{\frac{q-1}{3}}$, it follows that
at least one of $H_2$ and $H_3$ does not contain $p+q\mathbb{Z}$. If
$p+q\mathbb{Z}\notin H_3$ then again $G$ has index at most $2$, and we
conclude that $S=(\mathbb{Z}/q\mathbb{Z})^\times$ as above.
Hence, we may assume that $p+q\mathbb{Z}\notin H_2$,
so that $G$ has index dividing $3$.
If $S=(\mathbb{Z}/q\mathbb{Z})^\times$ then we are finished, so we may assume
that $G=H_3$ and $\#S<q-1$. By Lemma~\ref{l:Sq}, we must have $\#S>\#G$,
and it follows that $S=G\cup sG$ for some $s\in(\mathbb{Z}/q\mathbb{Z})^\times\setminus G$.
Going through the argument above with $H=H_3$, to avoid concluding that
there exists $p\in T$ such that $p+q\mathbb{Z}\notin H_3$, the function
$d\mapsto d+n/d$ must map $S$ 2--1 onto $H_3$. By the quadratic formula,
this means in particular that
$\left(\frac{h^2-4n}{q}\right)=1$ for every $h\in H_3$, and thus
$\left(\frac{x^6-4n}{q}\right)=1$ for every $x\in(\mathbb{Z}/q\mathbb{Z})^\times$.
However, that contradicts Lemma~\ref{l:hyp}(ii) for
$q\notin\{7,13\}$, and for $q\in\{7,13\}$ we verify directly
that $\#S_q>\frac23(q-1)$. This concludes the proof.
\end{proof}
\bibliographystyle{amsplain}
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{Introduction}
This paper forms part of a series by the authors \cites{GMV1,GMV3} concerning the structure theory of \emph{nilspaces}. Much of this is concerned with the approach of Szegedy \cite{S12} and Antol\'\i n Camarena and Szegedy \cite{CS12} to the inverse theorem for the Gowers norms, as well as with relations to dynamics and in particular work of Host and Kra \cite{HK05} and Host, Kra and Maass \cite{HKM10}.
The paper \cite{GMV1} contains an extensive introduction to this project from the viewpoint of higher order Fourier analysis and the inverse theorem for the Gowers norms. Similarly, \cite{GMV3} introduces the project from a dynamical perspective. We will not repeat the bulk of these introductions here, nor offer much motivation for the definition or study of nilspaces and related constructs, but instead refer the reader to these companion works.
Nilspaces originate in work of Host and Kra \cite{HK08}, where these objects appeared under the name of ``parallelepiped structures''. The study of these objects was furthered by Antol\'\i n Camarena and Szegedy \cite{CS12}, who in the same work formulated a strong structure theorem for nilspaces, subject to certain further hypotheses.
The papers of Candela \cites{Can1,Can2} expand on \cite{CS12}, providing more detailed proofs. He also includes
several additional results implicit in \cite{CS12}, particularly about continuous systems of measures.
The goals of this work are as follows.
\begin{itemize}
\item We prove a structure theorem for nilspaces with certain additional topological assumptions, which allow us to deduce that they are isomorphic (in a suitable sense) to nilmanifolds $G/\Gamma$. This is a key stepping stone towards the general structure theorem.
\item Along the way, we prove some rather technical results concerning ``cocycles'' on cubespaces (closely related to \cite{CS12}*{Section 3.6}). The resulting primitive ``cohomology'' theory is a powerful tool for deducing algebraic information from topological data, and will be invaluable both here and elsewhere in the project \cite{GMV3}.
\end{itemize}
The structural result we will prove here roughly has the following flavour: if a (compact, ergodic) nilspace $X = (X, C^n(X))$, satisfying some connectivity hypotheses, has any chance of being a nilmanifold topologically -- for instance, it had better be a topological manifold, i.e.~locally homeomorphic to a subset of $\RR^n$ -- then it is a nilmanifold.
One possible formulation of the statement is the following theorem.
(For the definition of nilspaces we refer the reader to \cite{GMV1}*{Section 3.1}
or \cite{GMV3}*{Section 1.3}, and for that of Host--Kra cubes,
see \cite{GMV1}*{Section 2 and Appendix A} or \cite{GMV3}*{Section 1.4}).
\begin{theorem}
\label{main-thm-simple}
Let $X = (X, C^n(X))$ be a \uppar{compact, ergodic}\footnote{The assumption that a nilspace be compact and ergodic is in force universally, and so it usually appears in parentheses to downplay its particular significance in any given context. However, these conditions are certainly not optional.} nilspace of degree $s$. Suppose $X$ is locally connected and of finite Lebesgue covering dimension, and further that all the spaces $C^n(X)$ are connected.
Then $X$ is isomorphic to a nilmanifold $G/\Gamma$. That is, there exists a filtered connected Lie group $G_\bullet$, a discrete co-compact subgroup $\Gamma$ of $G$, and a homeomorphism $\phi \colon X \leftrightarrow G/\Gamma$ that identifies the cubes $C^k(X)$ with the Host--Kra cubes $\HK^k(G_\bullet)/\Gamma$.
\end{theorem}
We note that the conditions on local connectedness and finite dimensionality hold in particular if $X$ is a topological manifold.
The topological conditions in the above theorem can be replaced by other sets of assumptions involving the so-called structure groups of the nilspace $X$. Such variants will be formulated below, including the one proved by Antol\'\i n Camarena and Szegedy \cite{CS12}*{Theorem 7}. In fact, we only prove one of these variants in this paper and the proof of Theorem \ref{main-thm-simple} is completed in \cite{GMV3}*{Theorem A.1}
using the general structure theory of nilspaces.
Both the overall structure of these arguments, and a good part of the fine detail, are modelled closely on the work of Antol\'\i n Camarena and Szegedy, and we will attempt to make this dependence explicit wherever possible. However, in some aspects we have deviated from their approach. Our reasons for doing so are some mixture of the following.
\begin{itemize}
\item In places, we obtain slightly stronger results, which will be useful especially in applications to dynamics. Obtaining these can require a modified approach.
\item Our arguments stay entirely within the topological category, avoiding reference to objects that are merely measurable rather than continuous. This eliminates some subtleties encountered in \cite{CS12} concerning the replacement of measurable objects by continuous ones; albeit arguably at the expense of introducing other subtleties in different places. In particular this strategy requires a different treatment of the part of the argument covered in Section \ref{sec-main-details}. Similarly, for some results in Section \ref{sec:cocycle}, we will have to prove strengthened versions, which are not required here but are needed to keep the arguments in \cite{GMV3} in the topological category.
\item In some cases, we find our alternatives simpler, easier to understand
or to yield a fuller understanding of the methods and structures involved.
\end{itemize}
\subsection{Structure of the paper}
As we have said, we will not repeat in detail the relevant definitions of cubespaces, nilspaces, ergodicity and so on, that appear in \cite{GMV1}*{Section 3.1},
referring the reader to that paper. Similarly, we will assume some familiarity with the crucial subject of the \emph{structure groups} $A_k(X)$ of a nilspace, and with the ``weak structure theorem'' of \cite{GMV1}*{Section 5.1}.
Alternatively, a reader mainly interested in dynamics may read \cite{GMV3}*{Sections 1.3--1.6}.
The paper \cite{GMV3} introduces the same notions and their properties motivated from a dynamical viewpoint.
However, the proofs are available only in \cite{GMV1}.
We will recall only some of the more specialized definitions relating to Theorem \ref{main-thm-simple} and its generalizations, in Section \ref{sec-statements} below. The same section will introduce formally the group $\Aut_k(X)$ of \emph{$k$-translations} of a nilspace, which will play a key role in the arguments. Section \ref{sec-statements} concludes with a discussion of the various variations on the statement of Theorem \ref{main-thm-simple}.
Section \ref{sec:main-outline} explains the high-level steps of the proof of the structural results.
This reduces the structural result to a statement (see Proposition \ref{lem:surj})
that a nilmanifold $X$ has ``enough automorphisms'' in a certain technical sense.
We establish this technical statement in Section \ref{sec-main-details}, conditional on a
``cohomological'' theorem (see Theorem \ref{thm:cocycle-special}) that will arise fairly naturally in the course of the proof.
We introduce the cocycle theory, and give a proof of this remaining theorem, in Section \ref{sec:cocycle}.
\subsection{Acknowledgments}
First and foremost we owe gratitude to Bernard Host who introduced us to the subject and to Omar Antol\'\i n Camarena and Bal\'azs Szegedy
whose groundbreaking work \cite{CS12} was a constant inspiration
for us.
We would like to thank Emmanuel Breuillard, J\'er\^ome Buzzi,
Yves de Cornulier, Sylvain Crovisier, Eli Glasner, Ben Green, Bernard Host, Micha\pol{} Rams, Bal\'azs Szegedy, Anatoly Vershik and Benjamin
Weiss for helpful discussions. We are grateful to Pablo Candela and Bryna Kra for a careful reading of a preliminary version. We are grateful to Jacob Rasmussen for suggesting the reference \cite{S51}.
We are grateful to the referee for her or his careful reading of our paper and for
her or his many helpful comments, which greatly improved the presentation of the paper.
\section{Definitions and statements}
\label{sec-statements}
The flavour of all our structural statements is to find conditions one can impose on a cubespace $X = (X, C^n(X))$ that are sufficient to ensure that it is actually a nilmanifold $G / \Gamma$.
The various ``algebraic'' constraints we impose are discussed at length in \cite{GMV1}. In short, it is fairly hopeless to ask for a rigid structure theorem unless we insist that $X$ is an ergodic nilspace.
However, we also require some topological input (beyond the standard assumption that $X$ is compact). For instance, a nilmanifold is certainly a smooth real manifold, so we need to rule out examples such as the solenoid\footnote{Here $\ZZ_2$ denotes the $2$-adic integers and $\ZZ$ is embedded diagonally in the product.} $(\RR \times \ZZ_2) / \ZZ$, which is a compact abelian group and hence a nilspace of degree $1$, but not a manifold.
In the previous section, we formulated Theorem \ref{main-thm-simple} under topological conditions on the space $X$ and the space of cubes $C^n(X)$. We will now state a similar result in which these conditions are replaced by conditions on the \emph{structure groups} of $X$. We recall that these are a sequence of compact abelian groups $A_k(X)$, defined canonically in terms of the cubestructure on $X$. The key topological hypothesis we can impose is that these are Lie groups. \footnote{Note we do not assume a Lie group is connected. So, a compact abelian Lie group is precisely of the form $(\RR/\ZZ)^d \times K$ where $K$ is a (discrete) finite group.}
These formulations turn out to fit more naturally in the theory, and we prove only these statements in this paper.
Unfortunately the condition that $A_k(X)$ are Lie still does not suffice. Our method also requires some fairly strong assumptions on connectivity, to which we now turn.
\subsection{Connectivity hypotheses}
The simplest connectivity hypothesis is just that $X$ itself is connected as a topological space. It turns out that this condition becomes inadequate fairly quickly, because it fails to establish connectedness ``on all levels'' that is required to explore the structure of $X$ inductively.
Sticking with the idea of imposing conditions on the structure groups $A_k(X)$, one can insist that these groups are connected for all $k$. Since we have already assumed that they are Lie, this is equivalent to asking that they be tori $(\RR/\ZZ)^d$ for some integer $d$.
\begin{definition}
We say a (compact, ergodic) nilspace $X$ is \emph{toral} if for each $k \ge 1$ the structure group\footnote{For a definition see \cite{GMV1}*{Theorem 5.4}.} $A_k(X)$ is isomorphic to a torus $(\RR/\ZZ)^d$ for some integer $d$ (which depends on $X$ and $k$).
\end{definition}
This is the approach taken in \cite{CS12}.
\begin{theorem}[\cite{CS12}*{Theorem 7}]\label{thm:main-toral}
Let $X = (X, C^n(X))$ be a \uppar{compact, ergodic} toral nilspace of degree $s$. Then $X$ is isomorphic to a nilmanifold $G/\Gamma$ in the sense of Theorem \ref{main-thm-simple}.
\end{theorem}
For the purposes of our core structural result, Theorem \ref{thm:main-sc} below, we will work with a slightly different topological condition, this time relating to the spaces of cubes $C^k(X)$. This is needed for compatibility with our statement of the final structure theorem for general nilspaces (\cite{GMV1}*{Theorem 4.1}).
\begin{definition}
We say a nilspace $X$ is \emph{strongly connected} if the topological space $C^k(X)$ is connected for all $k \ge 0$. (In particular, $X = C^0(X)$ itself is connected.)
\end{definition}
It turns out that these two connectivity conditions are equivalent.
While it is easy to see that a toral nilspace is strongly connected
(see Proposition \ref{prop-toral-sc}), the converse appears much more difficult.
In fact, the only proof we are aware of is based on the full force of our
structure theorem.
Indeed, Theorem \ref{thm:main-sc} below implies that a strongly connected nilspace with Lie structure
groups is isomorphic to a Host--Kra nilspace of a connected nilpotent Lie group endowed with a filtration
of connected subgroups, and hence is a toral nilspace \emph{a fortiori}.
\begin{proposition}
\label{prop-toral-sc}
A toral \uppar{compact, ergodic} nilspace is strongly connected.
\end{proposition}
\begin{proof}
By the weak structure theorem (\cite{GMV1}*{Theorem 5.4})
the space $C^n(X)$ is expressible as a tower of extensions (more correctly, principal bundles)
\[
C^n(X) = C^n(X_s) \to C^n(X_{s-1}) \to \dots \to C^n(X_1) \to \{\ast\}
\]
where each fiber is the compact abelian group $C^n(\cD_k(A_k(X)))$, which in turn is isomorphic to $A_k(X)^d$ for some $d \ge 0$, and in particular is connected.
Since a tower of extensions by connected fibers is connected, we deduce that $C^n(X)$ is connected.
\end{proof}
In fact our techniques allow us to say something even in the absence of any such connectivity hypothesis, and we will make statements in this setting as well (see Theorem \ref{weaker-main-thm} and Corollary \ref{cor-weaker-main-thm} in the sequel). These statements are especially interesting from the viewpoint of topological dynamics.
\subsection{Some notation}
We recall some miscellaneous notation from \cite{GMV1}. Given two configurations $c, c' \colon \{0,1\}^n \to X$, we denote by $[c, c']$ the ``concatenated'' configuration
\begin{align*}
[c,c'] \colon \{0,1\}^{n+1} &\to X \\
\omega &\mapsto \begin{cases} c(\omega_1,\dots,\omega_n) &\colon \omega_{n+1} = 0 \\ c'(\omega_1,\dots,\omega_n) &\colon \omega_{n+1} = 1 \end{cases} \ .
\end{align*}
We will not be too concerned about which coordinate in $\{1, \dots, n+1\}$ is the preferred one along which the concatenation occurs; i.e.~some other coordinate may play the role of $(n+1)$ in this expression, and we will still denote the final configuration by $[c, c']$.
We also use $\square^n(x)$ to denote the constant configuration $\{0,1\}^n \to X$ sending every coordinate to $x$. Similarly, the notation $\llcorner^n(x; y)$ denotes the configuration
\begin{align*}
\llcorner^n(x; y) \colon \{0,1\}^n &\to X \\
\omega &\mapsto \begin{cases} y &\colon \omega = \vec{1} \\ x &\colon \omega \ne \vec{1} \end{cases} \ .
\end{align*}
More generally, we use the notation $\square^n(c)$ where $c\in X^{\{0,1\}^k}$ to denote the $(n+k)$-configuration $(\omega_1,\ldots,\omega_{n+k})\mapsto c(\omega_1,\ldots,\omega_k)$. We use the notation $\llcorner^n(c_1; c_2)$ in a similar manner for $c_1,c_2\in X^{\{0,1\}^k}$.
We reserve the right to mix and abuse notation freely, as in $\llcorner^2(\square^3(x); \square^3(y))$. Hopefully the meaning will always be clear.
Finally, we use the standard notation $X \lesssim_{a,b,\dots} Y$ to denote that $X \le C(a,b,\dots) Y$ for some constant $C$ depending only on the variables $(a,b,\dots)$.
\subsection{Automorphisms and translations}
The heart of the problem of identifying a suitable nilspace $X$ with a nilmanifold $G / \Gamma$, is in recovering the group $G$ and its group law.
The approach to doing this taken in \cite{CS12}, which we follow, is to consider a suitable group of automorphisms of $X$. Since automorphism groups are clearly groups, we will be in good shape if we can take $G$ to be this group, and then identify $X$ with a suitable quotient of $G$.
However, we want to describe not just $X$ as a topological space, but also its cube structure. The cubes on a nilmanifold $G / \Gamma$ are given by the Host--Kra construction, which requires the additional data of a \emph{filtration} on the group $G$. (For a detailed account of this theory, again see \cite{GMV1}*{Section 2 and Appendix A}.)
So, we need to find not just a suitable group of automorphisms of $X$, but also a filtration on that group.
Fortunately, there are natural definitions of all of these objects, which we now describe. The automorphism group itself is straightforward.
\begin{definition}
Let $X$ be a compact cubespace. We write $\Aut(X)$ for the group of all automorphisms of $X$ in the category of cubespaces; that is, the collection of homeomorphisms $\phi \colon X \to X$ such that for any configuration $c \colon \{0,1\}^k \to X$, $c \in C^k(X)$ if and only if $\phi(c) \in C^k(X)$.
We endow $\Aut(X)$ with the usual supremum metric
\[
d(\phi,\psi) = \sup_{x \in X} d(\phi(x),\psi(x))=\|\psi\circ\phi^{-1}\|,
\]
where we used the notation:
\[
\|\phi\| = \sup_{x \in X} d(x, \phi(x)),
\]
which generate the usual compact-open topology.
\end{definition}
\begin{remark}
\label{rem:second-countable-auts}
It is straightforward to verify that $\Aut(X)$ is a closed subgroup of the group of homeomorphisms of $X$, and hence completely metrizable.\footnote{Note that we do not claim that the supremum metric itself on $\Aut(X)$ is complete; merely that $\Aut(X)$ is complete with respect to some metric which generates the same topology, e.g., $d'(\phi,\psi) = d(\phi,\psi) +d(\phi^{-1},\psi^{-1})$ (\cite{BK96}*{Corollary 1.2.2}).
The choice of the metric is not important, but we stick with the supremum metric for convenience of notation.}
It follows from a standard theorem (see \cite{BK96}*{Example 1.3(v)}), that $\Aut(X)$ is also separable, or equivalently second countable.
\end{remark}
We now consider the filtration.
\begin{definition}
\label{def:k-translation}
Again let $X$ be a compact cubespace, and fix $k \ge 0$. Given $\phi \in \Aut(X)$ and a face $F$ of $\{0,1\}^n$, let $[\phi]_{F}$ denote the element of $\Aut(X)^{\{0,1\}^n}$ given by
\[
\omega \mapsto \begin{cases} \phi &\colon \omega \in F \\ \id & \colon \omega \notin F \end{cases}
\]
We write $\Aut_k(X)$ for the collection of $\phi \in \Aut(X)$ with the following additional property. For any integer $n \ge k$, any face $F$ of $\{0,1\}^n$ of codimension $k$, and any $c \in C^n(X)$, the configuration $[\phi]_{F}. c$
\begin{align*}
\{0,1\}^n &\to X \\
\omega &\mapsto \begin{cases} \phi(c(\omega)) &\colon \omega \in F \\ c(\omega) &\colon \omega \notin F \end{cases}
\end{align*}
is in $C^n(X)$.
We refer to the elements of $\Aut_k(X)$ as \emph{$k$-translations}. Clearly $\Aut_0(X) = \Aut(X)$.
\end{definition}
The notion of translations originate from the work of Host and Kra \cite{HK08}*{Definition 6}
and they play a pivotal role in the programme of
Antol\'\i n Camarena and Szegedy \cite{CS12}, which we discuss now.
The motivation for this definition is perhaps not entirely clear.%
\footnote{
Below is one way to arrive at this definition. We stress that the following discussion is purely for motivation, and is not logically necessary for the argument (although some ideas discussed now will be relevant later; see for instance Proposition \ref{prop:evaluation-is-morphism}).
We define a \emph{cubegroup} to be an object $G$ that is both a topological group and a cubespace, with the added requirement that $C^k(G)$ is a (closed) subgroup of $G^{\{0,1\}^k}$ under pointwise operations, for all $k$.
It turns out that $\Aut(X)$ is a natural example of a cubegroup: that is, there is a canonical categorial notion of when $2^k$ automorphisms of $X$ form a $k$-cube, and this notion is closed under pointwise composition. The cubespace structure can be described informally as the largest possible one such that $\omega \mapsto \phi_\omega(c(\omega))$ is a cube of $X$ for every $(\omega \mapsto \phi_\omega) \in C^k(\Aut(X))$ and every $c \in C^k(X)$, i.e.~such that pointwise action of a cube on $X$ sends cubes to cubes. More precisely, one should define an element $\psi \colon \{0,1\}^k \to \Aut(X)$ to be in $C^k(\Aut(X))$ if and only if $\square^\ell(\psi)(c) \in C^{k+\ell}(X)$ for every $\ell \ge 0$ and $c \in C^{k+\ell}(X)$, and then one can check that this defines a cubegroup structure on $\Aut(X)$.
It is a fact (which we will not prove, because we do not need it anywhere in the paper) that -- in complete generality -- all cubegroups arise from the Host--Kra construction applied to a filtered group. That is, given a cubegroup $G$, there is a unique filtration $G_\bullet$ on $G$ such that $C^k(G) = \HK^k(G_\bullet)$ for all $k$, and so specifying a cubegroup structure on $G$ is equivalent to specifying a filtration. Under this correspondence, the filtration $\Aut_k(X)$ from Definition \ref{def:k-translation} is precisely the one giving rise to the cubegroup structure on $\Aut(X)$ discussed above.
In the interests of concreteness and simplicity we will suppress explicit discussion of the cube structure on $\Aut(X)$ in what follows (referring equivalently instead to $\HK^k(\Aut(X)_\bullet)$) and phrase everything in terms of filtrations, taking Definition \ref{def:k-translation} as the logical starting point.
}
We will illustrate it somewhat with some facts and examples.
\begin{proposition}
\label{translation-prefiltration}
Let $X$ be a compact cubespace. The groups $(\Aut_k(X))_{k \ge 0}$ form a \emph{filtration}. That is, $\Aut_k(X)$ are a decreasing sequence of closed subgroups of $\Aut(X)$, and for any $i, j$ we have have the commutator inclusion $[\Aut_i(X), \Aut_j(X)] \subseteq \Aut_{i+j}(X)$.
\end{proposition}
\begin{proof}
That $\Aut_k(X)$ forms a subgroup is clear, and it is similarly straightforward to argue that they are closed in $\Aut(X)$ (since $C^n(X)$ is closed in $X^{\{0,1\}^n}$ for all $n$). Similarly, by modifying the same cube twice on an adjacent pair of faces, it is easy to see that a $(k+1)$-translation is also a $k$-translation for all $k$.
For the commutator result, take any $\phi \in \Aut_i(X)$, $\psi \in \Aut_j(X)$ and $c \in \{0,1\}^n$ for some $n \ge i + j$, and let $F$ be a face of $\{0,1\}^n$ of codimension $(i+j)$. Pick faces $F_1, F_2 \subseteq \{0,1\}^n$ of codimensions $i, j$ respectively such that $F_1 \cap F_2 = F$ and note that
\[
[\phi]_{F_1} [\psi]_{F_2} [\phi]_{F_1}^{-1} [\psi]_{F_2}^{-1} = [\phi \psi \phi^{-1} \psi^{-1}]_F
\]
(with operations applied pointwise). But clearly the configuration
\[
[\phi]_{F_1} [\psi]_{F_2} [\phi]_{F_1}^{-1} [\psi]_{F_2}^{-1} . c
\]
is in $C^n(X)$ by hypothesis on $\phi$ and $\psi$, as required.
\end{proof}
\begin{proposition}
\label{prop-translations-terminate}
Suppose $X$ is a compact cubespace with $k$-uniqueness. Then $\Aut_k(X) = \{\id\}$.
\end{proposition}
\begin{proof}
For any $x \in X$ and $\phi \in \Aut_k(X)$, let $F = \{\vec{1}\} \subseteq \{0,1\}^k$, and apply $[\phi]_F$ to the constant cube $c = \square^k(x)$ in $C^k(X)$. Then $[\phi]_F . c$ is a cube, agreeing with $c$ on all but the topmost vertex of $\{0,1\}^k$, and so is equal to $c$ by $k$-uniqueness. Hence $\phi(x) = x$.
\end{proof}
Note that in general we may have that $\Aut_0(X) \ne \Aut_1(X)$, and hence even if $X$ is a nilspace of degree $s$, this filtration is not proper, hence it does not guarantee that $\Aut_0(X)$ is a nilpotent group: indeed, we will see below that it usually is not. However, it \emph{does} guarantee that $\Aut_1(X)$ is nilpotent (of nilpotency class at most $s$) under these conditions. Hence, the group of $1$-translations, equipped with the filtration
\begin{equation}\label{eq:trans-filtration}
\Aut_1(X) = \Aut_1(X) \supseteq \Aut_2(X) \supseteq \dots \supseteq \Aut_k(X) \supseteq \dots
\end{equation}
is a much better candidate for our group $G$ used to construct the nilmanifold $G / \Gamma$. We will never have much need to consider the full automorphism group $\Aut(X)$ again, restricting our attention to $\Aut_1(X)$.
In what follows, we denote by $\Aut_\bullet(X)$ the group $\Aut_1(X)$ endowed with the
filtration \eqref{eq:trans-filtration}.
\begin{example}
Suppose $X = (\RR/\ZZ)^2$, with cubespace structure coming from the usual degree $1$ filtration. Then $\Aut(X) \cong \GL_2(\ZZ) \ltimes (\RR/\ZZ)^2$, the group of \emph{affine automorphisms} of $(\RR/\ZZ)^2$, acting by
\[
(M,a) . x = M x + a \ .
\]
Indeed, since the $k$-cubes of $X$ are the $k$-dimensional parallelepipeds, i.e.~configurations of the form
\begin{align*}
\{0,1\}^k &\to (\RR/\ZZ)^2 \\
\omega &\mapsto x + \sum_{i=1}^k \omega_i h_i
\end{align*}
for some coefficients $x, h_i \in (\RR/\ZZ)^2$, it is easy to see that such maps send cubes to cubes. Conversely, since the $2$-cubes of $X$ are the configurations $[[x, y],[z,w]]$ such that $x+w=y+z$, we have that for any $\phi \in \Aut(X)$ and $x,y,z,w \in X$ such that $x+w=y+z$,
\[
\phi(x) + \phi(w) = \phi(y) + \phi(z)
\]
and so in particular taking $z=0$ we have
\[
(\phi(x+w) - \phi(0)) = (\phi(x) - \phi(0)) + (\phi(w) - \phi(0))
\]
so we deduce that $(\phi - \phi(0))$ is necessarily linear and hence $\phi$ is affine-linear.
The subgroup of $1$-translations is precisely $\Aut_1(X) = \{\id\} \ltimes (\RR/\ZZ)^2$. To see this, we observe that if $\phi \in \Aut^1(X)$ and $[[x,y],[z,w]] \in C^2(X)$ then so is $[[x,y],[\phi(z), \phi(w)]]$; equivalently, if $x,y,z,w \in (\RR/\ZZ)^2$ satisfy $x-y=z+w$ then $x-y=\phi(z)-\phi(w)$. So, $\phi(z)-\phi(w) = z - w$ for all $z,w \in (\RR/\ZZ)^2$ and hence $\phi$ is just a translation $x \mapsto x+t$.
It is easy to check that such maps obey the $1$-translation property for higher-dimensional cubes.
For $k \ge 2$, $\Aut_k(X)$ is the trivial group.
\end{example}
\begin{example}
Let $X = G / \Gamma$ where $G$ is the Heisenberg group
\[
G = \left\{ \heis{x}{y}{z} \colon x, y, z \in \RR \right\}
\]
with its central series filtration, and $\Gamma$ is
\[
\Gamma = \left\{ \heis{x}{y}{z} \colon x, y, z \in \ZZ \right\} \ ,
\]
the usual discrete co-compact subgroup.
Then for any $g \in G$, the map
\[
x \Gamma \mapsto g x \Gamma
\]
is a $1$-translation on $G / \Gamma$. If $g$ lies in $G_2$ (the center of $G$) then this map is moreover a $2$-translation. It turns out in this case that these are the only elements of $\Aut_1(X)$ and $\Aut_2(X)$.
However, if we replace $\Gamma$ by
\[
\Gamma' = \left\{ \heis{x}{y}{z} \colon x, y \in 2 \ZZ, \, z \in \ZZ \right\}
\]
then the map
\[
\heis{x}{y}{z} \Gamma' \mapsto \heis{x}{y+1}{z+x}\Gamma'=\heis{x}{y}{z}\heis{0}{1}{0}\Gamma'
\]
is a $1$-translation not of the form $x \Gamma' \mapsto g x \Gamma'$. However, the \emph{connected component of the identity} in $\Aut_1(G/\Gamma')$ still consists entirely of maps $x \Gamma' \mapsto g x \Gamma'$.
The full automorphism group $\Aut(X)$ is rather complicated (again, one can argue that it contains a copy of $\SL_2(\ZZ)$) and will not concern us.
\end{example}
\begin{remark}
\label{rem:generic-nilmanifold-auts}
Generalizing the previous example, for any filtered nilmanifold $G /\Gamma$ it turns out that maps $x \Gamma \mapsto g x \Gamma$ are always in $\Aut_k(X)$ provided $g \in G_k$. As we saw above, these need not be all the elements of $\Aut_k(G/\Gamma)$, but they will yield all of the connected component of the identity in $\Aut_k(G/\Gamma)$.
This is of course encouraging for the approach of using $\Aut_1(X)$ as a proxy for $G$.
\end{remark}
The definition of $k$-translation we have given is in some sense the most natural, but can be cumbersome in practice, because of the need to consider rather general configurations, and in particular cubes of arbitrarily large dimension.
The following proposition establishes a convenient alternative definition that is equivalent under certain circumstances.
\begin{proposition}
\label{prop:k-translation-equiv}
Suppose $X$ is a \uppar{compact, ergodic} nilspace of degree $s$, and fix $k$, $0 \le k \le s+1$. Then $\phi \in \Homeo(X)$ is a $k$-translation if and only if for any $c \in C^{s+1-k}(X)$ the configuration $\llcorner^k(c; \phi(c))$ is an $(s+1)$-cube.
\end{proposition}
Note this condition is a special case of that from Definition \ref{def:k-translation}, specializing to $n=s+1$ and $c = \square^k(c')$ for some $c' \in C^{s+1-k}(X)$. Hence, the ``only if'' direction is clear; the content is that it suffices to check the definition on configurations of this form.
Before we give the proof we recall some facts and terminology from \cite{GMV1}.
We say that a cubespace obeys the {\em glueing axiom} if the following holds:
for any triple $c_1,c_2,c_3$ of cubes, $[c_1,c_2]$ and $[c_2,c_3]$
being cubes imply that $[c_1,c_3]$ is a cube, also.
This can be visualized as the cubes $[c_1,c_2]$ and $[c_2,c_3]$ glued
along the common face $c_2$.
We recall from \cite{GMV1}*{Proposition 6.2}
that nilspaces obey the glueing axiom.
Readers unfamiliar with these ideas are advised to consult
\cite{GMV1}*{Section 6.1}.
\begin{proof}
We first recall \cite{GMV1}*{Proposition 3.11}
that in a nilspace of degree $s$, a configuration $\{0,1\}^n \to X$ for $n \ge s+1$ is a cube, if and only if every face of dimension $(s+1)$ is a cube.
Our first objective is to prove that it is enough to prove the condition from Definition \ref{def:k-translation} for cubes $c$ of dimension $s+1$.
Let $F$ be a face of $\{0,1\}^n$ of codimension $k$ and $c \in C^n(X)$ for some $n \ge s+1$. Every face of $[\phi]_F(c)$ of dimension $(s+1)$ has the form $[\phi]_{F'}(c')$ where $c'$ is a face of $c$ and $F'$ is a face of $\{0,1\}^{s+1}$ of codimension at most $k$. So, the condition from Definition \ref{def:k-translation} for $n=s+1$ implies all cases $n \ge s+1$.
If $n < s+1$ and $c \in C^n(X)$, we note that the duplicated configuration $\square^{s+1-n}([\phi]_F(c))$ has the form
$[\phi]_{F'}(\square^{s+1-n}(c))$ required by Definition \ref{def:k-translation}, where $F' = \square^{s+1-n}(F)$ (abusing notation somewhat) is a corresponding face of $\{0,1\}^{s+1}$ of codimension $k$, and so this configuration lies in $C^{s+1}(X)$.
Restricting to an appropriate face, we recover that $[\phi]_F(c) \in C^n(X)$. So, the condition for $n=s+1$ implies the condition for all $n$.
We now have to show that it suffices to consider cubes of the form $\square^k(c')$ for $c' \in C^{s+1-k}(X)$. This is by a ``glueing argument'', which is very much related to the ``universal replacement property'' of the canonical equivalence relation in \cite{GMV1}*{Proposition 6.3}.
There we prove that in a fibrant cubespace $Y$ the fact that $\llcorner^k(y;y')$ is a cube implies the following: if $c\in C^k(Y)$ is such that $c(\vec 1)=y$, then the configuration $c'$ defined by $c'(\omega)=c(\omega)$ for $\omega\neq\vec1$ and $c'(\vec1)=y'$, is also a cube. A full proof of this is given in the second claim in the proof of Proposition 7.12 in \cite{GMV1}.
One can turn that into a proof of our claim by replacing vertices by cubes of dimension $s+1-k$ in a straightforward manner.
Here we only give some diagrams for the case $s=1$ and $k=2$ for the reader's convenience. The general case is very similar but notationally awkward.
Let $c \in C^{s+1}(X) = C^2(X)$. We draw this as
\inlinetikz{
\singlesquare{$c(00)$}{$c(10)$}{$c(01)$}{$c(11)$}
}
and we wish to show that
\inlinetikz{
\singlesquare{$c(00)$}{$c(10)$}{$c(01)$}{$\phi(c(11))$}
}
is a cube. Our hypothesis states that for every $a\in C^{s+1-k}(X) = C^0(X)=X$ the following configuration is an $(s+1)$-cube, i.e.~a $2$-cube:
\inlinetikz{
\singlesquare{$a$}{$a$}{$a$}{$\phi(a)$}
}
We consider
\inlinetikz{
\begin{scope}[scale=1.5]
\draw[thin,black] (0,0) grid (2, 2);
\node [below left] at (0, 0) {$c(00)$};
\node [above left] at (0, 2) {$c(01)$};
\node [below right] at (2, 0) {$c(10)$};
\node [above right] at (2, 2) {$\phi(c(11))$};
\node [left] at (0, 1) {$c(01)$};
\node [right] at (2, 1) {$c(11)$};
\node [below] at (1, 0) {$c(10)$};
\node [above] at (1, 2) {$c(11)$};
\node [above right] at (1, 1) {$c(11)$};
\end{scope}
}%
and note that the upper right square is a cube by this hypothesis, the remaining small squares are cubes by the cubespace axioms, and hence the outer square is a cube by glueing (twice) as required.
\end{proof}
Recalling Remark \ref{rem:generic-nilmanifold-auts}, the connected component of the identity in $\Aut_k(X)$ may be a better object to work with than $\Aut_k(X)$ itself. We consider this object briefly now.
\begin{definition}
For any topological group $G$, we let $G^\circ \le G$ denote the connected component of the identity in $G$. In particular, $\Aut_k^\circ(X)$ denotes the connected component of the identity in the group of $k$-translations.
\end{definition}
\begin{remark}
Note that the connected component is taken in $\Aut_k(X)$; \emph{not} in the larger group $\Aut(X)$ or $\Aut_1(X)$. Indeed, these need not agree in general, i.e.~we may have $\Aut_k^\circ(X) \subsetneq \Aut_k(X) \cap \Aut_1^\circ(X)$.
\end{remark}
We should really check that this revised sequence of groups $\Aut_1^\circ(X) \supseteq \Aut_2^\circ(X) \supseteq \dots$ is still well-behaved.
\begin{proposition}
Let $X$ be a compact cubespace. Then the sequence $(\Aut_k^\circ(X))_{k \ge 0}$ is still a filtration of closed groups, in the sense of Proposition \ref{translation-prefiltration}.
\end{proposition}
\begin{proof}
Since $\Aut_{k+1}^\circ(X)$ is a connected subset of $\Aut_k(X)$ containing the identity, it is contained in $\Aut_k^\circ(X)$. Certainly all these groups are closed. Since the commutator map is continuous, the set
\[
\{ [g,h] \colon g \in \Aut_i^\circ(X), g \in \Aut_j^\circ(X) \}
\]
is a connected subset of $\Aut_{i+j}(X)$, so is contained in $\Aut_{i+j}^\circ(X)$, and hence so is the closed subgroup it generates.
\end{proof}
\subsection{A discrete subgroup and the evaluation map}
Having defined the group $\Aut_1^\circ(X)$ that we hope will take the role of $G$ in the construction of the nilmanifold $G / \Gamma$, we now consider the question of how $\Gamma$, and the isomorphism $X \leftrightarrow G / \Gamma$, will arise.
Suppose we fix an element $x_0 \in X$. There is a natural map
\begin{align*}
\ev_{x_0} \colon \Aut_1(X) &\to X \\
\phi &\mapsto \phi(x_0)
\end{align*}
and this gives rise to an identification of the orbit of $x$ under $\Aut_1(X)$ with $\Aut_1(X) / \stab(x_0)$ where $\stab(x) = \{ \phi \in \Aut_1(X) \colon \phi(x) = x \}$ denotes the stabilizer. Clearly the same story makes sense restricting to $\Aut_1^\circ(X)$. For notational convenience we denote by $\stab(x)$ also $\{ \phi \in \Aut_1^\circ(X) \colon \phi(x) = x \}$. The meaning of $\stab(x)$ is always clear from the context.
Suppose we knew that $\stab(x_0)$ were a discrete and co-compact subgroup of $\Aut_1^\circ(X)$, and also that the action of $\Aut_1^\circ(X)$ on $X$ were transitive. Then we would have a homeomorphism $X \cong \Aut_1^\circ(X) / \stab(x_0)$, which is already enough to identify $X$ -- as a topological space -- with a nilmanifold.
However, this transitivity assumption is a very significant one. So far, we have made no progress towards even showing that $\Aut_1(X)$ is non-trivial. Showing that $1$-translations (and more generally, $k$-translations) are fairly abundant, is both the core and the hardest aspect of the whole argument.
If the identification of $X$ with $\Aut_1^\circ(X) / \stab(x_0)$ should hold true in the category of cubespaces, we would need to know that $C^n(X)$ were identified under this bijection with the Host--Kra cubes $\HK^n(\Aut^\circ_\bullet(X)) / \stab(x_0)$ where here $\stab(x_0)=\{ \phi \in \HK^n(\Aut^\circ_\bullet(X)) \colon \phi(\square^n(x_0)) = \square^n(x_0) \}$. Unwrapping the definitions, this says that a configuration $c \colon \{0,1\}^n \to X$ is a cube if and only if it has the form $\omega \mapsto \phi_\omega . x_0$ for some $(\phi_\omega) \in \HK^k(\Aut_\bullet^\circ(X))$.
It is certainly not hard to show the ``if'' direction (see also footnote 6).
\begin{proposition}
\label{prop:evaluation-is-morphism}
Suppose $(\phi_\omega)_{\omega \in \{0,1\}^n} \in \HK^n(\Aut_\bullet^\circ(X))$ is an element of the Host--Kra cube group, and $c \in C^n(X)$ is a cube. Then the configuration
\[
\omega \mapsto (\phi_\omega(c(\omega)))
\]
is a cube in $C^n(X)$.
\end{proposition}
In particular, the map $\Aut_1^\circ(X) / \stab(x_0) \to X$ is a cubespace morphism.
\begin{proof}
The Host--Kra cube group is generated (by definition) by elements $[\phi]_F$ where $F \subseteq \{0,1\}^n$ has codimension $k$ and $\phi \in \Aut_k^\circ(X)$. By definition of $\Aut_k$, applying any such configuration pointwise to a cube $c \in C^n(X)$ yields another cube in $C^n(X)$. The result follows from repeated application of this fact.
\end{proof}
However, the ``only if'' direction is not clear. Indeed, the case $n = 0$ corresponds to the statement that the action of $\Aut_1^\circ(X)$ on $X$ is transitive; for higher $n$, this statement is some kind of assertion of ``higher transitivity'' or ``transitivity on all levels''. Proving this is strongly analogous to proving transitivity, although one needs an analogous strengthening of the hypotheses.
\subsection{Structural results}
\label{statements-subsec}
We are now in a position to state more precise versions of the main structural result (Theorem \ref{main-thm-simple}) whose proof will occupy us for the rest of this article.
The first is a very slight strengthening of Theorem \ref{thm:main-toral}, using the different connectivity hypothesis explained above. Part of our proof was reported in \cite{Gut_Oberwolfach}.
\begin{theorem}
\label{thm:main-sc}
Let $X = (X, C^n(X))$ be a \uppar{compact, ergodic} nilspace of degree $s$. Suppose that all the structure groups $A_k(X)$ are Lie groups. Finally, suppose $X$ is strongly connected.
Fix an element $x_0 \in X$. Then the following hold.
\begin{enumerate}
\item The group $\Aut_1^\circ(X)$ is Lie, and the filtration $(\Aut_k^\circ(X))_{k \ge 1}$ is Lie and
has degree at most $s$.
\item The subgroup $\stab(x_0) \le \Aut_1^\circ(X)$ is discrete and co-compact in $\Aut_1^\circ(X)$. Moreover, $\stab(x_0) \cap \Aut_k^\circ(X)$ is co-compact in $\Aut_k^\circ(X)$ for all $k\geq 1$.
\item The natural map
\begin{align*}
\Aut_1^\circ(X) &\to X \\
\phi &\mapsto \phi(x_0)
\end{align*}
gives rise to an isomorphism of cubespaces $\Aut_1^\circ(X) / \stab(x_0) \to X$. That is, this map is a homeomorphism, and $C^n(X)$ is identified with the Host--Kra cubes $\HK^n(\Aut_\bullet^\circ(X)) / \stab(x_0)$ on the nilmanifold $\Aut_1^\circ(X) / \stab(x_0)$.
\end{enumerate}
One can summarize these conclusions more briefly by saying that $X$ is a nilmanifold of degree at most $s$.
\end{theorem}
By Proposition \ref{prop-toral-sc}, this implies Theorem \ref{thm:main-toral}.
As mentioned above, it is instructive to record some of what our method provides when the connectivity hypothesis is not assumed.
\begin{theorem}
\label{weaker-main-thm}
Let $X = (X, C^n(X))$ be a compact ergodic nilspace of degree $s$, whose structure groups are Lie groups. Then $X$ has finitely many connected components, each of which is open in $X$, and each of which is homeomorphic to a nilmanifold.
Furthermore, for any $x_0 \in X$, the natural evaluation map
\begin{align*}
\ev_{x_0} \colon \Aut_1^\circ(X) &\to X \\
\phi &\mapsto \phi(x_0)
\end{align*}
induces an homeomorphism $S \cong \Aut_1^\circ(X) / \stab(x_0)$, where $\Aut_1^\circ(X)$ is a nilpotent Lie group of degree at most $s$, $\stab(x_0)$ is a discrete and co-compact subgroup and $S$ denotes the connected component of $x_0$ in $X$.
Finally, if $\Aut_1(X)$ is known \emph{a priori} to act transitively on the set of connected components of $X$, then there is a stronger identification \uppar{of topological spaces} $X \cong \Aut_1(X) / \stab(x_0)$ for any $x_0 \in X$, where again $\Aut_1(X)$ is a nilpotent Lie group of degree at most $s$ and $\stab(x_0)$ is a discrete co-compact subgroup.
\end{theorem}
Note this result provides information only about the topological structure of the nilspace $X$.
We are able to say something about the cubespace structure, as well, but we postpone this
discussion.
See Lemma \ref{lem:cube-eval-open} below.
The primary interest for the last statement of the above theorem comes from the case of dynamical systems, where this condition arises naturally. In particular we record the following corollary.
\begin{corollary}\label{cor-weaker-main-thm}
Let $(G,X)$ be a minimal topological dynamical system, where $X$ is a \uppar{compact, ergodic} nilspace of degree $s$, whose structure groups $A_k(X)$ are Lie groups; and suppose $G$ acts on $X$ through a continuous group homomorphism $\sigma \colon G \to \Aut_1(X)$.
Then $X$ is homeomorphic to a nilmanifold as in the previous theorem: that is, $X \cong \Aut_1(X) / \stab(x_0)$ for any $x_0 \in X$. Under this identification, $G$ acts by left translations $x \Gamma \mapsto \sigma(g) x \Gamma$.
\end{corollary}
\begin{proof}
Since $(G,X)$ is minimal, one may map any element of $X$ into any open set by an element of $G$. But the connected components of $X$ are open by the theorem, so we conclude that the group of translations $\Aut_1(X)$ acts transitively on the set of connected components of $X$.
\end{proof}
To motivate the above result further, we mention that for any minimal group action
$(G,X)$, one can associate a natural cubespace structure $\{C_G^{n}\}_{n\in\ZZ_{\ge 0}}$
such that the acting group $G$ immerses into $\Aut_1(X,C_G)$.
This construction is due to Host, Kra and Maass \cite{HKM10}.
Given a minimal group action $(G,X)$, Gutman, Glasner and Ye \cite{GGY} defined the regional proximal
relation $\RP_G^n(X)$, which is a closed equivalence relation on $X$ for each $n\in\ZZ_{\ge 0}$, and extends the definition of
Host, Kra and Maass \cite{HKM10} for actions of $\ZZ$ to general group actions.
It is proved in \cite{GGY} that the cubespace structure $(X,C_G)$ is a nilspace
of degree at most $s$ for some $s\in\ZZ_{\ge 0}$ if and only if $\RP_G^s(X)$ is trivial.
For a detailed exposition and further results we refer to the companion paper \cite{GMV3} and to \cite{GGY}.
Based on these results, the following is a special case of Corollary \ref{cor-weaker-main-thm}.
\begin{corollary}\label{cor:even-more-weaker}
Let $(G,X)$ be a minimal topological dynamical system.
Suppose that the regional proximal relation $\RP_G^s(X)$ is trivial for some $s\in\ZZ_{\ge 0}$
and the structure groups of $(X,C_G)$ are Lie groups.
Then $(G,X)$ is a nilsystem, that is, $X$ can be identified with a nilmanifold $N/\Gamma$, where $N$
is a nilpotent Lie group of step at most $s$ and the action of $G$ can be realized through a continuous homomorphism
$G\to N$.
\end{corollary}
We note that the hypothesis on the structure groups can be replaced by some assumptions on the topology of $X$
similarly to Theorem \ref{main-thm-simple}, and therefore, the corollary can be stated without any reference to the cubespace structure.
However, the proof of that result requires the structure theory for general nilspaces developed in \cite{GMV3}, and we refer
the interested reader to the Appendix of that paper.
\section{The main steps in the structural result}
\label{sec:main-outline}
We continue our outline of the proof of the main structure theorems of this paper
following \cite{CS12}.
From the discussion above, it is clear that our main task is to show that $X$ has ``enough $k$-translations'' in some sense. At this stage, it is not particularly obvious that even a single non-trivial $k$-translation exists for any $k$, and so we first need to find a source of these automorphisms.
The source is ultimately based on the weak structure theory expounded in \cite{GMV1}. We recall that if $X$ is a (compact, ergodic) nilspace of degree $s$, then there is a \emph{canonical factor}
\[
\pi_{s-1} \colon X \to \pi_{s-1}(X)
\]
where $\pi_{s-1}(X)$ is a (compact, ergodic) nilspace of degree at most $(s-1)$, and the fibers of $\pi_{s-1}$ are identified with the $s$-th structure group $A_s(X)$ of $X$, which is some compact abelian group. More precisely, there is an action of $A_s(X)$ on the whole space $X$, whose orbits are precisely these fibers. For full statements, see \cite{GMV1}*{Theorem 5.4}.
The key observation is the following.
\begin{proposition}
\label{prop:structure-group-translations}
Let $X$ be a compact ergodic nilspace of degree $s$, and let $a \in A_s(X)$. Write $t_a \colon X \to X$ for the action $x \mapsto a . x$ of $A_s$ on $X$ described above.
Then $t_a \in \Aut_s(X)$.
\end{proposition}
\begin{proof}
After unwrapping the definitions, this is immediate from the weak structure theory. Indeed, what we need -- that applying a fixed element $a \in A_s$ to a face of $c \in C^n(X)$ of codimension $s$ yields another cube -- is a special case of
\cite{GMV1}*{Theorem 5.4(ii)}.
\end{proof}
So, although we still cannot prove that $\Aut_1(X)$ acts transitively, we at least know that it acts transitively on each fiber of $\pi_{s-1}$ (and in fact this is true even of $\Aut_s(X)$).
The second idea is then to induct on the degree. Suppose we had already established the relevant transitivity statement for nilspaces of degree $(s-1)$; so in particular we knew that $\Aut_1^\circ(\pi_{s-1}(X))$ acts transitively on $\pi_{s-1}(X)$ (given suitable assumptions). Then we might hope to achieve transitivity on $X$ by first moving to the correct fiber of $\pi_{s-1}$ (by inductive assumption), and then moving to the correct point in that fiber using an element of $A_s$.
Of course, the challenge for this plan is that there is no obvious way to relate $\Aut_1(\pi_{s-1}(X))$ to $\Aut_1(X)$. Specifically, to make this approach work we will need to be able to \emph{lift} a $1$-translation of $\pi_{s-1}(X)$ to one of $X$, under suitable hypotheses; i.e.~given $\phi \in \Aut_1(\pi_{s-1}(X))$, to find $\tilde{\phi} \in \Aut_1(X)$ such that the diagram
\[
\begin{CD}
X @>\tilde{\phi}>> X \\
@VV\pi_{s-1}V @VV\pi_{s-1}V \\
\pi_{s-1}(X) @>\phi>> \pi_{s-1}(X)
\end{CD}
\]
commutes.
Going in the other direction -- i.e.~that given $\tilde{\phi} \in \Aut_k(X)$ we can push it down to $\phi \in \Aut_k(\pi_{s-1}(X))$ -- is straightforward, and we verify this now.
\begin{proposition}
Let $X$ be a compact ergodic nilspace of degree $s$, and fix $k \ge 0$. Then there is a canonical \uppar{continuous} group homomorphism $\pi_\ast \colon \Aut_k(X) \to \Aut_k(\pi_{s-1}(X))$ such that
\[
\begin{CD}
X @>\phi>> X \\
@VV\pi_{s-1}V @VV\pi_{s-1}V \\
\pi_{s-1}(X) @>\pi_\ast(\phi)>> \pi_{s-1}(X)
\end{CD}
\]
commutes for every $\phi \in \Aut_k(X)$.
\end{proposition}
The proof is completely mechanical.
\begin{proof}
Note that if $k > s$ the result is trivial (Proposition \ref{prop-translations-terminate}), so we assume $k \le s$. We do the only thing possible: given $y \in \pi_{s-1}(X)$, we pick $x \in \pi_{s-1}^{-1}(y)$ arbitrarily, and define $(\pi_\ast(\phi))(y) := \pi_{s-1}(\phi(x))$. It suffices to check that this is well-defined and a $k$-translation; continuity is then straightforward to check.
If $x' \in \pi_{s-1}^{-1}(y)$ is another lift of $y$ then there is some $a \in A_s$ such that $x' = a . x$ (see \cite{GMV1}*{Theorem 5.4(i)}).
But $t_a$ is an element of $\Aut_s(X)$ (Proposition \ref{prop:structure-group-translations}), and this commutes with $\phi$ (Proposition \ref{prop-translations-terminate} again, and Proposition \ref{translation-prefiltration}). So, $\phi(x') = a . \phi(x)$, and so $\pi_{s-1}(\phi(x')) = \pi_{s-1}(a . \phi(x)) = \pi_{s-1}(\phi(x))$, since $t_a$ preserves fibers of $\pi_{s-1}$. This shows well-definedness.
Now suppose $c \in C^n(\pi_{s-1}(X))$ and $F \subseteq \{0,1\}^n$ a face of codimension $k$ are given. By definition, there is a cube $\tilde{c} \in C^n(X)$ such that $\pi_{s-1}(\tilde{c}) = c$. So,
\[
[\pi_\ast(\phi)]_F . c = \pi_{s-1}([\phi]_F . \tilde{c})
\]
and the right hand side is a cube as required, since $\phi$ is a $k$-translation and $\pi_{s-1}$ is a cubespace morphism.
\end{proof}
The statement for identity components follows trivially from this, i.e.~$\pi_\ast(\Aut_k^\circ(X)) \subseteq \Aut_k^\circ(\pi_{s-1}(X))$.
The ``translation lifting'' statement we require is precisely that this group homomorphism is surjective. This is not quite true in general. However, it is the case that any \emph{sufficiently small} element of $\Aut_k(\pi_{s-1}(X))$ can be lifted to a $k$-translation on $X$, i.e.~lies in the image of $\pi_\ast$. In fact we will show the following.
\begin{proposition}
\label{lem:surj}
Let $X$ be a \uppar{compact, ergodic} nilspace of degree $s$ whose structure groups $A_r(X)$ are all Lie. For any $k \ge 1$, the homomorphism $\pi_\ast \colon \Aut_k(X) \to \Aut_k(\pi_{s-1}(X))$ is an open map.
\end{proposition}
In particular, the image of $\pi_\ast$ is an open subgroup of $\Aut_k(\pi_{s-1}(X))$, meaning exactly that all sufficiently small translations on $\pi_{s-1}(X)$ lift to $X$. More precisely: there exists $\delta > 0$ (depending only on $X$ and $k$) such that if $\phi \in \Aut_k(\pi_{s-1}(X))$ is a ``perturbation of the identity'' in the sense that $\|\phi\| \le \delta$, then there is a lift $\tilde{\phi} \in \Aut_k(X)$ such that $\pi_\ast(\tilde{\phi}) = \phi$, i.e.~$\pi_{s-1} \circ \tilde{\phi} = \phi \circ \pi_{s-1}$.
Proposition \ref{lem:surj} is the heart of the argument, and will occupy us for the majority of the rest of the paper. We will return to its proof in Section \ref{sec-main-details}.
\bigskip
For the remainder of this section, we will fill in the outstanding gaps in the deduction of Theorem \ref{thm:main-sc}, and the other results stated in Section \ref{statements-subsec}, from Proposition \ref{lem:surj}.
Some of this work is concerned with making the previous discussion rigorous, and also extending all mention of transitivity of the action of $\Aut_1(X)$ to the ``higher transitivity'' required to get the corresponding statement about cubes.
A large part of the rest is of a strongly topological nature: we will have to check that certain groups are Lie groups, certain maps are open maps and so forth. To some extent, this material is technical and could be skipped on first reading. However, we caution that it cannot easily be separated from the rest of the argument: without this topological input, we could not prove even a much weakened version of Theorem \ref{thm:main-sc}.
Here and later in the paper we will have to draw on a couple of fairly powerful results about topological groups: Gleason's lemma (Theorem \ref{gleason-lemma}) concerning principal bundles, and the Gleason--Kuranishi extension theorem (Theorem \ref{thm:gleason-extension}) that asserts that an extension of a Lie group by a Lie group is Lie.
However, we remark that we will not use the full power of either of these. Specifically, Gleason's lemma is only required in the compact abelian Lie setting, where the proof is essentially an application of Fourier analysis. For the Gleason--Kuranishi extension theorem, we require only the case of central extensions. Both of these facts are given a self-contained treatment in \cite{T14} which is substantially easier than the original references. See Remark \ref{remark-tao} for more details.
\subsection{The stabilizer is discrete}
\label{subsec-stab-discrete}
To construct our nilmanifold $G/\Gamma$, recall we plan to use $G = \Aut_1^\circ(X)$ and $\Gamma = \stab(x_0)$ for some $x_0 \in X$. So, we will need to know that this $\Gamma$ is discrete.
\begin{lemma}
\label{stab-discrete}
Let $X$ be a \uppar{compact, ergodic} nilspace of degree $s$ whose structure groups $A_r(X)$ are Lie, and fix any $x_0 \in X$. Then $\stab(x_0) \subseteq \Aut_1(X)$ is discrete.
\end{lemma}
Equivalently (since $\Aut_1(X)$ carries the supremum metric) we wish to show that if $\phi$ is a $1$-translation other than the identity and $\phi(x_0) = x_0$ then $\|\phi\| = \sup_{x \in X} d(x, \phi(x)) \ge \delta$ for some constant $\delta$ independent of $\phi$ (and, in fact, $x_0$).
\begin{proof}[Proof of Lemma \ref{stab-discrete}]
We proceed by induction on $s$. The case $s=0$ is a triviality (since a $0$-step compact ergodic nilspace is just the one-point space $\{\ast\}$).
Again we consider $\pi_\ast \colon \Aut_1(X) \to \Aut_1(\pi_{s-1}(X))$. It is clear from the definition that $\pi_\ast$ maps $\stab(x_0)$ into $\stab(\pi_{s-1}(x_0))$. By inductive hypotheses, $\stab(\pi_{s-1}(x_0))$ is discrete; so we conclude that if $\phi \in \Aut_1(X)$ and $\|\phi\|$ is sufficiently small, then $\phi \in \ker(\pi_\ast)$. Equivalently, $y$ and $\phi(y)$ lie in the same fiber for all $y \in X$.
By the weak structure theory, for each $y \in X$ there is an unique $a \in A_s$ such that $\phi(y) = a . y$. Hence there is a function $\tau \colon X \to A_s$ such that $\phi(y) = \tau(y) . y$. It is clear $\tau$ is continuous. Since $\phi$ commutes with the action of $A_s \subseteq \Aut_s(X)$ (by Proposition \ref{prop:structure-group-translations} and Proposition \ref{translation-prefiltration}), in fact $\tau$ factors through $\pi_{s-1}$, i.e.~as a map $X \to \pi_{s-1}(X) \to A_s$. By abuse of notation we write $\tau$ for this map $\pi_{s-1}(X) \to A_s$.
We now want to unwrap the fact that $\phi$ is a $1$-translation to a condition on $\tau$. By Proposition \ref{prop:k-translation-equiv}, this is equivalent to saying that $[c, \tau(c) . c] \in C^{s+1}(X)$ for every $c = C^s(X)$ (where $\tau(c)$ denotes pointwise application). By the weak structure theorem \cite{GMV1}*{Theorem 5.4},
this is equivalent to saying that
\[
[\square^s(0),\tau(c)] \in C^{s+1}(\cD_s(A_s(X)))
\]
for every $c \in C^n(X)$.
We recall that $\cD_s(A_s)$ is the cubespace whose base space is the group $A_s$, and whose cubes are $\HK^n(A_s)$ where $A_s$ is given the degree $s$ filtration $A_s = \dots = A_s \supseteq \{0\}$. We also recall (\cite{GMV1}*{Proposition 5.1})
that a configuration $a \colon \{0,1\}^{s+1} \to A_s$ lies in $C^{s+1}(\cD_s(A_s))$ if and only if
\[
\sum_{\omega \in \{0,1\}^{s+1}} (-1)^{|\omega|} a(\omega) = 0 \ .
\]
It follows that $\phi$ is a $1$-translation if and only if the function
\begin{align*}
\rho \colon C^s(\pi_{s-1}(X)) &\to A_s \\
c &\mapsto \sum_{\omega \in \{0,1\}^s} (-1)^{|\omega|} \tau(c(\omega))
\end{align*}
is identically zero.
We will see in Section \ref{sec:cocycle} that this is a natural statement in terms of the rudimentary cohomology theory expounded there. Moreover, if $\|\phi\|$ is sufficiently small (i.e.~less than some absolute constant), then $\tau$ also takes values close to $0$ in $A_s$ (i.e.~$d(0, \tau(y)) \le \eps$ for any fixed small $\eps$ and all $y$). Under these conditions, a rigidity result, Theorem \ref{thm:cocycle-uniqueness}, states that $\tau$ must in fact be constant.
Since $\phi(x_0) = x_0$ and hence $\tau(x_0) = 0$ by assumption, we must therefore have that $\tau$ is identically zero. But then $\phi = \id$, as required.
\end{proof}
\subsection{The group \texorpdfstring{$\Aut_1(X)$}{Aut1(X)} is Lie}
Recall that we wish to use $\Aut_1^{\circ}(X)$ as the group $G$ in the definition of a nilmanifold $G / \Gamma$.
Hence it will be important to verify that this is a Lie group.
In fact, we will prove that $\Aut_1(X)$ is a Lie group, which implies that $\Aut_1^{\circ}(X)$ is also a Lie group.
Once again we stress that our definitions do not require a Lie group to be connected.
\begin{lemma}
\label{lemma-auts-lie}
Let $X = (X, C^n(X))$ be a \uppar{compact, ergodic} nilspace of degree $s$, whose structure groups $A_k(X)$ are Lie. Then $\Aut_1(X)$ is a Lie group.
\end{lemma}
Our approach will be to apply induction on $s$, with the case $s=0$ a trivial base case. We again consider the map
\[
\pi_{\ast} \colon \Aut_1(X) \to \Aut_1(\pi_{s-1}(X)) \ .
\]
Since $\pi_{s-1}(X)$ is a compact ergodic nilspace of degree at most $(s-1)$, by inductive hypothesis $\Aut_1(\pi_{s-1}(X))$ is Lie.
By Proposition \ref{lem:surj}, the image of $\pi_\ast$ is an open subgroup, and so is also Lie.
We will use the following result.
\begin{theorem}[{Gleason--Kuranishi extension theorem}]
\footnote{This is Theorem 3.1 of \cite{G51}. The Kuranishi extension theorem
usually refers to the same statement where it is assumed in addition
that $G$ is locally compact (see \cite{I_MR}). As pointed out by
Gleason, Kuranishi actually proved a weaker statement in \cite{K50}.}
\label{thm:gleason-extension}
Let $G$ be a topological group, and suppose there exists a closed normal subgroup $N$ of $G$ such that $N$ and $G/N$ are Lie groups. Then $G$ is also Lie.
\end{theorem}
Given this, it suffices to verify that $\ker(\pi_\ast) \le \Aut_1(X)$ is a Lie group. We know that the copy of $A_s(X)$ in $\Aut_1(X)$ (acting by $t_a \colon x \mapsto a . x$) is contained in $\ker(\pi_\ast)$, since it acts on each fiber of $\pi_{s-1}$. Hence it will suffice to show:
\begin{lemma}
The subgroup $A_s(X) \le \ker(\pi_\ast)$ is open.
\end{lemma}
\begin{proof}
Suppose $\phi \in \ker(\pi_\ast) \setminus A_s$. Fix any point $y \in X$. Since $y$ and $\phi(y)$ lie in the same fiber of $\pi_{s-1}$, there is an unique $a \in A_s(X)$ such that $y = a . \phi(y)$. As before write $t_a$ for the translation $x \mapsto a . x$ on $X$, and define $\phi'(y) = t_a \circ \phi(y)$; hence, $\phi'(x) = x$ and so $\phi' \in \stab(x) \subseteq \Aut_1(X)$.
By Lemma \ref{stab-discrete}, $\|\phi'\| \ge \delta$ for some constant $\delta$ independent of $\phi$. If $\|t_a\| \le \delta / 2$, then $\|\phi\| \ge \delta - \delta / 2 = \delta / 2$ and we are done. If not, i.e.~$\|t_a\| > \delta/2$, then since $A_s(X)$ is compact and acts freely on $X$, we have that $d(x, t_a(x))$ is bounded \emph{below} by a constant depending only on $\delta/2$, and hence $\|\phi\|\ge d(x,t_a(x))$ is bounded below as required.
\end{proof}
We noted above (Remark \ref{rem:second-countable-auts}) that $\Aut(X)$ is second countable, and hence clearly so is the closed subgroup $\ker(\pi_\ast)$.
Moreover, $A_s(X)$ is an open subgroup of $\ker(\pi_\ast)$ that is Lie, and so by standard results in Lie theory we may extend the differentiable structure on $A_s(X)$ to one on all of $\ker(\pi_\ast)$, making the latter into a Lie group.
This completes the proof of Lemma \ref{lemma-auts-lie}.
\begin{remark}\label{remark-tao}
Noting that the copy of $A_s(X)$ in $\Aut_1(X)$ is central, we have only really used the fact that a central extension of a Lie group by a Lie group is Lie, and that a \emph{discrete} extension of a Lie group is Lie. The latter fact is straightforward given elementary Lie theory (e.g., assuming something of equivalent strength to Cartan's closed subgroup theorem). Hence we may rely on the proof in \cite{T14}*{Theorem 2.6.1} rather than \cite{G51}.
In fact we know rather more, namely that the middle group $\Aut_1(X)$ in the exact sequence $0\to \ker(\pi_\ast) \to \Aut_1(X) \to \Aut_1(\pi_{s-1}(X))\to 0$ is nilpotent. It is likely one can obtain a yet more direct and elementary proof of Theorem \ref{thm:gleason-extension} in this special case, but we have not pursued this.
\end{remark}
\subsection{Completing the proof of Theorem \ref{thm:main-sc}}
We now have all the ingredients in place to deduce Theorem \ref{thm:main-sc}, as well as the other results stated in Section \ref{statements-subsec}.
Recall that we wished to know that $\Aut_1(X)$ acts transitively on $X$, under suitable hypotheses. Again it turns out the most natural statement of this is in terms of openness of some map.
\begin{lemma}
\label{lem:open-evaluation-map}
Let $X$ be a \uppar{compact, ergodic} nilspace of degree $s$ whose structure groups $A_r(X)$ are Lie, and let $x_0 \in X$ be fixed. Then the evaluation map
\begin{align*}
\ev_{x_0} \colon \Aut_1(X) &\to X \\
\phi &\mapsto \phi(x_0)
\end{align*}
is continuous and open.
\end{lemma}
Equivalently, this states that if two points $x, y \in X$ are very close together, then there is a small $1$-translation $\phi$ such that $\phi(x) = y$.
For inductive reasons we will verify the following slight strengthening.
\begin{lemma}
\label{lem:inductive-open-evaluation-map}
Let $X$ be as in Lemma \ref{lem:open-evaluation-map}. Let $0 \le k < s$ be fixed. For all $\eps > 0$ there exists $\delta > 0$ such that the following holds: if $x, y \in X$, $d(x,y) \le \delta$ and $\pi_k(x) = \pi_k(y)$, then there exists $\phi \in \Aut_{k+1}(X)$, $\|\phi\| \le \eps$ such that $\phi(x) = y$.
\end{lemma}
\begin{proof}
Essentially this just combines Proposition \ref{prop:structure-group-translations} with Proposition \ref{lem:surj}. We proceed by induction on $s$, and note that the case $s=0$ is trivial.
If $k = s-1$, then the result is immediate from Proposition \ref{prop:structure-group-translations}: we know $\pi_{s-1}(x) = \pi_{s-1}(y)$, so we can set $\phi = t_a$ for the unique $a \in A_s$ such that $a.x = y$, and so $\phi \in \Aut_s(X)$ as required. The bound on $\|t_a\|$ follows from Proposition \ref{prop:robust-free-action}.
If $k < s-1$, we apply induction on $s$ to obtain $\tilde{\phi} \in \Aut_{k+1}(\pi_{s-1}(X))$ such that $\tilde{\phi}(\pi_{s-1}(x)) = \pi_{s-1}(y)$; and $\|\tilde{\phi}\|$ can be made arbitrarily small if $d(x,y)$ is small enough.
By Proposition \ref{lem:surj}, we may obtain a lift $\phi' \in \Aut_{k+1}(X)$ of $\tilde{\phi}$ (i.e.~such that $\pi_\ast(\phi') = \tilde{\phi}$) where again $\|\phi'\|$ can made as small as we like.
We now know that $\pi_{s-1}(\phi'(x)) = \pi_{s-1}(y)$. Finally, we note that $d(\phi'(x), y) \le d(x,y) + \|\phi'\|$ which is still as small as we like, so we again apply the case $k=s-1$ to obtain $\psi \in \Aut_{s}(X)$ such that $\psi(\phi'(x)) = y$. So, $\phi := \psi \circ \phi'$ is in $\Aut_{k+1}(X)$ and is still arbitrarily small.
\end{proof}
\begin{proof}[{Proof of Lemma \ref{lem:open-evaluation-map}}]
Continuity is immediate by the choice of metric on $\Aut_1(X)$. Since $\pi_0(X) = \{\ast\}$, openness follows immediately from the case $k=0$ of Lemma \ref{lem:inductive-open-evaluation-map}.
\end{proof}
Our final conclusions are phrased in terms of the identity component $\Aut_1^\circ(X)$. It is fairly painless to deduce useful facts about this from what we have already shown about $\Aut_1(X)$.
\begin{corollary}
Let $X$ be as in the statement of Lemma \ref{lem:open-evaluation-map}. Then the identity component $\Aut_1^\circ(X)$ acts transitively on each connected component of $X$.
Moreover, there are only finitely many connected components of $X$, and each one is open and closed in $X$.
\end{corollary}
\begin{proof}
Since $\Aut_1(X)$ is Lie (Lemma \ref{lemma-auts-lie}) its identity component $\Aut_1^\circ(X)$ is open. Therefore by Lemma \ref{lem:open-evaluation-map}, the orbits of $\Aut_1^\circ(X)$ in $X$ are open, and hence closed (since the orbits partition the space). It follows that any connected subset of $X$ is contained in a single orbit.
Since each orbit of $\Aut_1^\circ(X)$ is the image of a connected set under a continuous map and hence connected, we further deduce that every point of $X$ has a connected open neighbourhood, and so all connected components are open. Since they partition the space, they are therefore closed; and since $X$ is compact, it follows there are only finitely many.
\end{proof}
Combined with Lemma \ref{lemma-auts-lie} and Lemma \ref{stab-discrete}, this is enough to complete the proof of Theorem \ref{weaker-main-thm}.
\bigskip
We now turn to the outstanding statements from Theorem \ref{thm:main-sc}. Essentially, we still have said nothing about the cubes $C^n(X)$ in terms of the filtration $\Aut_k(X)$. To do so, we need to generalize several of the arguments above that concerned $1$-translations acting on $X$, to statements about the Host--Kra cube group $\HK^n(\Aut_\bullet(X))$ acting on $C^n(X)$.
For instance, the following generalizes Lemma \ref{lem:open-evaluation-map}.
\begin{lemma}
\label{lem:cube-eval-open}
Let $X$ be as in Lemma \ref{lem:open-evaluation-map}, and fix $c \in C^n(X)$. Then the evaluation map
\begin{align*}
\ev_c \colon \HK^n(\Aut_\bullet(X)) &\to C^n(X) \\
(\phi_\omega)_{\omega \in \{0,1\}^n} &\mapsto \left(\omega \mapsto \phi_\omega(c(\omega)) \right)
\end{align*}
is open.
\end{lemma}
\begin{remark}
Recall that this definition makes sense by virtue of Proposition \ref{prop:evaluation-is-morphism}.
\end{remark}
Note this lemma (as well as the previous ones) requires no connectivity assumptions.
Hence this lemma provides a lot of information about cubes in not necessarily strongly connected nilspaces.
In particular it implies that all sufficiently small cubes can be obtained from a constant cube using the action
of the Host Kra cubegroups of $\Aut_\bullet(X)$.
On the other hand, we have no information about large cubes, this is why we need the connectivity
hypotheses in the main results.
The missing ingredients that prevent us from simply repeating the proof of Lemma \ref{lem:open-evaluation-map} are all concerned with topological facts about Host--Kra cube groups. Everything we will need follows cleanly in turn from the algebraic theory of $\HK^n(G_\bullet)$ expounded in the appendix of \cite{GMV1}.
We recall these results now.
\begin{lemma}[\cite{GMV1}*{Lemma A.12}]\label{lm:hk-homomorphisms}
Let $G_\bullet$, $H_\bullet$ be two filtered topological groups.
Let $\tau:G\to H$ be a homomorphism such that
$\tau(G_i)\subseteq\tau(H_i)$.
Then $\tau$ induces a homomorphism $\tau: \HK^n(G_\bullet)\to\HK^n(H_\bullet)$
for each $n$ by pointwise application on the vertices.
If $\tau:G_i\to H_i$ is open for each $i$, then so is the induced homomorphism
$\tau: \HK^n(G_\bullet)\to\HK^n(H_\bullet)$
for each $n$.
\end{lemma}
\begin{lemma}[\cite{GMV1}*{Lemma A.13}]\label{lm:hk-subgroups}
Let $G_\bullet$, $H_\bullet$ be two filtered topological groups.
Suppose $G_i\subseteq H_i$ for each $i$.
If $G_i$ are open in $H_i$ (resp., connected) for each $i$, then
$\HK^n(G_\bullet)$ is also open in $\HK^n(H_\bullet)$ (resp., connected) for each $n$.
\end{lemma}
We note the following consequence.
\begin{lemma}
\label{cube-surj-lemma}
Let $X$ be a \uppar{compact, ergodic} nilspace of degree $s$ whose structure groups $A_r(X)$ are Lie groups. Let $\pi_{s-1} \colon X \to \pi_{s-1}(X)$ denote the canonical projection, and $\pi_\ast \colon \Aut_k(X) \to \Aut_k(\pi_{s-1}(X))$ the canonical homomorphism described previously.
Then
\[
\pi_\ast \colon \HK^n(\Aut_\bullet(X)) \to \HK^n(\Aut_\bullet(\pi_{s-1}(X)))
\]
is an open map.
\end{lemma}
\begin{proof}
This follows from Proposition \ref{lem:surj} and Lemma \ref{lm:hk-homomorphisms}.
\end{proof}
\begin{proof}[Proof of Lemma \ref{lem:cube-eval-open}]
We follow the proof of Lemma \ref{lem:open-evaluation-map} very closely; in fact, that result corresponds precisely to the case $n=0$ of this. As ever, we induct on $s$, the degree of $X$, the case $s=0$ being trivial.
It suffices to show that for any $\eps > 0$ there exists a $\delta > 0$ such that if $c_1, c_2 \in C^n(X)$ and $d(c_1,c_2) \le \delta$ then there exists $\phi \in \HK^n(\Aut_\bullet(X))$, $\|\phi\| \le \eps$ such that $\phi(c_1) = c_2$.
By inductive hypothesis, the map $\ev_{\pi_{s-1}(c)} \colon \HK^n(\Aut_\bullet(\pi_{s-1}(X))) \to C^n(\pi_{s-1}(X))$ is open, and by Lemma \ref{cube-surj-lemma} so is $\pi_\ast \colon \HK^n(\Aut_\bullet(X)) \to \HK^n(\Aut_\bullet(\pi_{s-1}(X)))$. As before, we consider the composite and deduce that, for suitably chosen $\delta$, there exists $\phi' \in \HK^n(\Aut_\bullet(X))$, $\|\phi'\| \le \eta$ such that $\pi_{s-1}(\phi'(c_1)) = \pi_{s-1}(c_2)$, where $0 < \eta \le \eps / 2$ is some parameter to be determined.
Now, we have that $d(\phi'(c_1), c_2) \le \delta + \eta$ and moreover these lie in the same fiber of $\pi_{s-1}$. But we recall that the fibers of $C^n(X) \to C^n(\pi_{s-1}(X))$ are completely described by the weak structure theory (\cite{GMV1}*{Theorem 5.4(ii)}):
specifically, $C^n(\cD_s(A_s(X)))$ acts simply transitively on each fiber. We again use $t_a$ to denote this action, and crucially observe that it corresponds to an element of $\HK^n(\Aut_\bullet(X))$ (since $C^n(\cD_s(A_s(X)))$ is defined as the Host--Kra group on $A_s(X)$ with the degree $s$ filtration, which is a sub-filtration of $\Aut_\bullet(X)$).
So, there is an unique $a \in C^n(\cD_s(A_s(X)))$ such that $t_a(\phi'(c_1)) = c_2$. By Proposition \ref{prop:robust-free-action}, for suitably chosen $\eta$ and $\delta$ we may ensure that $\|t_a\| \le \eps / 2$. Hence taking $\phi = t_a \circ \phi'$ we have $\phi(x) = y$ and $\|\phi\| \le \eps$ as required.
\end{proof}
Finally we can conclude the proof of Theorem \ref{thm:main-sc}.
\begin{proof}[Proof of what is left of Theorem \ref{thm:main-sc}]
What is left to prove is that $\HK^n(\Aut_\bullet^\circ(X))$ acts transitively on $C^n(X)$, using in particular that $X$ is strongly connected.
Since $\Aut_k(X)$ is a closed subgroup of $\Aut_1(X)$, it is Lie (by Cartan's closed subgroup theorem) and hence $\Aut_k^\circ(X)$ is open in $\Aut_k(X)$. By Lemma \ref{lm:hk-subgroups}, it follows that $\HK^n(\Aut_\bullet^\circ(X))$ is an open subgroup of $\HK^n(\Aut_\bullet(X))$. Hence, by Lemma \ref{lem:cube-eval-open}, the orbits of the action of this group on $C^n(X)$ are open, and hence closed (as they partition the space). But since $C^n(X)$ was assumed to be connected, this means the action is transitive, as required.
Given this, it follows that the map $\ev_{\square^n(x_0)} \colon \HK^n(\Aut_\bullet^\circ(X)) \to C^n(X)$ given by $(\phi_\omega)_{\omega \in \{0,1\}^n} \mapsto \left(\omega \mapsto \phi_\omega(x_0) \right)$ is continuous, open and surjective, and so induces an homeomorphism $\HK^n(\Aut_\bullet^\circ(X))/\stab(x_0)\cong C^n(X)$ for each $n\geq 0$. In particular when $n=0$ we have that $\Aut_1^\circ(X)/\stab(x_0)\cong X$.
Finally, we must check that $\stab(x_0) \cap \Aut_k^\circ(X)$ is co-compact in $\Aut_k^\circ(X)$ for all $k \ge 1$. By Lemma \ref{lem:inductive-open-evaluation-map}, the orbits of the action of $\Aut_k^\circ(X)$ on the closed equivalence class $\pi_{k-1}^{-1}(\pi_{k-1}(x_0))$ are open (noting $\Aut_k^\circ(X)$ is open in $\Aut_k(X)$ as above). Hence these orbits are also closed (as they partition the space), and therefore compact. If $S$ is the orbit containing $x_0$, then the restricted evaluation map $\ev_{x_0} \colon \Aut_k(X)^\circ \to S$ is continuous, open and surjective and so induces an homeomorphism $\Aut_k^\circ(X) / (\stab(x_0) \cap \Aut_k^\circ(X)) \cong S$, and so the left hand side is compact as required.
\end{proof}
\section{Enough \texorpdfstring{$k$}{k}-translations}
\label{sec-main-details}
We now return to the proof of Proposition \ref{lem:surj}, the ``translation lifting'' statement.
We recall the set-up. We have a (compact, ergodic) nilspace $X = (X, C^n(X))$ of degree $s$, and are considering the canonical factor map $\pi_{s-1} \colon X \to \pi_{s-1}(X)$. For notational brevity, we write $\pi$ for $\pi_{s-1}$
and $\pi(X)$ for $\pi_{s-1}(X)$, for the remainder of this section. We are assuming in particular that the top structure group $A_s(X)$ is a Lie group.
We wish to show the following: for every $\eps > 0$ there exists $\delta > 0$ such that for any $\phi \in \Aut_k(\pi(X))$ with $\|\phi\| \le \delta$, there exists $\tilde{\phi} \in \Aut_k(X)$ with $\|\tilde{\phi}\| \le \eps$ such that $\pi \circ \tilde{\phi} = \phi \circ \pi$.
The strategy is roughly as follows.
\begin{enumerate}[label=(\arabic*)]
\item We first seek \emph{some function} $\psi \colon X \to X$ such that $\pi \circ \psi = \phi \circ \pi$, and which behaves nicely with respect to the action of $A_s(X)$. We will also be able to choose $\psi$ to be small. Crucially, though, it may \emph{not} be a $k$-translation on $X$.
\item We then argue that we can ``repair'' $\psi$ to a genuine $k$-translation $\tilde{\phi}$. Since necessarily $\pi \circ \psi = \pi \circ \tilde{\phi}$, this amounts to finding a function $f \colon X \to A_s$ from $X$ to the structure group $A_s$, and setting $\tilde{\phi}(x) = f(x) . \psi(x)$ (where as usual this denotes the action of $A_s(X)$ on $X$).
\end{enumerate}
Here we deviate from the approach taken in \cite{CS12}.
Both proofs rely on the cocycle theory developed in \cite{CS12}
and expounded in Section \ref{sec:cocycle}.
(This is step (2) in the above programme.)
However the lift of small translations are constructed
in different ways. The main difference is that in \cite{CS12} a measurable, but
not necessarily continuous lift is constructed first, while our argument stays in
the continuous category.
We recall that $A_s(X)$ acts freely on $X$, with orbits precisely the fibers of $\pi$. The standard name for this set-up is that $X \to \pi(X)$ is a \emph{principal bundle} or $A_s$-principal bundle. However, this will not typically mean that $X$ is homeomorphic as a topological space to $A_s \times \pi(X)$, i.e.~this bundle need not be the trivial bundle. In our setting, it turns out that this bundle is nonetheless \emph{locally trivial} in the sense that it locally resembles such a direct product. This result is due to Gleason:
\begin{theorem}[{\cite{G50}*{Theorem 3.3}}]
\label{gleason-lemma}
Suppose $A$ is a compact Lie group acting freely on a completely regular topological space $Y$. Let $\tau \colon Y \to Y / A$ denote the quotient map. Then for every point $x \in Y / A$ there is a neighbourhood $U$ of $x$ and a local section $\sigma \colon U \to \tau^{-1}(U)$, i.e.~a continuous map such that $\tau \circ \sigma = \id_U$.
\end{theorem}
Although this is a significant result, we remark that the proof simplifies somewhat when the acting group $A$ is abelian, which is the only case we use.
The next result is the key ingredient for part (1) of our argument. It can be extracted from the proof of the first
covering homotopy theorem (\cite{S51}*{Theorem 11.3}), but we include a proof for completeness.
\begin{lemma}
\label{lem-bundle-lift}
Let $X$ be a \uppar{compact, ergodic} nilspace of degree $s$ whose structure group $A_s(X)$ is Lie. For all $\eps > 0$, there exists $\delta > 0$ such that the following holds. Given any homeomorphism $f \colon \pi(X) \to \pi(X)$ with $\|f\| \le \delta$, there exists a homeomorphism $f' \colon X \to X$ such that:
\begin{enumerate}
\item $\|f'\| \le \eps$,
\item $f'$ is a lift of $f$, i.e.~$\pi \circ f' = f \circ \pi$, and
\item $f'$ is a \emph{bundle map}, meaning $f'(a . x) = a . f'(x)$ for all $a \in A_s(X)$.
\end{enumerate}
\end{lemma}
\begin{proof}
Let $\eps > 0$ be fixed.
By Theorem \ref{gleason-lemma}, we may choose an open cover $\{U_i\}_{i=1}^m$ of $\pi(X)$ and a family of local sections $\sigma_i \colon U_i \to \pi^{-1}(U_i)$. We may assume that $\sigma_i$ are uniformly continuous, and so for any fixed $\eta > 0$ (to be specified later) we may refine the cover if necessary so that $d(\sigma_i(x), \sigma_i(x')) \le \eta$ for all $i$ and $x,x' \in U_i$.
We now choose $\delta$ to be a Lebesgue number for the cover $U_i$; i.e.~for all $x \in \pi(X)$ there is an $i$ such that $\{y \colon d(x,y) \le \delta\} \subseteq U_i$. Write $U'_i$ for the set of all $x$ such that $\{y \colon d(x,y) \le \delta\} \subseteq U_i$; this is another open cover of $\pi(X)$.
We would like to define $f'$ as follows. Given $x \in \pi^{-1}(U'_i)$, there is an unique element $a_i(x) \in A_s(X)$ such that $x = a_i(x) . \sigma_i(\pi(x))$.
(If we use the local section $\sigma_i$ to identify $\pi^{-1}(U'_i)$ with $U_i'\times A_s(X)$,
then $a_i$ is simply the projection to the $A_s(X)$ component.)
Then we set for $x\in U_i'$
\[
f'_i(x) = a_i(x) . \sigma_i(f(\pi(x))) \ .
\]
Note that $d(\pi(x),f(\pi(x)))\le\delta$, hence $f(\pi(x))\in U_i$ for all $x\in U_i'$.
Note that $d(\sigma_i(\pi(x)), \sigma_i(f(\pi(x)))) \le \eta$ by assumption, and by uniform continuity of the action of $A_s(X)$, we may choose $\eta$ so that
\[
d\left(a\, .\, \sigma_i(\pi(x)), a\, .\, \sigma_i(f(\pi(x)))\right) \le \eps' / 2
\]
for all $a \in A_s(X)$, where $\eps'$ is some constant to be determined; and so in particular when $a = a_i(x)$ we have
\[
d(x, f'_i(x)) \le \eps'/2 \ .
\]
Properties (ii) and (iii) for $f'_i$ are clear by construction.
Unfortunately we are not done: $f'_i$ is only defined on $U'_i$, and the functions for different $i$ are not compatible. So, we have not constructed a function on all of $X$ with the desired properties.
To fix this, we will take an average of the values $f_i'(x)$ at each point $x$, in a sense to be made precise. We first choose a continuous partition of unity $\nu_i \colon \pi(X) \to \RR_{\ge 0}$ adapted to $U'_i$; that is,
\[
\sum_{i=1}^m \nu_i(y) = 1
\]
for all $y \in \pi(X)$, and $\nu_i$ is supported on $U'_i$ for each $i$.
We now define
\[
f'(x) = \sum_{i=1}^m \nu_i(\pi(x)) f'_i(x) \ .
\]
Strictly speaking, this definition makes no sense whatsoever: $f_i'(x)$ is an element of $X$, on which neither addition nor multiplication by real numbers is legitimate. We now justify what is actually meant by it, as follows.
\begin{itemize}
\item For fixed $x$ all the points $\{ f'_i(x) \colon 1 \le i \le m,\, \pi(x) \in U_i \}$ lie in the same fiber of $\pi$, and hence are related to each other by the action of elements of $A_s(X)$. To fix notation, let $i_0$ be such that $\pi(x) \in U'_{i_0}$, and write $\alpha_i(x) \in A_s(X)$ for the unique element such that $\alpha_i(x) . f'_{i_0}(x) = f'_i(x)$ for each $i$ such that $\pi(x)\in U'_i$.
\item Since $d(f'_i(x), f'_j(x)) \le \eps'$ for all $i,j$, we find that the elements $\alpha_i(x) \in A_s(X)$ may be assumed to be small with respect to our preferred metric on $A_s(X)$ (by Proposition \ref{prop:robust-free-action}).
\item Write $A_s(X) \cong (\RR/\ZZ)^d \times K$ for some $d \ge 0$ and some finite group $K$. The $\alpha_i$ all lie in some small neighbourhood of the identity, which in turn is locally isomorphic to a small ball in $\RR^d \times \{0\}$, for small enough $\eps'$. Write $\tilde{\alpha_i}(x) \in \RR^d$ for the lift of $\alpha_i(x)$.
\item It now makes sense to define
\[
\tilde{\alpha}(x) = \sum_{i=1}^m \nu_i(x) \tilde{\alpha_i}(x) \in \RR^d
\]
which also lies in the same small neighbourhood of the identity; and then to let $\alpha(x) \in (\RR/\ZZ)^d \times \{0\}$ be the projection in $A_s(X)$. We let
\[
f'(x) = \alpha(x) . f'_1(x)
\]
for all $x$.
\item Finally, we remark that this process was independent of the choice of $i_0$, and hence $f'$ defines a function on all of $X$.
\end{itemize}
In short, we are doing a kind of integration of functions with values in a compact abelian Lie group $A_s(X)$, or more precisely on the space $\pi^{-1}(f(x))$ which is identified with $A_s(X)$, and claiming this makes sense whenever all the points being averaged are sufficiently close together. We will need to repeat this kind of process in what follows, and a further discussion of closely related issues appears in Section \ref{subsec:integration}.
We conclude by arguing that $f'$ as defined has the required properties. Since $f'(x)$ lies in the same fiber as each $f'_i(x)$ by construction, it is clear that (ii) holds. Since the elements $\alpha_i(x)$ and hence $\alpha(x)$ are all arbitrarily close to $0 \in A_s(X)$ provided $\eps'$ is sufficiently small, we can ensure that $d(f'_{i_0}(x), \alpha(x). f'_{i_0}(x)) \le \eps/2$ and hence (i) holds, given previous discussion. It is also not hard to argue that $\alpha_i(x)$ and $\alpha(x)$ are continuous functions of $x$ for each choice of $i_0$, and hence $f'$ is locally continuous and therefore continuous.
Property (iii) holds because this integration process commutes with translation by $A_s(X)$; indeed, it is clear from the definitions that $\alpha_i(x) = \alpha_i(a.x)$ for any $a \in A_s(X)$. It is a consequence of (ii) and (iii) that $f'$ is a bijection and hence a homeomorphism.
\end{proof}
We have now completed part (1) of our strategy: we write $\psi \colon X \to X$ for the function $f'$ returned by applying Lemma \ref{lem-bundle-lift} to $\phi$. To recap, we know that
\begin{itemize}
\item $\pi \circ \psi = \phi \circ \pi$;
\item $\|\psi\| \rightarrow 0$ as $\|\phi\| \rightarrow 0$; and
\item $\psi$ commutes with $A_s(X)$, i.e.~$\psi \circ t_a = t_a \circ \psi$ for all $a \in A_s(X)$.
\end{itemize}
We now wish to attack part (2); that is, ``fixing'' $\psi$ to make it into a $k$-translation.
Just as in the proof of Lemma \ref{stab-discrete}, it is helpful to unwrap what it would mean for a function $\psi$ with these properties to be a $k$-translation on $X$. More accurately, we want a condition that measures the \emph{failure} of $\psi$ to be a $k$-translation. The hope is then to construct a correction to this failure, and thereby construct $\tilde{\phi}$.
In the interests of generality, let $\chi \colon X \to X$ denote an arbitrary homeomorphism such that $\pi \circ \chi(x) = \phi \circ \pi(x)$ for all $x \in X$.
By Proposition \ref{prop:k-translation-equiv}, we have that $\chi$ is a $k$-translation if and only if for every $c \in C^{s+1-k}(X)$ we have that $\llcorner^k(c; \chi(c)) \in C^{s+1}(X)$.
By assumption, the projection $\pi(\llcorner^k(c; \chi(c)))$ is a cube of $\pi(X)$. Hence -- by the weak structure theory -- the obstructions to this configuration being a cube in $X$ can be expressed in terms of the structure group $A_s$. We recall the following precise statement of this.
\begin{proposition}
\label{structure-thm-facts}
Let $c \in C^{s+1}(X)$ be some cube, and $c' \colon \{0,1\}^{s+1} \to X$ a configuration such that $\pi(c) = \pi(c')$. Let $\alpha \colon \{0,1\}^{s+1} \to A_s(X)$ denote the unique configuration such that $\alpha\, .\, c = c'$.
Then $c' \in C^{s+1}(X)$ if and only if $\alpha \in C^{s+1}(\cD_s(A_s(X)))$. This holds if and only if the alternating sum
\[
\sum_{\omega \in \{0,1\}^{s+1}} (-1)^{|\omega|} \alpha(\omega)
\]
is zero.
\end{proposition}
\begin{proof}
See \cite{GMV1}*{Theorem 5.4 and Proposition 5.1}.
\end{proof}
We term this single element of $A_s(X)$ the \emph{discrepancy} of the configuration, which we now define formally.
\begin{definition}
Let $c' \colon \{0,1\}^{s+1} \to X$ be a configuration such that $\pi(c') \in C^{s+1}(\pi(X))$. Let $c$ be some element of $C^{s+1}(X)$ such that $\pi(c) = \pi(c')$ (which always exists
by the definition of quotient cubespaces -- see \cite{GMV1}*{Definition 5.2}).
As above, define $\alpha \colon \{0,1\}^{s+1} \to A_s(X)$ to be the unique configuration such that $\alpha \,.\, c = c'$. Then we define the \emph{discrepancy} $\Delta(c')$ of $c'$ by
\[
\Delta(c') = \sum_{\omega \in \{0,1\}^{s+1}} (-1)^{|\omega|} \alpha(\omega) \ .
\]
\end{definition}
\begin{proposition}
\label{prop:discrepancy}
The discrepancy is well-defined, and $\Delta(c') = 0$ if and only if $c' \in C^{s+1}(X)$.
\end{proposition}
\begin{proof}
The second statement is taken directly from Proposition \ref{structure-thm-facts}. To verify well-definedness, note that if $c_1, c_2 \in C^{s+1}(X)$ are cubes with $\pi(c_1) = \pi(c_2) = \pi(c')$ and $\alpha_1, \alpha_2 \colon \{0,1\}^{s+1} \to A_s(X)$ satisfy $\alpha_1 \,.\, c_1 = \alpha_2 \,.\,c_2 = c'$, then $(\alpha_1 - \alpha_2). c_1 = c_2$ and invoking the Proposition again we find that
\[
\sum_{\omega \in \{0,1\}^{s+1}} (-1)^{|\omega|} (\alpha_1(\omega) - \alpha_2(\omega)) = 0
\]
as required.
\end{proof}
We note one more elementary fact about discrepancies.
\begin{proposition}
\label{prop-discrepancy-additive}
The discrepancy is \emph{additive} in the following sense: if $c' = [c_0, c_1]$ and $c'' = [c_1, c_2]$ are configurations such that $\pi(c'), \pi(c'') \in C^{s+1}(\pi(X))$ are cubes, then $\pi([c_0,c_2]) \in C^{s+1}(\pi(X))$ and $\Delta([c_0, c_2]) = \Delta(c') + \Delta(c'')$.
\end{proposition}
\begin{proof}
The first fact is immediate from glueing (see \cite{GMV1}*{Proposition 6.2}).
The discrepancy identity is clear from considering the alternating sums.
\end{proof}
We now return to considering our function $\chi$. With the above discussion in mind, we define a function
\begin{align*}
\rho_\chi \colon C^{s+1-k}(X) &\to A_s(X) \\
c &\mapsto \Delta(\llcorner^k(c; \chi(c)))
\end{align*}
i.e.~the discrepancy of the configuration $\llcorner^k(c; \chi(c))$ that arose while characterizing $k$-translations.
Our observations can be summarized by the following result.
\begin{lemma}
Given $\chi$ as above, we have that $\chi$ is a $k$-translation if and only if $\rho_\chi$ is identically zero.
\end{lemma}
\begin{proof}
This is immediate from Proposition \ref{prop:k-translation-equiv}, Proposition \ref{prop:discrepancy} and Proposition \ref{structure-thm-facts}.
\end{proof}
So, the failure of $\chi$ to be a $k$-translation is entirely captured by the function $\rho_\chi$. Now suppose that we take our original function $\psi \colon X \to X$ and attempt to repair it by setting
\[
\tilde{\phi}(x) = f(x) \,.\, \psi(x)
\]
for some function $f \colon X \to A_s(X)$, as suggested originally. Then by the lemma, we will have succeeded if and only if $\rho_{\tilde{\phi}} \equiv 0$, which in turn holds if and only if for every $c \in C^{s+1-k}(X)$,
\[
\rho_\psi(c) + \sum_{\omega \in \{0,1\}^{s+1-k}} (-1)^{|\omega| + k} f(c(\omega)) = 0 \ .
\]
The crucial point is now the following. The function $\rho_\psi$ is some kind of ``cocycle'': a function on cubes with certain properties that we will describe. The above equation states that $\rho_\psi$ is some kind of ``coboundary'', i.e.~equal to the ``derivative''
\[
\pm \sum_{\omega \in \{0,1\}^{s+1-k}} (-1)^{|\omega|} f(c(\omega)) \ .
\]
So, what we want is to show some kind of ``cohomological triviality'': that \emph{every} cocycle $\rho_\psi$ will be a coboundary, and therefore $\psi$ can be corrected to a $k$-translation.
This will not be always true. However, it \emph{is} true provided $\rho_\psi$ is ``sufficiently small'' in the sense of its image being contained in a small ball in $A_s(X)$.\footnote{An equivalent statement is that the ``cohomology in $\RR^d$'' is trivial, or that the ``cohomology group is discrete'' in an appropriate sense. We will not make these statements rigorous, though one could certainly do so.}
It remains only to define what we mean by a ``cocycle'' and ``coboundary'', and to state the cohomological triviality result that we will prove in the next section.
\begin{definition}
\label{def:cocycle}
Let $\ell \ge 0$ be an integer, $X$ a cubespace and $A$ an abelian group. By an \emph{$\ell$-cocycle} on $X$ with values in $A$, we mean a continuous function
\[
\rho \colon C^\ell(X) \to A
\]
satisfying the following conditions:
\begin{enumerate}
\item (additivity) if $c' = [c_0, c_1]$, $c'' = [c_1, c_2]$ and $c = [c_0, c_2]$ are all $\ell$-cubes then $\rho(c) = \rho(c') + \rho(c'')$;
\item (reflections) if $c = [c_0, c_1]$ then $\rho(c) = -\rho([c_1, c_0])$;
\item (degenerate cubes) if $c = [c_0, c_0]$ then $\rho(c) = 0$.
\end{enumerate}
We say $\rho$ is a \emph{coboundary} if there exists a continuous function $f \colon X \to A$ such that
\[
\rho(c) = \partial^\ell f(c) := \sum_{\omega \in \{0,1\}^\ell} (-1)^{|\omega|} f(c(\omega)) \ .
\]
\end{definition}
\begin{remark}\
\label{rem:cocycle-defs}
\begin{itemize}
\item A $0$-cocycle is the same thing as a continuous function $X \to A$.
\item We stress that our rather vague notation allows the concatenation operation $[-,-]$ to occur on any coordinate $\{1, \dots, \ell\}$, not just the last one. For instance, for a $2$-cococyle $\rho$ it is the case that
\[
\rho([[a,b],[c,d]]) = -\rho([[c,d],[a,b]]) = -\rho([[b,a],[d,c]]) = \rho([[d,c],[b,a]]) \ .
\]
\item Note that properties (i), (ii), (iii) are not logically independent, i.e.~our definition is not minimal. Indeed, item (iii) is an immediate consequence of item (i) with the substitution $c_0=c_1=c_2=c_0$. Then item (ii) can also be deduced by taking $c_2=c_0$ in (i) and using (iii).
\item As well as taking the ``derivative'' $\partial^\ell$ of a function on $X$, one may analogously take the derivative of a function of cubes; e.g.~given $\rho \colon C^n(X) \to A$ we may write
\[
\partial \rho([c, c']) = \rho(c) - \rho(c') \ .
\]
This definition does now depend implicitly on which coordinate $\{1,\dots,n+1\}$ is used for the concatenation; we will introduce more precise conventions below when this is necessary to avoid confusion.
\item It is trivial to verify that any coboundary is indeed a cocycle. More generally, if $\rho$ is an $\ell$-cocycle then $\partial^k \rho$ is an $(\ell+k)$-cocycle.
\end{itemize}
\end{remark}
\begin{lemma}
For any $\chi$ having the properties discussed above, the function $\rho_\chi$ is indeed a cocycle.
\end{lemma}
\begin{proof}
Additivity follows directly from the additivity of discrepancies, proven in Proposition \ref{prop-discrepancy-additive}. The other two properties follow from this.
\end{proof}
In summary, at long last, it will suffice to prove the following fact, which is a special case of Theorem \ref{thm:cocycle-main}.
\begin{theorem} \cite{CS12}*{Lemma 3.19}
\label{thm:cocycle-special}
Suppose $X$ is a compact ergodic nilspace of degree $s$, $A$ is a compact abelian Lie group
(equipped with a metric $d_A$),
$\ell \ge 0$ is an integer and $\rho \colon C^\ell(X) \to A$ is an $\ell$-cocycle.
Then there exists $\eps = \eps(s, \ell, A) > 0$, depending only on $s$, $\ell$ and $A$ \uppar{but not on $X$}, such that the following holds. Suppose that $\delta \le \eps$ and $d_A(\rho(c), \rho(c')) \le \delta$ for all $c, c' \in C^\ell(X)$. Then there exists $f \colon X \to A$ such that $\rho = \partial^\ell f$ and also $d_A(f(x), f(x')) \lesssim_{s,\ell} \delta$ for all $x, x' \in X$.
\end{theorem}
We now bring everything together to conclude the proof of Proposition \ref{lem:surj}. If the original map $\phi$ has $\|\phi\|$ small enough, then $\|\psi\|$ is also arbitrarily small. Hence, the cocycle $\rho_\psi$ can be taken to be as small as we like with respect to the metric
\[
\sup_{c, c' \in C^\ell(X)} d_{A_s(X)}(\rho_\psi(c), \rho_\psi(c')) \ ;
\]
indeed, if $\psi$ is small then $\llcorner^k(c; \psi(c))$ is close to a genuine cube $\square^k(c)$, and since the discrepancy map $\Delta$ is continuous (where here we have used Proposition \ref{prop:robust-free-action}) it follows that $\rho_\psi(c)$ is close to $0$ for every $c$.
Hence, the correction $f$ from Theorem \ref{thm:cocycle-special} can be chosen to be arbitrarily small, meaning the final $k$-translation $\tilde{\phi} \in \Aut_k(X)$ that we obtain is also arbitrarily small, as required.
\section{Cocycle theory}
\label{sec:cocycle}
The purpose of this section is to prove Theorem \ref{thm:cocycle-special} concerning triviality of small cocycles.
It will be necessary elsewhere in this project to have a version of this result in greater generality than as stated therein. The proof of this more general version is very similar to a direct proof of the specialized version; the
only difference is that we need to draw on the \emph{relative} weak structure theory, as expounded in \cite{GMV1}*{Section 7},
which generalizes (again in a fairly routine way) the absolute version from \cite{CS12} (or \cite{GMV1}*{Section 6}).
The technical generalization is as follows.
The reader is advised to recall the notion of a fibration, and other related terminology, from
\cite{GMV1}*{Section 7}.
\begin{theorem}
\label{thm:cocycle-main}
Let $A$ be a compact abelian Lie group (equipped with a metric $d_A$), and let $s \ge 0$, $\ell \ge 1$ be given. Then there exists $\eps = \eps(s, \ell, A) > 0$ such that the following holds.
Let $\beta \colon X \to Y$ be any fibration of degree $s$ between compact ergodic cubespaces $X$ and $Y$ that obey the glueing axiom.
Now let $\rho$ be a continuous $\ell$-cocycle on $X$ with values in $A$, let $0 < \delta \le \eps$ be given and suppose that $d(\rho(c), \rho(c')) \le \delta$ whenever $\beta(c) = \beta(c')$.
Then $\rho = \partial^\ell f + \tilde{\rho} \circ \beta$, where $f \colon X \to A$ is continuous and satisfies $d(f(x), f(y)) \lesssim_{s,\ell} \delta$ \uppar{that is, there exists a constant $c=c(s,\ell)>0$ such that $d(f(x), f(y))\leq c\delta$} whenever $\beta(x) = \beta(y)$, and $\tilde{\rho}$ is a continuous $\ell$-cocycle on $Y$.
\end{theorem}
(Recall $\partial^\ell$ was defined in Definition \ref{def:cocycle}.)
We briefly pause to explain why this does in fact generalize Theorem \ref{thm:cocycle-special}.
Letting $X$, $A$ and $\rho$ be as in the statement of Theorem \ref{thm:cocycle-special}, we may invoke Theorem \ref{thm:cocycle-main} where:
\begin{itemize}
\item $X$ is still the same (compact, ergodic, degree $s$) nilspace $X$;
\item $Y$ is $\{\ast\}$, the one-point space;
\item $\beta \colon X \to \{\ast\}$ is the trivial map; and
\item the cocycle $\rho$ is unchanged.
\end{itemize}
The statement that $\beta$ is a fibration of finite degree is precisely saying that $X$ is a nilspace of finite degree. Hence, all the hypotheses of Theorem \ref{thm:cocycle-main} in this special case follow from those of Theorem \ref{thm:cocycle-special}. Given the conclusion of Theorem \ref{thm:cocycle-main}, note that the only $\ell$-cocycle $C^\ell(\{\ast\}) \to A$ is the zero function, and so $\tilde{\rho} = 0$ and we recover the conclusion of Theorem \ref{thm:cocycle-special}.
Even though Theorem \ref{thm:cocycle-main} does indeed generalize Theorem \ref{thm:cocycle-special}, since its proof will involve some additional notational and technical difficulties, we will continue to provide sketches of the argument in the special case of Theorem \ref{thm:cocycle-special}. The hope is that the intuition will be easier to grasp in this model setting.
\bigskip
Note that, in Section \ref{sec:main-outline}, we needed a kind of dual to Theorem \ref{thm:cocycle-special}, stating that the function $f \colon X \to A$ that is used to construct the coboundary $\partial^\ell f$ is itself essentially uniquely determined. The precise statement is the following.
\begin{theorem}
\label{thm:cocycle-uniqueness}
Let $A$ be a compact abelian Lie group, equipped with the metric $d_A$, and let $s \ge 0$, $\ell \ge 1$ be given. Then there exists $\eps = \eps(s, \ell, A) > 0$ such that the following holds.
Suppose $X = (X, C^n(X))$ is a compact ergodic nilspace of degree $s$, and $f \colon X \to A$ is a continuous function such that $d_A(f(x), f(x')) \le \eps$ for all $x, x' \in X$, and $\partial^\ell f \colon C^\ell(X) \to A$ is the zero function.
Then $f$ is constant.
\end{theorem}
Again, we would like a version that holds in the setting of Theorem \ref{thm:cocycle-main}, concluding that the function $f \colon X \to A$ obtained there is essentially unique. It turns out that in this case, this generalization follows easily from the original version.
\begin{corollary}
\label{cor:general-uniqueness}
Suppose $A$, $\beta$, $X$, $Y$ are as in the statement of Theorem \ref{thm:cocycle-main}. Then there exists $\eps = \eps(A) > 0$ such that the following holds.
Suppose $\ell \ge 0$ is an integer, and $f \colon X\to A$ is a continuous function such that $d_A(f(x), f(x')) \le \eps$ for all $x, x' \in X$ and furthermore $\partial^\ell f$ is constant on fibers of $\beta$ \uppar{meaning if $c, c' \in C^\ell(X)$ and $\beta(c) = \beta(c')$ then $\partial^\ell f(c) = \partial^\ell f(c')$}.
Then $f$ is also constant on fibers of $\beta$, i.e.~$f = \tilde{f} \circ \beta$ for some continuous $\tilde{f} \colon Y \to A$.
\end{corollary}
\begin{proof}[{Proof of Corollary \ref{cor:general-uniqueness} from Theorem \ref{thm:cocycle-uniqueness}}]
Given an $y \in Y$, its preimage $Z := f^{-1}(y) \subseteq X$ is a nilspace of degree $s$. Moreover, by hypothesis $\partial^\ell f$ is constant on $C^\ell (Z)$, and therefore zero on $C^\ell(Z)$, since clearly $\partial^\ell f(\square^\ell(z)) = 0$ for any $z \in Z$.
Applying Theorem \ref{thm:cocycle-uniqueness}, we deduce that $f$ is constant on $Z$. But since $y$ was arbitrary, this suffices.
\end{proof}
We will spend a while building up the proof of Theorem \ref{thm:cocycle-main} in stages. The uniqueness result (Theorem \ref{thm:cocycle-uniqueness}) is somewhat easier, and we return to this at the end.
\subsection{The case \texorpdfstring{$\ell = 1$}{l=1}}
\label{subsec:1-cocycle}
We briefly contemplate the first even slightly non-trivial case of Theorem \ref{thm:cocycle-special}, i.e.~when $\ell = 1$. Since $X$ is ergodic, $C^1(X) = X \times X$, and so this means we have a function
\[
\rho \colon X \times X \to A
\]
whose image is contained in a small ball of $A$, and such that
\[
\rho([x,y]) + \rho([y,z]) = \rho([x,z])
\]
for all $x,y,z \in X$ (this is just from the definition of a $1$-cocycle). Also, $\rho([x,y]) = -\rho([y,x])$ and $\rho([x,x]) = 0$ for all $x,y$.
Finally, recall we wish to construct $f \colon X \to A$ such that $\rho([x,y]) = f(x) - f(y)$ for all $x,y \in X$. It turns out this is not a terribly deep statement: one can simply fix $x_0 \in X$ and define
\[
f(x) := \rho([x, x_0])
\]
noting that
\[
f(x) - f(y) = \rho([x, x_0]) - \rho([y, x_0]) = \rho([x, x_0]) + \rho([x_0, y]) = \rho([x,y]) \ .
\]
Observe that we have not even really used our ``small image'' assumption on $\rho$, or that $A$ is Lie, in this argument.
\bigskip
When we come to prove more general cases of the result, it will be important for inductive reasons that we can find a \emph{canonical} function $f$ to define our coboundary: here, $f$ depends on the arbitrary choice of $x_0$. To set the scene for what follows, we note that we can achieve this if, rather than fixing one particular $x_0$, we instead \emph{average} over all choices. That is, we define
\begin{equation}\label{eq:f-def}
f(x) := \int_X \rho([x,y]) dy
\end{equation}
which is indeed canonical. To show this is still a valid choice, we work through the calculation
\begin{align*}
f(x) - f(y) &= \int_X \rho([x, z])\ dz - \int_X \rho([y, z])\ dz \\
&= \int_X \left[ \rho([x, z]) + \rho([z, y]) \right]\ dz\\
&= \int_X \rho([x,y])\ dz\\
&= \rho([x,y]) \ .
\end{align*}
However, we have overlooked two important subtleties here.
\begin{itemize}
\item We are integrating expressions with values in $A$, a compact abelian Lie group.
\item We have not defined a probability measure on the space $X$, so the integration is currently meaningless.
\end{itemize}
The first item will be discussed in Section \ref{subsec:integration} below.
In brief, we can lift the integrand to the universal cover of (the respective connected
component of) $A$, and then project back the result.
It turns out that this can be defined canonically provided the integrand takes values in a sufficiently
small ball in $A$.
The second item could be addressed using the weak structure theory, which implies that $X$ can
be obtained by successive principal bundle extensions
\[
X \to \pi_{s-1}(X) \to \dots \to \pi_{0}(X) = \{\ast\}.
\]
This allows the construction of a probability measure on $X$ from the Haar measures on the structure groups.
This is the approach taken in \cite{CS12}.
We found it easier to carry out the averaging in stages and prove Theorem \ref{thm:cocycle-main}
by a double induction on $s$ and $\ell$.
In the special case $\ell=1$ currently discussed, this takes the following shape.
We put
\[
f'=\int_{A_s} \rho([x,a.x])d\mu_{A_s}(a),
\]
i.e. we integrate only on the fibre of $\pi_s$ that contains $x$ using the Haar measure on the structure group $A_s$.
It is no longer reasonable to expect that $\partial f'$ agrees with $\rho$, since the definition of $f'$ depends
only on the values of $\rho$ on edges whose endpoints lie in the same fibre of $\pi_{s-1}$.
However, it turns out that $\rho-\partial f'$ is constant on the fibres of $\pi_{s-1}$ and hence it
can be pushed down to a cocycle on $\pi_{s-1}(X)$, and the argument can be completed by induction.
We omit the details, but they are given in Section \ref{subsec:general} in a greater generality.
We finish this section with a comment on the proof in the general case $\ell\ge 1$. Antol\'\i n Camarena and Szegedy define the function
\[
f(x)=\int_{c\in C^{\ell}(X):c(\vec0)=x}\rho(c) dc
\]
using a system of suitable probability measures that are defined on the space of cubes whose vertex
at $\vec 0$ is the point $x$. (Compare this formula with \eqref{eq:f-def}.)
Our proof carries out the above averaging in several steps using a double induction on $s$ and $\ell$. If one combines the averaging in the inductive steps, the resulting formula will be the same as in the approach of Antol\'\i n Camarena and Szegedy. However, we believe that the technical details of the argument become simpler in our approach. In particular, we can avoid any discussion of continuous systems of measures, and moreover the combinatorial properties we need to verify are also easier.
\subsection{A remark on integration}
\label{subsec:integration}
We discuss now the issue raised above about integration of functions taking values in Lie groups.
Suppose $(Y, \mu)$ is some probability space, and $f \colon Y \to \RR / \ZZ$ is some measurable function. It is clear that the integral $\int f(x) d\mu(x)$ does not make sense in general, since averaging on $\RR/\ZZ$ is not well-defined.
However, suppose we know that $f$ takes values in some specified interval $(a,b) \subseteq \RR/\ZZ$ of width at most $1/10$ (say). Then we can make sense of $\int f(x) d\mu(x)$ by identifying $(a,b) \subseteq \RR/\ZZ$ bijectively with a suitable interval $(\tilde{a}, \tilde{b}) \subseteq \RR$, lifting $f$ to a function $Y \to (\tilde{a}, \tilde{b})$, performing normal real integration, and then projecting the result (which lies in $(\tilde{a}, \tilde{b})$ by convexity) back down to $(a,b) \subseteq \RR/\ZZ$. It is straightforward to check that this operation does not depend on the choice of $\tilde{a}, \tilde{b}$.
We will abuse notation to write $\int f(x) d\mu(x)$ for this element of $\RR/\ZZ$, whenever it makes sense to do so. Specifically, we will only write this when $\sup \{|f(x) - f(y)| \colon x, y \in Y \} \le \delta$ for some suitably small absolute constant $\delta$ (in this case taken to be $1/10$).
More generally, the same remarks hold for any compact abelian Lie group $A = (\RR/\ZZ)^d \times K$ for $K$ a finite group. For any reasonable metric on this space, integration is well-defined for functions whose image lies in a sufficiently small $\delta$-ball, with $\delta$ depending only on $A$; and the integral will also lie in the same $\delta$-ball.
For this to hold, we do need to make some choices about the metric we impose on $A$. Specifically, it will be convenient to use a metric $d_A$ on $A \cong (\RR/\ZZ)^d \times K$ that is induced from the standard Euclidean metric on $\RR^d$ (and the standard discrete metric on $K$). This choice is not canonical, because it requires us to fix an isomorphism $A \cong (\RR/\ZZ)^d \times K$; though of course, all metrics on a compact space are equivalent, so we can simply fix a choice and the ambiguity need not concern us.
\begin{definition}
\label{def:torus-metric}
For any compact abelian Lie group $A$ that appears, we assume an identification $A \cong (\RR/\ZZ)^d \times K$ has been fixed and define
\[
d_A((t, k), (t',k')) = (1 - [k=k']) + \inf \left\{ \|x - x'\|_2 \colon x, x' \in \RR^d,\ x, x' \equiv t, t' \pmod{1} \right\} \ .
\]
Similarly, we write $\|\cdot\|_A$ for the ``norm'' $d_A(0, -)$.
\end{definition}
(Here the expression $[k=k']$ takes the value $1$ if $k=k'$ and $0$ otherwise.)
Finally, we observe that the integral discussed above is finitely additive wherever this makes sense; i.e.~if $f$, $f'$ and $f+f'$ all have images within (possibly different) $\delta$-balls in $A$, then $\int (f+f') = \int f + \int f'$.
\subsection{A slightly less easy case}
\label{subsec:ds-cocycle-fixing}
As one further stepping stone towards Theorem \ref{thm:cocycle-special} and Theorem \ref{thm:cocycle-main}, we consider the case where $X = \cD_s(H)$ for some compact abelian group $H$. Recall this means $H$ carries the degree $s$ filtration
\[
H = H_0 = H_1 = \dots = H_s \supseteq \{0\}
\]
and $C^k(X)$ are the Host--Kra cubes $\HK^k(H_\bullet)$ with respect to this filtration. (We use additive notation for $H$ and also for the Host--Kra groups we derive from it.)
We prove Theorem \ref{thm:cocycle-special} in the special case considered in this section by induction on $\ell$. In the inductive step, we are looking for an $(\ell-1)$-cocycle $\rho': C^{\ell-1}\to A$ such that $\partial\rho'=\rho$. Motivated by the previously discussed case of $\ell=1$, we intend to define $\rho'(c)$ by taking an average of $\rho([c,c'])$ over the space
\[
\left\{ c' \colon [c, c'] \in C^\ell(X) \right\}.
\]
In order to compute the average, we need to specify a probability measure. Luckily, the space can be identified with a compact abelian Lie group, and we can make use of the Haar measure on this group. We denote
\[
T_1^\ell := \left\{ c \colon \{0,1\}^{\ell-1} \to H \ \colon \ [\vec{0}, c] \in C^\ell(X) \right\}
\]
and observe that $[\vec{0}, T_1^\ell]$ is a closed subgroup of $\HK^\ell(H_\bullet)$. It is clear that $[c_1, c_2] \in C^\ell(\cD_s(H))$ if and only if $c_2 = c_1 + t$ for some $t \in T_1^\ell$: indeed, the Host--Kra cube group is a group, so since $[c_1, c_2]$ and $[c_1, c_1]$ are both cubes, so is $[c_1, c_2] - [c_1, c_1] = [\vec{0}, c_2 - c_1]$ and so $t=c_2 - c_1$ is in $T_1^\ell$.
In fact, it turns out that $T_1^\ell$ is just the Host--Kra cubespace of dimension $(\ell-1)$ of $H$ given the degree $(s-1)$ filtration, but we will not actually need to know this.
The special case of Theorem \ref{thm:cocycle-special} considered in this section follows from the following proposition by induction on $\ell$.
\begin{proposition}
\label{prop:ds-cocycle-fixing}
Suppose $X = \cD_s(H)$ as above, $A$ is a compact abelian Lie group, and $\ell \ge 1$ and $\rho \colon C^\ell(X) \to A$ is a cocycle such that $d_A(\rho(c_1), \rho(c_2)) \le \delta$ for all $c_1$, $c_2$ and some suitably small $\delta$ \uppar{i.e.~$\delta \le \eps(\ell, A)$}.
We define
\begin{align*}
\rho' \colon C^{\ell-1}(X) &\to A \\
c' &\mapsto \int_{T_1^\ell} \rho([c', c'+t]) d\mu_{T_1^\ell}(t) \ .
\end{align*}
Then
\begin{enumerate}
\item the function $\rho'$ is continuous and $d_A(\rho'(c_1'), \rho'(c_2')) \le \delta$ for all $c_1', c_2' \in C^{\ell-1}(X)$;
\item moreover, $\rho'$ is an $(\ell-1)$-cocycle; and
\item we have $\rho([c_1, c_2]) = \rho'(c_1) - \rho'(c_2)$ for all cubes $c = [c_1, c_2] \in C^\ell(X)$.
\end{enumerate}
\end{proposition}
\begin{proof}
Note we are drawing on Section \ref{subsec:integration} to make sense of the integral in the definition of $\rho'$. To see (i), we note that
\[
\rho'(c_1') - \rho'(c_2') = \int_{T_1^\ell} (\rho([c_1', c_1' + t]) - \rho([c_2', c_2'+t])) d \mu_{T_1^\ell}(t)
\]
and since $\|\rho(c_1) - \rho(c_2)\|_A \le \delta$ for all $c_1, c_2 \in C^\ell(X)$ by assumption, the integrand is bounded in norm pointwise by $\delta$ and hence so is the integral.
Moreover, since the maps
\[
c' \mapsto [c', c'+t]
\]
are equicontinuous for $t \in T_1^\ell$, and since $\rho$ is uniformly continuous, we deduce that the integrand becomes arbitrarily small when $d(c_1', c_2')$ is arbitrarily small. Again, an average of small values in $A$ is small, so $\rho'$ is (uniformly) continuous.
We now consider (ii). We need some further definitions. Let
\[
T_2^\ell := \left\{ c \colon \{0,1\}^{\ell-2} \to H \ \colon \ [[\vec{0}, \vec{0}], [\vec{0},c]] \in C^\ell(H) \right\}
\]
and note as before that $[t_0, t_1] \in T_1^\ell$ if and only if $t_1 = t_0 + u$ for some $u \in T_2^\ell$ (again, just using the fact that the Host--Kra cube group is a group). In other words, every $t \in T_1^\ell$ has a unique decomposition as
\[
t = [v, v] + [\vec{0}, u]
\]
for $v \in T_1^{\ell-1}$ and $u \in T_2^\ell$, and this establishes a group isomorphism $T_1^\ell \leftrightarrow T_1^{\ell-1} \times T_2^{\ell}$. (Indeed, it is easy to see that $[v,v]\in T_1^\ell$ if and only if $[[0,0],[v,v]]\in C^{\ell}(H)$ if and only if $[0,v]\in C^{\ell-1}(H)$ if and only if $v\in T_1^{\ell-1}$.)
To prove additivity of $\rho'$, suppose $[c_0, c_1], [c_1, c_2] \in C^{\ell-1}(X)$, and write
\begin{align*}
\rho'([c_0, c_2]) &= \int_{T_1^\ell} \rho([[c_0, c_2], [c_0, c_2] + t]) d\mu_{T_1^\ell}(t) \\
&= \int_{T_1^{\ell-1} \times T_2^\ell} \rho([[c_0, c_2], [c_0 + v, c_2 + u + v]]) d\mu_{T_1^{\ell-1}}(v) d\mu_{T_2^\ell}(u) \\
&= \int_{T_1^{\ell-1} \times T_2^{\ell}\times T_2^\ell} (\rho([[c_0, c_1], [c_0 + v, c_1 + u' + v]]) \\
&\qquad\qquad\qquad\qquad
+ \rho([[c_1, c_2], [c_1 + u'+ v, c_2 + u + v]]) ) d\mu_{T_1^{\ell-1}}(v) d\mu_{T_2^\ell}(u) d\mu_{T_2^\ell}(u')\\
\end{align*}
In the last line we used additivity of $\rho$ and that the cubes in the last line can be glued along the common face $[c_1,c_1+u'+v]$ to obtain the cube in the penultimate line.
We integrate out the first term and continue the calculation as follows.
\begin{align*}
\rho'([c_0, c_2]) &= \rho'([c_0, c_1]) + \int_{T_2^\ell}\Big(\int_{T_1^{\ell-1}\times T_2^\ell}\rho([[c_1, c_2], [c_1 + (v+u'),\\
&\qquad\qquad\qquad\qquad\qquad\qquad\qquad
c_2 + (u - u')+ (v+u')]]) d\mu_{T_1^{\ell-1}}(v) d\mu_{T_2^\ell}(u)\Big) d\mu_{T_2^\ell}(u')\\
&= \rho'([c_0, c_1]) + \int_{T_2^\ell}\Big(\int_{T_1^{\ell-1}\times T_2^\ell}\rho([[c_1, c_2], [c_1 + v'', c_2 + u''+ v'']]) d\mu_{T_1^{\ell-1}}(v'') d\mu_{T_2^\ell}(u'')\Big) d\mu_{T_2^\ell}(u')\\
&= \rho'([c_0, c_1])+\rho'([c_1, c_2])
\end{align*}
as required. For the penultimate equation we used the substitution $v''=v+u'$ and $u''=u-u'$ and translation invariance of the Haar measure.
We finish the proof with (iii). Given $[c_1, c_2] \in C^\ell(X)$ we may write $c_2 = c_1 + t$ for some $t \in T_1^\ell$ and then observe
\begin{align*}
\rho([c_1, c_2]) &= \int_{T_1^\ell} \left(\rho([c_1, c_1+w]) + \rho([c_1+w, c_1+t] )\right) d\mu_{T_1^\ell}(w) \\
&= \int_{T_1^\ell} \rho([c_1, c_1+w]) d\mu_{T_1^\ell}(w) - \int_{T_1^\ell} \rho([c_1+t, c_1+w]) d\mu_{T_1^\ell}(w)
\end{align*}
which once again is just $\rho'(c_1) - \rho'(c_2)$ after reparameterization of the last integral (using that $t \in T_1^\ell$ and Haar measure is translation-invariant).
\end{proof}
\subsection{The general case}\label{subsec:general}
The key ideas for the general case, Theorem \ref{thm:cocycle-main}, have already appeared in the special cases we have just discussed. The very vague strategy they suggest for proving the general case is as follows.
\begin{itemize}
\item As in Section \ref{subsec:ds-cocycle-fixing}, we will perform an averaging over a space of cubes
\[
\{c' \colon [c, c'] \in C^\ell(X) \}
\]
to obtain an $(\ell-1)$-cocycle $\rho'$ from an $\ell$-cocycle $\rho$, such that $\rho([c_1, c_2]) = \rho'(c_1) - \rho'(c_2)$. Iterating this argument $\ell$ times will give the result.
\item Unfortunately we do not currently have measures defined on these spaces, so we will do as is suggested in Section \ref{subsec:1-cocycle} and appeal to the weak structure theory. This will enable us to express the averaging as a sequence of integrals over each of the structure groups $A_s(X)$ in turn, or more accurately over Host--Kra configurations built out of these groups.
\end{itemize}
With some thought, this approach is seen to be essentially equivalent to performing a \emph{double} induction on both $\ell$ (the order of the cocycle) and $s$ (the degree of the nilspace $X$ in Theorem \ref{thm:cocycle-special}, or of the fibration $\beta$ in the case of Theorem \ref{thm:cocycle-main}). In each stage of the induction, we perform some integration over groups related to $A_s(X)$ in the spirit of the previous subsections, and are left at the end with some simpler objects on which to iterate.
We need to introduce some notation for technical reasons that will become clear later. Let $c_1,c_2:\{0,1\}^\ell\to X$ be two configurations. Until now, we defined the concatenation by the same symbol irrespective of the coordinate on which the concatenation takes place. Now we wish to designate this in our notation, and write:
\begin{align*}
[c_1,c_2]_k:\{0,1\}^{\ell+1}\to& X\\
\omega\mapsto&
\begin{cases}
c_1(\omega_1,\ldots,\omega_{k-1},\omega_{k+1},\ldots,\omega_{\ell+1}) & \text{if $\omega_k=0$}\\
c_2(\omega_1,\ldots,\omega_{k-1},\omega_{k+1},\ldots,\omega_{\ell+1}) & \text{if $\omega_k=1$}.
\end{cases}
\end{align*}
We also introduce some notation for the derivative along a specific coordinate.
If $\rho:C^{\ell}\to A$ is a function on cubes, we write
\[
\partial_k\rho([c_1,c_2]_k)=\rho(c_1)-\rho(c_2).
\]
Observe that the identity
\[
\partial^{\ell+1} f=\partial_k (\partial^{\ell} f)
\]
holds irrespective of the value of $k$.
Hence, if we are differentiating a coboundary $\partial^{\ell} f$, the derivative is independent of the direction $k$; however, this will not be the case in general.
Armed with this notation, we can now formulate the inductive step in the proof
of Theorem \ref{thm:cocycle-special}.
\begin{lemma}
\label{lem:inductive-cocycle-special}
Let $A$ be a compact abelian Lie group with the metric $d_A$ as above, and let $s \ge 1$, $\ell \ge 1$ be given. Then there exists $\eps = \eps(s, \ell, A)$ such that the following holds.
Suppose $X$ is a \uppar{compact, ergodic} nilspace of degree $s$, and $\rho \colon C^\ell(X) \to A$ is an $\ell$-cocycle such that $d_A(\rho(c), \rho(c')) \le \delta$ for all $c,c' \in C^\ell(X)$ and some $\delta \le \eps$.
Write $\pi_{s-1} \colon X \to \pi_{s-1}(X)$ for the canonical factor. Then we may decompose
\[
\rho(c) = \sum_{k=1}^{\ell}\partial_k\rho'_k(c) + \tilde{\rho}(\pi_{s-1}(c))
\]
where
\begin{itemize}
\item $\rho'_k \colon C^{\ell-1}(X) \to A$ is an $(\ell-1)$-cocycle for each $k$, and
\item $\tilde{\rho} \colon C^\ell(\pi_{s-1}(X)) \to A$ is an $\ell$-cocycle on $\pi_{s-1}(X)$,
\end{itemize}
such that both $\rho'_k$ and $\tilde{\rho}$ take images in a small ball in $A$, i.e.~$d_A(\rho'_k(c_1), \rho'_k(c_2)) \lesssim_{s, \ell} \delta$ for all $c_1, c_2 \in C^{\ell-1}(X)$ and $k$, and $d_A(\tilde{\rho}(c_1), \tilde{\rho}(c_2)) \lesssim_{s,\ell} \delta$ for all $c_1, c_2 \in C^{\ell}(\pi_{s-1}(X))$.
\end{lemma}
We verify that this is enough to prove Theorem \ref{thm:cocycle-special}.
\begin{proof}[Proof of Theorem \ref{thm:cocycle-special} assuming Lemma \ref{lem:inductive-cocycle-special}]
We remark again that an ergodic nilspace of degree $0$ is just the one-point space $\{\ast\}$, and so a cocycle of any positive order on this space is identically zero. Also recall that a $0$-cocycle is just the same thing as a continuous function.
So, we may proceed by induction on $s$ and $\ell$, where the cases $\ell = 0$ and $s=0$ are both clear. Given $s, \ell > 0$ and a cocycle $\rho \colon C^\ell(X) \to A$ as in Theorem \ref{thm:cocycle-special}, we decompose
\[
\rho(c) = \sum_{k=1}^{\ell}\partial_k\rho'_k(c) + \tilde{\rho}(\pi_{s-1}(c))
\]
as in Lemma \ref{lem:inductive-cocycle-special}.
Temporarily we will say that a function $f \colon Y \to A$ has \emph{small image} if $d_A(f(y), f(y')) \lesssim_{s, \ell} \delta$ for all $y, y' \in X$. By inductive hypothesis, $\rho'_k = \partial^{\ell-1} g_k$ for some continuous $g_k \colon X \to A$ with small image, and similarly $\tilde{\rho} = \partial^\ell h$ for some continuous $h \colon \pi_{s-1}(X) \to A$ with small image.
Setting $f = \sum_{k=1}^{\ell} g_k + h \circ \pi_{s-1}$, we see that this is again a continuous function $X \to A$ with small image, and moreover
\[
\partial^\ell f = \sum_{k=1}^\ell \partial^\ell g_k + \partial^\ell(h \circ \pi_{s-1})
= \sum_{k=1}^\ell\partial_k \rho'_k + \tilde{\rho} = \rho
\]
which completes the proof.
\end{proof}
Lemma \ref{lem:inductive-cocycle-special} is proved by iterating the following
result. Recall from Section \ref{subsec:ds-cocycle-fixing}, that we write $T_{1}^\ell$ for the
set of configurations $t:\{0,1\}^{\ell-1}\to A_s(X)$ that satisfy
\[
[0,t]\in C^\ell(\cD_s(A_s)).
\]
(Here we did not indicate in which coordinate the concatenation takes place, as the resulting group is independent of this choice.)
\begin{lemma}
\label{lem:inductive-cocycle-more-special}
Let $A$ be a compact abelian Lie group with the metric $d_A$ as above, and let $s \ge 1$, $\ell \ge 1$ be given. Then there exists $\eps = \eps(s, \ell, A)$ such that the following holds.
Suppose $X$ is a \uppar{compact, ergodic} nilspace of degree $s$, and $\rho \colon C^\ell(X) \to A$ is an $\ell$-cocycle such that $d_A(\rho(c), \rho(c')) \le \delta$ for all $c,c' \in C^\ell(X)$ and some $\delta \le \eps$.
Write $\pi_{s-1} \colon X \to \pi_{s-1}(X)$ for the canonical factor. Fix a number $1\le k\le \ell$. Then we may decompose
\[
\rho(c) = \partial_k\rho'(c) + \tilde{\rho}(c)
\]
where
\begin{itemize}
\item $\rho' \colon C^{\ell-1}(X) \to A$ is an $(\ell-1)$-cocycle, and
\item $\tilde{\rho} \colon C^\ell(X) \to A$ is an $\ell$-cocycle, which is invariant under the action of the group $[0,T_{1}^\ell]_k$,
\end{itemize}
such that both $\rho'$ and $\tilde{\rho}$ take images in a small ball in $A$, i.e.~$d_A(\rho'(c_1), \rho'(c_2)) \lesssim_{s, \ell} \delta$ for all $c_1, c_2 \in C^{\ell-1}(X)$ and $d_A(\tilde{\rho}(c_1), \tilde{\rho}(c_2)) \lesssim_{s,\ell} \delta$ for all $c_1, c_2 \in C^{\ell}(X)$.
\end{lemma}
The proof of Lemma \ref{lem:inductive-cocycle-more-special} follows Section \ref{subsec:ds-cocycle-fixing} very closely, which is possible by virtue of the weak structure theory. There will be some additional complications caused by the ``relative'' nature of Lemma \ref{lem:inductive-cocycle-more-special} as compared to Proposition \ref{prop:ds-cocycle-fixing}.
\begin{proof}[Proof of Lemma \ref{lem:inductive-cocycle-more-special}]
The definition of $\rho'$ will be similar to that in Proposition \ref{prop:ds-cocycle-fixing}. Instead of integrating over the set
\[
\left\{ c' \colon [c, c']_k \in C^\ell(X) \right\}
\]
we will integrate over all such $c'$ such that additionally $c, c'$ lie in the same fiber of $\pi_{s-1}$. By the weak structure theory, this is equivalent to saying
\[
\rho'(c) = \int_{T_1^\ell} \rho([c, t.c]_k) d\mu_{T_{1}^\ell}(t)
\]
where as always $t.c$ denotes the action of $A_s(X)$ on $X$, applied pointwise to the configurations $t \colon \{0,1\}^{\ell-1} \to A_s(X)$, $c \colon \{0,1\}^{\ell-1} \to X$.
It suffices to check that
\begin{enumerate}
\item the function $\rho' \colon C^{\ell-1}(X) \to A$ is continuous and has small image (in the above sense);
\item also, $\rho'$ is an $(\ell-1)$-cocycle on $X$; and
\item the function
\[
\tilde{\rho}([c_1, c_2]_k) := \rho([c_1, c_2]_k) - (\rho'(c_1) - \rho'(c_2))
\]
is a continuous $\ell$-cocycle on $X$ with small image, and furthermore is invariant under the action of $[0,T_{1}^{\ell}]_k$.
\end{enumerate}
Part (i) follows by the same arguments as in Proposition \ref{prop:ds-cocycle-fixing}. Furthermore, the proof of (ii) is identical to the corresponding part of Proposition \ref{prop:ds-cocycle-fixing}, where we simply replace all expressions of the form $(c + t)$ with $(t . c)$, when $c$ is a cube of $X$ and $t$ a configuration with values in $A_s(X)$.
There is a bit more to say for (iii). First, note that since $\rho$ and $\partial_k \rho'$ are continuous $\ell$-cocycles with small image, it is immediate that $\tilde{\rho}$ inherits these properties (see Remark \ref{rem:cocycle-defs}); the challenge is to show the invariance property.
We have
\[
\tilde\rho([c, t.c']_k) = \tilde\rho([c, c']_k) + \tilde\rho([c', t.c']_k)
\]
and so it suffices to show that
\[
\rho([c', t.c']_k) = \rho'(c') - \rho'(t.c')
\]
for all $c'$ and $t \in T_{1}^\ell$. (Indeed, this implies that $\tilde\rho([c',t.c']_k)=0$.) Again this is very similar to Proposition \ref{prop:ds-cocycle-fixing}: we compute
\begin{align*}
\rho([c', t.c']_k) &= \int_{T_{1}^\ell} \left(\rho([c', w.c']_k) + \rho([w.c', t.c']_k) \right) d\mu_{T_{1}^\ell}(w) \\
&= \int_{T_{1}^\ell} \rho([c', w.c']_k) d\mu_{T_{1}^\ell}(w) - \int_{T_{1}^\ell} \rho([t.c', w.c']_k) d\mu_{T_{1}^\ell}(w) \\
&= \rho'(c') - \rho'(t.c')
\end{align*}
as required, once again after reparameterization of the last integral.
\end{proof}
\begin{proof}[Proof of Lemma \ref{lem:inductive-cocycle-special}]
We iterate Lemma \ref{lem:inductive-cocycle-more-special}.
We first write $\rho=\partial_1\rho'_1+\tilde \rho_1$, then in the $k$-th step we
apply Lemma \ref{lem:inductive-cocycle-more-special} to the cocycle
$\tilde \rho_{k-1}$ and write $\tilde\rho_{k-1}=\partial_k\rho'_k+\tilde\rho_k$.
Combining these equations we obtain the desired decomposition
\[
\rho=\sum_{k=1}^{\ell}\partial_k\rho'_k+\tilde\rho_\ell.
\]
It remains to show that $\tilde \rho_\ell$ descends to a cocycle on $\pi_{s-1}(X)$.
We first observe that $\tilde \rho_k$ is invariant under $[0,T_{1}^\ell]_k$ and this property
is inherited by $\tilde\rho_j$ for all $j>k$. Indeed, the cocycle $\tilde\rho_j$ is constructed
in the proof of Lemma \ref{lem:inductive-cocycle-more-special} as the difference
of a $[0,T_{1}^\ell]_k$-invariant cocycle and an average of $[0,T_{1}^\ell]_k$-invariant cocycles.
Thus $\tilde\rho_\ell$ is invariant under $[0,T_{1}^\ell]_k$ for all $k$.
Recalling that reflection simply negates the cocycle, i.e.~$\tilde\rho_\ell([c_0,c_1]_k) = -\tilde\rho_\ell([c_1,c_0]_k)$, it follows that $\tilde\rho_\ell$ is also invariant under $[T_1^\ell, 0]_k$.
It is clear that these groups generate $\cD_s(A_s(X))$, and hence $\tilde{\rho}_\ell$ is invariant under the
action of this group.
Now it follows that $\tilde\rho_\ell$ descends to a function on $C^{\ell}(\pi_{s-1}(X))$. However, it is not completely obvious that this yields a cocycle. One needs to check that configurations $\tilde{c}_0, \tilde{c}_1, \tilde{c}_2$ on $\pi_{s-1}$ such that $[\tilde{c}_0, \tilde{c}_1]$ and $[\tilde{c}_1, \tilde{c}_2]$ are cubes of $\pi_{s-1}(X)$, \emph{lift} to corresponding configurations $c_0, c_1, c_2$ on $X$ in a compatible way. Fortunately this is always true; see \cite{GMV1}*{Lemma 7.5}.
\end{proof}
This completes the proof of Theorem \ref{thm:cocycle-special}. The generalization to a proof of Theorem \ref{thm:cocycle-main} involves no new ideas at all, and only a small amount of further justification.
We will have to recall some notions from the \emph{relative} structure theory expounded in \cite{GMV1}*{Section 7}.
What we shall need is that the canonical factor $X \to \pi_{s-1}(X)$ of a nilspace $X$ of degree $s$, has an analogue for fibrations.
\begin{proposition}
\label{prop:relative-weak-structure}
Let $\beta \colon X \to Y$ be a fibration of degree $s$ between compact ergodic cubespaces $X$ and $Y$ that obey the glueing axiom. Then there is a map $\pi_{\beta,s-1} \colon X \to \pi_{\beta,s-1}(X)$ where
\begin{itemize}
\item the fibration $\beta$ factors as $X \xrightarrow[\pi_{\beta,s-1}]{} \pi_{\beta,s-1}(X) \xrightarrow[\tilde{\beta}]{} Y$;
\item the map $\tilde{\beta}$ is a fibration of degree $(s-1)$;
\item there is a compact abelian group $A_s(\beta)$, the \emph{structure group of the fibration}, which acts continuously and freely on all of $X$ and whose orbits are precisely the fibers of $\pi_{\beta,s-1}$;
\item a similar statement holds for cubes; specifically, $C^k(\cD_s(A_s(\beta)))$ acts pointwise on $C^k(X)$, and its orbits are precisely the fibers of the map $\pi_{\beta,s-1} \colon C^k(X) \to C^k(\pi_{\beta,s-1}(X))$.
\end{itemize}
\end{proposition}
For the details, see \cite{GMV1}*{Proposition 7.12 and Theorem 7.19}.
(Apply both results with $(s+1)$ in place of $s$.)
We can now state the inductive step that provides a technical generalization of Lemma \ref{lem:inductive-cocycle-special}.
\begin{lemma}
\label{lem:inductive-cocycle-general}
Let $A$ be a compact abelian Lie group with the metric $d_A$ as above, and let $s \ge 1$, $\ell \ge 1$ be given. Then there exists $\eps = \eps(s, \ell, A)$ such that the following holds.
Suppose $X, Y, \beta, \rho$ are as in Theorem \ref{thm:cocycle-main}. Then there is a decomposition
\[
\rho(c) = \sum_{k=1}^\ell\partial_k\rho'_k(c) + \tilde{\rho}(\pi_{\beta,s-1}(c))
\]
where
\begin{itemize}
\item $\rho'_k \colon C^{\ell-1}(X) \to A$ is an $(\ell-1)$-cocycle for each $k$, and
\item $\tilde{\rho} \colon C^\ell(\pi_{\beta,s-1}(X)) \to A$ is an $\ell$-cocycle on $\pi_{\beta,s-1}(X)$,
\end{itemize}
such that both $\rho'$ and $\tilde{\rho}$ have small image in the usual sense.
\end{lemma}
The deduction of Theorem \ref{thm:cocycle-main} is again very similar.
\begin{proof}[Proof of Theorem \ref{thm:cocycle-main} assuming Lemma \ref{lem:inductive-cocycle-general}]
The only thing that changes is the base case; the rest of the argument is the same as for Theorem \ref{thm:cocycle-special}. If we have a fibration $\beta \colon X \to Y$ of degree $0$ between two ergodic cubespaces $X$ and $Y$, then $\beta$ is an isomorphism. In particular, an $\ell$-cocycle on $X$ is the same thing (under $\beta$) as an $\ell$-cocycle on $Y$, so again Theorem \ref{thm:cocycle-main} is trivial in this case.
\end{proof}
Finally, the proof of Lemma \ref{lem:inductive-cocycle-general} is unchanged from that of Lemma \ref{lem:inductive-cocycle-special}, replacing all appearances of $\pi_{s-1}$ with $\pi_{\beta,s-1}$ and $A_s(X)$ with $A_s(\beta)$. All appeals to the weak structure theory are legitimized by Proposition \ref{prop:relative-weak-structure}.
\subsection{The uniqueness result}
\label{sec:uniqueness}
We now prove Theorem \ref{thm:cocycle-uniqueness}. Our proof is very closely modelled on that of \cite{CS12}*{Lemma 3.25}.
We work up to the result in several incremental stages, each of which is in fact a special case of the theorem. The first is classical.
\begin{lemma}
\label{lem:homs-discrete}
Suppose $A$ is a compact abelian Lie group equipped with the metric $d_A$. Then there exists $\eps = \eps(A) > 0$ such that the following holds.
If $G$ is an abelian group and $\phi \colon G \to A$ is a group homomorphism such that $d_A(0, \phi(g)) \le \eps$ for all $g \in G$, then $\phi$ is the trivial homomorphism.
\end{lemma}
\begin{proof}
This is equivalent to saying that $A$ has ``no small subgroups'': there is some neighbourhood of the identity in $A$ containing no non-trivial subgroups.
Since we know $A$ is isomorphic to $(\RR/\ZZ)^d \times K$ for some finite $K$, we may argue this directly. Let
\[
\pi \colon \RR^d \times K \to (\RR/\ZZ)^d \times K
\]
denote the projection map. Choose some $\delta > 0$ such that $\pi$ identifies $B_\delta(0) \subseteq (\RR/\ZZ)^d \times K$ bijectively with with the corresponding open ball $B_\delta(0) \times \{0\}$ in $\RR^d \times K$.
Let $x \in B_{\delta/2}(0) \setminus \{0\} \subseteq (\RR/\ZZ)^d \times K$, and let $\tilde{x}$ be the corresponding lift to $\RR^d$. Then there is some positive integer $n$ such that $\delta/2 \le |n \tilde{x}| < \delta$. It follows that $n x \in B_\delta(0) \setminus B_{\delta/2}(0) \subseteq (\RR/\ZZ)^d \times K$, and therefore $x$ is not contained in any subgroup of $B_{\delta/2}(0)$.
\end{proof}
Note that if we interpret $G$ as a nilspace of degree $1$ (i.e.~considering $\cD_1(G)$) then any homomorphism $\phi \colon G \to A$ has $\partial^2 \phi = 0$, so this is a special case of Theorem \ref{thm:cocycle-uniqueness}.
Having shown that the space of \emph{linear} maps $G \to A$ is discrete, we bootstrap this to a statement about \emph{polynomial} maps.
\begin{lemma}
\label{lem:uniqueness-d1-case}
Suppose $A$ is a compact abelian Lie group equipped with the metric $d_A$. Fix $\ell \ge 1$. Then there exists $\eps = \eps(\ell, A) > 0$ such that the following holds.
Let $G$ be any abelian group, and consider the cubespace structure $\cD_1(G)$ on $G$. Let $\gamma \colon G \to A$ be a map such that $\partial^\ell \gamma \equiv 0$ and $d_A(\gamma(g), \gamma(g')) \le \eps$ for all $g, g' \in G$.
Then $\gamma$ is constant.
\end{lemma}
\begin{proof}
We proceed by induction on $\ell$. The case $\ell = 1$ is trivial: since $[x,y] \in C^1(G)$ for every $x,y \in G$ we have $\partial^1 \gamma([x,y]) = \gamma(x) - \gamma(y) = 0$ for all $x,y$, and hence $\gamma$ is constant.
Now take $\ell > 1$. Note that an $\ell$-cube in $\cD_1(G)$ has the form $[c, t+c]$ for some $t \in G$, where $t+c$ denotes the pointwise shift $\omega \mapsto t + c(\omega)$ of $c$. Defining the ``derivative''
\[
\gamma_t(x) = \gamma(x) - \gamma(x+t)
\]
we note that
\[
\partial^\ell \gamma([c, t+c]) = \partial^{\ell-1} \gamma_t(c) \ .
\]
By hypothesis, for all $t$ we have $\partial^{\ell-1} \gamma_t \equiv 0$ and hence, for an appropriate choice of $\eps$, by induction $\gamma_t \equiv \alpha(t)$ is a constant function for all $t$.
But now observe that
\begin{align*}
\alpha(t) + \alpha(t') &= \gamma_t(0) + \gamma_{t'}(t) \\
&= \gamma(0) - \gamma(t) + \gamma(t) - \gamma(t+t') \\
&= \alpha(t+t')
\end{align*}
and trivially $\alpha(0) = 0$, $\alpha(-t) = -\alpha(t)$, so $\alpha \colon G \to A$ is a group homomorphism. Moreover, by hypothesis $d_A(0, \alpha(t)) \le \eps$ for all $t$; so by Lemma \ref{lem:homs-discrete} $\alpha \equiv 0$. Hence, $\gamma(x) = \gamma(0) - \alpha(x) = \gamma(0)$ is constant.
\end{proof}
Once again it is clear that this is a special case of the general result. The point is that we can use the weak structure theory to decompose our nilspace $X$ of degree $s$ into a tower of extensions
\[
X \to \pi_{s-1}(X) \to \dots \to \pi_{0}(X) = \{\ast\}
\]
whose fibers are compact abelian groups $A_k(X)$ equipped with a nilspace structure $\cD_k(A_k)$. By applying Lemma \ref{lem:uniqueness-d1-case} on the fibers one at a time, we deduce the result for $X$.
\begin{proof}[{Proof of Theorem \ref{thm:cocycle-uniqueness}}]
We proceed by induction on $s$. The case $s=0$ is trivial, since an ergodic nilspace of degree $0$ is just the $1$-point space $\{\ast\}$.
Suppose $s>0$. Fix $y \in \pi_{s-1}(X)$ and consider the fiber $\pi_{s-1}^{-1}(y)$. Clearly the restriction of $\gamma$ to $\pi_{s-1}^{-1}(y)$ still has the same properties. By the weak structure theorem (\cite{GMV1}*{Theorem 5.4}),
$\pi_{s-1}^{-1}(y)$ is isomorphic as a cubespace to $\cD_s(A_s(X))$, and hence $\gamma|_{\pi_{s-1}^{-1}(y)}$ can be identified with a map $\gamma_y \colon A_s(X) \to A$,\footnote{On this occasion $\gamma_y$ does \emph{not} denote a derivative.} such that
\[
\partial^\ell \gamma_y \colon C^\ell(\cD_s(A_s(X))) \to A
\]
is identically zero.
Note, however, that $C^\ell(\cD_1(A_s(X)))$ is contained in $C^\ell(\cD_s(A_s(X)))$. Hence, \emph{a fortiori} $\gamma$ satisfies the hypotheses of Lemma \ref{lem:uniqueness-d1-case}, and so (for a suitable choice of $\eps$) is constant.
Since $y$ was arbitrary, we deduce that $\gamma$ is constant on fibers of $\pi_{s-1}$. Hence, $\gamma$ factors as a map $\gamma' \colon \pi_{s-1}(X) \to A$ on a nilspace of degree at most $(s-1)$, and it is clear that $\gamma'$ inherits all the properties of $\gamma$. By inductive hypothesis, $\gamma'$ is constant and hence $\gamma$ is constant.
\end{proof}
|
\section{Introduction: big data, parallel processing and nested models}
\label{s:intro}
Serial computing has reached a plateau and parallel, distributed architectures are becoming widely available, from machines with a few cores to cloud computing with 1000s of machines. The combination of powerful nested models with large datasets is a key ingredient to solve difficult problems in machine learning, computer vision and other areas, and it underlies recent successes in deep learning \citep{Hinton_12a,Le_12a,Dean_12a}. Unfortunately, parallel computation is not easy, and many good serial algorithms do not parallelise well. The cost of communicating (through the memory hierarchy or a network) greatly exceeds the cost of computing, both in time and energy, and will continue to do so for the foreseeable future \citep{FullerMillet11a,Graham_04a}. Thus, good parallel algorithms must minimise communication and maximise computation per machine, while creating sufficiently many subproblems (ideally independent) to benefit from as many machines as possible. The load (in runtime) on each machine should be approximately equal. Faults become more frequent as the number of machines increases, particularly if they are inexpensive machines. Machines may be heterogeneous and differ in CPU and memory; this is the case with initiatives such as SETI@home, which may become an important source of distributed computation in the future. Big data applications have additional restrictions. The size of the data means it cannot be stored on a single machine, so distributed-memory architectures are necessary. Sending data between machines is prohibitive because of the size of the data and the high communication costs. In some applications, more data is collected than can be stored, so data must be regularly discarded. In others, such as sensor networks, limited battery life and computational power imply that data must be processed locally.
In this paper, we focus on machine learning models of the form $\ensuremath{\mathbf{y}} = \ensuremath{\mathbf{f}}_{K+1}(\dots \ensuremath{\mathbf{f}}_2(\ensuremath{\mathbf{f}}_1(\ensuremath{\mathbf{x}}))\dots)$, i.e., consisting of a nested mapping from the input \ensuremath{\mathbf{x}}\ to the output \ensuremath{\mathbf{y}}. Such \emph{nested models} involve multiple parameterised layers of processing and include deep neural nets \citep{HintonSalakh06a}, cascades for object recognition in computer vision \citep{Serre_07a,Ranzat_07b} or for phoneme classification in speech processing \citep{GoldMorgan99a,SaonChien12a}, wrapper approaches to classification or regression \citep{KohaviJohn97a}, and various combinations of feature extraction/learning and preprocessing prior to some learning task. Nested and hierarchical models are ubiquitous in machine learning because they provide a way to construct complex models by the composition of simple layers. However, training nested models is difficult even in the serial case because \emph{function composition produces inherently nonconvex functions}, which makes gradient-based optimisation difficult and slow, and sometimes inapplicable (e.g.\ with nonsmooth or discrete layers).
Our starting point is a recently proposed technique to train nested models, the \emph{method of auxiliary coordinates (MAC)} \citep{CarreirWang12a,CarreirWang14a}. This reformulates the optimisation into an iterative procedure that alternates training submodels independently with coordinating them. It introduces significant model and data parallelism, can often train the submodels using existing algorithms, and has convergence guarantees with differentiable functions to a local stationary point, while it also applies with nondifferentiable or even discrete layers. MAC has been applied to various nested models \citep{CarreirWang14a,WangCarreir14a,CarreirRaziper15a,RaziperCarreir16a,CarreirVladym15a}. However, the original papers proposing MAC \citep{CarreirWang12a,CarreirWang14a} did not address how to run MAC on a distributed computing architecture, where communication between machines is far costlier than computation. This paper proposes \emph{ParMAC}, a parallel, distributed framework to learn nested models using MAC, implements it in Message Passing Interface (MPI) for the problem of learning binary autoencoders (BAs), and demonstrates its ability to train on large datasets and achieve large speedups on a distributed cluster. We first review related work (section~\ref{s:related}), describe MAC in general and for BAs (section~\ref{s:MAC}) and introduce the ParMAC model and some extensions of it (section~\ref{s:ParMAC}). Then, we analyse theoretically ParMAC's parallel speedup (section~\ref{s:speedup-th}) and convergence (section~\ref{s:conv}). Finally, we describe our MPI implementation of ParMAC for BAs (section~\ref{s:ParMAC-BAhash-implem}) and show experimental results (section~\ref{s:expts}). Although our MPI implementation and experiments are for a particular ParMAC algorithm (for binary autoencoders), we emphasise that our contributions (the definition of ParMAC and the theoretical analysis of its speedup and convergence) apply to ParMAC in general for any situation where MAC applies, i.e., nested functions with $K$ layers.
\section{Related work}
\label{s:related}
Distributed optimisation and large-scale machine learning have been steadily gaining interest in recent years. Most work has centred on \emph{convex} optimisation, particularly when the objective function has the form of empirical risk minimisation (data fitting term plus regulariser) \citep{Cevher_14a}. This includes many important models in machine learning, such as linear regression, LASSO, logistic regression or SVMs. Such work is typically based on stochastic gradient descent (SGD) \citep{Bottou10a}, coordinate descent (CD) \citep{Wright16a} or the alternating direction method of multipliers (ADMM) \citep{Boyd_11a}. This has resulted in several variations of parallel SGD \citep{Mcdonal_10a,Bertsek11a,Zinkev_10a,Gemull_11a,Niu_11a}, parallel CD \citep{Bradley_11a,RichtarTakac13a,LiuWright15a} and parallel ADMM \citep{Boyd_11a,Ouyang_13a,ZhangKwok14a}.
It is instructive to consider the parallel SGD case in some detail. Here, one typically runs SGD independently on data subsets (done by $P$ worker machines), and a parameter server regularly gathers the replica parameters from the workers, averages them and broadcasts them back to the workers. One can show \citep{Zinkev_10a} that, for a small enough step size and under some technical conditions, the distance to the minimum in objective function value satisfies an upper bound. The upper bound has a term that decreases as the number of workers $P$ increases, so that parallelisation helps, but it has another term that is independent of $P$, so that past a certain point parallelisation does not help. In practice, the speedups over serial SGD are generally modest. Also, the theoretical guarantees of parallel SGD are restricted to \emph{shallow} models, as opposed to \emph{deep} or \emph{nested} models, because the composition of functions is nearly always nonconvex. Indeed, parallel SGD can diverge with nonconvex models. The intuitive reason for this is that, with local minima, the average of two workers can have a larger objective value than each of the individual workers, and indeed the average of two minima need not be a minimum. In practice, parallel SGD can give reasonable results with nonconvex models if one takes care to average replica models that are close in parameter space and thus associated with the same optimum (e.g.\ eliminating ``stale'' models and other heuristics), but this is not easy \citep{Dean_12a}.
Little work has addressed nonconvex models. Most of it has focused on deep nets \citep{Dean_12a,Le_12a}. For example, Google's DistBelief \citep{Dean_12a} uses asynchronous parallel SGD, with gradients for the full model computed with backpropagation, to achieve data parallelism (with the caveat above), and some form of model parallelism. The latter is achieved by carefully partitioning the neural net into pieces and allocating them to machines to compute gradients. This is difficult to do and requires a careful match of the neural net structure (number of layers and hidden units, connectivity, etc.\@) to the target hardware. Although this has managed to train huge nets on huge datasets by using tens of thousands of CPU cores, the speedups achieved were very modest. Other work has used similar techniques but for GPUs \citep{Chen_12e,Zhang_13a,Coates_13a,Seide_14a}.
Another recent trend is on parallel computation abstractions tailored to machine learning, such as Spark \citep{Zaharia_10a}, GraphLab \citep{Low_12a}, Petuum \citep{Xing_15a} or TensorFlow \citep{Abadi_15a}, with the goal of making cloud computing easily available to train machine learning models. Again, this is often based on shallow models trained with gradient-based convex optimisation techniques, such as parallel SGD. Some of these systems implement some form of deep neural nets.
Finally, there also exist specific approximation techniques for certain types of large-scale machine learning problems, such as spectral problems, using the Nystr{\"o}m formula or other landmark-based methods \citep{WilliamSeeger01a,Bengio_04a,DrineasMahoney05a,Talwal_13a,VladymCarreir13a,VladymCarreir16a}.
ParMAC is specifically designed for nested models, which are typically nonconvex and include deep nets and many other models, some of which have nondifferentiable layers. As we describe below, ParMAC has the advantages of being simple and relatively independent of the target hardware, while achieving high speedups.
\section{Optimising nested models using the method of auxiliary coordinates (MAC)}
\label{s:MAC}
Many machine learning architectures share a fundamental design principle: \emph{mathematically, they construct a (deeply) nested mapping from inputs to outputs}, of the form $\ensuremath{\mathbf{f}}(\ensuremath{\mathbf{x}};\ensuremath{\mathbf{W}}) = \ensuremath{\mathbf{f}}_{K+1}(\dots \ensuremath{\mathbf{f}}_2(\ensuremath{\mathbf{f}}_1(\ensuremath{\mathbf{x}};\ensuremath{\mathbf{W}}_1);\ensuremath{\mathbf{W}}_2)\dots;\ensuremath{\mathbf{W}}_{K+1})$ with parameters \ensuremath{\mathbf{W}}, such as deep nets or binary autoencoders consisting of multiple processing layers. Such problems are traditionally optimised using methods based on gradients computed using the chain rule. However, such gradients may sometimes be inconvenient to use, or may not exist (e.g.\ if some of the layers are nondifferentiable). Also, they are hard to parallelise, because of the inherent sequentiality in the chain rule.
The \emph{method of auxiliary coordinates (MAC)} \citep{CarreirWang12a,CarreirWang14a} is designed to optimise nested models without using chain-rule gradients while introducing parallelism. It solves an equivalent but in appearance very different problem to the nested one, which affords embarrassing parallelisation. The idea is to break nested functional relationships judiciously by introducing new variables (the \emph{auxiliary coordinates}) as equality constraints. These are then solved by optimising a penalised function using alternating optimisation over the original parameters (which we call the \ensuremath{\mathbf{W}}\ step) and over the coordinates (which we call the \ensuremath{\mathbf{Z}}\ step). The result is a \emph{coordination-minimisation (CM) algorithm}: the minimisation (\ensuremath{\mathbf{W}}) step updates the parameters by splitting the nested model into independent submodels and training them using existing algorithms, and the coordination (\ensuremath{\mathbf{Z}}) step ensures that corresponding inputs and outputs of submodels eventually match.
MAC algorithms have been developed for several nested models so far: deep nets \citep{CarreirWang14a}, low-dimensional SVMs \citep{WangCarreir14a}, binary autoencoders \citep{CarreirRaziper15a}, affinity-based loss functions for binary hashing \citep{RaziperCarreir16a} and parametric nonlinear embeddings \citep{CarreirVladym15a}. In this paper we focus mostly on the particular case of binary autoencoders. These define a nonconvex nondifferentiable problem, yet its MAC algorithm is simple and effective. It allows us to demonstrate, in an actual implementation in a distributed system, the fundamental properties of ParMAC: how MAC introduces parallelism; how ParMAC keeps the communication between machines low; the use of stochastic optimisation in the \ensuremath{\mathbf{W}}\ step; and the tradeoff between the different amount of parallelism in the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps. It also allows us to test how good our theoretical model of the speedup is in experiments. We first give the detailed MAC algorithm for binary autoencoders, and then generalise it to $K>1$ hidden layers.
\subsection{Optimising binary autoencoders using MAC}
\label{s:BAhash}
A \emph{binary autoencoder (BA)} is a usual autoencoder but with a binary code layer. It consists of an \emph{encoder} $\ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}})$ that maps a real vector $\ensuremath{\mathbf{x}}\in\ensuremath{\mathbb{R}}^D$ onto a \emph{binary} code vector with $L<D$ bits, $\ensuremath{\mathbf{z}}\in\{0,1\}^L$, and a linear \emph{decoder} $\ensuremath{\mathbf{f}}(\ensuremath{\mathbf{z}})$ which maps \ensuremath{\mathbf{z}}\ back to $\ensuremath{\mathbb{R}}^D$ in an effort to reconstruct \ensuremath{\mathbf{x}}. We will call \ensuremath{\mathbf{h}}\ a \emph{binary hash function} (see later). Let us write $\ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}}) = s(\ensuremath{\mathbf{A}}\ensuremath{\mathbf{x}})$ (\ensuremath{\mathbf{A}}\ includes a bias by having an extra dimension $x_0=1$ for each \ensuremath{\mathbf{x}}) where $\ensuremath{\mathbf{A}}\in\ensuremath{\mathbb{R}}^{L\times (D+1)}$ and $s(t)$ is a step function applied elementwise, i.e., $s(t) = 1$ if $t\ge 0$ and $s(t) = 0$ otherwise. Given a dataset of $D$-dimensional patterns $\ensuremath{\mathbf{X}} = (\ensuremath{\mathbf{x}}_1,\dots,\ensuremath{\mathbf{x}}_N)$, our objective function, which involves the nested model $\ensuremath{\mathbf{y}} = \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}}))$, is the usual least-squares reconstruction error:
\begin{equation}
\label{e:BA-nested}
E_{\text{BA}}(\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}}) = \sum^N_{n=1}{ \norm{\ensuremath{\mathbf{x}}_n - \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}}_n))}^2 }.
\end{equation}
Optimising this nonconvex, nonsmooth function is NP-complete. Where the gradients do exist with respect to \ensuremath{\mathbf{A}}\ they are zero, so optimisation of \ensuremath{\mathbf{h}}\ using chain-rule gradients does not apply. We introduce as auxiliary coordinates the outputs of \ensuremath{\mathbf{h}}, i.e., the codes for each of the $N$ input patterns, and obtain the following equality-constrained problem:
\begin{equation}
\label{e:BA-MAC-constrained}
\min_{\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}},\ensuremath{\mathbf{Z}}}{ \sum^N_{n=1}{ \norm{\ensuremath{\mathbf{x}}_n - \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{z}}_n)}^2 } } \quad \text{s.t.} \quad \ensuremath{\mathbf{z}}_n = \ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}}_n),\ \ensuremath{\mathbf{z}}_n\in\{0,1\}^L,\ n=1,\dots,N.
\end{equation}
Note the codes are binary. We now apply the quadratic-penalty method (it is also possible to apply the augmented Lagrangian method; \citealp{NocedalWright06a}) and minimise the following objective function while progressively increasing $\mu$, so the constraints are eventually satisfied:
\begin{equation}
\label{e:BA-MAC-QP}
E_Q(\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}},\ensuremath{\mathbf{Z}};\mu) = \sum^N_{n=1}{ \left( \norm{\ensuremath{\mathbf{x}}_n - \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{z}}_n)}^2 + \mu \norm{\ensuremath{\mathbf{z}}_n - \ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}}_n)}^2 \right) } \text{ s.t.\ } \ensuremath{\mathbf{z}}_n\in\{0,1\}^L,\ n=1,\dots,N.
\end{equation}
Finally, we apply alternating optimisation over \ensuremath{\mathbf{Z}}\ and $\ensuremath{\mathbf{W}} = (\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}})$. This results in the following two steps:
\begin{itemize}
\item Over \ensuremath{\mathbf{Z}}\ for fixed $(\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}})$, this is a binary optimisation on $NL$ variables, but it separates into $N$ independent optimisations each on only $L$ variables, with the form of a binary proximal operator (where we omit the index $n$): $\min_{\ensuremath{\mathbf{z}}}{ \norm{\ensuremath{\mathbf{x}} - \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{z}})}^2 + \mu \norm{\ensuremath{\mathbf{z}} - \ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}})}^2 }$ s.t.\ $\ensuremath{\mathbf{z}}\in\{0,1\}^L$. After some transformations, this problem can be solved exactly for small $L$ by enumeration or approximately for larger $L$ by alternating optimisation over bits, initialised by solving the relaxed problem to $[0,1]$ and truncating its solution (see \citealp{CarreirRaziper15a} for details).
\item Over $\ensuremath{\mathbf{W}} = (\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}})$ for fixed \ensuremath{\mathbf{Z}}, we obtain $L+D$ independent problems: for each of the $L$ single-bit hash functions (which try to predict \ensuremath{\mathbf{Z}}\ optimally from \ensuremath{\mathbf{X}}), each solvable by fitting a linear SVM; and for each of the $D$ linear decoders in \ensuremath{\mathbf{f}}\ (which try to reconstruct \ensuremath{\mathbf{X}}\ optimally from \ensuremath{\mathbf{Z}}), each a linear least-squares problem. With linear \ensuremath{\mathbf{h}}\ and \ensuremath{\mathbf{f}}\ this simply involves fitting $L$ SVMs to $(\ensuremath{\mathbf{X}},\ensuremath{\mathbf{Z}})$ and $D$ linear regressors to $(\ensuremath{\mathbf{Z}},\ensuremath{\mathbf{X}})$.
\end{itemize}
The user must choose a schedule for the penalty parameter $\mu$ (sequence of values $0 < \mu_1 < \dots < \infty$). This should increase slowly enough that the binary codes can change considerably and explore better solutions before the constraints are satisfied and the algorithm stops. With BAs, MAC stops for a finite value of $\mu$ \citep{CarreirRaziper15a}. This occurs whenever \ensuremath{\mathbf{Z}}\ does not change compared to the previous \ensuremath{\mathbf{Z}}\ step, which gives a practical stopping criterion. Also, in order to generalise well to unseen data, we stop iterating for a $\mu$ value not when we (sufficiently) optimise $E_Q(\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}},\ensuremath{\mathbf{Z}};\mu)$, but when the precision of the hash function in a validation set decreases. This is a form of early stopping that guarantees that we improve (or leave unchanged) the initial \ensuremath{\mathbf{Z}}, and besides is faster. We also have to initialise \ensuremath{\mathbf{Z}}. This can be done by running PCA and binarising its result, for example. Fig.~\ref{f:BA-alg} gives the MAC algorithm for BAs.
\begin{figure}[t]
\centering
\setlength{\fboxsep}{1ex}
\framebox{%
\begin{minipage}[c]{0.70\linewidth}
\begin{tabbing}
n \= n \= n \= n \= n \= \kill
\underline{\textbf{input}} $\ensuremath{\mathbf{X}}_{D \times N} = (\ensuremath{\mathbf{x}}_1,\dots,\ensuremath{\mathbf{x}}_N)$, $L \in \ensuremath{\mathbb{N}}$ \\
Initialise $\ensuremath{\mathbf{Z}}_{L \times N} = (\ensuremath{\mathbf{z}}_1,\dots,\ensuremath{\mathbf{z}}_N) \in \{0,1\}^{LN}$ \\
\underline{\textbf{for}} $\mu = 0 < \mu_1 < \dots < \mu_{\infty}$ \+ \\
\underline{\textbf{for}} $l = 1,\dots,L$ \` {\small\textsf{\ensuremath{\mathbf{W}}\ step: \ensuremath{\mathbf{h}}}} \+ \\
$h_l \leftarrow$ fit SVM to $(\ensuremath{\mathbf{X}},\ensuremath{\mathbf{Z}}_{\cdot l})$ \- \\
$\ensuremath{\mathbf{f}} \leftarrow$ least-squares fit to $(\ensuremath{\mathbf{Z}},\ensuremath{\mathbf{X}})$ \` {\small\textsf{\ensuremath{\mathbf{W}}\ step: \ensuremath{\mathbf{f}}}} \\
\underline{\textbf{for}} $n = 1,\dots,N$ \` {\small\textsf{\ensuremath{\mathbf{Z}}\ step}} \+ \\
$\ensuremath{\mathbf{z}}_n \leftarrow \argmin_{\ensuremath{\mathbf{z}}_n\in\{0,1\}^L}{\norm{\ensuremath{\mathbf{x}}_n-\ensuremath{\mathbf{f}}(\ensuremath{\mathbf{z}}_n)}^2 + \mu \norm{\ensuremath{\mathbf{z}}_n-\ensuremath{\mathbf{h}}(\ensuremath{\mathbf{x}}_n)}^2}$ \- \\
\underline{\textbf{if}} no change in \ensuremath{\mathbf{Z}}\ and $\ensuremath{\mathbf{Z}} = \ensuremath{\mathbf{h}}(\ensuremath{\mathbf{X}})$ \underline{\textbf{then}} stop \- \\
\underline{\textbf{return}} \ensuremath{\mathbf{h}}, $\ensuremath{\mathbf{Z}} = \ensuremath{\mathbf{h}}(\ensuremath{\mathbf{X}})$
\end{tabbing}
\end{minipage}
}
\caption{Binary autoencoder MAC algorithm.}
\label{f:BA-alg}
\end{figure}
The BA was proposed as a way to learn good binary hash functions for fast, approximate information retrieval \citep{CarreirRaziper15a}. Binary hashing \citep{GraumanFergus13a} has emerged in recent years as an effective way to do fast, approximate nearest-neighbour searches in image databases. The real-valued, high-dimensional image vectors are mapped onto a binary space with $L$ bits and the search is performed there using Hamming distances at a vastly faster speed and smaller memory (e.g.\ $N=10^9$ points with $D=500$ take 2 TB, but only 8 GB using $L=64$ bits, which easily fits in RAM). As shown by \citet{CarreirRaziper15a}, training BAs with MAC beats approximate optimisation approaches such as relaxing the codes or the step function in the encoder, and yields state-of-the-art binary hash functions \ensuremath{\mathbf{h}}\ in unsupervised problems, improving over established approaches such as iterative quantisation (ITQ) \citep{Gong_13a}.
In this paper, we focus on linear hash functions because these are, by far, the most used type of hash functions in the literature of binary hashing, due to the fact that computing the binary codes for a test image must be fast at run time. We also provide an experiment with nonlinear hash functions (RBF network).
\subsection{MAC with $K$ layers}
\label{s:MAC:Klayers}
We now consider the more general case of MAC with $K$ hidden layers \citep{CarreirWang12a,CarreirWang14a}, inputs \ensuremath{\mathbf{x}}\ and outputs \ensuremath{\mathbf{y}}\ (for a BA, $\ensuremath{\mathbf{x}} = \ensuremath{\mathbf{y}}$). It helps to think of the case of a deep net and we will use it as a running example, but the ideas apply beyond deep nets. Consider a regression problem of mapping inputs \ensuremath{\mathbf{x}}\ to outputs \ensuremath{\mathbf{y}}\ (both high-dimensional) with a deep net $\ensuremath{\mathbf{f}}(\ensuremath{\mathbf{x}})$ given a dataset of $N$ pairs $(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n)$. We minimise the least-squares error (other loss functions are possible):
\begin{equation}
\label{e:nested}
E(\ensuremath{\mathbf{W}}) = \frac{1}{2} \sum^N_{n=1}{\norm{\ensuremath{\mathbf{y}}_n - \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{x}}_n;\ensuremath{\mathbf{W}})}^2} \qquad \ensuremath{\mathbf{f}}(\ensuremath{\mathbf{x}};\ensuremath{\mathbf{W}}) = \ensuremath{\mathbf{f}}_{K+1}(\dots \ensuremath{\mathbf{f}}_2(\ensuremath{\mathbf{f}}_1(\ensuremath{\mathbf{x}};\ensuremath{\mathbf{W}}_1);\ensuremath{\mathbf{W}}_2)\dots;\ensuremath{\mathbf{W}}_{K+1})
\end{equation}
where each layer function has the form $\ensuremath{\mathbf{f}}_k(\ensuremath{\mathbf{x}};\ensuremath{\mathbf{W}}_k) = \sigma(\ensuremath{\mathbf{W}}_k\ensuremath{\mathbf{x}})$, i.e., a linear mapping followed by a squashing nonlinearity ($\sigma(t)$ applies a scalar function, such as the sigmoid $1/(1+e^{-t})$, elementwise to a vector argument, with output in $[0,1]$). We introduce one auxiliary variable per data point and per hidden unit and define the following equality-constrained optimisation problem:
\begin{equation}
\label{e:mac}
\frac{1}{2} \sum^N_{n=1}{\norm{\ensuremath{\mathbf{y}}_n - \ensuremath{\mathbf{f}}_{K+1}(\ensuremath{\mathbf{z}}_{K,n};\ensuremath{\mathbf{W}}_{K+1})}^2} \text{ s.t.\ }
\renewcommand{\arraystretch}{0.5}
\left\{
\begin{array}{@{}l@{}}
\ensuremath{\mathbf{z}}_{K,n} = \ensuremath{\mathbf{f}}_K(\ensuremath{\mathbf{z}}_{K-1,n};\ensuremath{\mathbf{W}}_K) \\ \dots \\ \ensuremath{\mathbf{z}}_{1,n} = \ensuremath{\mathbf{f}}_1(\ensuremath{\mathbf{x}}_n;\ensuremath{\mathbf{W}}_1)
\end{array}
\right\} n=1,\dots,N.
\end{equation}
Each $\ensuremath{\mathbf{z}}_{k,n}$ can be seen as the coordinates of $\ensuremath{\mathbf{x}}_n$ in an intermediate feature space, or as the hidden unit activations for $\ensuremath{\mathbf{x}}_n$. Intuitively, by eliminating \ensuremath{\mathbf{Z}}\ we see this is equivalent to the nested problem~\eqref{e:nested}; we can prove under very general assumptions that both problems have exactly the same minimisers \citep{CarreirWang12a}. Applying the quadratic-penalty method, we optimise the following function:
\begin{equation}
\label{e:mac-quadpen}
E_Q(\ensuremath{\mathbf{W}},\ensuremath{\mathbf{Z}};\mu) = \frac{1}{2} \sum^N_{n=1}{\norm{\ensuremath{\mathbf{y}}_n - \ensuremath{\mathbf{f}}_{K+1}(\ensuremath{\mathbf{z}}_{K,n};\ensuremath{\mathbf{W}}_{K+1})}^2} + \frac{\mu}{2} \sum^N_{n=1}{\sum^K_{k=1}{\norm{\ensuremath{\mathbf{z}}_{k,n} - \ensuremath{\mathbf{f}}_k(\ensuremath{\mathbf{z}}_{k-1,n};\ensuremath{\mathbf{W}}_k)}^2}}
\end{equation}
over $(\ensuremath{\mathbf{W}},\ensuremath{\mathbf{Z}})$ and drive $\mu \rightarrow \infty$. This defines a continuous path $(\ensuremath{\mathbf{W}}^*(\mu),\ensuremath{\mathbf{Z}}^*(\mu))$ which, under mild assumptions \citep{CarreirWang12a}, converges to a minimum of the constrained problem~\eqref{e:mac}, and thus to a minimum of the original problem~\eqref{e:nested}. In practice, we follow this path loosely. The quadratic-penalty objective function can be seen as breaking the functional dependences in the nested mapping \ensuremath{\mathbf{f}}\ and unfolding it over layers. Every squared term involves only a shallow mapping; all variables $(\ensuremath{\mathbf{W}},\ensuremath{\mathbf{Z}})$ are equally scaled, which improves the conditioning of the problem; and the derivatives required are simpler: we require no backpropagated gradients over \ensuremath{\mathbf{W}}, and sometimes no gradients at all. We now apply alternating optimisation of the quadratic-penalty objective over \ensuremath{\mathbf{Z}}\ and \ensuremath{\mathbf{W}}:
\begin{description}
\item[\ensuremath{\mathbf{W}}\ step (submodels)] Minimising over \ensuremath{\mathbf{W}}\ for fixed \ensuremath{\mathbf{Z}}\ results in a separate minimisation over the weights of each hidden unit---each a single-layer, single-unit \emph{submodel} that can be solved with existing algorithms (logistic regression).
\item[\ensuremath{\mathbf{Z}}\ step (coordinates)] Minimising over \ensuremath{\mathbf{Z}}\ for fixed \ensuremath{\mathbf{W}}\ separates over the coordinates $\ensuremath{\mathbf{z}}_n$ for each data point $n=1,\dots,N$ and can be solved using the derivatives with respect to \ensuremath{\mathbf{z}}\ of the single-layer functions $\ensuremath{\mathbf{f}}_1,\dots,\ensuremath{\mathbf{f}}_{K+1}$ (omitting the subindex $n$): $\smash{\min_{\ensuremath{\mathbf{z}}}{ \norm{\ensuremath{\mathbf{y}} - \ensuremath{\mathbf{f}}_{K+1}(\ensuremath{\mathbf{z}}_K)}^2 + \mu \sum^K_{k=1}{\norm{\ensuremath{\mathbf{z}}_k - \ensuremath{\mathbf{f}}_k(\ensuremath{\mathbf{z}}_{k-1})}^2} }}$.
\end{description}
Thus, the \ensuremath{\mathbf{W}}\ step results in many independent, single-layer single-unit submodels that can be trained with existing algorithms, without extra programming cost. The \ensuremath{\mathbf{Z}}\ step is new and has the form of a ``generalised'' proximal operator \citep{Rockaf76b,CombetPesquet11a}. MAC reduces a complex, highly-coupled problem---training a deep net---to a sequence of simple, uncoupled problems (the \ensuremath{\mathbf{W}}\ step) which are coordinated through the auxiliary variables (the \ensuremath{\mathbf{Z}}\ step). For a large net with a large dataset, this affords an enormous potential for parallel computation.
\section{ParMAC: a parallel, distributed computation model for MAC}
\label{s:ParMAC}
We now turn to the contribution of this paper, the distributed implementation of MAC algorithms. As we have seen, a specific MAC algorithm depends on the model and objective function and on how the auxiliary coordinates are introduced. We can achieve steps that are closed-form, convex, nonconvex, binary, or others. However, the following always hold: (1) In the \ensuremath{\mathbf{Z}}\ step, \emph{the $N$ subproblems for $\ensuremath{\mathbf{z}}_1,\dots,\ensuremath{\mathbf{z}}_N$ are independent, one per data point}. Each $\ensuremath{\mathbf{z}}_n$ step depends on all or part of the current model. (2) In the \ensuremath{\mathbf{W}}\ step, there are $M$ \emph{independent submodels}, where $M$ depends on the problem. For example, $M$ is the number of hidden units in a deep net, or the number of hash functions and linear decoders in a BA. Each submodel depends on all the data and coordinates (usually, a given submodel depends, for each $n$, on only a portion of the vector $(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n,\ensuremath{\mathbf{z}}_n)$). We now show how to turn this into a distributed, low-communication \emph{ParMAC} algorithm, give an MPI implementation of ParMAC for BAs, and discuss the convergence of ParMAC. Throughout the paper, unless otherwise indicated, we will use the term ``machine'' to mean a single-CPU processing unit with its own local memory and disk, which can communicate with other machines in a cluster through a network or shared memory.
\subsection{Description of ParMAC}
The basic idea in ParMAC is as follows. With large datasets in distributed systems, it is imperative to minimise data movement over the network because the communication time generally far exceeds the computation time in modern architectures. In MAC we have 3 types of data: the original training data $(\ensuremath{\mathbf{X}},\ensuremath{\mathbf{Y}})$, the auxiliary coordinates \ensuremath{\mathbf{Z}}, and the model parameters (the submodels). Usually, the latter type is far smaller. \emph{In ParMAC, we never communicate training or coordinate data; each machine keeps a disjoint portion of $(\ensuremath{\mathbf{X}},\ensuremath{\mathbf{Y}},\ensuremath{\mathbf{Z}})$ corresponding to a subset of the points. Only model parameters are communicated, during the \ensuremath{\mathbf{W}}\ step, following a circular topology, which implicitly implements a stochastic optimisation}. The model parameters are the hash functions \ensuremath{\mathbf{h}}\ and the decoder \ensuremath{\mathbf{f}}\ for BAs, and the weight vector $\ensuremath{\mathbf{w}}_h$ of each hidden unit $h$ for deep nets. Let us see this in detail (refer to fig.~\ref{f:ParMAC}).
\begin{figure}[t]
\scriptsize
\psfrag{data}[][][1][90]{\normalsize Data}
\psfrag{model}[][][1][90]{\normalsize Model}
\psfrag{P1}[t][t]{\normalsize Machine 1}
\psfrag{P2}[t][t]{\normalsize Machine 2}
\psfrag{P3}[t][t]{\normalsize Machine 3}
\psfrag{P4}[t][t]{\normalsize Machine 4}
\psfrag{wm}[][]{\normalsize $\ensuremath{\mathbf{w}}_h$}
\psfrag{w01}[r][Br]{$1$}
\psfrag{w02}[r][Br]{$2$}
\psfrag{w03}[r][Br]{$3$}
\psfrag{w04}[r][Br]{$4$}
\psfrag{w05}[r][Br]{$5$}
\psfrag{w06}[r][Br]{$6$}
\psfrag{w07}[r][Br]{$7$}
\psfrag{w08}[r][Br]{$8$}
\psfrag{w09}[r][Br]{$9$}
\psfrag{w10}[r][Br]{$10$}
\psfrag{w11}[r][Br]{$11$}
\psfrag{w12}[r][Br]{$12$}
\psfrag{w13}[r][Br]{$13$}
\psfrag{w14}[r][Br]{$14$}
\psfrag{w15}[r][Br]{$15$}
\psfrag{w16}[r][Br]{$16$}
\psfrag{w17}[r][Br]{$17$}
\psfrag{w18}[r][Br]{$18$}
\psfrag{w19}[r][Br]{$19$}
\psfrag{w20}[r][Br]{$20$}
\psfrag{w21}[r][Br]{$21$}
\psfrag{w22}[r][Br]{$22$}
\psfrag{w23}[r][Br]{$23$}
\psfrag{w24}[r][Br]{$24$}
\psfrag{w25}[r][Br]{$25$}
\psfrag{w26}[r][Br]{$26$}
\psfrag{w27}[r][Br]{$27$}
\psfrag{w28}[r][Br]{$28$}
\psfrag{w29}[r][Br]{$29$}
\psfrag{w30}[r][Br]{$30$}
\psfrag{w31}[r][Br]{$31$}
\psfrag{w32}[r][Br]{$32$}
\psfrag{w33}[r][Br]{$33$}
\psfrag{w34}[r][Br]{$34$}
\psfrag{w35}[r][Br]{$35$}
\psfrag{w36}[r][Br]{$36$}
\psfrag{w37}[r][Br]{$37$}
\psfrag{w38}[r][Br]{$38$}
\psfrag{w39}[r][Br]{$39$}
\psfrag{w40}[r][Br]{$40$}
\psfrag{w41}[r][Br]{$41$}
\psfrag{w42}[r][Br]{$42$}
\psfrag{w43}[r][Br]{$43$}
\psfrag{w44}[r][Br]{$44$}
\psfrag{w45}[r][Br]{$45$}
\psfrag{w46}[r][Br]{$46$}
\psfrag{w47}[r][Br]{$47$}
\psfrag{w48}[r][Br]{$48$}
\psfrag{xn}[B][B]{\normalsize $\ensuremath{\mathbf{x}}_n$}
\psfrag{yn}[B][B]{\normalsize $\ensuremath{\mathbf{y}}_n$}
\psfrag{zn}[B][B]{\normalsize $\ensuremath{\mathbf{z}}_n$}
\psfrag{x01}[r][Br]{$1$}
\psfrag{x02}[r][Br]{$2$}
\psfrag{x03}[r][Br]{$3$}
\psfrag{x04}[r][Br]{$4$}
\psfrag{x10}[r][Br]{$10$}
\psfrag{x11}[r][Br]{$11$}
\psfrag{x12}[r][Br]{$12$}
\psfrag{x13}[r][Br]{$13$}
\psfrag{x14}[r][Br]{$14$}
\psfrag{x20}[r][Br]{$20$}
\psfrag{x21}[r][Br]{$21$}
\psfrag{x22}[r][Br]{$22$}
\psfrag{x23}[r][Br]{$23$}
\psfrag{x24}[r][Br]{$24$}
\psfrag{x30}[r][Br]{$30$}
\psfrag{x31}[r][Br]{$31$}
\psfrag{x32}[r][Br]{$32$}
\psfrag{x33}[r][Br]{$33$}
\psfrag{x34}[r][Br]{$34$}
\psfrag{x40}[r][Br]{$40$}
\includegraphics[width=\linewidth]{ParMAC3.eps}
\caption{ParMAC model with $P=4$ machines, $M=12$ submodels and $N=40$ data points. ``$\ensuremath{\mathbf{w}}_h$'' represents the submodels (hash functions and decoders for BAs, hidden unit weight vectors for deep nets). Submodels $h$, $h+M$, $h+2M$ and $h+3M$ are copies of submodel $h$, but only one of them is the most currently updated. At the end of the \ensuremath{\mathbf{W}}\ step all copies are identical.}
\label{f:ParMAC}
\end{figure}
Assume for simplicity we have $P$ identical processing machines, each with their own memory and CPU, which are connected through a network. The machines are connected in a circular (ring) unidirectional topology, i.e., machine $1$ $\rightarrow$ machine $2$ $\rightarrow \cdots \rightarrow$ machine $P$ $\rightarrow$ machine $1$, where ``machine $p$ $\rightarrow$ machine $q$'' means machine $p$ can send data directly to machine $q$ (and we say machine $q$ is the successor of machine $p$). Call $\ensuremath{\mathcal{D}} = \{(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n,\ensuremath{\mathbf{z}}_n)\mathpunct{:}\ n \in \{1,\dots,N\}\}$ the entire dataset and corresponding coordinates. Each machine $p$ will store a subset $\ensuremath{\mathcal{D}}_p = \{(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n,\ensuremath{\mathbf{z}}_n)\mathpunct{:}\ n \in \ensuremath{\mathcal{I}}_p\}$ such that the subsets are disjoint and their union is the entire data, i.e., the index sets satisfy $\ensuremath{\mathcal{I}}_p \cap \ensuremath{\mathcal{I}}_q = \varnothing$ if $p \neq q$ and $\cup^P_{p=1}{\ensuremath{\mathcal{I}}_p} = \{1,\dots,N\}$.
The \ensuremath{\mathbf{Z}}\ step is very simple. Before the \ensuremath{\mathbf{Z}}\ step starts%
\footnote{Also, the machines need not start all at the same time in the \ensuremath{\mathbf{Z}}\ step. A machine can start the \ensuremath{\mathbf{Z}}\ step on its data as soon as it has received all the updated submodels in the \ensuremath{\mathbf{W}}\ step. Likewise, as soon as a machine finishes its \ensuremath{\mathbf{Z}}\ step, it can start the \ensuremath{\mathbf{W}}\ step immediately, without waiting for all other machines to finish their \ensuremath{\mathbf{Z}}\ step. However, in our implementation we consider the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps as barriers, so that all machines start the \ensuremath{\mathbf{W}}\ or \ensuremath{\mathbf{Z}}\ step at the same time.},
each machine will contain all the (just updated) submodels. This means that in the \ensuremath{\mathbf{Z}}\ step each machine $p$ processes its auxiliary coordinates $\{\ensuremath{\mathbf{z}}_n\mathpunct{:}\ n\in\ensuremath{\mathcal{I}}_p\}$ independently of all other machines, i.e., no communication occurs.
The \ensuremath{\mathbf{W}}\ step is more subtle. At the beginning of the \ensuremath{\mathbf{W}}\ step, each machine will contain all the submodels and its portion of the data and (just updated) coordinates. Each submodel must have access to the entire data and coordinates in order to update itself and, since the data cannot leave its home machine, the submodel must go to the data (this contrasts with the intuitive notion of the model sitting in a computer while data arrive and are processed). We achieve this in the circular topology as follows. We assume synchronous processing for simplicity, but in practice one would implement this asynchronously. Assume arithmetic modulo $P$ and an imaginary clock whose period equals the time that any one machine takes to process its portion $M/P$ of submodels. At each clock tick, the $P$ machines update each a different portion $M/P$ of the submodels. For example, in fig.~\ref{f:ParMAC}, at clock tick 1 machine $1$ updates submodels $1$--$3$ using its data $\ensuremath{\mathcal{D}}_{\ensuremath{\mathcal{I}}_1}$ (where $\ensuremath{\mathcal{I}}_1 = \{1,\dots,10\}$); machine $2$ updates submodels $4$--$6$; machine $3$ updates submodels $7$--$9$; and machine $4$ updates submodels $10$--$12$. This happens in parallel. Then each machine sends the submodels updated to its successor, also in parallel. In the next tick, each machine updates the submodels it just received, i.e., machine $1$ updates $10$--$12$, machine $2$ updates submodels $1$--$3$, machine $3$ updates submodels $4$--$6$; and machine $4$ updates submodels $7$--$9$ (and each machine always uses its data portion, which never changes). This is repeated until each submodel has visited each machine and thus has been updated with the entire dataset \ensuremath{\mathcal{D}}. This happens after $P$ ticks, and we call this an \emph{epoch}. This process may be repeated for $e$ epochs in $eP$ ticks. At this time, each machine contains $M/P$ submodels that are finished (i.e., updated $e$ times over the entire dataset), and the remaining $M(1-1/P)$ submodels it contains are not finished, indeed the finished versions of those submodels reside in other machines. Finally, before starting with the \ensuremath{\mathbf{Z}}\ step, each machine must contain all the (just updated) submodels (i.e., the parameters for the entire nested model). We achieve this%
\footnote{In MPI, this can be directly achieved with \texttt{MPI\_Alltoall} broadcasting, which scatters/gathers data from all members to all members of a group (a complete exchange). However, in this paper we implement it using the circular topology mechanism described.}
by running a final round of communication without computation, i.e., each machine sends its just updated submodels to its successor. Thus, after one clock tick, machine $p$ sends $M/P$ final submodels to machine $p+1$ and receives $M/P$ submodels from machine $p-1$. After $P-1$ clock ticks, each machine has received the remaining $M(1-1/P)$ submodels that were finished by other machines, hence each machine contains a (redundant) copy of all the current submodels. Fig.~\ref{f:ParMAC-anim} illustrates the sequence of operations during one epoch for the example of fig.~\ref{f:ParMAC}.
\begin{figure}[p]
\psfrag{P1}{}
\psfrag{P2}{}
\psfrag{P3}{}
\psfrag{P4}{}
\psfrag{wm}{}
\psfrag{w01}{}
\psfrag{w02}{}
\psfrag{w03}{}
\psfrag{w04}{}
\psfrag{w05}{}
\psfrag{w06}{}
\psfrag{w07}{}
\psfrag{w08}{}
\psfrag{w09}{}
\psfrag{w10}{}
\psfrag{w11}{}
\psfrag{w12}{}
\psfrag{w13}{}
\psfrag{w14}{}
\psfrag{w15}{}
\psfrag{w16}{}
\psfrag{w17}{}
\psfrag{w18}{}
\psfrag{w19}{}
\psfrag{w20}{}
\psfrag{w21}{}
\psfrag{w22}{}
\psfrag{w23}{}
\psfrag{w24}{}
\psfrag{w25}{}
\psfrag{w26}{}
\psfrag{w27}{}
\psfrag{w28}{}
\psfrag{w29}{}
\psfrag{w30}{}
\psfrag{w31}{}
\psfrag{w32}{}
\psfrag{w33}{}
\psfrag{w34}{}
\psfrag{w35}{}
\psfrag{w36}{}
\psfrag{w37}{}
\psfrag{w38}{}
\psfrag{w39}{}
\psfrag{w40}{}
\psfrag{w41}{}
\psfrag{w42}{}
\psfrag{w43}{}
\psfrag{w44}{}
\psfrag{w45}{}
\psfrag{w46}{}
\psfrag{w47}{}
\psfrag{w48}{}
\begin{tabular}{@{}c@{}c@{}c@{}c@{}c@{}}
\psfrag{model}[][]{\caja{c}{c}{tick \\ 1}}
\includegraphics[width=\linewidth,height=0.25\linewidth]{ParMAC-anim1small.eps} \\
\psfrag{model}[][]{\caja{c}{c}{tick \\ 2}}
\includegraphics[width=\linewidth,height=0.25\linewidth]{ParMAC-anim2small.eps} \\
\psfrag{model}[][]{\caja{c}{c}{tick \\ 3}}
\includegraphics[width=\linewidth,height=0.25\linewidth]{ParMAC-anim3small.eps} \\
\psfrag{model}[][]{\caja{c}{c}{tick \\ 4}}
\includegraphics[width=\linewidth,height=0.25\linewidth]{ParMAC-anim4small.eps} \\
\psfrag{P1}[t][t]{\normalsize Machine 1}
\psfrag{P2}[t][t]{\normalsize Machine 2}
\psfrag{P3}[t][t]{\normalsize Machine 3}
\psfrag{P4}[t][t]{\normalsize Machine 4}
\psfrag{model}[][]{\caja{c}{c}{tick \\ 5}}
\includegraphics[width=\linewidth,height=0.25\linewidth]{ParMAC-anim5small.eps}
\end{tabular}
\caption{Illustration of one epoch of the synchronous version of ParMAC's \ensuremath{\mathbf{W}}\ step for the example of fig.~\ref{f:ParMAC} with $P=4$ machines and $M=12$ submodels (we only show the ``model'' part of the figure). Each row corresponds to one clock tick, within which each machine computes on its portion of $M/P$ submodels (coloured gold) and sends them to its successor. The last tick ($= P+1$) is the start of the next epoch, at which point all submodels (coloured green) have been updated over the entire dataset.}
\label{f:ParMAC-anim}
\end{figure}
In practice, we use an asynchronous implementation. Each machine keeps a queue of submodels to be processed, and repeatedly performs the following operations: extract a submodel from the queue, process it (except in epoch $e+1$) and send it to the machine's successor (which will insert it in its queue). If the queue is empty, the machine waits until it is nonempty. The queue of each machine is initialised with the portion of submodels associated with that machine. Each submodel carries a counter that is initially $1$ and increases every time it visits a machine. When it reaches $Pe$ then the submodel is in the last machine and the last epoch. When it reaches $P(e+1)-1$, it has undergone $e$ epochs of processing and all machines have a copy of it, so it has finished the \ensuremath{\mathbf{W}}\ step.
Since each submodel is updated as soon as it visits a machine, rather than computing the exact gradient once it has visited all machines and then take a step, the \ensuremath{\mathbf{W}}\ step is really carrying out \emph{stochastic steps for each submodel}. For example, if the update is done by a gradient step, we are actually implementing stochastic gradient descent (SGD) where the minibatches are of size $N/P$ (or smaller, if we subdivide a machine's data portion into minibatches, which should be typically the case in practice). From this point of view, we can regard the \ensuremath{\mathbf{W}}\ step as doing SGD on each submodel in parallel by having each submodel visit the minibatches in each machine.
In summary, using $P$ machines, ParMAC iterates as follows:
\begin{description}
\item[\ensuremath{\mathbf{W}}\ step] The submodels (hash functions and decoders for BAs) visit each machine. This implies we train them with stochastic gradient descent, where one ``epoch'' for a submodel corresponds to that submodel having visited all $P$ machines. All submodels are communicated in parallel, asynchronously with respect to each other, in a circular topology. With $e$ epochs, the entire model parameters are communicated $e+1$ times. The last round of communication is needed to ensure each machine has the most updated version of the model for the \ensuremath{\mathbf{Z}}\ step.
\item[\ensuremath{\mathbf{Z}}\ step] Identical to MAC, each data point's coordinates $\ensuremath{\mathbf{z}}_n$ are optimised independently, in parallel over machines (since each machine contains $\ensuremath{\mathbf{x}}_n$, $\ensuremath{\mathbf{y}}_n$, $\ensuremath{\mathbf{z}}_n$, and all the model parameters). No communication occurs at all.
\end{description}
\subsection{A \ensuremath{\mathbf{W}}\ step with only two rounds of communication}
As described, and as implemented in our experiments, running $e$ epochs in the \ensuremath{\mathbf{W}}\ step requires $e$ rounds of communication (plus a final round). However, we can run $e$ epochs with only 1 round of communication \emph{by having a submodel do $e$ consecutive passes within each machine's data}. In the example of fig.~\ref{f:ParMAC}, running $e=2$ epochs for submodel $\ensuremath{\mathbf{w}}_1$ means the following: instead of visiting the data as 1,\dots,10, 11,\dots,20, 21,\dots,30, 31,\dots,40, 1,\dots,10, 11,\dots,20, 21,\dots,30, 31,\dots,40, it visits the data as 1,\dots,10, 1,\dots,10, 11,\dots,20, 11,\dots,20, 21,\dots,30, 21,\dots,30, 31,\dots,40, 31,\dots,40. (We can also have intermediate schemes such as doing between 1 and $e$ within-machine passes.) This reduces the amount of shuffling, but should not be a problem if the data are randomly distributed over machines (and we can still do within-machine shuffling). This effectively reduces the total communication in the \ensuremath{\mathbf{W}}\ step to 2 rounds regardless of the number of epochs $e$ (with the second round needed to ensure each machine has the most updated submodels).
\subsection{Extensions of ParMAC}
\label{s:ParMAC:extensions}
In addition, the ParMAC model offers good potential for data shuffling, load balancing, streaming and fault tolerance, which make it attractive for big data. We describe these next.
\paragraph{Data shuffling}
\label{s:shuffling}
It is well known that shuffling (randomly reordering) the dataset prior to each epoch improves the SGD convergence speed. With distributed systems, this can sometimes be a problem and require data movement across machines. Shuffling is easy in ParMAC. Within a machine, we can simply access the local data (minibatches) in random order at each epoch. Across machines, we can simply reorganise the circular topology randomly (while still circular) at the beginning of each new epoch (by generating a random permutation and resetting the successor's address of each machine). We could even have each submodel follow a different, random circular topology. However, we do not implement this because it is unlikely to help (since the submodels are independent) and can unbalance the load over machines.
\paragraph{Load balancing}
\label{s:load-balancing}
This is simple because the work in both the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps is proportional to the number of data points $N$. Indeed, in the \ensuremath{\mathbf{W}}\ step each submodel must visit every data point once per epoch. So, even if the submodels differ in size, the training of any submodel is proportional to $N$. In the \ensuremath{\mathbf{Z}}\ step, each data point is a separate problem dependent on the current model (which is the same for all points), thus all $N$ problems are formally identical in complexity. Hence, in the assumption that the machines are identical and that each data point incurs the same runtime, load balancing is trivial: the $N$ points are allocated in equal portions of $N/P$ to each machine. If the processing power of machine $p$ is proportional to $\alpha_p > 0$ (where $\alpha_p$ could represent the clock frequency of machine $p$, say), then we allocate to machine $p$ a subset of the $N$ points proportional to $\alpha_p$, i.e., machine $p$ gets $N \alpha_p / (\alpha_1 + \dots + \alpha_P)$ data points. This is done once and for all at loading time.
In practice, we can expect some degradation of the parallel speedup even with identical machines and submodels of the same type. This is because machines do vary for various reasons, e.g.\ the runtime can be affected by differences in ventilation across machines located in different areas of a data centre, or because machines are running other user processes in addition to the ParMAC optimisation. Another type of degradation can happen if the submodels differ significantly in runtime (e.g.\ because there are different types of submodels): the runtime of the \ensuremath{\mathbf{W}}\ step will be driven by the slow submodels, which become a bottleneck. As discussed in section~\ref{s:speedup-th:practical}, we can group the $M$ submodels into a smaller number $M' < M$ of approximately equal-size aggregate submodels, for the purpose of estimating the speedup in theory. This need not be the fastest way to schedule the jobs, and in practice we still process the individual submodels asynchronously.
\paragraph{Streaming}
\label{s:streaming}
Streaming refers to the ability to discard old data and to add new data from training over time. This is useful in online learning, or to allow the data to be refreshed, but also may be necessary when a machine collects more data than it can store. The circular topology allows us to add or remove machines on the fly easily, and this can be used to implement streaming.
We consider two forms of streaming: (1) new data are added within a machine (e.g.\ as this machine collects new data), and likewise old data are discarded within a machine. And (2) new data are added by adding a new machine to the topology, and old data are discarded by removing an existing machine from the topology. Both forms are easily achieved in ParMAC. The first form, within-machine, is trivial: a machine can always add or remove data without any change to the system, because the data for each note is private and never interacts with other machines other than by updating submodels. Adding or discarding data is done at the beginning of the \ensuremath{\mathbf{Z}}\ step. Discarding data simply means removing the corresponding $\{(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n,\ensuremath{\mathbf{z}}_n)\}$ from that machine. Adding data means inserting $\{(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n)\}$ in that machine and, if necessary, creating within that machine coordinate values $\{\ensuremath{\mathbf{z}}_n\}$ (e.g.\ by applying the nested model to $\ensuremath{\mathbf{x}}_n$). We never upload or send any \ensuremath{\mathbf{z}}\ values over the network.
The second form, creating a new machine or removing an existing one, is barely more complicated, assuming some support from the parallel processing library. We describe it conceptually. Imagine we currently have $P$ machines. We can add a new machine, with its own preloaded data $\{(\ensuremath{\mathbf{x}}_n,\ensuremath{\mathbf{y}}_n)\}$, as follows. Adding it to the circular topology simply requires connecting it between any two machines (done by setting the address of their successor): before we have ``machine $p$ $\rightarrow$ machine $p+1$'', afterwards we have ``machine $p$ $\rightarrow$ new machine $\rightarrow$ machine $p+1$''. We add it in the \ensuremath{\mathbf{W}}\ step, making sure it receives a copy of the final model that has just been finished. The easiest way to do this is by inserting it in the topology at the end of the \ensuremath{\mathbf{W}}\ step, when each machine is simply sending along a copy of the final submodels. In the \ensuremath{\mathbf{Z}}\ step, we proceed as usual, but with $P+1$ machines. Removing a machine is easier. To remove machine $p$, we do so in the \ensuremath{\mathbf{Z}}\ step, by reconnecting ``machine $p-1$ $\rightarrow$ machine $p+1$'' and returning machine $p$ to the cluster. That is all. In the subsequent \ensuremath{\mathbf{W}}\ step, all machines contain the full model, and the submodels will visit the data in each machine, thus not visiting the data in the removed machine.
\paragraph{Fault tolerance}
\label{s:fault}
This situation is similar to discarding a machine in streaming, except that the fault can occur at any time and is not intended. We can handle it with a little extra bookkeeping, and again assuming some support from the parallel processing library. Imagine a fault occurs at machine $p$ and we need to remove it. If it happens during the \ensuremath{\mathbf{Z}}\ step, all we need to do is discard the faulty machine and reconnect the circular topology. If it happens during the \ensuremath{\mathbf{W}}\ step, we also discard and reconnect, but in addition we need to rescue the submodels that were being updated in $p$, which we lose. To do this, we revert to the previously updated copy of them, which resides in the predecessor of $p$ in the circular topology (if no predecessor, we are at the beginning of the \ensuremath{\mathbf{W}}\ step and we can use any copy in any machine). As for the remaining submodels being updated in other machines, some will have already been updated in $p$ (which require no action) and some will not have been updated in $p$ yet (which should not visit $p$ anymore). We can keep track of this information by tagging each submodel with a list of the machines it has not yet visited. At the beginning of the \ensuremath{\mathbf{W}}\ step the list of each submodel contains $\{1,\dots,P\}$, i.e., all machines. When this list is empty, for a submodel, then that submodel is finished and needs no further updates.
Essentially, the robustness of ParMAC to faults comes from its in-built redundance. In the \ensuremath{\mathbf{Z}}\ (and \ensuremath{\mathbf{W}}) step, we can do without the data points in one machine because a good model can still be learned from the remaining data points in the other machines. In the \ensuremath{\mathbf{W}}\ step, we can revert to older copies of the lost submodels residing in other machines.
The asynchronous implementation of ParMAC we described earlier relied on tagging each submodel with a counter in order to know whether it needs processing and communicating. A more general mechanism to run ParMAC asynchronously is to tag each submodel with a list (per epoch) of machines it has to visit. All a machine $p$ needs to do upon receiving a submodel is check its list: if $p$ is not in the list, then the submodel has already visited machine $p$ and been updated with its data, so machine $p$ simply sends it along to its successor without updating it again. If $p$ is in the list, then machine $p$ updates the submodel, removes $p$ from its list, and sends it along to its successor. This works even if we use a different communication topology for each submodel at each epoch.
\section{A theoretical model of the parallel speedup for ParMAC}
\label{s:speedup-th}
In this section we give a theoretical model to estimate the computation and communication times and the parallel speedup in ParMAC. Specifically, eq.~\eqref{e:speedup} gives the speedup $S(P)$ as a function of the number of machines $P$ and other parameters, which seems to agree well with our experiments (section~\ref{s:expts:speedup}). In practice, this model can be used to estimate the optimal number of machines $P$ to use, or to explore the effect on the speedup of different parameter settings (e.g.\ the number of submodels $M$). Throughout the rest of the paper, we will call ``speedup'' $S(P)$ the ratio of the runtime using a single machine (i.e., the serial code) vs using $P > 1$ machines (the parallel code), and ``perfect speedup'' when $S(P) = P$. Our theoretical model applies to the general ParMAC case of $K$ layers, whether differentiable or not; it only assumes that the resulting submodels after introducing auxiliary coordinates are of the same ``size,'' i.e., have the same computation and communication time (this assumption can be relaxed, as we discuss at the end of the section).
We can obtain a quick, rough understanding of the speedup appealing to (a generalisation of) Amdahl's law \citep{GoedecHoisie01a}. ParMAC iterates the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps as follows (where $M$ is the number of submodels and $N$ the number of data points):
\begin{center}
\psfrag{W}[][B]{\caja{c}{c}{\ensuremath{\mathbf{W}}\ step: \\ $M$ problems}}
\psfrag{Z}[][B]{\caja{c}{c}{\ensuremath{\mathbf{Z}}\ step: \\ $N \gg M$ problems}}
\includegraphics[width=0.7\linewidth]{simple-ParMAC.eps}
\end{center}
Roughly speaking, the \ensuremath{\mathbf{W}}\ step has $M$ independent problems so its speedup would be $\min(M,P)$, while the \ensuremath{\mathbf{Z}}\ step has $N$ independent problems so its speedup would be $\min(N,P) = P$ (because in practice $N \ge P$). So the overall speedup would be between $M$ and $P$ depending on the relative runtimes of the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps. This suggests we would expect a nearly perfect speedup $S \approx P$ with $P \le M$ and diminishing returns for $P > M$. This simplified picture ignores important factors such as the ratio of computation vs communication (which our model will make more precise), but it does capture the basic, qualitative behaviour of the speedup.
\subsection{The theoretical model of the speedup}
Let us now develop a more precise, quantitative model. Consider a ParMAC algorithm, operating synchronously, such that there are $M$ independent submodels of the same size in the \ensuremath{\mathbf{W}}\ step, on a dataset with $N$ training points, distributed over $P$ identical machines (each with $N/P$ points). The ParMAC algorithm runs a certain number of iterations, each consisting of a \ensuremath{\mathbf{W}}\ and a \ensuremath{\mathbf{Z}}\ step, so if we ignore small overheads (setup and termination), we can estimate the total runtime as proportional to the number of iterations. Hence, we consider a theoretical model of the runtime of one iteration of the ParMAC algorithm, given the following parameters:
\begin{itemize}
\item $P$: number of machines.
\item $N$: number of training points. \\
We assume $N > P$ is divisible by $P$. This is not a problem because $N \gg P$ in practice (otherwise, there would be no reason to distribute the optimisation).
\item $M$: number of submodels in the \ensuremath{\mathbf{W}}\ step. \\
This may be smaller than, equal to or greater than $P$.
\item $e$: number of epochs in the \ensuremath{\mathbf{W}}\ step.
\item $t^{\ensuremath{\mathbf{W}}}_r$: computation time per submodel and data point in the \ensuremath{\mathbf{W}}\ step. \\
This is the time to process (within the current epoch) one data point by a submodel, i.e., the time do an SGD update to a weight vector, per data point (if we use minibatches, then this is the time to process one minibatch divided by the size of the minibatch).
\item $t^{\ensuremath{\mathbf{W}}}_c$: communication time per submodel in the \ensuremath{\mathbf{W}}\ step. \\
This is the time to send one submodel from one machine to another, including overheads such as buffering, partitioning into messages or waiting time. We assume communication does not overlap with computation, i.e., a machine can either compute or communicate at a given time but not both. Also, communication involves time spent both by the sender and the receiver; we interpret $t^{\ensuremath{\mathbf{W}}}_c$ as the time spent by a given machine in first receiving a submodel and then sending it.
\item $t^{\ensuremath{\mathbf{Z}}}_r$: computation time per data point in the \ensuremath{\mathbf{Z}}\ step. \\
This is the time to finish one data point entirely, using whatever optimisation algorithm performs the \ensuremath{\mathbf{Z}}\ step.
\end{itemize}
$P$, $N$, $M$ and $e$ are integers greater or equal than 1, and $t^{\ensuremath{\mathbf{W}}}_r$, $t^{\ensuremath{\mathbf{W}}}_c$ and $t^{\ensuremath{\mathbf{Z}}}_r$ are real values greater than 0. This model assumes that $t^{\ensuremath{\mathbf{W}}}_r$, $t^{\ensuremath{\mathbf{W}}}_c$ and $t^{\ensuremath{\mathbf{Z}}}_r$ are constant and equal for every submodel or and data point. In reality, even if the submodels are of the same mathematical form and dimension (e.g.\ each submodel is a weight vector of a linear SVM of dimension $D$), the actual times may vary somewhat due to many factors. However, as we will show in section~\ref{s:expts:speedup}, the model does agree quite well with the experimentally measured speedups.
Let us compute the runtimes in the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ step under these model assumptions. The runtime in the \ensuremath{\mathbf{Z}}\ step equals the time for any one machine to process its $N/P$ points on all $M$ submodels, i.e.,
\begin{equation}
\label{e:runtime-Z}
\textstyle T^{\ensuremath{\mathbf{Z}}}(P) = M \frac{N}{P} t^{\ensuremath{\mathbf{Z}}}_r
\end{equation}
since all machines start and end at the same time and do the same amount of computation, without communication. To compute the runtime in the \ensuremath{\mathbf{W}}\ step, we again consider the synchronous procedure of section~\ref{s:ParMAC}. At each tick of an imaginary clock, each machine processes its portion $M/P$ of submodels and sends it to its successor. After $P$ ticks, this concludes one epoch. This is repeated for $e$ epochs, followed by a final round of communication of all the submodels. If $M$ is not divisible by $P$, say $M = Q P + R$ with $Q,R \in \ensuremath{\mathbb{N}}$ and $0 < R < P$, we can apply this procedure pretending there are $P-R$ fictitious submodels%
\footnote{This means that our estimated runtime is an upper bound, because when $M$ is not divisible by $P$, there may be a better way to organise the computation in the \ensuremath{\mathbf{W}}\ step that reduces the time when any machine is idle. In practice this is irrelevant because we implement the computation asynchronously. Each machine keeps a queue of incoming submodels it needs to process, from which it repeatedly takes one submodel, processes it and sends it to the machine's successor.}
(on which machines do useless work). Then, the runtime in each tick is $\ceil{M/P} \frac{N}{P} t^{\ensuremath{\mathbf{W}}}_r$ (time for any one machine to process its $N/P$ points on its portion $\ceil{M/P}$ of submodels) plus $\ceil{M/P} t^{\ensuremath{\mathbf{W}}}_c$ (time for any one machine to send its portion of submodels). The total runtime of the \ensuremath{\mathbf{W}}\ step is then $Pe$ times this plus the time of the final round of computation:
\begin{equation}
\label{e:runtime-W}
\textstyle T^{\ensuremath{\mathbf{W}}}(P) = \ceil{M/P} \left( t^{\ensuremath{\mathbf{W}}}_r \frac{N}{P} + t^{\ensuremath{\mathbf{W}}}_c \right) P e + \ceil{M/P} t^{\ensuremath{\mathbf{W}}}_c P.
\end{equation}
(The final round actually requires $P-1$ ticks, but we take it as $P$ ticks to simplify the equation a bit.) Finally, the total runtime $T^{\ensuremath{\mathbf{W}}}(P) + T^{\ensuremath{\mathbf{Z}}}(P)$ of one ParMAC iteration (\ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ step) with $P$ machines is:
\begin{align}
\label{e:runtime}
T(P) &= \textstyle M \frac{N}{P} t^{\ensuremath{\mathbf{Z}}}_r + P \ceil{M/P} \left( e \left( t^{\ensuremath{\mathbf{W}}}_r \frac{N}{P} + t^{\ensuremath{\mathbf{W}}}_c \right) + t^{\ensuremath{\mathbf{W}}}_c \right),\ P > 1 \\
T(1) &= M N t^{\ensuremath{\mathbf{Z}}}_r + M N e t^{\ensuremath{\mathbf{W}}}_r
\end{align}
where for $P=1$ machine we have no communication ($t^{\ensuremath{\mathbf{W}}}_c = 0$). Hence, the parallel speedup is
\begin{equation}
\label{e:speedup1}
S(P) = \frac{T(1)}{T(P)} = \frac{M/P}{\ceil{M/P}} \frac{e t^{\ensuremath{\mathbf{W}}}_r + t^{\ensuremath{\mathbf{Z}}}_r}{\frac{1}{P} \left(e t^{\ensuremath{\mathbf{W}}}_r + \frac{M/P}{\ceil{M/P}} t^{\ensuremath{\mathbf{Z}}}_r \right) + \frac{1}{N} (e+1) t^{\ensuremath{\mathbf{W}}}_c} = \frac{\frac{1}{\ceil{M/P}} M (e t^{\ensuremath{\mathbf{W}}}_r + t^{\ensuremath{\mathbf{Z}}}_r) P}{\frac{1}{N} (e+1) t^{\ensuremath{\mathbf{W}}}_c P^2 + e t^{\ensuremath{\mathbf{W}}}_r P + \frac{1}{\ceil{M/P}} M t^{\ensuremath{\mathbf{Z}}}_r}
\end{equation}
which can be written more conveniently as
\begin{equation}
\label{e:speedup}
S(P) = \frac{\rho \frac{1}{\ceil{M/P}} M P}{\frac{1}{N} P^2 + \rho_2 P + \rho_1 \frac{1}{\ceil{M/P}} M}
\end{equation}
by defining the following constants:
\begin{equation}
\label{e:ratio-comp-comm}
\rho_1 = \frac{t^{\ensuremath{\mathbf{Z}}}_r}{(e+1) t^{\ensuremath{\mathbf{W}}}_c} \qquad \rho_2 = \frac{e t^{\ensuremath{\mathbf{W}}}_r}{(e+1) t^{\ensuremath{\mathbf{W}}}_c} \qquad \rho = \rho_1 + \rho_2 = \frac{e t^{\ensuremath{\mathbf{W}}}_r + t^{\ensuremath{\mathbf{Z}}}_r}{(e+1) t^{\ensuremath{\mathbf{W}}}_c}.
\end{equation}
These constants can be understood as ratios of computation vs communication, independent of the training set size, number of submodels and number of machines. These ratios depend on the actual computation within the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ step, and on the performance of the distributed system (computation power of each machine, communication speed over the network or shared memory, efficiency of the parallel processing library that handles the communication between machines). The value of these ratios can vary considerably in practice, but it will typically be quite smaller than 1 (say, $\rho \in [10^{-4},1]$), because communication is much slower than computation in current computer architectures.
\subsection{Analysis of the speedup model}
We can characterise the speedup $S(P)$ of eq.~\eqref{e:speedup} in the following three cases:
\begin{itemize}
\item \emph{If $M \ge P$ and $M$ is divisible by $P$}, then we can write the speedup as follows:
\begin{equation}
\label{e:speedup-divisible}
\text{if $M$ divisible by $P$:} \quad S(P) = 1 / \left( \frac{1}{P} + \frac{1}{\rho N} \right) = P / \left( 1 + \frac{P}{\rho N} \right) \le P.
\end{equation}
Here, the function $S(P)$ is independent of $M$ and monotonically increasing with $P$. It would asymptote to $\lim_{P\rightarrow\infty}{S(P) = \rho N}$, but the expression is only valid up to $P=M$. From~\eqref{e:speedup-divisible} we derive the following condition for perfect speedup to occur (in the limit)%
\footnote{Note that if $t^{\ensuremath{\mathbf{W}}}_c = 0$ (no communication overhead) then $\rho = \infty$ and there is no upper bound in~\eqref{e:speedup-divisible-perfect}, but $P \le N$ still holds, because we have to have at least one data point per machine.}:
\begin{equation}
\label{e:speedup-divisible-perfect}
S \approx P \Longleftrightarrow P \ll \rho N.
\end{equation}
This gives an upper bound on the number $P$ of machines to achieve an approximately perfect speedup. Although $\rho$ is quite small in practice, the value of $N$ is very large (typically millions or greater), otherwise there would be no need to distribute the data. Hence, we expect $\rho N \gg 1$, so $P$ could be quite large. In fact, the limit in how large $P$ can be does not come from this condition (which assumes $P \le M$ anyway) but from the number of submodels, as we will see next. \\
In summary, we conclude that if $M \ge P$ and $M$ is divisible by $P$ then the speedup $S(P)$ is given by~\eqref{e:speedup-divisible}, and in practice $S(P) \approx P$ typically.
\item \emph{If $M \ge P$ and $M$ is not divisible by $P$}, then $S(P)$ is given by the full expression~\eqref{e:speedup}, which is studied in appendix~\ref{s:speedup-app}. $S(P)$ is piecewise continuous on $M$ intervals of the form
\begin{equation}
\label{e:speedup-intervals}
\textstyle\big[1,\frac{M}{M-1}\big),\ \big[\frac{M}{M-1},\frac{M}{M-2}\big),\ \dots,\ \big[\frac{M}{2},M\big),\ [M,\infty).
\end{equation}
Within each interval $P \in \big[\frac{M}{k},\frac{M}{k-1}\big)$ for $k = 1,2,3\dots,M$ we have $\ceil{M/P} = k$ and we obtain that $S(P)$ either is monotonically increasing, or is monotonically decreasing, or achieves a single maximum at
\begin{equation}
\label{e:speedup-max}
P^*_k = \sqrt{\rho_1 M N / k} \qquad S^*_k = S(P^*_k) = \frac{\rho M / k}{\rho_2 + 2 \sqrt{\rho_1 M / N k}}.
\end{equation}
The parallelisation ability in this case is less than if $M$ is divisible by $P$, since now some machines are idle at some times during the \ensuremath{\mathbf{W}}\ step.
\item \emph{If $M < P$}, then we can write the speedup as follows:
\begin{equation}
\label{e:speedup-largeP}
\text{if $M < P$:} \quad S(P) = \rho / \left( \frac{\rho_1}{P} + \frac{\rho_2}{M} + \frac{P}{M N} \right) = \rho M / \left( \rho_2 + \rho_1 \frac{M}{P} + \frac{P}{N} \right)
\end{equation}
which corresponds to the last interval $P \in [M,\infty)$ (for $k=1$) over which $S$ is continuous. We obtain that $S(P)$ either is monotonically decreasing (if $M \ge P^*_1$), or it increases from $P = M$ up to a single maximum at $P = P^*_1$ and then decreases monotonically, with
\begin{equation}
\label{e:speedup-largeP:max}
P^*_1 = \sqrt{\rho_1 M N} \qquad S^*_1 = S(P^*_1) = \frac{\rho M}{\rho_2 + 2 \sqrt{\rho_1 M / N}}.
\end{equation}
As $P \rightarrow \infty$ we have that $S(P) \approx \rho N M / P \rightarrow 0$ (assuming $t^{\ensuremath{\mathbf{W}}}_c > 0$ so $\rho < \infty$). This decrease of the speedup for large $P$ is caused by the communication overhead in the \ensuremath{\mathbf{W}}\ step, where $P - M$ machines are idle at each tick in the \ensuremath{\mathbf{W}}\ step. \\
In the impractical case where there is no communication cost ($t^{\ensuremath{\mathbf{W}}}_c = 0$ so $\rho = \infty$) then $S(P)$ is actually monotonically increasing and $\lim_{P \rightarrow \infty}{S(P)} = S^*_1 = \frac{\rho}{\rho_2} M > M$, so the more machines the larger the speedup, although with diminishing returns.
\end{itemize}
Theorem~\ref{th:speedup-charact} shows that $S(P)$ at the beginning of each interval is greater than anywhere before that interval, i.e., $S(M/k) > S(P)$ $\forall P < M/k$, for $k=1,2,\dots,M$. That is, although the speedup $S(P)$ is not necessarily monotonically increasing for $P \ge 1$, it is monotonically increasing for $P \in \big\{1,\frac{M}{M-1},\frac{M}{M-2},\dots,\frac{M}{2},M\big\}$. This suggests selecting values of $P$ that make $M/P$ integer, in particular when $M$ is divisible by $P$.
\paragraph{Globally maximum speedup $S^* = \max_{P \ge 1}{S(P)}$}
This is given by (see appendix~\ref{s:speedup-app}):
\begin{itemize}
\item If $M \ge \rho_1 N$: $S^* = M / \left( 1 + \frac{M}{\rho N} \right) \le M$, achieved at $P = M$.
\item If $M < \rho_1 N$: $S^* = S^*_1 = \frac{\rho M}{\rho_2 + 2 \sqrt{\rho_1 M / N}} > M$, achieved at $P = P^*_1 = \sqrt{\rho_1 M N} > M$.
\end{itemize}
In practice, with large values of $N$, the more likely case is $S^* = S^*_1 > M$ for $P = P^*_1 > M$. In this case, the maximum speedup is achieved using more machines than submodels (even though this means some machines will be idle at some times in the \ensuremath{\mathbf{W}}\ step), and is bigger than $M$. Since diminishing returns occur as we approach the maximum, the practically best value of $P$ will be somewhat smaller than $P^*_1$.
\paragraph{The ``large dataset'' case}
The case where $N$ is large is practically important because the need for distributed optimisation arises mainly from this. Specifically, if we take $P \ll \rho_2 N$, the speedup becomes (see appendix~\ref{s:speedup-app}):
\begin{equation}
\label{e:speedup-largeN}
\text{if $M$ divisible by $P$:} \quad S(P) \approx P; \qquad \text{if $M > P$:} \quad S(P) \approx \rho / \left( \frac{\rho_1}{P} + \frac{\rho_2}{M} \right)
\end{equation}
so that the speedup is almost perfect up to $P = M$, and then it is approximately the weighted harmonic mean of $M$ and $P$ (hence, $S(P)$ is monotonically increasing and between $M$ and $P$). For $P \gg \rho_1$, we have $S(P) \approx \frac{\rho}{\rho_2} M > M$.
\paragraph{The ``dominant \ensuremath{\mathbf{Z}}\ step'' case}
If we take $t^{\ensuremath{\mathbf{Z}}}_r \gg t^{\ensuremath{\mathbf{W}}}_r, t^{\ensuremath{\mathbf{W}}}_c$ or equivalently $\rho \approx \rho_1$ very large, which means the \ensuremath{\mathbf{Z}}\ step dominates the runtime, then $S(P) \approx P$. This is because the \ensuremath{\mathbf{Z}}\ step parallelises perfectly (as long as $P < N$).
\paragraph{Transformations that keep the speedup invariant}
We can rewrite the speedup of eq.~\eqref{e:speedup} as:
\begin{equation}
\label{e:speedup-indepN}
S(P) = \frac{\rho' \frac{1}{\ceil{M/P}} M P}{P^2 + \rho'_2 P + \rho'_1 \frac{1}{\ceil{M/P}} M}
\end{equation}
with
\begin{equation}
\label{e:ratio-comp-comm2}
\rho' = \rho N = (\rho_1+\rho_2) N = \frac{N(e t^{\ensuremath{\mathbf{W}}}_r + t^{\ensuremath{\mathbf{Z}}}_r)}{(e+1) t^{\ensuremath{\mathbf{W}}}_c} \qquad \rho'_1 = \rho_1 N = \frac{N t^{\ensuremath{\mathbf{Z}}}_r}{(e+1) t^{\ensuremath{\mathbf{W}}}_c} \qquad \rho'_2 = \rho_2 N = \frac{N e t^{\ensuremath{\mathbf{W}}}_r}{(e+1) t^{\ensuremath{\mathbf{W}}}_c}
\end{equation}
so that $S(P)$ is independent of $N$, which has been absorbed into the communication-computation ratios. This means that $S(P)$ depends on the dataset size ($N$) and computation/communication times ($t^{\ensuremath{\mathbf{W}}}_r$, $t^{\ensuremath{\mathbf{Z}}}_r$, $t^{\ensuremath{\mathbf{W}}}_c$) only through $\rho'$, $\rho'_1$ and $\rho'_2$, and is therefore invariant to parameter transformations that leave these ratios unchanged. Such transformations are the following (where $\alpha > 0$):
\begin{itemize}
\item Scaling $N$, $t^{\ensuremath{\mathbf{W}}}_r$ and $t^{\ensuremath{\mathbf{Z}}}_r$ as $\alpha N$, $\frac{1}{\alpha} t^{\ensuremath{\mathbf{W}}}_r$ and $\frac{1}{\alpha} t^{\ensuremath{\mathbf{Z}}}_r$. \\
``Larger dataset, faster computation,'' or ``smaller dataset, slower computation.''
\item Scaling $N$ and $t^{\ensuremath{\mathbf{W}}}_c$ as $\alpha N$ and $\alpha t^{\ensuremath{\mathbf{W}}}_c$ \\
``Larger dataset, slower communication,'' or ``smaller dataset, faster communication.''
\item Scaling $t^{\ensuremath{\mathbf{W}}}_r$, $t^{\ensuremath{\mathbf{Z}}}_r$ and $t^{\ensuremath{\mathbf{W}}}_c$ as $\alpha t^{\ensuremath{\mathbf{W}}}_r$, $\alpha t^{\ensuremath{\mathbf{Z}}}_r$ and $\alpha t^{\ensuremath{\mathbf{W}}}_c$ \\
``Faster computation, faster communication,'' or ``slower computation, slower communication.''
\end{itemize}
\subsection{Discussion and examples}
\begin{figure}[b!]
\centering
\psfrag{processor}[t][]{number of machines $P$}
\psfrag{speedup}[][t]{speedup $S(P)$}
\psfrag{PleM}[b][b]{\caja[1.5]{c}{c}{$P \le M$: \\ $S(P) = \displaystyle\frac{P}{1 + \frac{P}{\rho N}} \approx P$}}
\psfrag{MleP}[t][t]{\caja[1.5]{c}{c}{$P > M$: \\ $S(P) = \displaystyle\frac{\rho}{\frac{\rho_1}{P} + \frac{\rho_2}{M} + \frac{P}{M N}} \approx \frac{\rho}{\frac{\rho_1}{P} + \frac{\rho_2}{M}}$}}
\psfrag{maxS}[t][t]{$S(P^*_1)$}
\begin{tabular}[c]{@{}c@{}}
\includegraphics[width=\linewidth]{speedup-typical-edited.eps}
\end{tabular}
\caption{Typical form of the theoretical speedup curve for realistic parameter settings, specifically $N = 10^6$ data points, $M = 512$ submodels, $e=1$ epoch in the \ensuremath{\mathbf{W}}\ step, and $t^{\ensuremath{\mathbf{W}}}_r = 1$ (this sets the units of time), $t^{\ensuremath{\mathbf{Z}}}_r = 5$ and $t^{\ensuremath{\mathbf{W}}}_c = 10^3$ (so $\rho_1 = 0.0025$, $\rho_2 = 0.0005$ and $\rho = 0.003$). Some of the discontinuities of the curve (where $\ceil{M/P}$ is discontinuous) are visible. We mark the values for $P$ such that $M$ is divisible by $P$ ($\circ$) and the maximum speedup ($\ast$), which occurs for $P = P^*_1 > M$.}
\label{f:speedup-typical}
\end{figure}
Fig.~\ref{f:speedup-typical} plots a ``typical'' speedup curve $S(P)$, obtained with a realistic choice of parameter values. It displays the prototypical speedup shape we should expect in practice (the experimental speedups of fig.~\ref{f:speedup} confirm this). For $P \le M$ the curve is very close to the perfect speedup $S(P) = P$, slowly deviating from it as $P$ approaches $M$. For $P > M$, the curve continues to increase until it reaches its maximum at $P = P^*_1$, and decreases thereafter.
\begin{figure}[p]
\centering
\psfrag{processor}{}
\begin{tabular}{@{}c@{\hspace{0.05\linewidth}}c@{\hspace{0.05\linewidth}}c@{}}
& $e = 1$ epoch & $e = 8$ epochs \\
\raisebox{0.16\linewidth}{\caja{c}{l}{$t^{\ensuremath{\mathbf{W}}}_c = 1$ \\ $t^{\ensuremath{\mathbf{Z}}}_r = 1$}} &
\psfrag{speedup}[][t]{speedup $S(P)$}
\includegraphics[height=0.30\linewidth]{speedup_e1_twc1_tzr1.eps} &
\psfrag{speedup}{}
\includegraphics[height=0.30\linewidth]{speedup_e8_twc1_tzr1.eps} \\[-1ex]
\raisebox{0.16\linewidth}{\caja{c}{l}{$t^{\ensuremath{\mathbf{W}}}_c = 1$ \\ $t^{\ensuremath{\mathbf{Z}}}_r = 100$}} &
\psfrag{speedup}[][t]{speedup $S(P)$}
\includegraphics[height=0.30\linewidth]{speedup_e1_twc1_tzr100.eps} &
\psfrag{speedup}{}
\includegraphics[height=0.30\linewidth]{speedup_e8_twc1_tzr100.eps} \\[-1ex]
\raisebox{0.16\linewidth}{\caja{c}{l}{$t^{\ensuremath{\mathbf{W}}}_c = 100$ \\ $t^{\ensuremath{\mathbf{Z}}}_r = 1$}} &
\psfrag{speedup}[][t]{speedup $S(P)$}
\includegraphics[height=0.30\linewidth]{speedup_e1_twc100_tzr1.eps} &
\psfrag{speedup}{}
\includegraphics[height=0.30\linewidth]{speedup_e8_twc100_tzr1.eps} \\[-1ex]
\raisebox{0.16\linewidth}{\caja{c}{l}{$t^{\ensuremath{\mathbf{W}}}_c = 1\,000$ \\ $t^{\ensuremath{\mathbf{Z}}}_r = 100$}} &
\psfrag{processor}[t][]{number of machines $P$}
\psfrag{speedup}[][t]{speedup $S(P)$}
\includegraphics[height=0.30\linewidth]{speedup_e1_twc1000_tzr100.eps} &
\psfrag{processor}[t][]{number of machines $P$}
\psfrag{speedup}{}
\includegraphics[height=0.30\linewidth]{speedup_e8_twc1000_tzr100.eps}
\end{tabular}
\caption{Theoretical speedup $S(P)$ as a function of the number of machines $P$ for various settings of the parameters of ParMAC with a binary autoencoder. The parameters are: dataset size $N = 50\,000$ training points; number of submodels $M \in \{1,2,4,8,16,32,64,128,256,512\}$; number of epochs in the \ensuremath{\mathbf{W}}\ step $e \in \{1,8\}$; \ensuremath{\mathbf{W}}\ step computation time (per submodel and data point) $t^{\ensuremath{\mathbf{W}}}_r = 1$ (this sets the units of time); \ensuremath{\mathbf{W}}\ step communication time (per submodel) $t^{\ensuremath{\mathbf{W}}}_c \in \{1,100,1\,000\}$; \ensuremath{\mathbf{Z}}\ step computation time (per submodel and data point) $t^{\ensuremath{\mathbf{Z}}}_r \in \{1,100\}$. Within each plot, each curve corresponds to one value of $M$, indicated on the right end of the plot.}
\label{f:speedup-th}
\end{figure}
Fig.~\ref{f:speedup-th} plots $S(P)$ for a wider range of parameter settings. We set the dataset size to a practically small value ($N = 50\,000$), otherwise the curves tend to look like the typical curve from fig.~\ref{f:speedup-typical}. The parameter settings are representative of different, potential practical situations (some more likely than others). We note the following observations:
\begin{itemize}
\item Again, the most important observation is that the number of submodels $M$ is the parameter with the most direct effect on the speedup: near-perfect speedups ($S \approx P$) occur if $M \ge P$, otherwise the speedups are between $M$ and $P$ (and eventually saturate if $P \gg M$).
\item When the time spent on communication is large in relative terms, the speedup is decreased. This can happen when the runtime of the \ensuremath{\mathbf{Z}}\ step is low (small $t^{\ensuremath{\mathbf{Z}}}_r$), when the communication cost is large (large $t^{\ensuremath{\mathbf{W}}}_c$), or with many epochs (large $e$). Indeed, since the \ensuremath{\mathbf{Z}}\ step is perfectly parallelisable, any decrease of the speedup should come from the \ensuremath{\mathbf{W}}\ step.
\item Some of the curves display noticeable discontinuities, caused by the function $(M/P)/\ceil{M/P}$, occurring at values of $P$ of the form $M/k$ for $k \in \{1,\dots,M\}$. At each such value, $S(P)$ is greater than for any smaller value of $P$, in accordance with theorem~\ref{th:speedup-charact}. This again suggests selecting values of $P$ that make $M/P$ integer, in particular when $M$ is a multiple of $P$ ($P$, $2P$, $3P$\dots). This achieves the best speedup efficiency in that machines are never idle (in our theoretical model). \\
Also, for fixed $P$, the function $(M/P)/\ceil{M/P}$ can take the same value for different values of $M$ (e.g.\ $M=32$ and $M=64$ for $P=60$). This explains why some curves (for different $M$) partly overlap.
\item The maximum speedup is typically larger than $M$ and occurs for $P > M$. It is possible to have the maximum speedup be smaller than $M$ (in which case it occurs at $P=M$); an example appears in fig.~\ref{f:speedup-th}, row 3, column 2 (for the larger $M$ values). But this happens only when $M \ge \rho_1 N$, which requires an unusually small dataset and an impractically large number of submodels. Generally, we should expect $P > M$ to be beneficial. This is also seen experimentally in fig.~\ref{f:speedup}.
\end{itemize}
\subsection{Practical considerations}
\label{s:speedup-th:practical}
In practice, given a specific problem (with a known number of submodels $M$, epochs $e$ and dataset size $N$), our theoretical speedup curves can be used to determine optimal values for the number of machines $P$ to use. As seen in section~\ref{s:expts:speedup}, the theoretical curves agree quite well with the experimentally measured speedups. The theoretical curves do need estimates for the computation time and communication times of the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps. These are hard to obtain a priori; the computational complexity of the algorithm in \ensuremath{\mathcal{O}}-notation ignores constant factors that affect significantly the actual times. Their estimates should be measured from test runs.
As seen from eq.~\eqref{e:speedup-indepN}, we can leave the speedup unchanged by trading off dataset size ($N$) and computation/communication times ($t^{\ensuremath{\mathbf{W}}}_r$, $t^{\ensuremath{\mathbf{Z}}}_r$, $t^{\ensuremath{\mathbf{W}}}_c$) in various ways, as long as one of the three following holds: the products $N t^{\ensuremath{\mathbf{W}}}_r$ and $N t^{\ensuremath{\mathbf{Z}}}_r$ remain constant; or the quotient $N/t^{\ensuremath{\mathbf{W}}}_c$ remains constant; or the quotients $t^{\ensuremath{\mathbf{W}}}_r/t^{\ensuremath{\mathbf{W}}}_c$ and $t^{\ensuremath{\mathbf{Z}}}_r/t^{\ensuremath{\mathbf{W}}}_c$ remain constant.
Theoretically, the most efficient operating points for $P$ are values such that $M$ is divisible by $P$, because this means no machine is ever idle. In practice with an asynchronous implementation and with $t^{\ensuremath{\mathbf{Z}}}_r$, $t^{\ensuremath{\mathbf{W}}}_r$ and $t^{\ensuremath{\mathbf{W}}}_c$ exhibiting some variation over submodels and data points, this is not true anymore. Still, if in a given application one is constrained to using $P \le M$ machines, choosing $P$ close to a divisor of $M$ would probably be preferable.
One assumption in our speedup model is that the $P$ machines are identical in processing power. The model does extend to the case where the machines are different, as noted in our discussion of load balancing (section~\ref{s:ParMAC:extensions}). This is because the work performed by each machine is proportional to the number of data points it contains: in the \ensuremath{\mathbf{W}}\ step, because every submodel runs (implicitly) SGD, and every submodel must visit each machine; in the \ensuremath{\mathbf{Z}}\ step, because each data point is a separate problem, and involves all submodels (which reside in each machine). Hence, we can equalise the work over machines by loading each machine with an amount of data proportional to its processing speed, independent of the number of submodels.
Another assumption in our model (in the \ensuremath{\mathbf{W}}\ step) is that all $M$ submodels are identical in ``size'' (computation and communication time). This is sometimes not true. For example, in the BA, we have submodels of two types: the $L$ encoders (each a binary linear SVM operating on a $D$-dimensional input) and the $D$ decoders (each a linear regressor operating on an $L$-dimensional input). Since $D > L$, the encoders are bigger than the decoders and take longer to train and communicate. We can still apply our speedup model if we ``group'' smaller submodels into a single submodel of size comparable to the larger submodels, so as to equalise as much as possible the submodel sizes (the actual implementation does not need to group submodels, of course). For the BA, under the reasonable assumption that the ratio of computation times $t^{\ensuremath{\mathbf{W}}}_r$ (and communication times $t^{\ensuremath{\mathbf{W}}}_c$) of decoder vs encoder is $L/D < 1$, we can group the $D$ decoders into $L$ groups of $D/L$ decoders each. Each group of decoders has now a computation and communication time equal to that of one encoder. This gives an effective number of independent submodels $M = 2 L$, and this is what we use when applying the model to the experimental speedups in section~\ref{s:expts:speedup}.
Finally, we emphasise that the goal of this section was to characterise the parallel speedup of ParMAC quantitatively and demonstrate the runtime gains that are achievable by using $P$ machines. In practice, other considerations are also important, such as the economic cost of using $P$ machines (which may limit the maximum $P$ available); the type of machines (obviously, we want all the computation and communication times as small as possible); the choice of optimisation algorithm in the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps; the fact that, because of its size, the dataset may need to be, or already is, distributed across $P$ machines; etc. It is also possible to combine ParMAC with other, orthogonal techniques. For example, if each of the submodels in the \ensuremath{\mathbf{W}}\ step is a convex optimisation problem (as is the case with the linear SVMs with the binary autoencoder), we could use the techniques described in section~\ref{s:related} for distributed convex optimisation to each submodel. This would effectively allow for larger speedups when $P > M$.
\section{Convergence of ParMAC}
\label{s:conv}
The only approximation that ParMAC makes to the original MAC algorithm is using SGD in the \ensuremath{\mathbf{W}}\ step. Since we can guarantee convergence of SGD under certain conditions, we can recover the original convergence guarantees for MAC. Let us see this in more detail. Convergence of MAC to a stationary point is given by theorem B.3 in \citet{CarreirWang12a}, which we quote here:
\begin{thm}
\label{th:MACQP}
Consider the constrained problem of eq.~\eqref{e:mac} and its quadratic-penalty function $E_Q(\ensuremath{\mathbf{W}},\ensuremath{\mathbf{Z}};\mu)$ of eq.~\eqref{e:mac-quadpen}. Given a positive increasing sequence $(\mu_k) \rightarrow \infty$, a nonnegative sequence $(\tau_k) \rightarrow 0$, and a starting point $(\ensuremath{\mathbf{W}}^0,\ensuremath{\mathbf{Z}}^0)$, suppose the quadratic-penalty method finds an approximate minimiser $(\ensuremath{\mathbf{W}}^k,\ensuremath{\mathbf{Z}}^k)$ of $E_Q(\ensuremath{\mathbf{W}}^k,\ensuremath{\mathbf{Z}}^k;\mu_k)$ that satisfies $\norm{\nabla_{\ensuremath{\mathbf{W}},\ensuremath{\mathbf{Z}}}{E_Q(\ensuremath{\mathbf{W}}^k,\ensuremath{\mathbf{Z}}^k;\mu_k)}} \le \tau_k$ for $k=1,2,\dots$ Then, $\lim_{k\rightarrow\infty}{(\ensuremath{\mathbf{W}}^k,\ensuremath{\mathbf{Z}}^k)} = (\ensuremath{\mathbf{W}}^*,\ensuremath{\mathbf{Z}}^*)$, which is a KKT point for the problem~\eqref{e:mac}, and its Lagrange multiplier vector has elements $\ensuremath{\boldsymbol{\lambda}}^*_n = \lim_{k\rightarrow\infty}{-\mu_k \, (\ensuremath{\mathbf{Z}}^k_n - \ensuremath{\mathbf{F}}(\ensuremath{\mathbf{Z}}^k_n,\ensuremath{\mathbf{W}}^k;\ensuremath{\mathbf{x}}_n))}$, $n=1,\dots,N$.
\end{thm}
This theorem applies to the general case of $K$ differentiable layers, where the standard Karush-Kuhn-Tucker (KKT) conditions hold \citep{NocedalWright06a}. It relies on a standard condition for penalty methods for nonconvex problems \citep{NocedalWright06a}, namely that we must be able to reduce the gradient of the penalised function $E_Q$ below an arbitrary tolerance $\tau_k \ge 0$ for each value $\mu_k$ of the penalty parameter (in MAC iterations $k=1,2,\dots$). This can be achieved by running a suitable (unconstrained) optimisation method for sufficiently many iterations. How does this change in the case of ParMAC? The \ensuremath{\mathbf{Z}}\ step remains unchanged with respect to MAC (the fact that the optimisation is distributed is irrelevant since the $N$ subproblems $\ensuremath{\mathbf{z}}_1,\dots,\ensuremath{\mathbf{z}}_N$ are independent). The \ensuremath{\mathbf{W}}\ step does change, because we are now obliged to use a distributed, stochastic training. What we need to ensure is that we can reduce the gradient of the penalised function with respect to each submodel (since they are independent subproblems in the \ensuremath{\mathbf{W}}\ step) below an arbitrary tolerance. This can also be guaranteed under standard conditions. In general, we can use convergence conditions from stochastic optimisation \citep{Benven_90a,KushnerYin03a,Pflug96a,Spall03a,BertsekTsitsik00a}. Essentially, these are Robbins-Monro schedules, which require the learning rate $\eta_t$ of SGD to decrease such that $\lim_{t\rightarrow\infty} \eta_t = 0$, $\sum^{\infty}_{t=1}{\eta_t} = \infty$, $\sum^{\infty}_{t=1}{\eta^2_t} < \infty$, where $t$ is the epoch number (SGD iteration, or pass over the entire dataset%
\footnote{Note that it is not necessary to assume that the points (or minibatches) are sampled at random during updates. Various results exist that guarantee convergence with deterministic errors (e.g.\ \citealp{BertsekTsitsik00a} and references therein), rather than stochastic errors. These results assume a bound on the deterministic errors (rather than a bound on the variance of the stochastic errors), and apply to general, nonconvex objective functions with standard conditions (Lipschitz continuity, Robbins-Monro schedules for the step size, etc.). They apply as a particular case to the ``incremental gradient'' method, where we cycle through the data points in a fixed sequence.}).
We can give much tighter conditions on the convergence and the convergence rate when the subproblems in the \ensuremath{\mathbf{W}}\ step are convex (which is often the case, as with logistic or linear regression, linear SVMs, etc.). This is a topic that has received much attention recently (see section~\ref{s:related}), and many such conditions exist, often based on techniques such as Nesterov accelerated algorithms and stochastic average gradient \citep{Cevher_14a}. They typically bound the distance to the minimum in objective function value as $\ensuremath{\mathcal{O}}(1/t^{\alpha})$ or $\ensuremath{\mathcal{O}}(1/\beta^t)$ where the coefficients $\alpha > 0$, $0< \beta < 1$ and the constant factors in the \ensuremath{\mathcal{O}}-notation depend on the (strong) convexity properties of the problem, Lipschitz constant, etc.
In summary, \emph{convergence of ParMAC to a stationary point is guaranteed by the same theorem as MAC, with an added SGD-type condition for the \ensuremath{\mathbf{W}}\ step}. This convergence guarantee is independent of the number of layers and submodels (since they are independent in the \ensuremath{\mathbf{W}}\ step) and the number of machines $P$ (since effectively we are doing SGD on shuffled datasets of size $N$, even if they are partitioned on portions of size $N/P$).
We can also guarantee ParMAC's convergence with only the original MAC theorem, without SGD-type conditions, while still in the distributed setting and achieving significant parallelism. This can be done by computing the gradient in the \ensuremath{\mathbf{W}}\ step exactly (as MAC assumes). First, each machine $p = 1,\dots,P$ computes the exact sum of per-point gradients for each submodel (by summing over its data portion), in parallel. Then, we aggregate these $P$ partial gradients into one exact gradient, for each submodel. This could be done via a parameter server, or by having each machine act as the parameter server for one submodel, and could be easily implemented with MPI functions. However, as is well known, this is far slower than using SGD.
With nondifferentiable layers, the convergence properties of MAC (and ParMAC) are not well understood. In particular, for the binary autoencoder the encoding layer is discrete and the problem is NP-complete. But, again, the only modification of ParMAC over MAC is the fact that the encoder and decoder are trained with SGD in the \ensuremath{\mathbf{W}}\ step, whose convergence tolerance can be achieved with SGD-type conditions. Indeed, our experiments show ParMAC gives almost identical results to MAC.
While convergence guarantees are important theoretically, in practical applications with large datasets in a distributed setting one typically runs SGD for just a few epochs, even one or less than one (i.e., we stop SGD before passing through all the data). This typically reduces the objective function to a good enough value as fast as possible, since each pass over the data is very costly. In our experiments, one to two epochs in the \ensuremath{\mathbf{W}}\ step make ParMAC very similar to MAC using an exact step.
\section{Implementation of ParMAC for binary autoencoders}
\label{s:ParMAC-BAhash-implem}
We have implemented ParMAC for binary autoencoders in C/C++ using the GNU Scientic Library (GSL) (\url{http://www.gnu.org/s/gsl}) and Basic Linear Algebra Subroutines (BLAS) library (\url{http://www.netlib.org}) for mathematical operations and linear algebra, and the Message Passing Interface (MPI) \citep{Gropp_99a,Gropp_99b,MPI12a} for interprocess communication.
GSL and BLAS provide a wide range of mathematical routines such as basic matrix operations, various matrix decompositions and least-squares fitting. We used the versions of GSL and BLAS that come with our Linux distribution (Ubuntu 14.04). Considerably better performance could be achieved by using LAPACK and an optimised version of BLAS (such as ATLAS, or as provided by a computer vendor for their specific architecture).
MPI is one of the most widely used frameworks for high-performance parallel computing today, and is the best option for ParMAC because of its support for distributed-memory machines and SPMD (single program, multiple data) model, its language independence, and its availability in multiple machines, from small shared-memory multiprocessor machines to hybrid clusters. In MPI, different processes cannot directly access each other's memory space, but data can be transferred by sending messages from one process to another, or collectively among multiple processes. The SPMD model, very useful in distributed machine learning, means that all processes share the same code (and executable file), and each of them can operate on different data with flow control using its individual process id.
MPI is an industry standard for which there are many implementations, such as MPICH or OpenMPI, mostly compatible with each other. We used MPICH on our UC Merced shared-memory cluster and OpenMPI on the UCSD TSCC distributed cluster (see section~\ref{s:expts:setup}). Our ParMAC C++ code compiles and runs with both implementations. We used the highest compiler optimisation level, specifically we ran \texttt{mpicc -O3 -lgsl -lgslcblas -lm}. This calls the GNU C compiler with option \texttt{-O3}, which turns on all the available code optimisation flags. It results in a longer compilation time but more efficient code.
\begin{figure}[p]
\small
\begin{tabbing}
n \= n \= n \= n \= n \= \kill
\texttt{MPI\_Init(\&argc, \&argv);} \` // initialise the MPI execution environment \\
\texttt{MPI\_Comm\_rank(MPI\_COMM\_WORLD, \&mpirank);} \` // get the rank of the calling MPI process \\
\texttt{MPI\_Comm\_size(MPI\_COMM\_WORLD, \&mpisize);} \` // get the total number of MPI processes \\
\texttt{loadsettings();} \` // load parameters ($\mu$, epochs, dataset path, etc.) \\
\texttt{loaddatasets();} \` // load input and output datasets and initial auxiliary coordinates \\
\texttt{initializelayers();} \` // allocate memory and initialise \ensuremath{\mathbf{f}}, \ensuremath{\mathbf{h}}\ and \ensuremath{\mathbf{Z}}\ steps \\
// we use \texttt{MPI\_Bsend} to avoid managing send buffers, so we need to allocate the required \\
// amount of buffer space into which data can be copied until it is delivered \\
\texttt{MPI\_Pack\_size(commbuffsize, MPI\_CHAR, MPI\_COMM\_WORLD, \&mpi\_attach\_buff\_size);} \\
// allocate enough memory so it can store the whole model \\
\texttt{mpi\_attach\_buff = malloc(totalsubmodelcount*(mpi\_attach\_buff\_size+MPI\_BSEND\_OVERHEAD));} \\
\texttt{MPI\_Buffer\_attach(mpi\_attach\_buff, mpi\_attach\_buff\_size);} \` // attach the allocated buffer \\
\\
\texttt{for (iter=1 to length($\mu$)) \{} \` // iterate over all the values of $\mu$ \+ \\
// begin \ensuremath{\mathbf{W}}-step \\
\texttt{visitedsubmodels = 0;} \\
// each process visits all the submodels, epochs + 1 times \\
\texttt{while (visitedsubmodels <= totalsubmodelcount*epochs) \{} \+ \\
// \texttt{stepcounter} is a number that each submodel carries and increases by one in each step. \\
// Once it reaches a certain value we stop sending the submodel around and it stops. \\
// We reset \texttt{stepcounter} for all the submodels in the beginning of each \ensuremath{\mathbf{W}}-step. \\
\texttt{if (stepcounter > 0) \{} \` // if this is not the first submodel to train in the iteration, we wait to receive \+ \\
// \texttt{MPI\_Recv} blocks until the requested data is available in the application buffer in the receiving task \\
\texttt{\textcolor{red}{MPI\_Recv}(receivebuffer, commbuffsize, MPI\_CHAR, MPI\_ANY\_SOURCE, MODEL\_MSG\_TAG,} \\
\hspace{20mm}\texttt{MPI\_COMM\_WORLD, \&recvStatus);} \\
\texttt{savesubmodel(receivebuffer);} \` // save the received buffer into a suitable struct \- \\
\texttt{\}} \\
\texttt{if (stepcounter < epochs*mpisize) \{} \` // we don't train the submodels in the last update round \+ \\
\texttt{switch(submodeltype)} \` // train each submodel according to its type \\
\texttt{case 'SVM': HtrainSGD();} \\
\texttt{case 'linlayer': FtrainSGD();} \- \\
\texttt{\}} \\
\texttt{if (stepcounter < (ringepochs+1)*mpisize) \{} \` // we still need to send this submodel around \+ \\
// the lookup table is created randomly and stores the path of each submodel over epochs and iterations \\
\texttt{successor = next\_in\_lookuptable();} \` // pick the successor process from the lookup table \\
\texttt{loadsubmodel(sendbuffer);} \` // load the submodel from its struct into the send buffer \\
// \texttt{MPI\_Bsend} returns after the data has been copied from application buffer space to the allocated send buffer \\
\texttt{\textcolor{red}{MPI\_Bsend}(sendbuffer, taskbufsize*sizeof(double), MPI\_CHAR, successor, MODEL\_MSG\_TAG,} \\
\hspace{20mm}\texttt{MPI\_COMM\_WORLD);} \- \\
\texttt{\}} \\
\texttt{visitedsubmodels++;} \- \\
\texttt{\}} \\
// end \ensuremath{\mathbf{W}}-step \\
\\
// begin \ensuremath{\mathbf{Z}}-step \\
\texttt{updateZ\_relaxed();} \` // initialise auxiliary coordinates based on a truncated, relaxed solution \\
\texttt{updateZ\_alternate();} \` // update auxiliary coordinates by alternating optimisation over bits \\
// end \ensuremath{\mathbf{Z}}-step \- \\
\texttt{\}} \\
\\
\texttt{MPI\_Buffer\_detach(\&mpi\_attach\_buff, \&mpi\_attach\_buff\_size);} \` // detach the allocated buffer \\
\texttt{free(mpi\_attach\_buff);} \` // free the allocated memory \\
\texttt{MPI\_Finalize();} \` // terminate the MPI execution environment
\end{tabbing}
\caption{Binary autoencoder ParMAC algorithm (fragment), showing important MPI calls.}
\label{f:BA-ParMAC-alg}
\end{figure}
The code snippet in figure~\ref{f:BA-ParMAC-alg} shows the main steps of the ParMAC algorithm for the BA. All the functions starting with \texttt{MPI\_} are API calls from the MPI library. As with all MPI programs, we start the code by initialising the MPI environment and end by finalising it. To receive data we use the synchronous%
\footnote{Note that the word ``synchronous'' here does not refer to how we process the different submodels, which as we stated earlier are not synchronised to start or end at specific clock ticks, hence are processed asynchronously with respect to each other. The word ``synchronous'' here refers to MPI's handling of an \emph{individual} receive function (see appendix~\ref{s:MPI}). This can be done either by calling \texttt{MPI\_Recv}, which will block until the data is received (synchronous blocking function), as in the pseudocode in fig.~\ref{f:BA-ParMAC-alg}; or by calling \texttt{MPI\_Irecv} (asynchronous nonblocking function) followed by a \texttt{MPI\_Wait}, which will block until the data is received, like this:
\begin{tabbing}
n \= n \= n \= n \= n \= \+ \kill
\texttt{MPI\_Irecv(receivebuffer, commbuffsize, MPI\_CHAR, MPI\_ANY\_SOURCE, MODEL\_MSG\_TAG, MPI\_COMM\_WORLD, \&recvRequest);} \\
\texttt{MPI\_Wait(\&recvRequest, \&recvStatus);}
\end{tabbing}
Both options are equivalent for our purpose, which is to ensure we receive the submodel before starting to train it. The \texttt{MPI\_Irecv}/\texttt{MPI\_Wait} option is slightly more flexible in that it would allow us to do some additional processing between \texttt{MPI\_IRecv} and \texttt{MPI\_Wait} and possibly achieve some performance gain.},
blocking MPI receive function \texttt{MPI\_Recv}. The process calling this blocks until the data arrives. To send data we use the buffered blocking version of the MPI send functions, \texttt{MPI\_Bsend}. This requires that we allocate enough memory and attach it to the system in advance. The process calling \texttt{MPI\_Bsend} blocks until the buffer is copied to the MPI internal memory; after that, the MPI library takes care of sending the data appropriately. The benefit of using this version of send is that the programmer can send messages without worrying about where they are buffered, so the code is simpler. Appendix~\ref{s:MPI} briefly describes important MPI functions and their arguments.
\section{Experiments}
\label{s:expts}
\subsection{Setup}
\label{s:expts:setup}
\paragraph{Computing systems}
We used two different computing systems, to which we will refer as \emph{distributed} and \emph{shared-memory}:
\begin{description}
\item[Distributed-memory] This used General Computing Nodes from the UCSD Triton Shared Computing Cluster (TSCC), available to the public for a fee. Each node contains 2 8-core Intel Xeon E5-2670 processors (16 cores in total), 64GB DRAM (4GB/core) and a 500GB hard drive. The nodes are connected through a 10GbE network. We used up to $P=128$ processors. Detailed specs are in table~\ref{t:specs} (obtained by running \texttt{dmidecode} in the actual processor) and \url{http://idi.ucsd.edu/computing}.
\item[Shared-memory] This is a 72-processor machine (36 physical cores with hyperthreading) with 256GB RAM located at UC Merced. The processors communicate through shared memory. We used this only for the large-scale experiment, and we used 64 of the 72 processors. Detailed specs are in table~\ref{t:specs} (obtained by running \texttt{dmidecode} in the actual processor).
\end{description}
In both systems, the interprocess communication is handled by MPI (OpenMPI on the TSCC cluster and MPICH on our shared-memory cluster). The shared-memory system has both faster processors and faster communication than the distributed one and this is seen in our experiments (3--4 times faster). This does not imply that shared-memory systems are necessarily superior in practice, it simply reflects characteristics of the equipment we had access to. The ParMAC speedups as a function of the number of processors are comparable in both systems.
\begin{table}
\caption{Detailed hardware specification of the two machines used in our experiments.}
\centering
\begin{tabular}{@{}lll@{}}
\toprule
& Distributed-memory (TSCC at UCSD) & Shared-memory (cluster at UC Merced) \\
\midrule
CPU & Intel(R) Xeon(R) CPU E5-2670 0 & Intel(R) Xeon(R) CPU E5-2699 v3 \\
CPU cache & 20 MB & 45 MB \\
CPU max frequency & 3.3 GHz & 3.6 GHz \\
Cores/threads & 8/16 & 8/16 \\
Memory types & DDR3 800/1066/1333/1600 & DDR4 1600/1866/2133 \\
RAM bandwidth & 51.2 GB/s & 68 GB/s \\
Processor connection & 10GbE & shared memory \\
\bottomrule
\end{tabular}
\label{t:specs}
\end{table}
\paragraph{Datasets}
We have used 4 datasets commonly used as image retrieval benchmarks. (1) CIFAR \citep{Krizhev09a} contains $60\,000$ $32\times 32$ colour images in 10 object classes. We ignore the labels in this paper and use $N = 50\,000$ images as training set and $10\,000$ as test set. We extract $D = 320$ GIST features \citep{OlivaTorral01a} from each image. (2) SIFT-10K \citep{Jegou_11a} contains $N = 10\,000$ training high-resolution colour images and $100$ test images, each represented by $D = 128$ SIFT features. (3) SIFT-1M \citep{Jegou_11a} contains $N = 10^6$ training and $10^4$ test images. (3) SIFT-1B (\citealp{Jegou_11b}; \url{http://corpus-texmex.irisa.fr}) has three subsets: $10^9$ base vectors where the search is performed, $N = 10^8$ learning vectors used to train the model and $10^4$ query vectors.
\paragraph{Performance measures}
Regarding the quality of the BA and hash functions learnt, we report the following. (1) The binary autoencoder error $E_{\text{BA}}(\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}})$ which we want to minimise, eq.~\eqref{e:BA-nested}. (2) The quadratic-penalty function $E_Q(\ensuremath{\mathbf{h}},\ensuremath{\mathbf{f}},\ensuremath{\mathbf{Z}};\mu)$ which we actually minimise for each value of $\mu$, eq.~\eqref{e:BA-MAC-QP}. (3) The retrieval precision (\%) in the test set using as true neighbours the $K$ nearest images in Euclidean distance in the original space, and as retrieved neighbours in the binary space we use the $k$ nearest images in Hamming distance. We set $(K,k) = (1\,000,100)$ for CIFAR, $(100,100)$ for SIFT-10K and $(10\,000,10\,000)$ for SIFT-1M. For SIFT-1B, as suggested by the dataset creators, we report the recall@R: the average number of queries for which the nearest neighbour is ranked within the top $R$ positions (for varying values of $R$); in case of tied distances, we place the query as top rank. All these measures are computed offline once the BA is trained.
\paragraph{Models and their parameters}
We use BAs with linear encoders (linear SVM) except with SIFT-1B, where we also use kernel SVMs. The decoder is always linear. We set $L = 16$ bits (hash functions) for CIFAR, SIFT-10K and SIFT-1M and $L = 64$ bits for SIFT-1B. The \ensuremath{\mathbf{Z}}\ step uses enumeration for SIFT-10K and SIFT-1M, and alternating optimisation (initialised by a truncated relaxed solution) otherwise. We initialise the binary codes from truncated PCA ran on a subset of the training set (small enough that it fits in one machine).
To train the encoder ($L$ SVMs) and decoder ($D$ linear mappings) with stochastic optimisation, we used the SGD code from \citet{BottouBousquet08a} (\url{http://leon.bottou.org/projects/sgd}), using its default parameter settings. The SGD step size is tuned automatically in each iteration by examining the first $1\,000$ datapoints.
We use a multiplicative $\mu$ schedule $\mu_i = \mu_0 a^i$ where the initial value $\mu_0$ and the factor $a > 1$ are tuned offline in a trial run using a small subset of the data. For CIFAR we use $\mu_0 = 0.005$ and $a = 1.2$ over 26 iterations ($i = 0,\dots,25$). For SIFT-10K and SIFT-1M we use $\mu_0 = 10^{-6}$ and $a = 2$ over 20 iterations. For SIFT-1B we use $\mu_0 = 10^{-4}$ and $a = 2$ over 10 iterations.
\subsection{Effect of stochastic steps in the \ensuremath{\mathbf{W}}\ step}
\label{s:expts:stoch}
Figures~\ref{f:sift10k-epochs} to~\ref{f:cifar-epochs-shuffle} show the effect on SIFT-10K and CIFAR of varying the number of epochs and shuffling the data as a function of the number of machines $P$ on the learning curves (errors $E_Q$ and $E_{\text{BA}}$, precision). As the number of epochs increases, the \ensuremath{\mathbf{W}}\ step is solved more exactly (8 epochs is practically exact in these datasets). Fewer epochs, even just one, cause only a small degradation. The reason is that, although these are relatively small datasets, they contain sufficient redundance that few epochs are sufficient to decrease the error considerably. This is also helped by the accumulated effect of epochs over MAC iterations. Running more epochs increases the runtime and lowers the parallel speedup in this particular model, because we use few bits ($L = 16$) and therefore few submodels ($M = 2L = 32$) compared to the number of machines (up to $P = 128$), so the \ensuremath{\mathbf{W}}\ step has less parallelism.
Fig.~\ref{f:cifar-epochs-shuffle} shows the positive effect of data shuffling in the \ensuremath{\mathbf{W}}\ step. To shuffle the minibatches, the successor of a machine is given by a random lookup table. Shuffling generally reduces the error (this is particularly clear in $E_Q$, which is what we actually minimise) and increases the precision with no increase in runtime.
Note that, even if we keep the circular topology fixed throughout the \ensuremath{\mathbf{W}}\ step (i.e., we do not randomise the topology at each epoch), there is still a small amount of shuffling. This occurs because, although all submodels process the data minibatches in the same ``direction'', submodels in different machines start at different minibatches. Were it not for this (and if no shuffling was done within-machine), ParMAC would give an identical result no matter the number of machines. However, as seen from figures~\ref{f:sift10k-epochs} to~\ref{f:cifar-epochs-shuffle}, it seems that this small intrinsic shuffling simply randomises the learning curves, but does not make them better than the learning curve for one machine.
\begin{figure}[t]
\psfrag{iteration}{}
\psfrag{time}{}
\psfrag{QPerror}[][t]{\footnotesize$E_Q$}
\psfrag{BAerror}[][t]{\footnotesize$E_{\text{BA}}$}
\psfrag{precision}[][t]{precision}
\begin{tabular}{@{}c@{\hspace{0\linewidth}}c@{\hspace{0.015\linewidth}}c@{\hspace{0\linewidth}}c@{}}
\multicolumn{2}{c}{\makebox[0.45\linewidth][c]{\dotfill Number of epochs in \ensuremath{\mathbf{W}}\ step\dotfill}} & \multicolumn{2}{c}{\makebox[0.45\linewidth][c]{\dotfill Number of machines $P$\dotfill}} \\
error-iteration view & error-time view & 1 epoch in \ensuremath{\mathbf{W}}\ step & 8 epochs in \ensuremath{\mathbf{W}}\ step \\[-1ex]
\includegraphics[width=0.274\linewidth]{sift10k/sift10k_iter_objerr_1proc.eps} &
\includegraphics[width=0.24\linewidth]{sift10k/sift10k_time_objerr_1proc.eps} &
\includegraphics[width=0.235\linewidth]{sift10k/sift10k_iter_objerr_1epoch.eps} &
\includegraphics[width=0.235\linewidth]{sift10k/sift10k_iter_objerr_8epoch.eps} \\[-2.5ex]
\includegraphics[width=0.274\linewidth]{sift10k/sift10k_iter_nest_1proc.eps} &
\includegraphics[width=0.24\linewidth]{sift10k/sift10k_time_nest_1proc.eps} &
\includegraphics[width=0.235\linewidth]{sift10k/sift10k_iter_nest_1epoch.eps} &
\includegraphics[width=0.235\linewidth]{sift10k/sift10k_iter_nest_8epoch.eps} \\[-2.5ex]
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.274\linewidth]{sift10k/sift10k_iter_prec_1proc.eps} &
\psfrag{time}[t][]{time}
\includegraphics[width=0.24\linewidth]{sift10k/sift10k_time_prec_1proc.eps} &
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.235\linewidth]{sift10k/sift10k_iter_prec_1epoch.eps} &
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.235\linewidth]{sift10k/sift10k_iter_prec_8epoch.eps}
\end{tabular}
\caption{SIFT-10K dataset. \emph{Left two columns}: single machine ($P = 1$) and different number of epochs $e$ in the \ensuremath{\mathbf{W}}\ step. \emph{Right two columns}: fixed number of epochs (either 1 or 8) but different number of machines $P$.}
\label{f:sift10k-epochs}
\end{figure}
\begin{figure}[t]
\psfrag{iteration}{}
\psfrag{time}{}
\psfrag{QPerror}[][t]{\footnotesize$E_Q$}
\psfrag{BAerror}[][t]{\footnotesize$E_{\text{BA}}$}
\psfrag{precision}[][t]{precision}
\begin{tabular}{@{}c@{\hspace{0\linewidth}}c@{\hspace{0.015\linewidth}}c@{\hspace{0\linewidth}}c@{}}
\multicolumn{2}{c}{\makebox[0.45\linewidth][c]{\dotfill Number of epochs in \ensuremath{\mathbf{W}}\ step\dotfill}} & \multicolumn{2}{c}{\makebox[0.45\linewidth][c]{\dotfill Number of machines $P$\dotfill}} \\
error-iteration view & error-time view & 2 epochs in \ensuremath{\mathbf{W}}\ step & 8 epochs in \ensuremath{\mathbf{W}}\ step \\[-1ex]
\includegraphics[width=0.274\linewidth]{cifarpca/cifar_iter_objerr_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_time_objerr_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_iter_objerr_2epoch.eps} &
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_iter_objerr_8epoch.eps} \\[-2.5ex]
\includegraphics[width=0.274\linewidth]{cifarpca/cifar_iter_nest_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_time_nest_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_iter_nest_2epoch.eps} &
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_iter_nest_8epoch.eps} \\[-2.5ex]
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.274\linewidth]{cifarpca/cifar_iter_prec_1proc.eps} &
\psfrag{time}[t][]{time}
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_time_prec_1proc.eps} &
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_iter_prec_2epoch.eps} &
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.235\linewidth]{cifarpca/cifar_iter_prec_8epoch.eps}
\end{tabular}
\caption{CIFAR dataset. \emph{Left two columns}: single machine ($P = 1$) and different number of epochs $e$ in the \ensuremath{\mathbf{W}}\ step. \emph{Right two columns}: fixed number of epochs (either 2 or 8) but different number of machines $P$.}
\label{f:cifar-epochs}
\end{figure}
\begin{figure}[t]
\psfrag{iteration}{}
\psfrag{time}{}
\psfrag{QPerror}[][t]{\footnotesize$E_Q$}
\psfrag{BAerror}[][t]{\footnotesize$E_{\text{BA}}$}
\psfrag{precision}[][t]{precision}
\begin{tabular}{@{}c@{\hspace{0\linewidth}}c@{\hspace{0.015\linewidth}}c@{\hspace{0\linewidth}}c@{}}
\multicolumn{2}{c}{\makebox[0.45\linewidth][c]{\dotfill Number of epochs in \ensuremath{\mathbf{W}}\ step\dotfill}} & \multicolumn{2}{c}{\makebox[0.45\linewidth][c]{\dotfill Number of machines $P$\dotfill}} \\
error-iteration view & error-time view & 2 epochs in \ensuremath{\mathbf{W}}\ step & 8 epochs in \ensuremath{\mathbf{W}}\ step \\[-1ex]
\includegraphics[width=0.274\linewidth]{cifarpcashuffle/cifar_iter_objerr_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_time_objerr_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_iter_objerr_2epoch.eps} &
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_iter_objerr_8epoch.eps} \\[-2.5ex]
\includegraphics[width=0.274\linewidth]{cifarpcashuffle/cifar_iter_nest_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_time_nest_1proc.eps} &
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_iter_nest_2epoch.eps} &
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_iter_nest_8epoch.eps} \\[-2.5ex]
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.274\linewidth]{cifarpcashuffle/cifar_iter_prec_1proc.eps} &
\psfrag{time}[t][]{time}
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_time_prec_1proc.eps} &
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_iter_prec_2epoch.eps} &
\psfrag{iteration}[t][]{iteration}
\includegraphics[width=0.235\linewidth]{cifarpcashuffle/cifar_iter_prec_8epoch.eps}
\end{tabular}
\caption{Like fig.~\ref{f:cifar-epochs} but with and without minibatch shuffling in the \ensuremath{\mathbf{W}}\ step (solid and dashed lines, resp.).}
\label{f:cifar-epochs-shuffle}
\end{figure}
\subsection{Speedup}
\label{s:expts:speedup}
The fundamental advantage of ParMAC and distributed optimisation in general is the ability to train on datasets that do not fit in a single machine, and the reduction in runtime because of parallel processing. Fig.~\ref{f:speedup} (top row) shows the strong scaling%
\footnote{In ``strong scaling'', the total problem size is fixed and the problem size on each machine is inversely proportional to the number of machines $P$. In ``weak scaling'', the problem size on each machine is fixed, so the total problem size is proportional to $P$. High speedups are easier to obtain in weak scaling than in strong scaling \citep{GoedecHoisie01a}.}
speedups achieved experimentally, as a function of the number of machines $P$ for fixed problem size (dataset and model), in CIFAR and SIFT-1M ($N =$ 50K and 1M training points, respectively). Even though these datasets and especially the number of independent submodels ($M = 2L = 32$ effective submodels of the same size, as discussed in section~\ref{s:speedup-th:practical}) are relatively small, the speedups we achieve are nearly perfect for $P \le M$ and hold very well for $P > M$ up to the maximum number of machines we used ($P = 128$ in the distributed system). The speedups flatten as the number of epochs (and consequently the amount of communication) increases, because for this experiment the bottleneck is the \ensuremath{\mathbf{W}}\ step, whose parallelisation ability (i.e., the number of concurrent processes) is limited by $M = 2L$ (the \ensuremath{\mathbf{Z}}\ step has $N$ independent processes and is never a bottleneck, since $N$ is very large). However, as noted earlier, using 1 to 2 epochs gives a good enough result, very close to doing an exact \ensuremath{\mathbf{W}}\ step. The runtime for SIFT-1M on $P = 128$ machines with 8 epochs was 12 minutes and its speedup 100$\times$. This is particularly remarkable given that the original, nested model did not have model parallelism. These speedups are vastly larger than those achieved by earlier large-scale nonconvex optimisation approaches such as Google's DistBelief system \citep{Le_12a,Dean_12a}, although admittedly the deep nets trained there were far larger than our BAs.
Fig.~\ref{f:speedup} (bottom) shows the speedups predicted by our theoretical model of section~\ref{s:speedup-th}. We set the parameters $e$ and $N$ to their known values, and $M = 2L = 32$ for CIFAR and SIFT-1M and $M = 2L = 128$ for SIFT-1B (effective number of independent equal-size submodels). For the time parameters, we set $t^{\ensuremath{\mathbf{W}}}_r = 1$ to fix the time units, and we set $t^{\ensuremath{\mathbf{W}}}_c$ and $t^{\ensuremath{\mathbf{Z}}}_r$ by trial and error to achieve a reasonably good fit to the experimental speedups. Specifically, we set $t^{\ensuremath{\mathbf{W}}}_c = 10^4$ for both datasets, and $t^{\ensuremath{\mathbf{Z}}}_r = 200$ for CIFAR and $40$ for SIFT-1M. Although these are fudge factors, they are in rough agreement with the fact that communicating a weight vector over the network is orders of magnitude slower than updating it with a gradient step, and that the \ensuremath{\mathbf{Z}}\ step is quite slower than the \ensuremath{\mathbf{W}}\ step because of the binary optimisation it involves.
Fig.~\ref{f:speedup} (bottom, right plot) also shows the theoretical prediction for the SIFT-1B dataset ($N = 10^8$, $M = 128$), using the same parameters as in SIFT-1M (again assuming the distributed memory system): $t^{\ensuremath{\mathbf{W}}}_c = 10^4$ and $t^{\ensuremath{\mathbf{Z}}}_r = 40$. Since here $M$ is quite larger and $N$ is much larger, the speedup is nearly perfect over a very wide range (note the plot goes up to $P = 1\,024$ machines, even though our experiments are limited to $P = 128$).
\begin{figure}[t]
\centering
\psfrag{processor}{}
\begin{tabular}{@{}c@{\hspace{0\linewidth}}c@{\hspace{0\linewidth}}c@{}}
CIFAR & SIFT-1M & SIFT-1B \\
\psfrag{speedup}[][t]{speedup $S(P)$ (experiment)}
\includegraphics[height=0.25\linewidth]{cifar_proc_speedup.eps} &
\psfrag{speedup}{}
\includegraphics[height=0.25\linewidth]{sift1m_proc_speedup-miguel.eps} &
\raisebox{0.125\linewidth}{\caja{c}{c}{too long to run}} \\[2ex]
\psfrag{processor}[t][]{number of machines $P$}
\psfrag{speedup}[][t]{speedup $S(P)$ (theory)}
\includegraphics[height=0.25\linewidth]{cifar_proc_speedup-theory.eps} &
\psfrag{processor}[t][]{number of machines $P$}
\psfrag{speedup}{}
\includegraphics[height=0.25\linewidth]{sift1m_proc_speedup-theory.eps} &
\psfrag{processor}[t][]{number of machines $P$}
\psfrag{speedup}{}
\includegraphics[height=0.25\linewidth]{sift1b_proc_speedup-theory.eps}
\end{tabular}
\caption{Speedup $S(P)$ as a function of the number of machines $P$ for CIFAR, SIFT-1M and SIFT-1B. Top: experimental result in the distributed memory system. Bottom: theoretical result predicted by the model of section~\ref{s:speedup-th}. The dataset size and number of submodels $(N,M)$ is $(50\,000,32)$ for CIFAR, $(10^6,32)$ for SIFT-1M and $(10^8,128)$ for SIFT-1B.}
\label{f:speedup}
\end{figure}
\subsection{Large-scale experiment: SIFT-1B dataset}
\label{s:expts:large-scale}
SIFT-1B is one of the largest datasets, if not the largest one, that are publicly available for comparing nearest-neighbour search algorithms with known ground-truth (i.e., precomputed exact Euclidean distances for each query to its $k$ nearest vectors in the base set). The training set contains $N = 100$M vectors, each consisting of 128 SIFT features.
Handling the SIFT-1B dataset required special care because of its size and the limited amount of memory (total 512GB for 128 processors in the distributed system and 256GB for 64 processors in the shared-memory one). Each vector has 128 SIFT features and each feature in the original dataset is stored in a single byte rather than as double-precision floats (8 bytes), as in our other experiments, totalling 12.8GB for the training set if using a linear hash function and 200GB if using an RBF one (see below). Rather than converting it to floats, which would exceed 1TB, we modified our code to convert each feature only as needed. In the \ensuremath{\mathbf{Z}}\ step each datapoint is processed independently and the conversion to double is done one point at a time. In the \ensuremath{\mathbf{W}}\ step it is done one minibatch at a time. It is of course possible to use hard disk as additional storage but this would slow down training. The auxiliary coordinates, which must be stored in MAC algorithms, take only 6.25\% the memory of the data (64 bits per datapoint compared to 128 bytes).
We used $L = 64$ bits (hash functions). As hash function, we trained a linear SVM as before, and a kernel SVM using $m$ Gaussian radial basis functions (RBF) with fixed bandwidth $\sigma$ and centres. This means the only trainable parameters are the weights, so the MAC algorithm does not change except that it operates on an $m$-dimensional input vector of kernel values (stored as one byte each), instead of the 128 SIFT features. The Gaussian kernel values are in $(0,1]$ but, as before, to save memory we store them as an unsigned one-byte integer value in $[0,255]$. We used $m = 2\,000$ centres, the maximum we could fit in memory, picked at random from the training set. In trials with a subset of the training set, we set $\sigma = 160$. This worked well and was wide enough to ensure that, with our limited one-byte precision, no data point would produce $m$ zeros as kernel values.
On trials on a subset of the training dataset, we set the number of epochs to $e = 2$ with shuffling (we observed no improvements by using more epochs, which is understandable given the size of the dataset). We initialised the binary codes from truncated PCA trained on a subset of size 1M (recall@R=100: 55.2\%), which gave results comparable to the baseline in \citet{Jegou_11b} if using 8 bytes per indexed vector (without postprocessing by reranking as done in that paper).
We ran ParMAC on the whole training set in the distributed system with 128 processors for 6 iterations and in the shared-memory one with 64 processors for 10 iterations. The results are given in the following table and figures~\ref{f:sift1m-iters}--\ref{f:sift1m-recall}:
\begin{center}
\begin{tabular}{@{}lccc@{}}
\toprule
\raisebox{-0.7ex}[0pt][0pt]{\caja{t}{l}{Hash function \\ (encoder)}} & \raisebox{-0.7ex}[0pt][0pt]{\caja{t}{c}{Recall \\ @R=100}} & \multicolumn{2}{c}{Time (hours)} \\
\cmidrule{3-4}
& & distrib. & shared \\
\midrule
linear SVM & 61.5\% & 29.30 & 11.04 \\
kernel SVM & 66.1\% & 83.44 & 32.19 \\
\bottomrule
\end{tabular}
\end{center}
The learning curves (fig.~\ref{f:sift1m-iters}) are essentially identical over both systems. The nonlinear RBF hash function outperforms the linear one in recall, as one would expect. The improvement occurs across the whole range of $R$ recall values (fig.~\ref{f:sift1m-recall}). Note the error in the nested model, $E_{\text{BA}}$, does not decrease monotonically. This is because MAC optimises instead the penalised function $E_Q$, in an effort the minimise $E_{\text{BA}}$ as $\mu$ increases.
Based on our previous results, the small number of epochs and the larger number of submodels in the \ensuremath{\mathbf{W}}\ step, we expect nearly perfect speedups. We cannot compute the actual speedup because the single-machine runtime is enormous. Using a scaled-down model and training set, we estimated that training in one machine (with enough RAM to hold the data and parameters) would take months.
Although the speedups are comparable on both the distributed and the shared-memory system, the former is 3--4 times slower. The reason is the distributed system has both slower processors and slower interprocessor communication (across a network); see also fig.~\ref{f:shared_dist}.
\begin{figure}[p]
\centering
\psfrag{recallR}[][t]{recall@R=100}
\psfrag{AEerr}[][t]{$E_{\text{BA}}$}
\psfrag{iteration}[][]{iteration}
\begin{tabular}{@{}c@{\hspace{0.05\linewidth}}c@{}}
\includegraphics[height=0.33\linewidth]{sift1b/iter_err.eps} &
\includegraphics[height=0.33\linewidth]{sift1b/iter_recallR.eps}
\end{tabular}
\caption{SIFT-1B dataset, using the shared-memory and distributed system (solid and dashed lines, resp.).}
\label{f:sift1m-iters}
\end{figure}
\begin{figure}[p]
\centering
\begin{tabular}{@{}c@{}c@{}c@{}}
\psfrag{recallR}[][]{recall@R}
\psfrag{R}[t][]{$R$}
\includegraphics[height=0.26\linewidth]{sift1b/r_recallR_comp.eps} &
\psfrag{recallR}{}
\psfrag{R}[t][]{$R$ (linear hash function)}
\includegraphics[height=0.26\linewidth,bb=5 158 588 616,clip]{sift1b/linmodified.eps} &
\psfrag{recallR}{}
\psfrag{R}[t][]{$R$ (RBF hash function)}
\includegraphics[height=0.26\linewidth,bb=5 158 588 616,clip]{sift1b/rbfmodified.eps}
\end{tabular}
\caption{Recall@R on the SIFT-1B dataset for truncated PCA (initialisation), linear and kernel hash functions (left plot: final result; right two plots: over iterations, as labelled).}
\label{f:sift1m-recall}
\end{figure}
\begin{figure}[p]
\centering
\psfrag{settings}[t][]{\# nodes $\times$ \# processors}
\psfrag{time (s)}[][t]{time (s)}
\includegraphics[width=0.40\linewidth]{shard_vs_dist.eps}
\caption{Time spent on communication and computation as a function of the number of processors per node in the TSCC cluster. The time for our shared-memory UC Merced cluster (also using 16 processors, i.e., corresponding to 1$\times$16) is 2.57 and 8.76 seconds for communication and computation, respectively.}
\label{f:shared_dist}
\end{figure}
\subsection{Shared-memory vs distributed systems}
The TSCC distributed cluster consists of nodes containing 16 processors and 64GB RAM. These processors communicate through shared-memory within a node, and across a network otherwise (which is slower). In all our experiments up to now, we always allocated processors within the same node if possible. But, depending on whether a user requests processors within or across nodes, there is then a tradeoff between both communication modes. A full study of this issue is beyond our scope, which is to understand the ParMAC algorithm in general rather than for specific computer architectures. However, we ran a small experiment to evaluate the computation and communication time spent as a function of the number of processors per node. We set the total number of processors to $P = 16$ but allocated them in the following configurations: from a single node with all 16 processors (1$\times$16, pure shared-memory) to 16 nodes each with 1 processor (16$\times$1, pure distributed), and intermediate configurations such as 2 nodes each with 8 processors (2$\times$8). We used the RBF hash function from the SIFT-1B experiment with all settings as before and trained it on a subset of 20K points for a single iteration. Figure~\ref{f:shared_dist} shows the resulting times. While the computation time remains constant, the communication time increases as we move from shared-memory to distributed settings, as expected. Hence, the effect on the ParMAC algorithm would be to increase the \ensuremath{\mathbf{W}}\ step runtime correspondingly and lower the parallel speedup.
\section{Discussion}
\label{s:discussion}
Developing parallel, distributed optimisation algorithms for nonconvex problems in machine learning is challenging, as shown by recent efforts by large teams of researchers \citep{Le_12a,Dean_12a}. One important advantage of ParMAC is its simplicity. Data and model parallelism arise naturally thanks to the introduction of auxiliary coordinates. The corresponding optimisation subproblems can often be solved reusing existing code as a black box (as with the SGD training of SVMs and linear mappings in the BA). A circular topology is sufficient to achieve a low communication between machines. There is no close coupling between the model structure and the distributed system architecture. The development and implementation of ParMAC for binary autoencoders on large datasets in a distributed cluster was achieved in a few months by the PI and one junior PhD student (both without prior experience in MPI).
Rather than an algorithm, MAC is a meta-algorithm that can produce a specific optimisation algorithm for a given nested problem, depending on how the auxiliary coordinates are introduced and on how the resulting subproblems are solved (in this sense, MAC is similar to expectation-maximisation (EM) algorithms; \citealp{Dempst_77a}). For example, in the low-dimensional SVMs of \citet{WangCarreir14a}, the \ensuremath{\mathbf{Z}}\ step is a small quadratic program for each data point. However, regardless of these specifics, the resulting MAC algorithm typically exhibits a \ensuremath{\mathbf{W}}\ step with $M$ independent submodels and a \ensuremath{\mathbf{Z}}\ step with $N$ independent coordinates for the data points. Likewise, the specifics of a ParMAC algorithm (how the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ steps are optimised) will depend on its corresponding MAC algorithm. However, it will always split the data and auxiliary coordinates over machines and consist of a \ensuremath{\mathbf{Z}}\ step with no communication between machines, and a \ensuremath{\mathbf{W}}\ step where submodels visit machines in a circular topology, effectively training themselves by stochastic optimisation.
Further improvements can be made in specific problems. For example, it is possible to have further parallelisation or less dependencies (e.g.\ the weights of hidden units in layer $k$ of a neural net depend only on auxiliary coordinates in layers $k$ and $k+1$). This may reduce the communication in the \ensuremath{\mathbf{W}}\ step, by sending to a given machine only the model portion that it needs, or by allocating cores within a multicore machine accordingly. Also, the \ensuremath{\mathbf{W}}\ and \ensuremath{\mathbf{Z}}\ step optimisations can make use of further parallelisation by GPUs or by distributed convex optimisation algorithms. And, if the submodels are small in size, it may be better for each machine to operate on a set of submodels and then send them all together in a larger message (rather than sending each submodel as it is finished), since this will reduce latency overheads (i.e., the setup cost of a message). Many more refinements can (and should) be done in an industrial implementation. For example, one can store and communicate reduced-precision values for data and parameters with little effect of the accuracy, as has been done in neural nets (e.g.\ \citealp{Gupta_15a,Han_15a,Han_16a}). Various system-dependent optimisations may be possible (beyond those that a good compiler may be able to do, such as loop unrolling or code inlining), such as improving the spatial or temporal locality of the code given the type and size of the cache in each machine. In this paper, we have tried to keep our implementation as simple as possible, because our goal was to understand the parallelisation speedups of ParMAC in a setting as general as possible, rather than trying to achieve the very best performance for a particular dataset, model or distributed system.
ParMAC is very efficient in communication: no data or coordinates are ever sent, only the entire model $e+1$ times per iteration. Using one epoch (which is sufficient in large datasets), or using $e$ epochs but performing them within each machine, the entire model is sent twice per iteration. This is near optimal if we note the following. If the data cannot be communicated, then at every iteration each submodel must visit each machine (for it to be trained on the entire data). Hence, any correct algorithm will have to send the entire model at least once; ParMAC does so twice. Also, the circular topology is the minimal topology (considered as a directed graph on the $P$ machines) that is necessary to be able to optimise a global model on the entire dataset with $P$ machines, because each machine must be able to communicate with some other machine. It has $P$ edges and is truly distributed, with each machine having the same importance.
A popular model for distributed optimisation (e.g.\ with parallel SGD) uses a worker-server connectivity: $W \gg 1$ workers that do actual parameter optimisation and $S \ge 1$ ``parameter servers'' that collect and broadcast parameters to worker machines. This is a bipartite graph with bidirectional edges between servers and workers, having $SW$ edges, which is quite larger than the number of machines $P = S+W$. The entire model must be sent twice per iteration, to and from the parameter server, but this creates a bottleneck when multiple workers send data to the same server, and $S \ll W$ in practice. No such bottleneck occurs in ParMAC.
Parallel and distributed computing systems have been around for decades. One important class are supercomputers, which are carefully designed in terms of the processors, memory system and connection network. They have been traditionally used to solve a wide variety of large-scale scientific computation problems, such as weather prediction, nuclear reactor modelling, or astrophysical or molecular simulations. Another important class are clusters of inexpensive, heterogenous workstations connected through an Ethernet network, with workstations differing in speed, memory/disk capacity, number of cores/GPUs, etc. This is used in data centres in Google, Amazon and other companies, and also in distributed computation models such as SETI@home that capitalise on the computation and Internet connectivity available to individuals, and their willingness to donate them to projects they find worthy. In these systems, the machine learning task may be one of other tasks running concurrently, such as web searches or email in a data centre (which may operate on the same data as the machine learning task), or personal applications in an individual's workstation. Supercomputers and clusters differ considerably across important factors: suitability for a particular problem, computation and communication speed, size of memory and disk, connection network, fault tolerance, load, cost, energy consumption, etc. At present it is unclear what the best choices will be for machine learning models (which exhibit a wide variety themselves), and we expect to see many different possibilities been researched in the immediate future. We suggest that ParMAC, by itself or in combination with other techniques, may play an important role with nested models because of the embarrassing parallelism it introduces and its loose demands on the underlying distributed system.
\section{Conclusion}
\label{s:concl}
We have proposed ParMAC, a distributed model for the method of auxiliary coordinates for training nested, nonconvex models in general, analysed its parallel speedup and convergence, and demonstrated it with an MPI-based implementation for a particular case, to train binary autoencoders. MAC creates parallelism by introducing auxiliary coordinates for each data point to decouple nested terms in the objective function. ParMAC is able to translate the parallelism inherent in MAC into a distributed system by 1) using data parallelism, so that each machine keeps a portion of the original data and its corresponding auxiliary coordinates; and 2) using model parallelism, so that independent submodels (weight vectors of a hash function or hidden unit) visit every machine in a circular topology, effectively executing epochs of a stochastic optimisation, without the need for a parameter server and therefore no communication bottlenecks. This keeps the communication between machines to a minimum within each iteration. In this sense, ParMAC can be seen as a strategy to be able to use existing, well-developed (convex) distributed optimisation techniques---applicable to simple functions---to a setting where simple functions are coupled by nesting into a nonconvex function whose training data is distributed over machines. The convergence properties of MAC (to a stationary point of the objective function) remain essentially unaltered in ParMAC. The parallel speedup can be theoretically predicted to be nearly perfect when the number of submodels is comparable or larger than the number of machines, and to eventually saturate as one continues to increase the number of machines, and indeed this was confirmed in our experiments. ParMAC also makes it easy to account for data shuffling, load balancing, streaming and fault tolerance. Hence, we expect that ParMAC could be a basic building block, in combination with other techniques, for the distributed optimisation of nested models in big data settings.
\subsubsection*{Acknowledgements}
Work supported by a Google Faculty Research Award and by NSF award IIS--1423515. We thank Dong Li (UC Merced) for useful discussions about MPI and performance evaluation on parallel systems, and Quoc Le (Google) for useful discussion about Google's DistBelief system.
\clearpage
|
\section{{\bf Introduction}} \label{sect.a}
Given a convex optimization problem possessing little more structure than having a known strictly-feasible point, we provide a simple transformation to an equivalent convex optimization problem which has only linear equations as constraints, and has Lipschitz-continuous objective function defined on the whole space. Virtually any subgradient method can be applied to solve the equivalent problem. Relying only on a line-search during an iteration, the resulting iterate is made to be a feasible point for the original problem. Moreover, this sequence of feasible points converges to within user-specified error of optimality. \vspace{1mm}
The algorithm can be implemented directly in terms of the original problem, leaving the equivalent problem hidden from view.
In this introductory section we present two examples of algorithms and complexity bounds that ensue when standard subgradient methods are applied to the equivalent problem. Not until subsequent sections is the equivalent problem revealed. \vspace{1mm}
Let $ {\mathcal E} $ denote a finite-dimensional real vector space with inner product $ \langle \; , \; \rangle $ and associated norm $ \| \, \, \| $. \vspace{1mm}
Consider an optimization problem
\begin{equation} \label{eqn.aa}
\begin{array}{rl} \min & f(x) \\
\textrm{s.t.} & x \in \mathrm{Feas} \; . \end{array}
\end{equation}
Assume $ f: {\mathcal E} \rightarrow (- \infty, \infty] $ is an extended-valued, lower-semicontinous convex function. Equivalently, assume $ f: {\mathcal E} \rightarrow \mathbb{R} $ has closed and convex epigraph, $ \mathrm{epi}(f) := \{ (x,t) \mid x \in \dom(f), \, t \geq f(x) \} $, (where $ \dom(f) $ is the (effective) domain of $ f $ (i.e., the set on which $ f $ is finite)). \vspace{1mm}
Let
\[ \mathrm{Feas} = \{ x \in S \mid Ax = b \} \; , \]
where $ S $ is a closed, convex set with nonempty interior. \vspace{1mm}
Assume $ \bar{e} $ is a known feasible point contained in $ \mathrm{int}( S \cap \mathrm{\mathrm{dom}(f)}) $. \vspace{1mm}
Let $ \bar{D} $ denote the diameter of the sublevel set $ \{ x \in \mathrm{Feas}: f(x) \leq f( \bar{e} ) \} $ (i.e., the supremum of distances between pairs of points in the set). We assume $ \bar{D} $ is finite, which together with previous assumptions implies the optimal value of (\ref{eqn.aa}) is attained at some feasible point. Let $ f^* $ denote the optimal value. \vspace{1mm}
Let $ \hat{f} $ be a user-chosen constant satisfying $ \hat{f} > f( \bar{e} ) $ (thus, $ (\bar{e}, \hat{f}) \in \mathrm{int}(\mathrm{epi}(f)) $). Define
\[
\hat{r} := \sup \{ r \mid \| x - \bar{e} \| \leq r \textrm{ and } Ax = b \, \, \Rightarrow \, \, x \in S \textrm{ and } f(x) \leq \hat{f} \} \; . \]
Note $ \hat{r} > 0 $, because $ \bar{e} \in \mathrm{int}\big( S \cap \mathrm{dom}(f)\big) $. \vspace{1mm}
The scalars $ \bar{D} $ and $ \hat{r} $ appear in the complexity bounds for the representative algorithms presented below, but their values are not assumed to be known. We now introduce notation to be used in specifying the algorithms. \vspace{1mm}
For fixed $ x \in {\mathcal E} $ and $ t \in \mathbb{R} $, and for scalars $ \alpha $, let
\[ x(\alpha) := \bar{e} + \alpha \cdot (x - \bar{e} ) \quad \textrm{and} \quad t( \alpha) := \hat{f} + \alpha \cdot ( t - \hat{f}) \; . \]
Define
\[ \alpha_1(x) := \sup \{ \alpha \mid x(\alpha) \in S \} \; , \]
and
\[ \alpha_2(x,t) := \sup \{ \alpha \mid f(x( \alpha)) \leq t( \alpha) \} \; . \]
Note that $ \alpha_1(x), \alpha_2(x,t) > 0 $, due to our requirements for $ \bar{e} $ and $ \hat{f} $. \vspace{1mm}
We are concerned only with pairs $ (x,t) $ for which $ Ax = b $ and $ t < \hat{f} $. (Observe then, $ A x( \alpha) = b $ and $ t(\alpha) < \hat{f} $ for all $ \alpha \geq 0 $.) We claim in this setting, at least one of the values $ \alpha_1(x) $, $ \alpha_2(x,t) $ is finite. To establish the claim, assume $ \alpha_2(x,t) = \infty $ -- we show it easily follows that $ \alpha_1(x) $ is finite. \vspace{1mm}
Since $ \alpha_2(x,t) = \infty $, the convex univariate function
\begin{equation} \label{eqn.ab}
\phi(\alpha) := f( x( \alpha)) - t( \alpha)
\end{equation}
has negative value for all $ \alpha \geq 0 $. Hence, since $ t < \hat{f} $ (by assumption) -- and thus $ t( \alpha) \rightarrow - \infty $ as $ \alpha \rightarrow \infty $ -- the function $ \alpha \mapsto f( x( \alpha)) $ is unbounded below on the interval $ [0, \infty ) $. In particular, $ f(x( \alpha)) < f^* $ for some $ \alpha > 0 $, implying for this value of $ \alpha $, $ x( \alpha) \notin \mathrm{Feas} $, that is, $ \alpha_1(x) < \alpha $, establishing the claim. \vspace{1mm}
Let
\[ \alpha(x,t) := \min \{ \alpha_1(x), \alpha_2(x,t) \} \; , \]
a value we now know to be positive and finite (assuming $ Ax = b $ and $ t < \hat{f} $). For given pairs $ (x,t) $, the algorithms require that the scalar $ \alpha(x,t) $ be computed. (This is the line search referred to above.) In general, of course, $ \alpha(x,t) $ cannot be computed exactly. However, there is hope for quickly gaining a ``good enough'' approximation to $ \alpha(x,t) $. \vspace{1mm}
To understand, first observe for many optimization problems, for each $ x $ the value $ \alpha_1(x) $ can readily be accurately approximated, or determined to be $ \infty $. All that is required here is finding whether a given half-line, with endpoint in $ \mathrm{int}(S) $, intersects the boundary of $ S $, and if so, accurately approximating where the intersection occurs. For most feasible regions, this is far more easily accomplished than one of the main operations underlying traditional subgradient methods, where an (arbitrary) point outside $ \mathrm{Feas} $ must be orthogonally projected onto $ \mathrm{Feas} $. \vspace{1mm}
Assume for the moment that $ \alpha_1(x) $ has been accurately approximated or determined to equal $ \infty $. For definiteness, assume $ \alpha_1(x) = \infty $, in which case we know $ \alpha_2(x,t) $ is finite. Approximating $ \alpha_2(x,t) $ is the same as approximating the solution to $ \phi(\alpha) = 0 $, where $ \phi $ is the function (\ref{eqn.ab}). As $ \phi $ is a convex function satisfying $ \phi (0) < 0 $, approximating the root can be accomplished by first finding a value $ \alpha $ sufficiently large so as to satisfy $ \phi(\alpha) > 0 $, and then proceeding to isolate the root using bisection. \vspace{1mm}
On the other hand, if $ \alpha_1(x) $ is finite, then to determine whether $ \alpha_2(x,t) $ even needs to be computed, check whether $ \phi( \alpha_1(x)) \leq 0 $. If so, there is no need to compute $ \alpha_2(x,t) $, because $ \alpha(x,t) = \alpha_1(x) $. If instead, $ \phi(\alpha_1(x)) > 0 $, then to compute $ \alpha_2(x,t) $ proceed by bisection, with initial \vspace{1mm} interval $ [0, \alpha_1(x)] $. \vspace{1mm}
That bisection can be applied gives reason to hope a good approximation to $ \alpha(x,t) $ can be quickly computed. Nonetheless, questions abound regarding what constitutes a ``good enough'' approximation. In what follows, we duck the issue, simply assuming $ \alpha(x,t) $ can be computed exactly. \vspace{1mm}
For $ x,t $ satisfying $ Ax = b $ and $ t < \hat{f} $, let
\begin{equation} \label{eqn.ac}
\pi'(x,t) := (\bar{e} , \hat{f} ) + \alpha(x,t) \cdot ( (x,t) - (\bar{e} , \hat{f})) \; .
\end{equation}
Thus, in traveling from $ (\bar{e} , \hat{f}) $ in the direction $ (x,t) - (\bar{e} , \hat{f}) $, $ \pi'(x,t) $ is the first point $ (x',t') $ encountered for which either $ x' \in \mathrm{bdy}(S) $ (boundary) or $ (x',t') \in \bdy(\epi(f)) $. Note in particular that if $ (x',t') = \pi' (x,t) $, then $ x' $ is both feasible and lies in the domain of $ f $.
\vspace{1mm}
For $ x' \in \mathrm{bdy}(S) $, let
\[ G_1(x') := \left\{ \frac{-1}{ \langle v , \bar{e} - x' \rangle } \, v \mid \, \vec{0} \neq v \in N_S(x') \right\} \; ,
\]
where $ N_S(x') $ is the normal cone to $ S $ at $ x' $ ($\{ v \mid \forall x \in S, \langle v, x - x' \rangle \leq 0 \} $). The denominator is negative, because $ \bar{e} \in \mathrm{int}(S) $. \vspace{1mm}
For $ (x',t') \in \mathrm{\bdy}(\mathrm{epi}(f)) $, let
\[ G_2(x',t') := \left\{ \frac{-1}{\langle v, \bar{e} - x' \rangle + ( \hat{f} - t') \delta } \, v \mid \, (\vec{0},0) \neq (v,\delta) \in N_{\mathrm{epi}(f)}(x',t') \right\} \; , \]
where here, normality is with respect to the inner product that assigns pairs $ (x_1, t_1) $, $ (x_2, t_2) $ the value $ \langle x_1, x_2 \rangle + t_1 t_2 $. \vspace{1mm}
If $ (x',t') \in \mathrm{bdy}(\mathrm{dom}(f)) $ and $ x' \in \mathrm{int}(\mathrm{dom}(f)) $ -- hence $ t' = f(x') $ -- a more concrete description applies:
\begin{align*}
& x' \in \mathrm{int}(\mathrm{dom}(f)) \, \\
& \qquad \Rightarrow \, G_2(x', f(x')) =
\left\{ \frac{1}{\hat{f} - ( f(x') + \langle v, \bar{e} - x' \rangle )}\, v \mid \, v \in \partial f(x') \right\} \; ,
\end{align*}
where $ \partial f(x') $ is the subdifferential of $ f $ at $ x' $ (set of subgradients). We rely only on sets $ G_2(x',t') $ for which $ x' \in \mathrm{Feas} $ -- thus, if $ \mathrm{Feas} \cap \mathrm{dom}(f) \subseteq \mathrm{int}(\mathrm{dom}(f)) $, the more concrete description applies. (In the traditional literature on subgradient methods, the stronger condition $ \mathrm{Feas} \subset \mathrm{int}(\mathrm{dom}(f)) $ is generally assumed, and hence the more concrete description certainly applies.) \vspace{1mm}
Finally, letting $ \bdy_1 := \{ (x',t') \mid x' \in \bdy(S) \} $ and $ \bdy_2 := \bdy(\epi(f)) $, define for $ (x',t') \in \bdy_1 \cup \bdy_2 \; , $
\[ G(x',t') := \begin{cases} G_1(x') & \textrm{if $ (x',t') \in \bdy_1 \setminus \bdy_2 \; , $} \\
G_2(x',t') & \textrm{if $ (x',t') \in \bdy_2 \setminus \bdy_1\; , $} \\
\textrm{hull}(G_1(x') \, \cup \, G_2(x',t') ) & \textrm{if $ (x',t') \in \bdy_1 \cap \bdy_2 \; , $}
\end{cases} \]
where ``hull'' denotes the convex hull. \vspace{1mm}
Let $ P $ denote orthogonal projection onto the kernel of $ A $. \vspace{1mm}
Following is a representative algorithm that arises from the framework developed in subsequent sections.
\noindent
\hrulefill
\noindent {\bf Algorithm A} \vspace{1mm}
\noindent
\begin{tabular}{lll}
(0) & Input: & $ 0 < \epsilon < 1 $ , \\
& & $ \bar{e} \in \mathrm{int}(S \cap \mathrm{dom}(f)) $ satisfying $ A \bar{e} = b $, and \\
&& $ \hat{f} $, a scalar satisfying $ \hat{f} > f( \bar{e} ) $. \\
& Initialize: & $ (x_0, t_0) := (\bar{e} , f( \bar{e} )) $ and $ (x_0',t_0') := (\bar{e} , f( \bar{e} )) $. \\
(1) & Iterate: & Compute $ \tilde{x}_{k+1} := x_k - \smfrac{\epsilon }{2 \| P(g') \|^2} P(g') $, \, where $ g' \in G( x_k', t_k') $. \\
&& Let $ \alpha_{k+1} := \alpha( \tilde{x}_{k+1}, t_k) $ and $ (x_{k+1}', t_{k+1}') := \pi' (\tilde{x}_{k+1} , t_k) $. \\
&& If $ \alpha_{k+1} \geq 4/3 $, then let $ (x_{k+1}, t_{k+1}) := (x_{k+1}', t_{k+1}') $, \\
&& $ \textrm{~} $ \quad else let $ (x_{k+1}, t_{k+1}) := (\tilde{x}_{k+1} , t_k) $. \vspace{1mm}
\end{tabular}
\noindent
\hrulefill
\vspace{2mm}
Critical to understanding the algorithm is the identity $ (x_k', t_k') = \pi'(x_k,t_k) $. Iterates in the sequence $ \{ x_k \} $ need not be feasible nor in the domain of $ f $, whereas $ x_k' $ is feasible and in the domain. \vspace{1mm}
Unlike traditional subgradient methods, no orthogonal projections onto $ \mathrm{Feas} $ are required.
The only projections are onto the kernel of $ A $, projections which are computed efficiently if the range space is low dimensional, and if as a preprocessing step the linear operator $ (A A^*)^{-1} $ is computed and stored in memory (where $ A^* $ is the adjoint of $ A $, and where we are assuming, of course, that $ A $ is surjective, for the inverse to exist). \vspace{1mm}
\begin{thm} \label{thm.aa}
For the feasible sequence $ \{ x_k' \} $ generated by Algorithm A,
\begin{align}
& \ell \geq 8 \, \left( \frac{\bar{D} }{\hat{r} } \right)^2 \, \left( \, \frac{1}{\epsilon^2} \, + \, \frac{1}{\epsilon} \, \log_{4/3} \left( 1 + \frac{\bar{D} }{\hat{r} } \right) \, \right) \label{eqn.ad} \\
& \quad \quad \Rightarrow \quad \min_{k \leq \ell } \, \frac{f(x_k') - f^*}{\hat{f} - f^*} \, \leq \, \epsilon \; . \label{eqn.ae}
\end{align}
\end{thm}
\vspace{2mm}
This result differs markedly from the traditional literature, in that there appears no Lipschitz constant for $ f $. The theorem applies even when for no neighborhood of an optimal solution does a (local) Lipschitz constant exist, such as occurs, for example, when $ \mathrm{Feas} = \mathbb{R}^{2} $ and
\[ f(x_1,x_2) = \begin{cases} 0 & \textrm{if $ x = \vec{0} $}, \\
x_1^2 + x_2^2/x_1 & \textrm{if $ x_1 > 0 $}, \\
\infty & \textrm{otherwise}.
\end{cases} \]
(For every $ r > 0 $, the function fails to be Lipschitz continuous on the relatively-open set $ \{ x \in \dom(f) \mid \| x \| < r \} $.)
\vspace{1mm}
The equivalent problem is hidden from view, the problem to which a standard subgradient method is applied and then translated as Algorithm A. The objective function for the equivalent problem has Lipschitz constant bounded above by $ 1/ \hat{r} $. Thus, a Lipschitz constant does in fact appear in (\ref{eqn.ad}), even though $ f $ need not be Lipschitz continuous. \vspace{1mm}
A tradeoff to gaining independence from Lipschitz continuity is that the error (\ref{eqn.ae}) is measured relatively rather than absolutely. Dependence on $ 1/ \epsilon^2 $ occurs both in (\ref{eqn.ad}) and in the \vspace{1mm} traditional literature. \vspace{1mm}
The criterion ``$ \alpha_{k+1} \geq 4/3 $'' is satisfied at most $ \log_{4/3}(1 + \bar{D}/ \hat{r}) $ times, leading to the logarithmic term in (\ref{eqn.ad}). The primary importance of using the criterion to create two cases is due to the accuracy needed in solving the equivalent problem being dependent not only on $ \epsilon $, but also on $ f^* $. Use of the criterion provides a means to compensate for not knowing $ f^* $. \vspace{1mm}
When $ f^* $ is known, a streamlined algorithm can be devised. This algorithm does not require $ \epsilon $ as input.
\vspace{1mm}
\noindent
\hrulefill
\noindent {\bf Algorithm B} \vspace{1mm}
\noindent
\begin{tabular}{lll}
(0) & Input: & $ f^* $, the optimal value, \\
& & $ \bar{e} \in \mathrm{int}( S \cap \mathrm{dom}(f)) $ satisfying $ A \bar{e} = b $, and \\
&& $ \hat{f} $, a scalar satisfying $ \hat{f} > f( \bar{e} ) $. \\
& Initialize: & $ x_0 := \bar{e} $, $ (x_0',t_0') := (\bar{e} , f( \bar{e} )) $ and $ \alpha_0 = \alpha(x_0,f^*) $ \\
(1) & Iterate: & Compute $ x_{k+1} := x_k + \smfrac{\alpha_k - 1 }{ \alpha_k \| P(g') \|^2} P(g') $, \, where $ g' \in G( x_k', t_k') $. \\
&& Let $ (x_{k+1}', t_{k+1}') := \pi' (x_{k+1}, f^*) $ and $ \alpha_{k+1} := \alpha(x_{k+1}, f^*) $. \vspace{1mm}
\end{tabular}
\noindent
\hrulefill
\vspace{2mm}
\begin{thm} \label{thm.ab}
For the feasible sequence $ \{ x_k' \} $ generated by Algorithm B, and for $ 0 < \epsilon < 1 $,
\begin{align}
& \ell \geq 4 \, \left( \frac{\bar{D}}{\hat{r}} \right)^2 \, \cdot \, \left( \, \frac{4}{ 3 } \left( \frac{1 - \epsilon }{ \epsilon } \right)^2 + 4 \left( \frac{1 - \epsilon}{\epsilon} \right) \right. \nonumber \\
& \qquad \qquad \qquad \qquad \qquad \qquad \left. + \log_2 \left( \frac{1 - \epsilon }{ \epsilon } \right) + \log_2 \left( \frac{\bar{D}}{\hat{r}} \right) \, + 1 \, \right) \nonumber \\ & \qquad \quad \Rightarrow \quad \min_{k \leq \ell } \, \frac{f(x_k') - f^*}{\hat{f} - f^*} \, \leq \, \epsilon \; , \label{eqn.af} \end{align}
\end{thm}
\vspace{2mm}
When $ S $ is polyhedral and $ f $ is piecewise linear, then in addition to the implication (\ref{eqn.af}), for some constants $ C_1 $ and $ C_2 $ there holds
\begin{equation} \label{eqn.ag}
\ell \geq C_1 \log(1/\epsilon ) + C_2 \quad \Rightarrow \quad \min_{k \leq \ell} \frac{f(x_k') - f^*}{\hat{f} - f^*} \leq \epsilon \; , \end{equation}
i.e., linear convergence. The constants, however, are highly dependent on $ \mathrm{Feas} $ and $ f $, making the linear convergence mostly a curiosity. \vspace{1mm}
In the following four sections, focus is exclusively on convex conic linear optimization problems. In the first of these sections (\S \ref{sect.b}), the equivalent problem is explained and the key theory is developed. Perhaps most surprising is that the transformation to an equivalent problem is simple and the theory is elementary. \vspace{1mm}
Representative algorithms for the conic setting are developed and analyzed in Sections~\ref{sect.d} and \ref{sect.e}, first in the ideal case where the optimal value is known, and then generally. (Groundwork for the development and analysis is laid in \S \ref{sect.c}) \vspace{1mm}
In Section~\ref{sect.f}, the general convex optimization problem (\ref{eqn.aa}) is recast into conic form, to which algorithms and analyses from earlier sections can be applied. The consequent complexity results are stated directly in terms of the original optimization problem (\ref{eqn.aa}), but the algorithms remain partially abstract in that some key computations are not expressed directly in terms of the original problem. \vspace{1mm}
Finally, in Section \ref{sect.g}, the remaining ties to the original optimization problem are established, allowing the algorithms applied to the conic recasting to instead be entirely expressed in terms of the original problem, resulting in Algorithms A and B above. The paper closes with the proofs of Theorems~\ref{thm.aa} and \ref{thm.ab}, and the proof of the claim regarding linear convergence (i.e., (\ref{eqn.ag})).
\vspace{1mm}
We do not discuss how to compute appropriate input $ \bar{e} $ when such a point exists but is unknown, because at present we do not see with high generality how to {\em cleanly} apply the framework to accomplish this, let alone see how to apply the framework to compute a ``good'' choice for $ \bar{e} $ from among the points in $ \int(S \cap \dom(f)) $. (Still, compared to the traditional literature where it is assumed $ \mathrm{Feas} \subset \int(\dom(f)) $ and (arbitrary) points can readily be orthogonally projected onto $ \mathrm{Feas} $, assuming $ \bar{e} $ is known and line-searches can be done seems to us considerably less restrictive.) \vspace{1mm}
The core of this paper, \S \S \ref{sect.b}--\ref{sect.f}, is virtually identical with our arXiv posting \cite{renegar2015framework}, a paper focused on convex conic linear optimization problems rather than general convex optimization problems, and having only at the end a brief discussion of the relevance to general convex optimization. (Motivated by reviewers' comments, we decided in revising the work for publication to give primary focus to general convex optimization, while still utilizing the conic setting as the natural venue for developing the equivalent problem to which traditional subgradient methods are applied.) Closely related to \cite{renegar2015framework} is recent work of Freund and Lu \cite{freund2015new}, who develop first-order methods for convex optimization problems in which the objective function has a known lower bound and is assumed to satisfy a Lipschitz condition. Their perspective provides an interesting juxtaposition to the conic-oriented presentation in \cite{renegar2015framework}. Also related is \cite{renegar2015accelerated}, in which accelerated methods for hyperbolic programming are presented.
\vspace{1mm}
We close the introduction by emphasizing the intent is to develop a framework, not to prescribe specific algorithms. Different subgradient methods applied in the framework yield different algorithms. Changing how a general optimization problem is recast into conic form also results in a different algorithm.
\vspace{2mm}
\section{{\bf Key Theory}} \label{sect.b}
The key theory is elementary and yet has been overlooked in the literature, a blind spot. \vspace{1mm}
Continue to let $ {\mathcal E} $ denote a finite-dimensional Euclidean space. \vspace{1mm}
Let $ {\mathcal K} \subset {\mathcal E} $ be a proper, closed, convex cone with nonempty interior. \vspace{1mm}
Fix a vector $ e \in \int( {\mathcal K} ) $. We refer to $ e $ as the ``distinguished direction.'' For each $ x \in {\mathcal E} $, let
\[ \lambda_{\min}(x) := \inf \{ \lambda \mid x - \lambda \, e \notin {\mathcal K} \} \; , \]
that is, the scalar $ \lambda $ for which $ x - \lambda e $ lies in the boundary of $ {\mathcal K} $. (Existence and uniqueness of $ \lambda_{\min}(x) $ follows from $ e \in \int({\mathcal K} ) \neq {\mathcal E} $ and the assumption that $ {\mathcal K} $ is a closed, convex cone.) \vspace{1mm}
If, for example, $ {\mathcal E} = \mathbb{S}^{n} $ ($ n \times n $ symmetric matrices), $ {\mathcal K} = \Symp $ (cone of positive semidefinite matrices), and $ e = I $ (the identity), then $ \lambda_{\min}(X) $ is the minimum eigenvalue of $ X $. \vspace{1mm}
On the other hand, if $ {\mathcal K} = \mathbb{R}^n_{{\scriptscriptstyle +}} $ (non-negative orthant) and $ e $ is a vector with all positive coordinates, then $ \lambda_{\min}(x) = \min_j x_j/e_j $ for $ x \in \mathbb{R}^n $. Clearly, the value of $ \lambda_{\min}(x) $ depends on the distinguished direction $ e $ (a fact the reader should keep in mind since the notation does not reflect the dependence). \vspace{1mm}
Obviously, $ {\mathcal K} = \{ x \mid \lambda_{\min}(x) \geq 0 \} $ and $ \int({\mathcal K} ) = \{ x \mid \lambda_{\min}(x) > 0 \} $. Also,
\begin{equation} \label{eqn.ba}
\lambda_{\min}(sx + te) = s \, \lambda_{\min}(x) + t \quad \textrm{for all $ x \in {\mathcal E}$ and scalars $ s \geq 0 $, $ t $} \; .
\end{equation}
Let
\[ \bar{{\mathcal B}} := \{ v \in {\mathcal E} \mid e + v \in {\mathcal K} \textrm{ and } e - v \in {\mathcal K} \} \; , \]
a closed, centrally-symmetric, convex set with nonempty interior.
Define the gauge (\cite[\S 15]{rockafellar1970convex}) on $ {\mathcal E} $ by
\[ \| u \|_{\infty} := \inf \{ t \geq 0 \mid u = tv \textrm{ for some $ v \in \bar{{\mathcal B}} $} \} \; , \]
which is easily shown to be a norm if and only if $ {\mathcal K} $ is a pointed cone.
Let $ \bar{B}_{\infty}(x,r) $ denote the closed ``ball'' centered at $ x $ and of radius $ r $. Clearly, $ \bar{B}_{\infty}(0,1) = \bar{{\mathcal B}} $, and $ \bar{B}_{\infty}(e,1) $ is the largest subset of $ {\mathcal K} $ that has symmetry point $ e $, i.e., for each $ v $, either both points $ e + v $ and $ e - v$ are in the set, or neither point is in the set.
\vspace{1mm}
\begin{prop} \label{prop.ba}
The function $ x \mapsto \lambda_{\min}(x) $ is concave and Lipschitz continuous:
\[ | \lambda_{\min}(x) - \lambda_{\min}(y) | \leq \| x - y \|_{\infty} \quad \textrm{for all $ x,y \in {\mathcal E} $} \; . \]
\end{prop}
\noindent {\bf Proof:} Concavity follows easily from the convexity of $ {\mathcal K} $, so we focus on establishing Lipschitz continuity. \vspace{1mm}
Let $ x, y \in {\mathcal E} $. According to (\ref{eqn.ba}), the difference $ \lambda_{\min}(x + te) - \lambda_{\min}(y+te) $ is independent of $ t $, and of course so is the quantity $ \| (x + te) - (y + te) \|_{\infty} \; . $ Consequently, in proving the Lipschitz continuity, we may assume $ x $ lies in the boundary of $ {\mathcal K} $, that is, we may assume $ \lambda_{\min}(x) = 0 $. The goal, then, is to prove
\begin{equation} \label{eqn.bb}
| \lambda_{\min}(x + v) | \leq \| v \|_{ \infty} \quad \textrm{for all $ v \in {\mathcal E} $} \; . \end{equation}
We consider two cases. First assume $ x + v $ does not lie in the interior of $ {\mathcal K} $, that is, assume $ \lambda_{\min}(x+v) \leq 0 $. Then, to establish (\ref{eqn.bb}), it suffices to show $ \lambda_{\min}(x + v) \geq - \| v \|_{\infty} \; , $ that is, to show
\begin{equation} \label{eqn.bc}
x + v + \| v \|_{\infty} \, e \in {\mathcal K} \; .
\end{equation}
However,
\begin{equation} \label{eqn.bd}
v + \| v \|_{\infty} \, e \in \bar{B}_{\infty}(\| v \|_{\infty} \, e, \| v \|_{\infty}) \subseteq {\mathcal K} \; ,
\end{equation}
the set containment due to $ {\mathcal K} $ being a cone and, by construction, $ \bar{B}_{\infty}(e,1) \subseteq {\mathcal K} $. Since $ x \in {\mathcal K} $ (indeed, $ x $ is in the boundary of $ {\mathcal K} $), (\ref{eqn.bc}) follows. \vspace{1mm}
Now consider the case $ x + v \in {\mathcal K} $, i.e., $ \lambda_{\min}(x+v) \geq 0 $. To establish (\ref{eqn.bb}), it suffices to show $ \lambda_{\min}(x + v) \leq \|v\|_{\infty} \; , $ that is, to show
\[ x + v - \|v\|_{\infty} \, e \notin \int({\mathcal K} ) \; . \]
Assume otherwise, that is, assume
\[ x = w + \| v \|_{\infty} \, e - v \quad \textrm{for some $ w \in \int({\mathcal K} ) $} \; . \]
Since $ \| v \|_{\infty} \, e - v \in {\mathcal K} $ (by the set containment on the right of (\ref{eqn.bd})), it then follows that $ x \in \int({\mathcal K} ) $, a contradiction to $ x $ lying in the boundary of $ {\mathcal K} $. \hfill $ \Box $
\vspace{3mm}
In this section, the inner product is used primarily for expressing a conic optimization problem. To allow a distinction between inner products, we denote evaluation for the ``modeling inner product'' on a pair $ u,v \in {\mathcal E} $ by $ u \cdot v $, whereas in later sections we denote -- as was done implicitly in the introduction -- evaluation for the ``computational inner product'' by $ \langle u, v \rangle $. \vspace{1mm}
Let $ \mathrm{Affine} \subseteq {\mathcal E} $ be an affine space, i.e., the translate of a subspace. For fixed $ c \in {\mathcal E} $, consider the conic program
\[
\left. \begin{array}{rl}
\inf & c \cdot x \\
\textrm{s.t.} & x \in \mathrm{Affine} \\
& x \in {\mathcal K} \; . \end{array} \right\} \cp \]
Assume $ \mathrm{Affine} \cap \int({\mathcal K} ) $ -- the set of strictly feasible points -- is nonempty. Let $ z^* $ denote the optimal value. \vspace{1mm}
Assume $ c $ is not orthogonal to the subspace of which $ \mathrm{Affine} $ is a translate, since otherwise all feasible points are optimal. This assumption implies that all optimal solutions for CP lie in the boundary of $ {\mathcal K} $. \vspace{1mm}
Fix a strictly feasible point, $ e $. The point $ e $ serves as the distinguished direction. \vspace{1mm}
For scalars $ z \in \mathbb{R} $, let
\begin{gather*} \mathrm{Affine}_{z} := \{ x \in \mathrm{Affine} \mid c \cdot x = z \}, \textrm{ and} \\
\textrm{let $ {\mathcal L} $ denote the subspace of which these affine spaces are translates.}
\end{gather*}
Presently we show that for any choice of $ z $ satisfying $ z < c \cdot e \; , $ CP can be easily transformed into an equivalent optimization problem in which the only constraint is $ x \in \mathrm{Affine}_{z} \; . $ We make a simple observation. \vspace{1mm}
\begin{lemma} \label{lem.bb}
Assume $ \mathrm{CP} $ has bounded optimal value.
$ \textrm{~} $ \qquad \qquad \qquad If $ x \in \mathrm{Affine} $ satisfies $ c \cdot x < c \cdot e $, then $ \lambdamin(x) < 1 \; . $
\end{lemma}
\noindent {\bf Proof:} It follows from (\ref{eqn.ba}) that if $ \lambda_{\min}(x) \geq 1 $, then $ e + t( x - e) $ is feasible for all $ t \geq 0 $. As the function $ t \mapsto c \cdot \big( e + t(x- e) \big) $ is strictly decreasing (because $ c \cdot x < c \cdot e $), this implies CP has unbounded optimal value, contrary to assumption.~$ \Box $
\vspace{3mm}
For $ x \in {\mathcal E} $ satisfying $ \lambda_{\min}(x) < 1 $, let $ \pi(x) $ denote the point where the half-line beginning at $ e $ in direction $ x - e $ intersects the boundary of $ {\mathcal K} $:
\begin{equation} \label{eqn.bda}
\pi(x) := e + \smfrac{1}{1 - \lambda_{\min}(x)} (x - e)
\end{equation}
(to verify correctness of the expression, observe (\ref{eqn.ba}) implies $ \lambda_{\min}(\pi(x)) = 0 $).
We refer to $ \pi(x) $ as the ``radial projection'' of $ x $. \vspace{1mm}
The proof of the following result is straightforward, but because it is central to our development, we label the result as a theorem. \vspace{1mm}
\begin{thm} \label{thm.bc}
Let $ z $ be any value satisfying \, $ z < c \cdot e \; . $ If $ x^* $ solves
\begin{equation} \label{eqn.be}
\begin{array}{rl}
\sup & \lambdamin(x) \\
\mathrm{s.t.} & x \in \mathrm{Affine}_{z} \; , \end{array}
\end{equation}
then $ \pi( x^* ) $ is optimal for $ \mathrm{CP} $. Conversely, if $ \pi^* $ is optimal for $ \mathrm{CP} $, then $ x^* := e + \frac{c \cdot e - z }{c \cdot e - z^* } ( \pi^* - e) $ is optimal for (\ref{eqn.be}), and $ \pi^* = \pi( x^* ) $.
\end{thm}
\noindent {\bf Proof:} Fix a value $ z $ satisfying $ z < c \cdot e $. It is easily proven from the convexity of $ {\mathcal K} $ that $ x \mapsto \pi(x) $ gives a one-to-one map from $ \mathrm{Affine}_{z} $ onto
\begin{equation} \label{eqn.bf}
\{ \pi \in \mathrm{Affine} \cap \mathrm{bdy}({\mathcal K} ) \mid c \cdot \pi < c \cdot e \} \; ,
\end{equation}
where $ \mathrm{bdy}({\mathcal K} ) $ denotes the boundary of $ {\mathcal K} $. \vspace{1mm}
For $ x \in \mathrm{Affine}_{z} \; , $ the CP objective value of $ \pi(x) $ is
\begin{align}
c \cdot \pi(x) & = c \cdot \big( e+ \smfrac{1}{1 - \lambda_{ \min}(x)} (x - e) \big) \nonumber \\
& = c \cdot e + \smfrac{1}{1 - \lambda_{ \min }(x)} ( z - c \cdot e ) \; , \label{eqn.bg}
\end{align}
a strictly-decreasing function of $ \lambda_{\min}(x) $.
Since the map $ x \mapsto \pi(x) $ is a bijection between $ \mathrm{Affine}_{z} $ and the set (\ref{eqn.bf}), the theorem readily follows. \hfill $ \Box $
\vspace{3mm}
CP has been transformed into an equivalent linearly-constrained maximization problem with concave, Lipschitz-continuous objective function. Virtually any subgradient method -- rather, {\em supgradient} method -- can be applied to this problem, the main cost per iteration being in computing a supgradient and projecting it onto the subspace $ {\mathcal L} $. \vspace{1mm}
For illustration, we digress to interpret the implications of the development thus far for the linear program
\[
\left. \begin{array}{rl}
\min_{x \in \mathbb{R}^n} & c^T x \\
\textrm{s.t.} & Ax = b \\
& x \geq 0 \; , \end{array} \right\} \, \mathrm{LP}
\]
assuming $ e = {\bf 1} $ (the vector of all ones), in which case $ \lambda_{\min}(x) = \min_j x_j \; , $ and $ \| \, \, \|_{\infty} $ is the $ \ell_{\infty} $ norm, i.e., $ \| v \|_{\infty} = \max_j |v_j| $. Let the number of rows of $ A $ be $ m \geq 1 $, assume the rows are linearly independent, and assume $ c $ is not a linear combination of the rows (otherwise all feasible points are optimal). \vspace{1mm}
For any scalar $ z < c^T {\bf 1} $, Theorem~\ref{thm.bc} asserts that LP is equivalent to
\begin{equation} \label{eqn.bh}
\begin{array}{rl}
\max_x & \min_j x_j \\
\textrm{s.t.} & Ax = b \\
& c^T x = z \; , \end{array} \end{equation}
in that when $ x $ is feasible for (\ref{eqn.bh}), $ x $ is optimal if and only if the projection \\ $ \pi(x) = \mathbf{1} + \smfrac{1}{1 - \min_j x_j} (x - \mathbf{1} ) $ is optimal for LP.
The setup is shown schematically in the following figure:
$ \textrm{~} $ \quad \qquad \quad \qquad \qquad \includegraphics[scale=.26]{figure2015_05.png}
\vspace{4mm}
Proposition~\ref{prop.ba} asserts that, as is obviously true, $ x \mapsto \min_j x_j $ is $ \ell_{\infty} $-Lipschitz continuous with constant 1. Consequently, the function also is $ \ell_2$-Lipschitz continuous with constant 1, as is relevant if supgradient methods rely on the standard inner product in computing supgradients. \vspace{1mm}
With respect to the standard inner product, the supgradients of $ x \mapsto \min_j x_j $ at $ x $ are the convex combinations of the standard basis vectors $ e(k) $ for which $ x_k = \min_j x_j $. Consequently, the projected supgradients at $ x $ are the convex combinations of the vectors $ \bar{P}_k $ for which $ x_k = \min_j x_j $, where $ \bar{P}_k $ is the $ k^{th} $ column of the matrix projecting $ \mathbb{R}^n $ onto the nullspace of $ \bar{A} = \left[ \begin{smallmatrix} A \\ c^T \end{smallmatrix} \right] $, that is
\[
\bar{P} := I- \bar{A}^T (\bar{A} \, \bar{A}^T)^{-1} \bar{A} \; . \]
If $ m \ll n $, then $ \bar{P} $ is not computed in its entirety, but instead the matrix $ \bar{M} = ( \bar{A} \bar{A}^T)^{-1} $ is formed as a preprocessing step, at cost $ O(m^2 n) $ (the inverse exists because the set consisting of $ c $ and the rows of $ A $ has been assumed linearly independent). Then, for any iterate $ x $ and an index $ k $ satisfying $ x_k = \min_j x_j $, the projected supgradient $ \bar{P}_k $ is computed according to
\[ u = \bar{M} \, \bar{A}_k \quad \rightarrow \quad v = \bar{A}^T u \quad \rightarrow \quad \bar{P}_k = e(k) - v \; , \]
for a cost of $ \, O(m^2 \, + \, \# \mathrm{non\_zero\_entries\_in\_} A ) $ per iteration. \vspace{1mm}
Before returning to the general theory, we note that if the choices are $ {\mathcal E} = \mathbb{S}^{n} $, $ {\mathcal K} = \Symp $ and $ e = I $ (and thus $ \lambda_{\min}(X) $ is the minimum eigenvalue of $ X $), then with respect to the trace inner product, the supgradients at $ X $ for the function $ X \mapsto \lambda_{\min}(X) $ are the convex combinations of the matrices $ v v^T $, where $ Xv = \lambda_{\min}(X) v $ and $ \| v \|_2 = 1 $. \vspace{1mm}
{\em Assume, henceforth, that CP has at least one optimal solution, and that $ z $ is a fixed scalar satisfying \, $ z < c \cdot e $.} Then the equivalent problem (\ref{eqn.be}) has at least one optimal solution. Let $ x^*_{z} $ denote any of the optimal solutions for the equivalent problem, and recall
$ z^* $ denotes the optimal value of CP. A useful characterization of the optimal value for the equivalent problem is easily provided. \vspace{1mm}
\begin{lemma} \label{lem.bd}
\[ \lambda_{\min}(x^*_{z }) = \frac{z - z^* }{c \cdot e -
z^* } \]
\end{lemma}
\noindent {\bf Proof:}
By Theorem~\ref{thm.bc}, $ \pi(x^*_{z}) $ is optimal for CP -- in particular, $ c \cdot \pi(x^*_{z}) = z^* $. Thus, according to (\ref{eqn.bg}),
\[ z^* = c \cdot e + \smfrac{1}{1 - \lambda_{\min}(x^*_{z })} \, ( z - c \cdot e) \; . \]
Rearrangement completes the proof. \hfill $ \Box $
\vspace{3mm}
We focus on the goal of computing a point $ \pi $ which is feasible for CP and has better objective value than $ e$ in that
\begin{equation} \label{eqn.bi}
\frac{c \cdot \pi - z^* }{c \cdot e - z^* } \leq \epsilon \; \; ,
\end{equation}
where $ 0 < \epsilon < 1 $. Thus, for the problem of primary interest, CP, the focus is on relative improvement in the objective value.
\vspace{1mm}
The following proposition provides a useful characterization of the accuracy needed in approximately solving the CP-equivalent problem (\ref{eqn.be}) so as to ensure that for the computed point $ x $, the projection $ \pi = \pi(x) $ satisfies (\ref{eqn.bi}). \vspace{1mm}
\begin{prop} \label{prop.be}
If $ x \in \mathrm{Affine}_z $ and $ 0 < \epsilon < 1 $, then
\begin{align*}
& \frac{ c \cdot \pi(x) - z^* }{c \cdot e - z^* } \, \leq \, \epsilon
\\ & \qquad \qquad \qquad \textrm{if and only if} \\ & \qquad \qquad \qquad \qquad
\lambdamin( x^*_{ z} ) - \lambdamin( x) \, \leq \, \frac{\epsilon }{1 - \epsilon } \, \, \frac{c \cdot e - z }{\, \, \, c \cdot e - z^* } \; .
\end{align*}
\end{prop}
\noindent {\bf Proof:} Assume $ x \in \mathrm{Affine}_{z} $. For $ y = x $ and $ y = x^*_{z} \; , $ we have the equality (\ref{eqn.bg}), that is,
\[ c \cdot \pi(y) = c \cdot e + \smfrac{1}{1 - \lambda_{\min}(y)} ( z - c \cdot e ) \; .
\]
Thus,
\begin{align*}
\frac{ c \cdot \pi(x) - z^* }{c \cdot e - z^* } & = \frac{ c \cdot \pi(x) - c \cdot \pi(x^*_{z} ) }{c \cdot e - c \cdot \pi(x^*_{z} ) } \\
& = \frac{ \smfrac{1}{1 - \lambdamin(x)} - \smfrac{1}{1 - \lambdamin(x^*_{z} )} }{ - \smfrac{1}{1 - \lambdamin(x^*_{z} )}} \\
& = \frac{ \lambdamin(x^*_{z} ) - \lambdamin(x) }{ 1 - \lambdamin(x)} \; .
\end{align*}
Hence,
\begin{gather*}
\frac{c \cdot \pi(x) - z^* }{c \cdot e - z^* } \leq \epsilon \\
\Leftrightarrow \\
\lambdamin(x^*_{z} ) - \lambdamin(x) \leq \epsilon \, ( 1 - \lambdamin(x)) \\
\Leftrightarrow \\
(1 - \epsilon ) ( \lambdamin(x^*_{z} ) - \lambdamin(x) ) \leq \epsilon ( 1 - \lambdamin(x^*_{z} )) \\
\Leftrightarrow \\ \lambdamin(x^*_{z} ) - \lambdamin(x) \leq \smfrac{\epsilon }{1 - \epsilon } ( 1 - \lambdamin(x^*_{z} )) \; .
\end{gather*}
Using Lemma~\ref{lem.bd} to substitute for the rightmost occurrence of $ \lambda_{\min}(x^*_{z} ) $ completes the proof. \hfill $ \Box $
\vspace{3mm}
\section{ {\bf Groundwork for Algorithm Design and Analysis}} \label{sect.c}
In Sections \ref{sect.d} and \ref{sect.e}, we show how the key theory leads to algorithms and complexity results regarding the solution of the conic program CP. In this section, groundwork is laid for the development in those and subsequent sections.
\vspace{1mm}
Continue to assume CP has an optimal solution, denote the optimal value by $ z^* $, and let $ e$ be a strictly feasible point, the distinguished direction.
Given $ \epsilon > 0 $ and a value $ z $ satisfying $ z < c \cdot e $, the approach is to apply supgradient methods to approximately solve
\begin{equation} \label{eqn.ca}
\begin{array}{rl}
\max & \lambda_{ \min}(x) \\
\textrm{s.t.} & x \in \mathrm{Affine}_{z} \; , \end{array}
\end{equation}
where by ``approximately solve'' we mean that $ x \in \mathrm{Affine}_{z} $ is computed for which
\[
\lambda_{\min}(x^*_{z}) - \lambda_{ \min}(x) \, \leq \, \frac{\epsilon}{1 - \epsilon } \, \, \frac{c \cdot e - z }{\, \, \, c \cdot e - z^* } \; .
\]
Indeed, according to Proposition~\ref{prop.be}, the projection $ \pi = \pi(x) $ will then satisfy
\[
\frac{c \cdot \pi - z^* }{c \cdot e - z^* } \leq \epsilon \; .
\]
\vspace{1mm}
Not until results regarding CP are applied to general convex optimization -- Sections \ref{sect.f} and \ref{sect.g} -- do we require a characterization of supgradients of the function $ x \mapsto \lambdamin(x) $, but as the characterization is interesting in itself, we present it as the first piece of groundwork laid in this section. \vspace{1mm}
Let $ P_{{\mathcal L}} $ denote orthogonal projection onto $ {\mathcal L} $ with respect to $ \langle \; , \; \rangle $ (the computational inner product), and let $ \hat{\partial} \lambda_{\min}(x) $ denote the supdifferential at $ x $. The following proposition relates supdifferentials to normal cones (normal with respect to $ \langle \; , \; \rangle $). \vspace{1mm}
\begin{prop} \label{prop.ca}
For all $ x \in {\mathcal E} $,
\[ \hat{\partial}\lambda_{\min}(x) = \{ v \mid -v \in N_{{\mathcal K}}( x - \lambda_{\min}(x) e) \textrm{ and } \langle e, v \rangle = 1 \} \; . \]
\end{prop}
\noindent {\bf Proof:} We know
\begin{equation} \label{eqn.cb}
\lambda_{\min}(y + re) = \lambda_{\min}(y) + t \quad \textrm{for all $ y \in {\mathcal E} $, $ t \in \mathbb{R} $} \; ,
\end{equation}
from which follows for every $ x $ and $ t $,
\begin{equation} \label{eqn.cc}
\hat{\partial}\lambda_{\min}(x + te) = \hat{\partial} \lambda_{\min}(x)
\end{equation}
and
\begin{equation} \label{eqn.cd}
v \in \hat{\partial} \lambda_{\min}(x) \quad \Rightarrow \quad \langle e, v \rangle = 1 \; .
\end{equation}
Due to (\ref{eqn.cc}), in proving the lemma we may assume $ \lambdamin(x) = 0 $, i.e., $ x \in \mathrm{bdy}({\mathcal K}) $.
\vspace{1mm}
Since $ \lambdamin(x) = 0 $, and the value of $ \lambdamin $ is non-negative on $ {\mathcal K} $,
\[ g \in \hat{\partial} \lambdamin(x) \quad \Rightarrow \quad \forall y \in {\mathcal K}, \, \langle g, y - x \rangle \geq 0 \; . \]
Thus, with (\ref{eqn.cd}), we have
\begin{equation} \label{eqn.ce}
\hat{\partial}\lambdamin(x) \subseteq \{ v \mid \langle e, v \rangle = 1 \textrm{ and } - v \in N_{{\mathcal K}}(x) \} \; .
\end{equation}
On the other hand, if $ -v \in N_{{\mathcal K}}(x) $, then
\[ y \in \mathrm{bdy}({\mathcal K}) \quad \Rightarrow \quad 0 = \lambdamin(y) \leq \lambdamin(x) + \langle v, y - x \rangle \]
(using $ \lambdamin(x) = 0 $). Thus, if in addition, $ \langle e, v \rangle = 1 $,
\[ t \in \mathbb{R} \textrm{ and } y \in \mathrm{bdy}({\mathcal K}) \quad \Rightarrow \quad \lambdamin(y + te) \leq \lambdamin(x) + \langle v, (y + te) - x \rangle \]
(using \ref{eqn.cb}). Since $ {\mathcal E} = \{ y + te \mid y \in \mathrm{bdy}( {\mathcal K}) \textrm{ and } t \in \mathbb{R} \} $, the reverse inclusion to (\ref{eqn.ce}) thus holds, establishing the proposition. \hfill $ \Box $
\vspace{3mm}
Let $ \| \, \, \| $ be the norm associated with $ \langle \; , \; \rangle $, and for $ z \in \mathbb{R} $, let
\[ M_z := \sup \left\{ \smfrac{| \lambda_{\min}(x) - \lambda_{\min}(y)|}{ \| x - y \| } \mid x,y \in \mathrm{Affine}_z \textrm{ and } x \neq y \right\} \; , \]
the Lipschitz constant for the map $ x \mapsto \lambda_{\min}(x) $ restricted to $ \mathrm{Affine}_z $. Proposition~\ref{prop.ba} implies $ M_z $ is well-defined (finite), although unlike the Lipschitz constant for the norm appearing there (i.e., $ \| \, \, \|_{\infty} $), $ M_z $ might exceed 1, depending on $ \| \, \, \| $. The next piece of groundwork to be laid is a geometric characterization of an upper bound on $ M_z $. \vspace{1mm}
We claim the values $ M_z $ are identical for all $ z $. To see why, consider that for $ z_1, z_2 < c \cdot e \; , $ a bijection from $ \mathrm{Affine}_{z_1} $ onto $ \mathrm{Affine}_{z_2} $ is provided by the map
\[ x \mapsto y(x) := \smfrac{z_2 - z_1}{c \cdot e - z_1} e + \smfrac{c \cdot e - z_2}{c \cdot e - z_1} x \; . \]
Observe, using (\ref{eqn.ba}),
\[ \lambda_{\min}(y(x)) = \smfrac{c \cdot e - z_2}{c \cdot e - z_1} \lambda_{\min}(x) + \smfrac{z_2 - z_1}{c \cdot e - z_1} \; , \]
and thus
\[ \lambda_{\min}(y(x)) - \lambda_{\min}(y( \bar{x})) = \smfrac{c \cdot e - z_2}{c \cdot e - z_1} \left( \lambda_{\min}(x) - \lambda_{\min}( \bar{x}) \right) \quad \textrm{for $ x, \bar{x} \in \mathrm{Affine}_{z_1} $} \; . \]
Since, additionally, $ \| y(x) - y( \bar{x}) \| = \smfrac{c \cdot e - z_2}{c \cdot e - z_1} \| x - \bar{x} \| $, it is immediate that the values $ M_z $ are identical for all $ z < c \cdot e $. A simple continuity argument then implies this value is equal to $ M_{c \cdot e} $. Analogous reasoning shows $ M_z = M_{ c \cdot e} $ for all $ z > c \cdot e $. In all, $ M_z $ is independent of $ z $, as claimed. \vspace{1mm}
Let $ M $ denote the common value, i.e., $ M = M_z $ for all $ z $. \vspace{1mm}
The following proposition can be useful in modeling and in choosing the computational inner product. This result plays a central role in our complexity bounds for general convex optimization. \vspace{1mm}
Let $ \bar{B}(e,r) := \{ x \mid \| x - e \| \leq r \} $.
\begin{prop} \label{prop.cb} $ M \leq 1/ r_e \; , $ where $ r_e := \max \{ r \mid \bar{B}(e,r) \cap \mathrm{Affine}_{c \cdot e} \subseteq {\mathcal K} \} $
\end{prop}
\noindent {\bf Proof:} According to Proposition~\ref{prop.ba},
\[ | \lambda_{\min}(x) - \lambda_{\min}(y)| \leq \| x - y \|_{\infty} \quad \textrm{for all $ x, y $} \; . \]
Consequently, it suffices to show $ \| x - y \| \geq r_e \| x - y \|_{\infty} $ for all $ x,y \in \mathrm{Affine}_{c \cdot e} \; , $ i.e., it suffices to show for all $ v \in {\mathcal L} $ that $ \| v \| \geq r_e \| v \|_{\infty} $. \vspace{1mm}
However, according to the discussion just prior to Proposition~\ref{prop.ba}, $ \bar{B}_{\infty}(e,1) $ is the largest set which both is contained in $ {\mathcal K} $ and has symmetry point $ e $, from which follows that $ \bar{B}_{\infty}(e,1) \cap \mathrm{Affine}_{c \cdot e} $ is the largest set which is both contained in $ {\mathcal K} \cap \mathrm{Affine}_{c \cdot e} $ and has symmetry point $ e $. Hence
\[ \bar{B}(e, r_e) \cap \mathrm{Affine}_{ c \cdot e} \subseteq \bar{B}_{\infty}(e,1) \cap \mathrm{Affine}_{c \cdot e} \; , \]
implying $ \| v \| \geq r_e \| v \|_{\infty} $ for all $ v \in {\mathcal L} $. \hfill $ \Box $ \vspace{3mm}
Towards considering specific supgradient methods, we recall the following standard and elementary result, rephrased for our setting:
\begin{lemma} \label{lem.cc}
Assume $ z \in \mathbb{R} $, $ x, y \in \mathrm{Affine}_z $ and $ g \in \hat{\partial}\lambdamin(x) $.
For all scalars $ \alpha $,
\[
\| ( x + \alpha P_{{\mathcal L}} (g)) - y \|^2 \leq \| x - y \|^2 - 2 \alpha \left( \lambda_{\min}(y) - \lambda_{\min}(x) \right) + \alpha^2 \| P_{{\mathcal L}} (g) \|^2 \; . \]
\end{lemma}
\noindent {\bf Proof:} Letting $ \tilde{g} := P_{{\mathcal L}} (g )$, simply observe
\begin{align*}
\| ( x + \alpha \tilde{g} ) - y \|^2 & = \| x - y \|^2 + 2 \alpha \langle \tilde{g} , x - y \rangle + \alpha^2 \| \tilde{g} \|^2 \\
& = \| x - y \|^2 - 2 \alpha \langle g, y - x \rangle + \alpha^2 \| \tilde{g} \|^2 \quad \textrm{(by $ x - y \in {\mathcal L} $)} \\
& \leq \| x - y \|^2 - 2 \alpha \left( \lambda_{\min}(y) - \lambda_{\min}(x) \right) + \alpha^2 \| \tilde{g} \|^2 \; ,
\end{align*}
the inequality due to concavity of the map $ x \mapsto \lambda_{\min}(x) \; . $ \hfill $ \Box $ \vspace{2mm}
\section{{\bf Algorithm 1: When the Optimal Value is Known}} \label{sect.d}
Knowing $ z^* $ is not an entirely implausible situation. For example, if strict feasibility holds for a primal conic program and for its dual
\[ \begin{array}{rl}
\min & \bar{c}^T x \\
\textrm{s.t.} & \bar{A}x = \bar{b} \\
& x \in \bar{{\mathcal K} } \end{array} \qquad \begin{array}{rl}
\max & \bar{b}^T y \\
\textrm{s.t.} & \bar{A}^T y + s = \bar{c} \\
& s \in \bar{{\mathcal K} }^* \; , \end{array} \]
then the combined primal-dual conic program is known to have optimal value equal to zero:
\[ \begin{array}{rl}
\min & \bar{c}^T x - \bar{b}^T y \\
\textrm{s.t.} & \bar{A} x = \bar{b} \\
& \bar{A}^T y + s = \bar{c} \\
& (x,s) \in \bar{{\mathcal K} } \times \bar{{\mathcal K} }^* \; . \end{array} \]
\noindent
\hrulefill
\noindent {\bf Algorithm 1} \vspace{1mm}
\noindent
\noindent
\begin{tabular}{lll}
(0) &Input: & $ z^* $, the optimal value of CP, \\
&& $ e $, a strictly feasible point for CP, and \\
&& $ \bar{x} \in \mathrm{Affine} $ satisfying $ c \cdot \bar{x} < c \cdot e \; . $ \\
& Initialize: & Let $ x_0 = e + \frac{c \cdot e - z^*}{c \cdot e - c \cdot \bar{x} }(\bar{x} -e) $ \, \, (thus, $ c \cdot x_0 = z^* $), \\
&& and let $ \pi_0 = \pi(x_0) $ \, \, ($ = \pi(\bar{x} ) $). \\
(1) & Iterate: & Compute $ x_{k+1} = x_k - \smfrac{\lambda_{\min}(x_k)}{\| P_{{\mathcal L}} ( g_k ) \|^2} P_{{\mathcal L}} ( g_k ) $, \, where $ g_k \in \hat{\partial}\lambdamin(x_k) $. \\
&& Let $ \pi_{k+1} = \pi(x_{k+1}) $. \vspace{1mm}
\end{tabular}
\noindent
\hrulefill
\vspace{2mm}
All of the iterates $ x_k $ lie in $ \mathrm{Affine}_{z^*} $, and hence, $ \lambda_{\min}(x_k) \leq 0 $, with equality if and only if $ x_k $ is feasible (and optimal) for CP. \vspace{1mm}
For all scalars $ z < c \cdot e $ and for $ x \in \mathrm{Affine}_z \; , $ define
\[
\mathrm{dist}_z(x) := \min \{ \| x - x_z^* \| \mid x_z^* \textrm{ is optimal for (\ref{eqn.ca})} \} \; . \]
\begin{prop} \label{prop.da}
The iterates for Algorithm 1 satisfy
\[
\max \{ \lambda_{\min}(x_k) \mid k = \ell, \ldots, \ell + m \} \geq -M \, \mathrm{dist}_{z^*}(x_{\ell})/ \sqrt{m+1} \; .
\]
\end{prop}
\noindent {\bf Proof:} Letting $ \tilde{g}_k := P_{{\mathcal L}} (g_k) $, Lemma~\ref{lem.cc} implies
\begin{align*}
& \mathrm{dist}_{z^*}(x_{k+1})^2 \\ & \leq \mathrm{dist}_{z^*}(x_k)^2 - 2 (-\lambda_{\min}(x_k)/\| \tilde{g}_k \|^2) \, ( 0 - \lambda_{\min}(x_k)) \, + \, (\lambda_{\min}(x_k)/\| \tilde{g}_k \|)^2 \\
& = \mathrm{dist}_{z^*}(x_k)^2 - (\lambda_{\min}(x_k)/\| \tilde{g}_k \|)^2 \; ,
\end{align*}
and thus by induction (and using $ \| \tilde{g}_k \| \leq M $),
\begin{align*}
\mathrm{dist}_{z^*}(x_{\ell + m + 1 })^2 & \leq \mathrm{dist}_{z^*}(x_{\ell})^2 - \sum_{k= \ell }^{\ell + m} ( \lambda_{\min}(x_k)/M)^2 \\
& \leq \mathrm{dist}_{z^*}(x_{\ell})^2 - \smfrac{m + 1}{M^2} \min \{ \lambda_{\min}(x_k)^2 \mid k = \ell, \ldots, \ell + m \} \; ,
\end{align*}
implying the proposition (keeping in mind $ \lambda_{\min}(x_k) \leq 0 $). \hfill $ \Box $
\vspace{3mm}
We briefly digress to consider the case of $ {\mathcal K} $ being polyhedral, where already an interesting result is easily proven. The following corollary is offered only as a curiosity, as the constants typically are so large as to render the bound on $ \ell$ meaningless except for minuscule $ \epsilon$.
\vspace{1mm}
\begin{cor} \label{cor.db}
Assume $ {\mathcal K} $ is polyhedral. There exist constants $ C_1 $ and $ C_2 $ (dependent on CP, $ e $, $ \bar{x} $ and the computational inner product), such that for all $ 0 < \epsilon < 1 $,
\[ \ell \geq C_1 + C_2 \log(1/ \epsilon ) \quad \Rightarrow \quad \min_{k \leq \ell} \frac{c \cdot \pi_k - z^*}{c \cdot e - z^*} \, \leq \, \epsilon \; . \]
\end{cor}
\vspace{2mm}
For first-order methods, such a logarithmic bound in $ \epsilon $ was initially established by Gilpin, Pe\~{n}a and Sandholm \cite{gilpin2012first}. They did not assume an initial feasible point $ e $ was known, but neither did they require the computed approximate solution to be feasible (instead, constraint residuals were required to be small). They relied on an accelerated gradient method, along with the smoothing technique of Nesterov \cite{nesterov2005smooth}. As is the case for the above result, they assumed the optimal value of CP to be known apriori, and they restricted $ {\mathcal K} $ to be polyhedral. \vspace{2mm}
The proof of the corollary depends on the following simple lemma. \vspace{1mm}
\begin{lemma} \label{lem.dc}
For Algorithm 1, the iterates satisfy
\[ \frac{c \cdot \pi_k - z^*}{c \cdot e - z^*} = \frac{- \lambda_{\min}(x_k) }{1 - \lambda_{\min}(x_k)} \; . \]
\end{lemma}
\noindent {\bf Proof:} Immediate from $ \pi_k = e + \frac{1}{1 - \lambda_{\min}(x_k)} (x_k - e) $ and $ c \cdot x_k = z^* $. \hfill $ \Box $
\vspace{3mm}
\noindent {\bf Proof of Corollary~\ref{cor.db}:}
With $ {\mathcal K} $ being polyhedral, the concave function $ x \mapsto \lambda_{\min}(x) $ is piecewise linear, and thus there exists a positive constant $ C $ such that
\[ \mathrm{dist}_{z^*}(x) \leq - C \, \lambda_{\min}(x) \quad \textrm{for all $ x \in \mathrm{Affine}_{z^*} $} \; . \]
Then Proposition~\ref{prop.da} gives
\[
\max \{ \lambda_{\min}(x_k) \mid k = \ell, \ldots, \ell + m \} \geq C \, M \, \lambda_{\min}(x_{\ell})/ \sqrt{m+1} \; , \]
from which follows
\[ \max \{ \lambda_{\min}(x_k) \mid k = \ell, \ldots, \ell + \lceil (2CM)^2 \rceil \} \geq \smfrac{1}{2} \lambda_{\min}(x_{\ell}) \; , \]
i.e., $ \lambda_{\min}(x_{\ell}) $ is ``halved'' within $ \lceil (2CM)^2 \rceil $ iterations. The proof is easily completed using Lemma~\ref{lem.dc}. \hfill $ \Box $ \vspace{3mm}
We now return to considering general convex cones $ {\mathcal K} $. \vspace{1mm}
The iteration bound provided by Proposition~\ref{prop.da} bears little obvious connection to the geometry of the conic program CP, except in that the constant $ M $ is related to the geometry by Proposition~\ref{prop.cb}. The other constant -- $ \mathrm{dist}_{z^*}(x_k) $ -- does not at present have such a clear geometrical connection to CP. We next observe a meaningful connection. \vspace{1mm}
The {\em level sets} for CP are the sets
\[ \mathrm{Level}_{z} = \mathrm{Affine}_{z} \cap {\mathcal K} \; , \]
that is, the largest feasible sets for CP on which the objective function is constant\footnote{There is possibility of confusion here, as in the optimization literature, the terminology ``level set'' is often used for the portion of the feasible region on which the (convex) objective function does not exceed a specified value rather than -- as for us -- exactly equals the value. Our terminology is consistent with the general mathematical literature, where the region on which a function does not exceed a specified value is referred to as a sublevel set, not a level set.}. If $ z < z^* $, then $ \mathrm{Level}_{z} = \emptyset \; . $ \vspace{1mm}
If some level set is unbounded, then either CP has unbounded optimal value or can be made to have unbounded value with an arbitrarily small perturbation of $ c $. Thus, in developing numerical optimization methods, it is natural to focus on the case that level sets for CP are bounded. \vspace{1mm}
For scalars $ z $, define the diameter of $ \mathrm{Level}_z $ by
\[ \diamval := \sup \{ \| x - y \| \mid x,y \in \mathrm{Level}_{z} \} \; , \]
the diameter of $ \mathrm{Level}_{z} $. If $ \mathrm{Level}_z = \emptyset $, let $ \diamval := - \infty $. \vspace{1mm}
\begin{lemma} \label{lem.dd}
Assume $ x \in \mathrm{Affine}_{z^*} $, and let $ \pi = \pi(x) $. Then
\[ \mathrm{dist}_{z^*}(x) \, = \, (1 - \lambda_{\min}(x)) \, \mathrm{dist}_{c \cdot \pi}(\pi) \, = \, \frac{ \mathrm{dist}_{c \cdot \pi}(\pi)}{1 - \frac{c \cdot \pi - z^*}{c \cdot e - z^*}} \, \leq \, \frac{ \mathrm{diam}_{c \cdot \pi}}{1 - \frac{c \cdot \pi - z^*}{c \cdot e - z^*}} \; . \]
\end{lemma}
\noindent {\bf Proof:} Since
\begin{equation} \label{eqn.da}
\pi = e + \smfrac{1}{1 - \lambda_{\min}(x)}(x-e) \; , \end{equation}
Theorem~\ref{thm.bc} implies that the maximizers of the map $ y \mapsto \lambda_{\min}(y) $ over $ \mathrm{Affine}_{ c \cdot \pi} $ are precisely the points of the form
\[ x_{c \cdot \pi}^* = e + \smfrac{1}{1 - \lambda_{\min}(x)}(x_{z^*}^*-e) \; , \]
where $ x_{z^*}^* $ is a maximizer of the map when restricted to $ \mathrm{Affine}_{z^*} $ (i.e., is an optimal solution of CP). Observing
\[ \pi - x_{c \cdot \pi}^* = \smfrac{1}{1 - \lambda_{\min}(x)} (x - x_{z^*}^*) \; , \]
it follows that
\[ \mathrm{dist}_{c \cdot \pi}( \pi) = \smfrac{1}{1 - \lambda_{\min}(x)} \, \mathrm{dist}_{z^*}(x) \; , \]
establishing the first equality in the statement of the lemma. The second equality then follows easily from (\ref{eqn.da}) and $ c \cdot x = z^* $. The inequality is due simply to $ \pi, x_{c \cdot \pi}^* \in \mathrm{Level}_{ c \cdot \pi} $, for all optimal solutions $ x_{c \cdot \pi}^* $ of the CP-equivalent problem (\ref{eqn.ca}) (with $ z = c \cdot \pi $). \hfill $ \Box $
\vspace{3mm}
For scalars $ z $, define
\[ \mathrm{Diam}_z := \max \{ \mathrm{diam}_{z'} \mid z' \leq z \} \; , \]
the ``horizontal diameter'' of the sublevel set consisting of points $ x $ that are feasible for CP and satisfy $ c \cdot x \leq z $. For $ z^* < z < c \cdot e $, the value $ \mathrm{Diam}_{z} $ can be thought of as a kind of condition number for CP, because $ \mathrm{Diam}_{z} $ being large is an indication that the optimal value for CP is relatively sensitive to perturbations in the objective vector $ c $. \vspace{1mm}
For $ z^* \leq z < c \cdot e $, define
\[ \mathrm{Dist}_z := \sup \{ \mathrm{dist}_{z'}(x) \mid z' \leq z \textrm{ and } x \in \mathrm{Level}_{z'} \} \; . \]
Clearly, there holds the relation
\[ \mathrm{Dist}_z \leq \mathrm{Diam}_z \; , \]
and hence if the ``condition number'' $ \mathrm{Diam}_z $ is only of modest size, so is the value $ \mathrm{Dist}_z $. \vspace{1mm}
Following is our main result for Algorithm 1. By substituting $ \mathrm{Diam}_{c \cdot \pi_0} $ for $ \mathrm{Dist}_{ c \cdot \pi_0} $, and $ 1/r_e$ for $ M $ (where $ r_e $ is as in Proposition~\ref{prop.cb}), the statement of the theorem becomes phrased in terms clearly reflecting the geometry of CP. \vspace{1mm}
\begin{thm} \label{thm.de}
Assume $ 0 < \epsilon < \frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} $, where $ \pi_0 = \pi(x_0) $ is the initial $ \mathrm{CP} $-feasible point for Algorithm 1 (i.e., assume $ \pi_0 $ does not itself satisfy the desired accuracy). Then
\begin{align*}
& \ell \geq (2M \, \mathrm{Dist}_{c \cdot \pi_0})^2 \, \, \left( \, \frac{4}{ 3 } \left( \frac{1 - \epsilon }{ \epsilon } \right)^2 + 4 \left( \frac{1 - \epsilon}{\epsilon} \right) \right. \\
& \qquad \qquad \qquad \qquad \qquad \qquad \left. + \log_2 \left( \frac{\frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} }{\epsilon} \right) + \log_2 \left( \frac{1 - \epsilon }{1 - \frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} } \right) \, + 1 \, \right) \\ & \qquad \quad \Rightarrow \quad \min_{k \leq \ell } \, \frac{c \cdot \pi_k - z^*}{c \cdot e - z^*} \, \leq \, \epsilon \; . \end{align*}
\end{thm}
\noindent {\bf Proof:} To ease notation, let $ \lambda_k := \lambda_{\min}(x_k) \; . $ \vspace{1mm}
Let $ k_0 = 0 $ and recursively define $ k_{i+1} $ to be the first index for which $ \lambda_{k_{i+1}} \geq \lambda_{k_i}/2 $ (keeping in mind $ \lambda_k \leq 0 $ for all $ k $). Proposition~\ref{prop.da} implies
\begin{align}
k_{i+1} - k_i + 1 & \leq \left( \frac{2M \, \mathrm{dist}_{z^*}(x_{k_i})}{\lambda_{k_i}} \right)^2 \nonumber \\
& = \left( \, 2 M \, \mathrm{dist}_{c \cdot \pi_{k_i}}(\pi_{k_i}) \, \frac{ 1 - \lambda_{k_i}}{ \lambda_{k_i}} \, \right)^2 \quad \textrm{(by Lemma~\ref{lem.dd})} \nonumber \\
& \leq \left( \, 2 M \, \mathrm{Dist}_{c \cdot \pi_0} \, \frac{ 1 - \lambda_{k_i}}{ \lambda_{k_i}} \, \right)^2 \; , \label{eqn.db}
\end{align}
where the final inequality is due to $ c \cdot \pi_{k_i} $ ($ i = 0, 1, \ldots $) being a decreasing sequence (using Lemma~\ref{lem.dc}). \vspace{1mm}
Let $ i' $ be the first sub-index for which $ \lambda_{k_{i'}} \geq - \epsilon/(1 - \epsilon) $. Lemma~\ref{lem.dc} implies
\[ \frac{c \cdot \pi_{k_{i'}} - z^*}{ c \cdot e - z^* } \, \leq \, \epsilon \; . \]
Thus, to prove the theorem, it suffices to show $ \ell = k_{i'} $ satisfies the inequality in the statement of the theorem. \vspace{1mm}
Note $ i' > 0 $ (because, by assumption, $ \epsilon < \frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} $). Observe, then,
\begin{align}
i' & < 1 + \log_2 \left( \frac{ \lambda_0}{- \epsilon/(1 - \epsilon )} \right) \nonumber \\
& = 1 + \log_2 \left( \frac{\frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} }{ \epsilon} \right) + \log_2 \left( \frac{1 - \epsilon }{1 - \frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} } \right) \label{eqn.dc}
\end{align}
(again using Lemma~\ref{lem.dc}).
\vspace{1mm}
Additionally,
\begin{align*}
k_{i'} & = \sum_{i=0}^{i'-1} k_{i+1} - k_i \\ &
\leq (2M \, \mathrm{Dist}_{c \cdot \pi_0})^2 \, \sum_{i=0}^{i'-1} \left( \frac{1 - \lambda_{k_i}}{\lambda_{k_i}} \right)^2 \quad \textrm{(by (\ref{eqn.db}))} \\
& \leq (2M \, \mathrm{Dist}_{c \cdot \pi_0})^2 \, \sum_{i=0}^{i'-1} \left( \frac{1 - 2^i \lambda_{k_{i'-1}}}{2^i \lambda_{k_{i'-1}}} \right)^2 \\
& \leq (2M \, \mathrm{Dist}_{c \cdot \pi_0})^2 \, \sum_{i=0}^{i'-1} \left( \frac{1 + 2^i \epsilon /(1 - \epsilon)}{2^i \epsilon /(1 - \epsilon) } \right)^2 \\
& = (2M \, \mathrm{Dist}_{c \cdot \pi_0})^2 \, \sum_{i=0}^{i'-1} \left( 1 + \frac{1}{2^i} \, \frac{1 - \epsilon}{\epsilon } \right)^2 \\
& \leq (2M \, \mathrm{Dist}_{c \cdot \pi_0})^2 \, \left( i' + 4 \frac{1- \epsilon }{ \epsilon } + \frac{4}{3} \left( \frac{1- \epsilon }{ \epsilon } \right)^2 \right) \; .
\end{align*}
Using (\ref{eqn.dc}) to substitute for $ i' $ completes the proof. \hfill $ \Box $
\section{ {\bf Algorithm 2: When The Optimal Value Is Unknown}} \label{sect.e}
For the second algorithm, we discard the requirement of knowing $ z^* $. Now $ \epsilon $ (the desired relative-accuracy) is required as input. \vspace{2mm}
\noindent
\hrulefill
\noindent {\bf Algorithm 2} \vspace{1mm}
\noindent
\begin{tabular}{lll}
(0) & Input: & $ 0 < \epsilon < 1 $ ,\\
&& $ e $, a strictly feasible point for CP, and \\
&& $ \bar{x} \in \mathrm{Affine} $ satisfying $ c \cdot \bar{x} < c \cdot e $. \\
& Initialize: & $ x_0 = \pi_0 = \pi(\bar{x} ) $ \\
(1) &Iterate: & Compute $ \tilde{x}_{k+1} := x_k + \smfrac{ \epsilon }{2 \| P_{{\mathcal L}} g_k \|^2} P_{{\mathcal L}} g_k $, \, where $ g_k \in \hat{\partial}\lambdamin(x_k) $. \\
&& Let $ \pi_{k+1} := \pi(\tilde{x}_{k+1}) \; . $ \\
&& If $ c \cdot (e - \pi_{k+1}) \geq \smfrac{4}{3} \, c \cdot (e - \tilde{x}_{k+1}) $, let $ x_{k+1} = \pi_{k+1} $; \\
&& \qquad else, let $ x_{k+1} = \tilde{x}_{k+1} \; . $ \vspace{1mm}
\end{tabular}
\noindent
\hrulefill
\vspace{2mm}
Unsurprisingly, the iteration bound we obtain for Algorithm 2 is worse than the result for Algorithm 1, but perhaps surprisingly, the bound is not excessively worse, in that the factor for $ 1/ \epsilon^2 $ is essentially unchanged (it's the factor for $ 1/\epsilon $ that increases, although typically not by a large amount). \vspace{2mm}
\begin{thm} \label{thm.ea}
Assume $ 0 < \epsilon < \frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*} $. For the iterates of Algorithm 2,
\begin{align*}
& \ell \geq 8 \, (M \, \mathrm{Dist}_{c \cdot \pi_0} )^2 \, \left( \, \frac{1}{\epsilon^2} \, + \, \frac{1}{\epsilon} \, \log_{4/3} \left( \frac{ 1 }{1 - \frac{c \cdot \pi_0 - z^*}{c \cdot e - z^*}} \right) \, \right) \\ & \qquad \quad \Rightarrow \quad \min_{k \leq \ell} \, \frac{c \cdot \pi_k - z^*}{c \cdot e - z^*} \, \leq \, \epsilon \; . \end{align*}
\end{thm}
\noindent {\bf Proof:} In order to distinguish the iterates obtained by projecting to the boundary, we record a notationally-embellished rendition of the algorithm which introduces a distinction between ``inner iterations'' and ``outer iterations'':
\vspace{2mm}
\noindent
\hrulefill
\noindent {\em Algorithm 2 (notationally-embellished version):} \vspace{1mm}
\noindent
\begin{tabular}{l}
\begin{tabular}{lll}
(0) & Input: & $ 0 < \epsilon < 1 \; , $ $ e $ and $ \bar{x} $. \\
& Initialize: & $ y_{1,0} = \pi(\bar{x} ) \; , $ \\
&& $ i = 1$ (outer iteration counter), \\
&& $ j = 0 $ (inner iteration counter).
\end{tabular} \\
\begin{tabular}{ll}
(1) & Compute $ y_{i,j+1} = y_{i,j} + \smfrac{\epsilon }{2 \| \tilde{g}_{i,j} \|^2} \, \tilde{g}_{i,j} \; , $ \\
& \qquad \qquad \qquad \qquad \qquad \qquad \quad where $ \tilde{g}_{i,j} = P_{{\mathcal L}}g_{i,j} $ and $ g_{i,j} \in \hat{\partial}\lambdamin(y_{i,j}) $. \\
(2) & If $ c \cdot (e - \pi(y_{i,j+1})) \geq \smfrac{4}{3} \, c \cdot (e - y_{i,j+1} ) \; , $ \\
& \qquad then let \, $ y_{i+1,0} = \pi(y_{i,j+1}) $, \, $ i \leftarrow i + 1 $ \, and \, $ j \leftarrow 0 \; $;\\
& \quad else, let \, $ j \leftarrow j + 1 \; . $ \\
(3) & Go to step 1. \vspace{1mm}
\end{tabular}
\end{tabular}
\noindent
\hrulefill
\vspace{2mm}
For each outer iteration $ i $, all of the iterates $ y_{i,j} $ have the same objective value. Denote the value by $ z_i $. Obviously, $ z_1 $ is equal to the value $ c \cdot \pi_0 $ appearing in the statement of the theorem. Let
\[ \mathrm{Dist} := \mathrm{Dist}_{c \cdot \pi_0} \, = \mathrm{Dist}_{z_1} \; . \]
Step 2 ensures
\begin{equation} \label{eqn.ea}
c \cdot e - z_{i+1} \geq \smfrac{4}{3} ( c \cdot e - z_i) \; .
\end{equation}
Thus, $ z_1, z_2, \ldots $ is a strictly decreasing sequence. Consequently, as $ y_{i,0} \in \mathrm{Level}_{z_i} $, we have $ \mathrm{dist}_{z_i}(y_{i,0}) \leq \mathrm{Dist} $ for all $ i $. \vspace{1mm}
From (\ref{eqn.ea}) we find for scalars $ \delta > 0 $ that
\[
\frac{c \cdot e - z_{i+1}}{ c \cdot e - z^*} < \delta \quad \Rightarrow \quad i < \log_{4/3} \left( \frac{\delta}{ \frac{c \cdot e - z_1}{c \cdot e - z^*}} \right) \, = \log_{4/3} \left( \frac{\delta}{ 1 - \frac{z_1 - z^*}{c \cdot e - z^*}} \right) \; ,
\]
and thus, for $ \epsilon < 1 $,
\begin{equation} \label{eqn.eb}
\frac{z_i - z^*}{c \cdot e - z^*} > \epsilon \quad \Rightarrow \quad i < 1 + \log_{4/3} \left( \frac{1 - \epsilon}{ 1 - \frac{z_1 - z^*}{c \cdot e - z^*}} \right) \; .
\end{equation}
Hence, if an outer iteration $ i $ fails to satisfy the inequality on the right, the initial inner iterate $ y_{i,0} $ fulfills the goal of finding a CP-feasible point $ \pi $ satisfying $ \frac{c \cdot \pi - z^*}{c \cdot e - z^*} \leq \epsilon $ (i.e., the algorithm has been successful no later than the start of outer iteration $ i $). Also observe that (\ref{eqn.eb}) provides (letting $ \epsilon \downarrow 0 $) an upper bound on $ I $, the total number of outer iterations:
\begin{equation} \label{eqn.ec}
I \leq 1 + \log_{4/3} \left( \frac{1}{ 1 - \frac{z_1 - z^*}{c \cdot e - z^*}} \right) \; .
\end{equation}
\vspace{1mm}
For $ i = 1, \ldots, I $, let $ J_i $ denote the number of inner iterates computed during outer iteration $ i $, that is, $ J_i $ is the largest value $ j $ for which $ y_{i,j} $ is computed. Clearly, $ J_I = \infty $, whereas $ J_1, \ldots, J_{I-1} $ are finite. \vspace{1mm}
To ease notation, let $ \lambda_{i,j} := \lambda_{\min}(y_{i,j}) $, and let $ \lambda_i^* := \lambda_{\min}(x_{z_i}^*) $, the optimal value of
\[ \begin{array}{rl}
\max & \lambda_{\min}(x) \\
\textrm{s.t.} & x \in \mathrm{Affine}_{z_i} \; . \end{array} \]
According to Lemma~\ref{lem.bd},
\begin{equation} \label{eqn.ed}
\lambda_i^* = \frac{z_i - z^*}{c \cdot e - z^*} \; .
\end{equation}
It is thus valid, for example, to substitue $ \lambda_i^* $ for $ \frac{z_i - z^*}{c \cdot e - z^*} $ in (\ref{eqn.eb}). Additionally, (\ref{eqn.ed}) implies (\ref{eqn.ea}) to be equivalent to
\begin{equation} \label{eqn.ee}
1 - \lambda_{i+1}^* \geq \smfrac{4}{3} ( 1 - \lambda_i^* ) \; .
\end{equation}
For any point $ y $, we have $ \pi(y) = e + \smfrac{1}{1 - \lambda_{\min}(y)} ( y - e) $, and thus,
\[ \frac{ c \cdot e - c \cdot \pi(y)}{c \cdot e - c \cdot y} = \frac{1}{1 - \lambda_{\min}(y)} \; . \]
Hence,
\[ \frac{c \cdot e - c \cdot \pi(y)}{c \cdot e - c \cdot y} \geq \frac{4}{3} \quad \Leftrightarrow \quad \lambda_{\min}(y) \geq 1/4 \; . \]
Consequently,
\begin{equation} \label{eqn.ef}
\lambda_{i,j} < 1/4 \, \, \textrm{ for $ j < J_i $} \; .
\end{equation}
We use the following relation implied by Lemma~\ref{lem.cc}:
\begin{equation} \label{eqn.eg}
\mathrm{dist}_{z_i}(y_{i,j+1})^2 \leq \mathrm{dist}_{z_i}(y_{i,j})^2 - \smfrac{\epsilon }{\| \tilde{g}_{i,j} \|^2 } ( \lambda_i^* - \lambda_{i,j} ) + ( \smfrac{\epsilon}{2 \| \tilde{g}_{i,j} \| })^2 \; .
\end{equation}
We begin bounding the number of inner iterations by showing
\begin{equation} \label{eqn.eh}
\lambda_i^* \geq \max \{ \smfrac{1}{2} , \epsilon \} \quad \Rightarrow \quad J_i \leq \frac{8 (M \, \mathrm{Dist})^2}{ \epsilon } \; .
\end{equation}
Indeed, for $ j < J_i \; , $
\begin{align*}
& - \epsilon \, ( \lambda_i^* - \lambda_{i,j} ) + \smfrac{1}{4} \, \epsilon^2 \\
& < - \epsilon \, ( \max \{ \smfrac{1}{2}, \epsilon \} - \smfrac{1}{4} ) + \smfrac{1}{4} \epsilon^2 \quad \textrm{(using (\ref{eqn.ef}))} \\
& = \min \left\{ \smfrac{1}{4} ( \epsilon^2 - \epsilon), \smfrac{1}{4} \epsilon - \smfrac{3}{4} \epsilon^2 \right\} \\
& \leq \smfrac{3}{4} \, \smfrac{1}{4} ( \epsilon^2 - \epsilon) + \smfrac{1}{4} ( \smfrac{1}{4} \epsilon - \smfrac{3}{4} \epsilon^2 ) \\
& = - \smfrac{1}{8} \epsilon \; .
\end{align*}
Thus, according to (\ref{eqn.eg}), for $ j < J_i $,
\[ \mathrm{dist}_{z_i}(y_{i,j+1})^2 \leq \mathrm{dist}_{z_i}(y_{i,j})^2 - \frac{ \epsilon}{8 M^2} \; , \]
inductively giving
\begin{align*}
\mathrm{dist}_{z_i}(y_{i,j+1})^2 & \leq \mathrm{dist}_{z_i}(y_{i,0})^2 - \frac{(j+1) \, \epsilon}{8 \, M^2 } \\
& \leq \mathrm{Dist}^2 - \frac{ (j+1) \, \epsilon}{ 8 M^2} \; .
\end{align*}
The implication (\ref{eqn.eh}) immediately follows. \vspace{1mm}
The theorem is now readily established in the case $ \epsilon \geq 1/2 $. Indeed, because of the identity (\ref{eqn.ed}), the quantity on the right of (\ref{eqn.eb}) provides an upper bound on the number of outer iterations $ i $ for which $ \lambda_i^* > \epsilon $, whereas the quantity on the right of (\ref{eqn.eh}) gives, assuming $ \epsilon \geq 1/2 $, an upper bound on the number of inner iterations for each of these outer iterations. However, for the first outer iteration satisfying $ \lambda_i^* \leq \epsilon $, the initial iterate $ y_{i,0} $ ($ = \pi(y_{i,0})) $ itself achieves the desired accuracy $ \frac{c \cdot \pi - z^*}{c \cdot e - z^*} \leq \epsilon $. Thus, the total number of inner iterations made before the algorithm is successful is at most the product of the two quantities which is seen not to exceed the iteration bound in the statement of the theorem (using $ \log_{4/3}(1- \epsilon) \leq \log_{4/3}(1/2) < -1 $). \vspace{1mm}
Before considering the remaining case, $ \epsilon < 1/2 $, we establish a relation applying for all $ \epsilon $. For any outer iteration $ i $ for which $ \lambda_i^* < 3/4 $, and for any $ 0 < \epsilon < 1 $, let
\[ \widehat{J}_i := \left\lceil \frac{1}{\smfrac{3}{4} - \lambda_i^*} \, \left( \frac{M \, \mathrm{Dist}}{ \epsilon } \right)^2 - 1 \right\rceil \; . \]
We claim that either
\begin{equation} \label{eqn.ei}
J_i \leq \widehat{J}_i \quad \textrm{or} \quad \min \left\{ \frac{c \cdot \pi(y_{i, j }) - z^*}{ c \cdot e - z^* } \mid j = 0, \ldots, \widehat{J}_i \right\} \, \leq \, \epsilon \; .
\end{equation}
Consequently, if $ J_i > \widehat{J}_i $, the algorithm will achieve the goal of computing a point $ y $ satisfying $ \frac{c \cdot \pi(y) - z^*}{c \cdot e - z^*} \leq \epsilon $ within $ \widehat{J}_i $ inner iterations during outer iteration $ i $.
\vspace{1mm}
To establish (\ref{eqn.ei}), assume $ \widehat{J}_i < J_i $ and yet the inequality on the right of (\ref{eqn.ei}) does not hold. (We obtain a contradiction.) For every $ j \leq \widehat{J}_i $, Proposition~\ref{prop.be} then implies
\begin{align*}
\lambda_i^* - \lambda_{i,j} & > \epsilon \, \, \frac{c \cdot e - z_i}{c \cdot e - z^*} \\
& = (1 - \lambda_i^*) \, \epsilon \quad \textrm{(by (\ref{eqn.ed}))} \; ,
\end{align*}
and hence, using (\ref{eqn.eg}),
\[ \mathrm{dist}_{z_i}(y_{i,j+1})^2 < \mathrm{dist}_{z_i}(y_{i,j})^2 - ( \smfrac{3}{4} - \lambda_i^*) \, \left( \epsilon / M \right)^2 \; , \]
from which inductively follows
\begin{align*}
\mathrm{dist}_{z_i}( y_{i,\widehat{J}_i+1})^2 & < \mathrm{dist}_{z_i}( y_{i,0})^2 - (\widehat{J}_i +1) \, (\smfrac{3}{4} - \lambda_i^* ) \, ( \epsilon/M)^2 \\
& \leq \mathrm{Dist}^2 - (\widehat{J}_i +1) \, (\smfrac{3}{4} - \lambda_i^* ) \, ( \epsilon/M)^2 \\
& \leq 0 \; ,
\end{align*}
a contradiction. The claim is established. \vspace{1mm}
Assume $ \epsilon < 1/2 $, the case remaining to be considered. \vspace{1mm}
As each outer iteration $ i $ satisfying $ \lambda_i^* \geq 1/2 $ has only finitely many inner iterations, there must be at least one outer iteration $ i $ satisfying $ \lambda_i^* < 1/2 $.
Let $ i $ be the first outer iteration for which $ \lambda_i^* < 1/2 $. From (\ref{eqn.ec}) and (\ref{eqn.eh}), the total number of inner iterations made before reaching outer iteration $ i $ is at most
\begin{equation} \label{eqn.ej}
\frac{8 (M \, \mathrm{Dist})^2}{ \epsilon } \, \, \log_{4/3} \left( \frac{1}{ 1 - \frac{z_1 - z^*}{c \cdot e - z^*}} \right) \; .
\end{equation}
According to (\ref{eqn.ei}), during outer iteration $ i $, the algorithm either achieves its goal within $ \widehat{J}_i $ inner iterations, or the algorithm makes no more than $ \widehat{J}_i $ inner iterations before starting a new outer iteration. Assume the latter case. Then, for outer iteration $ i + 1 $, the algorithm either achieves its goal within $ \widehat{J}_{i+1} $ inner iterations, or the algorithm makes no more than $ \widehat{J}_{i+1} $ inner iterations before starting a new outer iteration. Assume the latter case. In iteration $ i + 2 $, the algorithm definitely achieves its goal within $ \widehat{J}_{i+2} $ inner iterations, because there cannot be a subsequent outer iteration due, by (\ref{eqn.ee}), to
\[ \smfrac{4}{3} (1 - \lambda_{i+2}^*) \geq \left( \smfrac{4}{3} \right)^3 (1 - \lambda_i^*) > \left( \smfrac{4}{3} \right)^3 \smfrac{1}{2} > 1 \; . \]
The total number of inner iterations made before the algorithm achieves its goal is thus bounded by the sum of the quantity (\ref{eqn.ej}) and
\begin{align*}
& \widehat{J}_i + \widehat{J}_{i+1} + \widehat{J}_{i+2} \\
& < \left( \frac{1}{(1-\frac{1}{2}) - \frac{1}{4} } + \frac{1}{\frac{4}{3} \, (1 - \smfrac{1}{2}) - \smfrac{1}{4} } + \frac{1}{\frac{4}{3} \, \frac{4}{3} \, (1 - \smfrac{1}{2}) - \smfrac{1}{4}} \right) \, \, \left( \frac{M \, \mathrm{Dist}}{ \epsilon } \right)^2 \\
& \qquad \textrm{(using $ \smfrac{3}{4} - \lambda_j^* = (1 - \lambda_j^*) - \smfrac{1}{4} $)} \\
& < 8 \, \left( \frac{M \, \mathrm{Dist}}{ \epsilon } \right)^2 \; ,
\end{align*}
completing the proof of the theorem. \hfill $ \Box $
\section{{\bf Application to General Convex Optimization}} \label{sect.f}
We now return to the setting of Section 1, considering optimization problems of the form
\begin{equation} \label{eqn.fa}
\begin{array}{rl} \min & f(x) \\
\textrm{s.t.} & x \in \mathrm{Feas} \; , \end{array}
\end{equation}
where $ f: {\mathcal E} \rightarrow (- \infty , \infty ] $ is an extended-valued and lower-semicontinuous convex function, where $ \mathrm{Feas} = \{ x \in S \mid Ax = b \} $, with $ S $ being a closed convex set, and where there is known a point $ \bar{e} \in \mathrm{int}(S \cap \dom(f)) $.
\vspace{1mm}
In this section we recast (\ref{eqn.fa}) into conic form, then interpret the complexity results obtained by applying Algorithms 1 and 2 to the conic refomulation. Before proceeding, we recall notation from \S \ref{sect.a}. \vspace{1mm}
Let $ \langle \; , \; \rangle $ be the (computational) inner product on $ {\mathcal E} $, and $ \| \, \, \| $ the associated norm. Let $ P $ be the linear operator orthogonally projecting $ {\mathcal E} $ onto the kernel of $ A $. \vspace{1mm}
Recall $ \bar{D} $ denotes the diameter of the sublevel set $ \{ x \in \mathrm{Feas} \mid f(x) \leq f( \bar{e} ) \} $. The diameter is assumed to be finite, implying the optimal value $ f^* $ is attained at some feasible point. \vspace{1mm}
Recall $ \hat{f} $ is a user-chosen scalar satisfying $ \hat{f} > f( \bar{e} ) $ (hence, $ ( \bar{e}, \hat{f}) \in \int(\epi(f)) $), and recall
\begin{equation} \label{eqn.fb}
\hat{r} := \sup \{ r \mid \| x - \bar{e} \| \leq r \textrm{ and } Ax = b \, \, \Rightarrow \, \, x \in S \textrm{ and } f(x) \leq \hat{f} \} \; ,
\end{equation}
a positive scalar (because $ \bar{e} \in \mathrm{int}\big( S \cap \mathrm{dom}(f)\big) $). \vspace{1mm}
The values $ \bar{D} $ and $ \hat{r} $ are not assumed to be known, but do appear in complexity results. \vspace{1mm}
For later reference, observe the convexity of $ f $ implies
\[ f( \bar{e} ) \leq \smfrac{\hat{r} }{ \bar{D} + \hat{r} } f^* + \smfrac{\bar{D} }{\bar{D} + \hat{r} } \hat{f} \; , \]
which in turn implies
\begin{equation} \label{eqn.fc}
\frac{1}{1 - \frac{f( \bar{e} ) - f^*}{ \hat{f} - f^*}} \, \leq \, 1 + \bar{D}/\hat{r} \; .
\end{equation}
As $ S $ and $ \epi(f) $ are closed and convex, there exist closed, convex cones $ {\mathcal K}_1, {\mathcal K}_2 \subseteq {\mathcal E} \times \mathbb{R} \times \mathbb{R} $ for which
\[ S \times \mathbb{R} = \{ (x,t) \mid (x,1, t) \in {\mathcal K}_1 \} \quad \textrm{and} \quad \mathrm{epi}(f) = \{ (x,t) \mid (x,1,t) \in {\mathcal K}_2 \} \; . \]
Letting $ {\mathcal K} := {\mathcal K}_1 \cap {\mathcal K}_2 $,
clearly the optimization problem (\ref{eqn.da}) is equivalent to
\begin{equation} \label{eqn.fd}
\begin{array}{rl}
\min_{x,s,t} & t \\
\textrm{s.t.} & Ax = b \\
& s = 1 \\
& (x,s,t) \in {\mathcal K} \; , \end{array} \end{equation}
and has the same optimal value, $ f^* $. The conic program (\ref{eqn.fd}) is of the same form as CP, the focus of preceding sections. Clearly,
\[ \mathrm{Affine} = \{ (x,1,t) \mid Ax = b \} \; , \textrm{ and for scalars $ z $, } \mathrm{Affine}_z = \{ (x,1,z) \mid Ax = b \} \; . \]
For distinguished direction, choose $ e = ( \bar{e} , 1, \hat{f} \, ) $, which clearly lies in the interiors of $ {\mathcal K}_1 $ and $ {\mathcal K}_2 $, and thus lies in $ \int( {\mathcal K}) $. This distinguished direction, along with the cone $ {\mathcal K} $, determines the map $ (x,s,t) \mapsto \lambda_{\min}(x,s,t) $ on $ {\mathcal E} \times \mathbb{R} \times \mathbb{R} $. \vspace{1mm}
For $ z < f^* $, the conic problem \ref{eqn.fd} -- and hence the problem \ref{eqn.fa} -- is, by Theorem~\ref{thm.bc}, equivalent to
\begin{equation} \label{eqn.fe}
\begin{array}{rl}
\max_{x,s,t} & \lambda_{\min}(x,s,t) \\
\textrm{s.t.} & (x,s,t) \in \mathrm{Affine}_z \; ,
\end{array}
\end{equation}
It is to this problem that supgradient methods are applied.
\vspace{1mm}
Choose the computational inner product on $ {\mathcal E} \times \mathbb{R} \times \mathbb{R} $ to to be
\[ \blin (x_1, s_1, t_1), (x_2, s_2, t_2) \brin := \langle x_1, x_2 \rangle + s_1 s_2 + t_1 t_2 \; , \]
in which case
\[ P_{{\mathcal L}}(x,s,t) = (P(x),0,0) \; . \]
\vspace{1mm}
Observe for all scalars $ z $,
\begin{equation} \label{eqn.ff}
\mathrm{Level}_{z} = \{ (x,1, z ) \mid x \in \mathrm{Feas} \textrm{ and } f(x) \leq z \} \; . \end{equation}
Recall that the Lipschitz constant $ M $ for the map $ (x,s,t) \mapsto \lambda_{\min}(x,s,t) $ restricted to $ \mathrm{Affine}_z $ is independent of $ z $. Since by
(\ref{eqn.fb}),
\[ \mathrm{Affine}_{\hat{f}} \cap \bar{B}(e,\hat{r} ) \subseteq {\mathcal K} \; , \]
Proposition~\ref{prop.cb} implies
\begin{equation} \label{eqn.fg}
M \leq 1/ \hat{r} \; .
\end{equation}
Choose the input $ \bar{x} $ to Algorithms 1 and 2 as $ \bar{x} = ( \bar{e} , 1, f( \bar{e} )) $, which clearly is feasible for the conic program (\ref{eqn.fd}). Note that (\ref{eqn.ff}) then implies the horizontal diameter of the relevant sublevel set for the conic program satisfies
\begin{equation} \label{eqn.fh}
\mathrm{Diam}_{f(\bar{e} )} = \bar{D} \; .
\end{equation}
Algorithm 2 requires input $ 0 < \epsilon < 1 $, but not $ f^* $.
Applying Algorithm 2 results in a sequence of iterates $ (x_k, 1, t_k) $ for which the projections $ (x_k',1,t_k') := \pi(x_k,1,t_k) $ satisfy $ x_k' \in \mathrm{Feas} $ and $ f(x_k') \leq t_k' $ (simply because $ (x_k',1,t_k') $ is feasible for the conic program (\ref{eqn.fd})). \vspace{1mm}
Since $ (x_0,1,t_0) = (\bar{e} ,1,f(\bar{e} )) \in \bdy({\mathcal K}) $, we have $ \pi(x_0, 1, t_0) = (x_0,1,t_0) $ -- in particular, the objective value of $ \pi(x_0,1,t_0) $ is $ f(\bar{e}) $. Consequently, the sequence of points $ x_k' $ not only lie in $ \mathrm{Feas} $, but by Theorem~\ref{thm.ea} satisfies
\begin{align}
& \ell \geq 8 \, \left( \frac{\bar{D}}{\hat{r}} \right)^2 \, \left( \, \frac{1}{\epsilon^2} \, + \, \frac{1}{\epsilon} \, \log_{4/3} \left( 1 + \frac{\bar{D}}{\hat{r}} \right) \, \right) \label{eqn.fi} \\
& \quad \quad \Rightarrow \quad \min_{k \leq \ell } \, \frac{f(x_k') - f^*}{\hat{f} - f^*} \, \leq \, \epsilon \; , \label{eqn.fj}
\end{align}
where for (\ref{eqn.fi}) we have used (\ref{eqn.fc}), (\ref{eqn.fg}), (\ref{eqn.fh}), and for (\ref{eqn.fj}) have used $ f(x_k') \leq t_k' $. \vspace{1mm}
For Algorithm 1, which requires input $ f^* $ but not $ \epsilon $, points $ (x_k', 1, t_k') $ are generated for which the sequence $ \{ x_k' \} $ is feasible and, by Theorem~\ref{thm.de}, satisfies
\begin{align}
&
\ell \geq 4 \, \left( \frac{\bar{D}}{\hat{r}} \right)^2 \, \cdot \, \left( \, \frac{4}{ 3 } \left( \frac{1 - \epsilon }{ \epsilon } \right)^2 + 4 \left( \frac{1 - \epsilon}{\epsilon} \right) \right. \label{eqn.fk} \\
& \qquad \qquad \qquad \qquad \qquad \qquad \left. + \log_2 \left( \frac{1 - \epsilon }{ \epsilon } \right) + \log_2 \left( \frac{\bar{D}}{\hat{r}} \right) \, + 1 \, \right)
\label{eqn.fl} \\ & \qquad \quad \Rightarrow \quad \min_{k \leq \ell } \, \frac{f(x_k') - f^*}{\hat{f} - f^*} \, \leq \, \epsilon \; . \label{eqn.fm} \end{align}
We prove Theorem~\ref{thm.aa} by showing Algorithm A is ``equivalent'' to applying Algorithm 2 as above. Likewise, we prove Theorem~\ref{thm.ab} by showing algorithm B is equivalent to Algorithm 1. To be in position to establish these equivalences, however, we need a characterization of the supdifferentials $ \hat{\partial} \lambdamin(x,s,t) $, a characterization in terms of the original problem (\ref{eqn.fa}). \vspace{1mm}
\section{{\bf A Practical Characterization of Supdifferentials, \\ and the Proofs of Theorems~\ref{thm.aa} and \ref{thm.ab} }} \label{sect.g}
We continue with the setting and notation developed in Section \ref{sect.f}.
\vspace{1mm}
We provide a practical characterization of the supgradients for the ``$ \lambda_{\min} $ function,'' a characterization expressed in terms of the original problem (\ref{eqn.fa}). The characterization provides the means for other algorithms designed using the framework to be implemented directly in terms of (\ref{eqn.fa}), avoiding computation of the conic reformulation. \vspace{1mm}
For the reader's convenience, we recall key sets appearing in Section~\ref{sect.a} that also appear in the characterization below. \vspace{1mm}
For $ x' \in \mathrm{bdy}(S) $, let
\[ G_1(x') := \left\{ \frac{-1}{ \langle v , \bar{e} - x' \rangle } \, v \mid \, \vec{0} \neq v \in N_S(x') \right\} \; .
\]
For $ (x',t') \in \mathrm{\bdy}(\mathrm{epi}(f)) $, let
\[ G_2(x',t') := \left\{ \frac{-1}{\langle v, \bar{e} - x' \rangle + ( \hat{f} - t') \delta } \, v \mid \, (\vec{0},0) \neq (v,\delta) \in N_{\mathrm{epi}(f)}(x',t') \right\} \; , \]
where here, normality is with respect to the inner product that assigns pairs $ (x_1, t_1) $, $ (x_2, t_2) $ the value $ \langle x_1, x_2 \rangle + t_1 t_2 $.
\vspace{1mm}
Finally, letting $ \bdy_1 := \{ (x',t') \mid x' \in \bdy(S) \} $ and $ \bdy_2 := \bdy(\epi(f)) $, define for $ (x',t') \in \bdy_1 \cup \bdy_2 \; , $
\[ G(x',t') := \begin{cases} G_1(x') & \textrm{if $ (x',t') \in \bdy_1 \setminus \bdy_2 \; , $} \\
G_2(x',t') & \textrm{if $ (x',t') \in \bdy_2 \setminus \bdy_1 \; , $} \\
\textrm{hull}(G_1(x') \, \cup \, G_2(x',t') ) & \textrm{if $ (x',t') \in \bdy_1 \cap \bdy_2 \; . $}
\end{cases} \]
\begin{prop} \label{prop.ga}
Assume $ (x,1,t) \in \mathrm{Affine} $, and assume $ t < \hat{f} $. If $ (x',1,t') = \pi(x,1,t) $, then
\[ P_{{\mathcal L}}( \, \hat{\partial}\lambda_{\min}(x,1,t) \, ) = \{ (P(v),0,0) \mid -v \in G(x',t') \} \; . \]
\end{prop}
\noindent {\bf Proof:} Proposition \ref{prop.ca} implies for all $ y \in {\mathcal E} \times \mathbb{R} \times \mathbb{R} $,
\begin{align}
\hat{\partial} \lambda_{\min}(y) & = \{ \upsilon \in - N_{{\mathcal K}}(y - \lambda_{\min}(y) e ) \mid \blin e, \upsilon \brin = 1 \} \nonumber \\
& = - \{ \upsilon \in N_{{\mathcal K}}(y - \lambda_{\min}(y) e ) \mid \blin e, \upsilon \brin = - 1 \} \; . \label{eqn.ga}
\end{align}
However,
\begin{equation} \label{eqn.gb}
\textrm{if $ y = (x,1,t) \in \mathrm{Affine} $ and $ t < \hat{f} $,}
\end{equation}
then $ \pi (y) $ is a positive multiple of $ y - \lambda_{\min}(y) e $ (using $ \lambda_{\min}(y) < 1 $ (Lemma~\ref{lem.bb}), and $ \pi(y) = e - \frac{1}{1 - \lambda_{\min}(y)}(y - e) $), and thus has the same normal cone to $ {\mathcal K} $. Hence, in this case, (\ref{eqn.ga}) is equivalent to
\begin{equation} \label{eqn.gc}
\hat{\partial} \lambda_{\min}(y) = - \{ \upsilon \in N_{{\mathcal K}}(\, \pi (y) \, ) \mid \blin e, \upsilon \brin = -1 \} \; .
\end{equation}
As $ \mathrm{int}( {\mathcal K}) \neq \emptyset $ and $ {\mathcal K} = {\mathcal K}_1 \cap {\mathcal K}_2 $, if $ y \in {\mathcal K} $ then
\[ N_{{\mathcal K}}(y) = \mathrm{hull}( \, N_{{\mathcal K}_1}(y) \cup N_{{\mathcal K}_2}(y) \, ) \; .
\]
Thus, for $ y $ as in (\ref{eqn.gb}), using (\ref{eqn.gc}) we have
\begin{equation} \label{eqn.gd}
\hat{\partial} \lambda_{\min}(y) = - \{ \upsilon \in \mathrm{hull}( \, N_{{\mathcal K}_1}(\pi(y)) \cup N_{{\mathcal K}_2}(\pi(y)) \, ) \mid \blin e, \upsilon \brin = -1 \} \; .
\end{equation}
\vspace{1mm}
Since $ e \in \mathrm{int}( {\mathcal K}_i) $ ($ i = 1,2 $),
\[ y \in {\mathcal K}_i \, \wedge \, \upsilon \in N_{{\mathcal K}_i}(y) \, \wedge \, \blin e, \upsilon \brin = 0 \quad \Rightarrow \quad \upsilon = 0 \; . \]
Consequently, it follows from (\ref{eqn.gd}) that for $ y $ as in (\ref{eqn.gb}),
\[
\hat{\partial} \lambda_{\min}(y) = -\mathrm{hull}( {\mathcal N}_1( \pi(y)) \cup {\mathcal N}_2(\pi (y)) ) \; ,
\]
where for $ i = 1,2 $ and $ y' \in {\mathcal K}_i $,
\[ {\mathcal N}_i(y') := \begin{cases} \{ v \in N_{{\mathcal K}_i}(y') \mid \blin e, \upsilon \brin = - 1 \} & \textrm{if $ y' \in \mathrm{bdy}({\mathcal K}_i$)}, \\ \{ \vec{0} \} & \textrm{if $ y' \in \mathrm{int}({\mathcal K}_i $)}.
\end{cases} \]
Clearly, then,
\begin{equation} \label{eqn.ge}
P_{{\mathcal L}}( \hat{\partial} \lambda_{\min}(y)) = - \mathrm{hull} \big( \, P_{\mathcal L}( {\mathcal N}_1( \pi(y))) \, \cup \, P_{{\mathcal L}}( {\mathcal N}_2( \pi(y))) \, \big) \; .
\end{equation}
Since $ S \times \mathbb{R} = \{ (x,t) \mid (x,1,t) \in {\mathcal K}_1 \} \; , $ if $ y' = (x',1,t') \in \bdy( {\mathcal K}_1) $ then
\[ N_{{\mathcal K}_1}(y') = \{ (v, \gamma, 0) \mid v \in N_S(x') \, \wedge \langle x', v \rangle = 0 \} \; , \]
the equation to ensure perpendicularity to the ray through $ y' $. Thus, a vector $ (v, \gamma, \delta) $ is an element of $ {\mathcal N}_1(y') $ if and only if $ v \in N_S(x') $, $ \delta = 0 $ and
\[ \begin{array}{ccccc}
\langle x', v \rangle & + & \gamma & = & 0 \\
\langle \bar{e} , v \rangle & + & \gamma & = & -1
\end{array}
\]
(note $ v $ cannot be $ \vec{0} $).
However, for $ \vec{0} \neq v \in N_S(y') $, there is a unique scaling of $ v $ for which there exist $ \gamma $ and $ \delta $ satisfying the three equations, namely,
\[ v \mapsto \smfrac{-1}{\langle v, \bar{e} - x' \rangle } (v, \, - \langle x', v \rangle , \, 0) \; . \]
The denominator in the scaling is negative, because $ \bar{e} \in \int(S) $. It follows that
\begin{equation} \label{eqn.gf}
P_{\mathcal L} ( {\mathcal N}_1 ) = \{ \smfrac{-1}{\langle v, \bar{e} - x' \rangle } ( P(v), 0, 0) \mid \vec{0} \neq v \in N_S(x') \} \; .
\end{equation}
Similarly, for $ y' = (x',1,t') \in \bdy( {\mathcal K}_2) $, a vector $ (v,\gamma, \delta) $ is in $ {\mathcal N}_2(y') $ if and only $ ( v, \delta ) \in N_{ \epi(f)}( x', t') $ and
\[ \begin{array}{ccccccc}
\langle x', v \rangle & + & \gamma & + & \delta t' & = & 0 \\
\langle \bar{e}, v \rangle & + & \gamma & + & \delta \hat{f} & = & -1 \
\end{array}
\]
(note $ (v, \delta) $ cannot be $ ( \vec{0},0) $).
For $ ( \vec{0},0) \neq (v,\delta) \in N_{ \epi(f)}(x',t') $, the unique scaling for which there exists $ \gamma $ satisfying the equations is
\[ (v, \delta) \mapsto \smfrac{-1}{\langle v, \bar{e} - x' \rangle + (\hat{f} - t') \delta} (v, \, - \langle x', v \rangle - t' \delta , \, \delta) \; . \]
The denominator is negative, because $ (\bar{e}, \hat{f}) \in \int(\epi(f)) $. It follows that
\begin{equation} \label{eqn.gg}
P_{\mathcal L} ( {\mathcal N}_2 ) = \{ \smfrac{-1}{\langle v, \bar{e} - x' \rangle + (\hat{f} - t') \delta} ( P(v), 0, 0) \mid (\vec{0},0) \neq (v, \delta) \in N_{\epi(f)}(x',t') \} \; . \end{equation}
Together, (\ref{eqn.ge}), (\ref{eqn.gf}) and (\ref{eqn.gg}) establish the proposition. \hfill $ \Box $
\vspace{5mm}
\noindent
{\bf Proofs of Theorems \ref{thm.aa} and \ref{thm.ab}, and of (\ref{eqn.ag}):} We first establish Theorem~\ref{thm.aa}, and then remark on the few changes required to establish Theorem~\ref{thm.ab} by the same approach. \vspace{1mm}
By ``Algorithm 2,'' we mean the application of Algorithm 2 to the conic optimization problem (\ref{eqn.fd}), as presented in the preceding section. \vspace{1mm}
Due to complexity bound (\ref{eqn.fi}), (\ref{eqn.fj}) established for Algorithm 2, it suffices to show the two algorithms are equivalent in that the iterates generated by Algorithm 2 are ``identical'' to those generated by Algorithm A. \vspace{1mm}
The initial iterates for Algorithm 2 are $ (x_0, 1, t_0) = (x_0',1,t_0') = (\bar{e}, 1, f( \bar{e}) $, and for Algorithm A are $ (x_0, t_0) = (x_0',t_0') = (\bar{e}, f(\bar{e})) $. For beginning an inductive proof showing the two algorithms are equivalent, observe that the initial iterate $ (x_0, t_0) $ (resp., $ (x_0', t_0') $) for Algorithm A is obtained simply by eliminating the ``1'' from the initial iterate $ (x_0,1,t_0) $ (resp., $ (x_0',1,t_0') $) for Algorithm 2. \vspace{1mm}
For the inductive step, assume eliminating ``1'' from the iterate $ (x_k,1,t_k) $ (resp., $ (x_k',1,t_k') $) computed by Algorithm 2 results in the iterate $ (x_k,t_k) $ (resp., $ (x_k',t_k') $) computed by Algorithm A.
\vspace{1mm}
The first iterate computed by Algorithm 2 in Step 1 is
\[ (\tilde{x}_{k+1}, 1, t_k) = (x_k, 1, t_k) + \smfrac{\epsilon }{2 \| P_{\mathcal L} \bar{g} \|^2} P_{{\mathcal L}} \bar{g} \quad \textrm{where } \bar{g} \in \hat{\partial} \lambdamin(x_k', 1, t_k') \; , \]
whereas Algorithm A computes
\[ \tilde{x}_{k+1} = x_k - \smfrac{\epsilon }{2 \| P g \|^2} Pg \quad \textrm{where } g \in G(x_k', t_k') \]
(and thereafter relies on the pair $ ( \tilde{x}_{k+1}, t_k) $ in the same manner that Algorithm 2 relies on the triple $ (\tilde{x}_{k+1}, 1, t_k) $). Thus, due to the identity provided by Proposition~\ref{prop.ga}, the algorithms can be considered equivalent in this computation.
\vspace{1mm}
Algorithm 2 next computes $ (x_{k+1}', 1, t_{k+1}') := \pi(\tilde{x}_{k+1},1, t_k) $, the point in the boundary of $ {\mathcal K} $ encountered when moving from $ (\bar{e}, 1, \hat{f}) $ in direction $ (\tilde{x}_{k+1},1 , t_k) - (\bar{e},1,\hat{f}) $. Due to the definition of $ {\mathcal K} $, however, this is the (only) feasible point $ (x,1,t) $ for which there exists $ \alpha \geq 0 $ such that
\[ x = x( \alpha) := \bar{e} + \alpha \cdot ( \tilde{x}_{k+1} - \bar{e}) \quad \textrm{and} \quad t = t( \alpha) := \hat{f} + \alpha \cdot ( t_k - \hat{f}) \]
and either $ x \in \mathrm{bdy}(S) $ or $ (x,t) \in \bdy(\epi(f)) $. Thus, for the iterate $ (x_{k+1}', 1, t_{k+1}') $ computed by Algorithm 2, we have $ (x_{k+1}', t_{k+1}') = \pi'( \tilde{x}_{k+1}, t_k) $, where $ \pi' $ is defined by (\ref{eqn.ac}). Thus, eliminating ``1'' from the iterate $ (x_{k+1}', 1, t_{k+1}' ) $ computed by Algorithm 2 gives the iterate $ (x_{k+1}', t_{k+1}') $ computed by Algorithm A. \vspace{1mm}
Observe, moreover, for Algorithm A, $ t_{k+1}' = \hat{f} + \alpha_{k+1} \cdot (t_k - \hat{f}) $, and thus,
\begin{equation} \label{eqn.gh}
\alpha_{k+1} = \frac{\hat{f} - t_{k+1}'}{\hat{f} - t_k} \; .
\end{equation}
Algorithm A decides how to define its iterate $ (x_{k+1}, t_{k+1}) $ based on whether $ \alpha_{k+1} \geq 4/3 $, whereas Algorithm 2 decides how to define its iterate $ (x_{k+1},1,t_{k+1}) $ based on whether $ \hat{f} - t_{k+1}' \geq \smfrac{4}{3} ( \hat{f} - t_k) $. Thus, due to (\ref{eqn.gh}) and the equivalence of the two algorithms up until now, it is simple to verify that removing ``1'' from the iterate $ (x_{k+1},1,t_{k+1}) $ for Algorithm 2 gives the iterate $ (x_{k+1}, t_{k+1}) $ for Algorithm A. \vspace{1mm}
Similarly, the proof of Theorem~\ref{thm.ab} is accomplished by showing Algorithm B is ``equivalent'' to Algorithm 1, and making use of the complexity result (\ref{eqn.fk}), (\ref{eqn.fl}), (\ref{eqn.fm}) for Algorithm 1. The proof of equivalence is virtually identical to the one above, the main difference being that the scalar $ \alpha_k $ used by Algorithm B has to be related to the analogous scalar used by Algorithm 1 -- in particular, for equivalence, the identity $ \frac{\alpha(x_k, f^*) - 1}{\alpha(x_k, f^*)} = \lambda_{\min}(x_k,1,f^*) $ is needed. \vspace{1mm}
The identity, however, is easily established. Indeed, $ \alpha(x_k,f^*) $ is the smallest scalar $ \alpha $ for which the point $ (x,t) = ( \bar{e}, \hat{f}) + \alpha \cdot ( (x_k, f^*) - (\bar{e}, \hat{f})) $ satisfies either $ x \in \mathrm{bdy}(S) $ or $ (x,t) \in \bdy(\epi(f)) $. Due to the definition of $ {\mathcal K} $, however, this is readily seen to equal the smallest scalar $ \alpha $ for which $ e + \alpha \cdot ((x_k,1,f^*) - e) $ lies in the boundary of $ {\mathcal K} $. Hence, from (\ref{eqn.bda}), $ \alpha(x_k,f^*) = \frac{1}{1 - \lambda_{\min}(x_k,1,f^*)} $. Rearrangement establishes the desired identity. \vspace{1mm}
Finally, immediately after the statement of Theorem~\ref{thm.ab}, it is claimed in (\ref{eqn.ag}) that Algorithm B converges linearly when $ f $ is a piecewise linear function and $ \mathrm{Feas} $ is polyhedral. This is immediate from the fact that Algorithm 1 converges linearly when $ {\mathcal K} $ is a polyhedral cone (Corollary~\ref{cor.db}).
\hfill $ \Box $
\vspace{3mm}
{\bf Acknowledgements:} The author expresses gratitude to the reviewers and the associate editor, for careful attention to detail and for suggestions on how to position the paper. The author thanks Yurii Nesterov for encouragement at an early, critical stage of the research, Rob Freund for discussions and feedback influencing the overall manner of presentation, and Wei Qian for conversations affecting the deduction and presentation of results for general convex optimization.
\bibliographystyle{plain}
|
\section{Introduction}
\subsection{Background and Motivation}
Derivatives and integrals of non-integer order, often referred to as
\textit{fractional}, are natural extensions of the standard
integer-order ones which enjoy certain favourable properties:
they are linear operators, preserve analyticity, and have the
semigroup property~\cite{Pod99,Hil00}.
Nonetheless, fractional derivatives are non-local operators, that is,
unlike integer-order ones, they cannot be evaluated
at a given point by mere knowledge of the function in
a neighbourhood of this point and for that reason they are
suitable for describing phenomena with infinite memory~\cite{Pod99}.
Fractional dynamics seems to be omnipresent in nature.
Examples of fractional systems include, but are not limited to,
semi-infinite transmission
lines with losses~\cite{ClaNarHan04}, viscoelastic
polymers~\cite{Hil00}, magnetic core coils~\cite{Magnetism},
anomalous diffusion in semi-infinite bodies~\cite{GuoLiWan15}
and biomedical applications~\cite{Magin2010} for which
Magin \textit{et al.} provided a thorough
review~\cite{MagOrtPodTru11}.
A good overview of the applications of fractional systems in
physics is given in~\cite{Hil00} and~\cite{TarBook}.
A shift towards fractional-order dynamics in the field of pharmacokinetics
may be observed after the classical \textit{in-vitro-in-vivo correlations} theory
proved to have faced its limitations~\cite{KytMacDok10}.
Non-linearities, anomalous diffusion, deep tissue trapping,
diffusion across capillaries, synergistic and competitive
action and other phenomena give rise to fractional-order
pharmacokinetics~\cite{DokMac08}.
In fact, Pereira derived fractional-order diffusion laws for
media of fractal geometry~\cite{Per10}.
Increasing attention has been drawn on
modelling and control of such systems
\cite{DokMac11,DokMagMac10,SopSar14}, especially in presence of
state and input constraints.
\red
Model predictive control (MPC) is an advanced, successful and well recognized control
methodology and its adaptation to fractional systems is of particular interest.
The current model predictive control framework for fractional-order systems has been
developed in a series of papers where integer-order approximations are used to formulate the
control problem~\cite{RomDeMad+13,RomDeMad+10,BouBou10,DenCao+10,RomDeMadManVin10}.
CARIMA (controlled auto-regressive moving average) models are often used in
predictive control formulations for the approximation of the fractional dynamics~\cite{RomDeMad+13,RomDeMad+10,RomDeMadManVin10,Joshi2015}.
The CARIMA-based approach has been used in various applications such as the
heating control of a semi-infinite rod~\cite{RhoBou14},
the power regulation of a solid oxide fuel cell~\cite{DenCao+10}
and various applications in automotive technology~\cite{RomDeMad+12}.
The celebrated Oustaloup approximation has also been used in MPC settings~\cite{RomDeMad+13}.
It should, however, be noted that such approximations aim at capturing the system
dynamics in a range of operating frequencies and, as a result, are not suitable for a
rigorous analysis and design of controllers for constrained systems.
Additionally, all of the aforementioned works provide examples of unconstrained
systems; this shortcoming was in fact identified in the recent paper~\cite{Joshi2015}.
}
Nevertheless, this profusion of purportedly successful paradigms of MPC for
fractional-order systems is not accompanied by a proper stability
analysis especially when input and state constraints are present.
A common denominator of all approaches in the literature is
that they approximate the actual fractional dynamics by
integer-order dynamics and design controllers for the
approximate system using standard techniques.
No stability and constraint satisfaction guarantees can be
deduced for the original fractional-order system.
Currently, one of the very few works on constrained
control for fractional-order systems is due to Mesquine \textit{et al.}
where, however, only input constraints are taken into account \red{for
the design of a linear feedback controller}~\cite{MesHmaBenBenTad15}.
Hitherto, two approaches can be found in the literature in
regard to the stability analysis of discrete-time fractional
systems. The first one considers the stability of a
finite-dimensional linear time-invariant (LTI) system, known as
\textit{practical stability}, but fails to provide
conditions for the actual fractional-order system to be
(asymptotically) stable~\cite{BusKac09,GueDjeBet12}. This
approach is tacitly pursued in many applied papers \red{where stability is
established only for a finite-dimensional approximation of the fractional-order
system
~\cite{RomDeMad+13,RomTejSuar09}.
On the other hand, fractional systems can be treated as
infinite-dimensional systems for which various stability
conditions can be derived (See for example~\cite[Thm.~2]{GueDjeBetMaa10}),
but conditions are difficult to verify in
practice let alone to use for the design of model predictive --- or other ---
controllers.
\subsection{Contribution}
In this paper we describe a stabilising MPC framework
for fractional-order systems (of the Gr\"{u}{nwald-Letnikov type)
subject to state and input constraints.
We discretise linear continuous-time fractional dynamics
using the Gr\"{u}{nwald-Letnikov scheme which leads to infinite-dimensional linear
systems. Using a finite-dimensional approximation
we arrive at a linear time-invariant system with an additive
uncertainty term which casts the discrepancy with the
infinite-dimensional system. We then introduce a tube-based MPC
control scheme which is known to steer the state to a
neighbourhood of the origin which can become arbitrarily
small as the order of the approximation of the
fractional-order system increases.
In our analysis, we consider both state and input constraints
which we show that are respected by the MPC-controlled system.
We finally prove that under a certain contraction-type condition
the origin is an asymptotically stable
equilibrium point for the MPC-controlled fractional-order
system (see Section~\ref{sec:stabilisation}). \red{In this work we provide,
for the first time, asymptotic stability conditions (Theorem~\ref{thm:stability-2}) and we propose a
control methodology which guarantees the satisfaction of the prescribed state
and input constraints.}
This paper builds up on~\cite{SopNtoSar15} where the unmodelled
part of the system dynamics was cast as a bounded
additive uncertainty term and used existing MPC theory to
drive the system's state in a neighbourhood of the origin
without, however, providing any (asymptotic) stability conditions
for the origin.
\subsection{Mathematical preliminaries}\label{sec:mathematics}
The following definitions and notation will be used throughout
the rest of this paper.
Let $\N$, $\Re^{n}$, $\Re_+$, $\Re^{m\times n}$
denote the set of non-negative integers,
the set of column real vectors of length
$n$, the set of non-negative numbers and the set of
$m$-by-$n$ real matrices respectively.
For any nonnegative integers $k_1\leq k_2$ the finite
set $\{k_1,\ldots,k_2\}$ is denoted by $\N_{[k_1,k_2]}$.
Let $x$ be a sequence of real vectors of $\Re^n$.
The $k$-th vector of the sequence is denoted by
$x_k$ and its $i$-th element is denoted by $x_{k,i}$.
We denote by $\mathcal{B}_\epsilon^n=\{x\in\Re^n:\|x\|<
\epsilon\}$ the \emph{open ball} of $\Re^n$ with radius
$\epsilon$ and we use the shorthand $\mathcal{B}^n=\mathcal{B}_1^n$.
We define the \emph{point-to-set distance}
of a point $z\in X$ from $A$ as
$\dist(z,A)=\inf_{a\in A}\|z-a\|$.
The space of bounded real sequences is denoted by
$\ell^\infty$.
We define the space $\ell^\infty_n$
of all sequences of real $n$-vectors $z$ so that
$(z_{k,i})_{k}\in\ell^\infty$ for $i\in\N_{[1,n]}$.
Let $\Gamma$ be a topological real vector space and $A,B\subseteq \Gamma$.
For $\lambda\in\Re$
we define the scalar product $\lambda C=\{\lambda c: c\in C\}$
and the \emph{Miknowski sum} $A\oplus B=\{a+b: a\in A, b\in B\}$.
The Minkowski sum of a finite family of sets $\{A_i\}_{i=1}^{k}$
will be denoted by $\bigoplus_{i=1}^{k}A_i$. The Minkowski
sum of a sequence of sets $\{A_i\}_{i\in\N}$ is denoted by
$\bigoplus_{i\in\N}A_i$ or $\bigoplus_{i=0}^{\infty}A_i$ and is defined as the \emph{Painlev{\'e}-Kuratowski
limit} of $\bigoplus_{i=1}^{k}A_i$ as
$k{\to}\infty$~\cite{RocWet98}. The \emph{Pontryagin difference} between two sets
$A,B\subseteq \Gamma$ is defined as $A\ominus B=
\{a\in A:a+b\in A, \forall b\in B\}$. A set $C$ is called
\emph{balanced} if for every $x\in C$, $-x\in C$.
\section{Fractional-order Systems}
\subsection{Discrete-time fractional-order systems}
Let $x:\Re\to\Re^n$ be a uniformly bounded function,
\textit{i.e.}, there is a $M>0$ so that $\|x(t)\|\leq M$
for all $t\in\Re$.
The Gr\"{u}nwald-Letnikov fractional-order difference of $x$ of order $\alpha>0$
and step size $h>0$ at $t$
is defined as the linear operator~\cite{OrtCoiTru15,RhoBouBou14} $\Delta^\alpha_h:\ell_n^\infty\to\ell_n^\infty$:
\begin{align}\label{eq:gl-diff}
\Delta^\alpha_h x(t) = \sum_{j=0}^{\infty}(-1)^j{\alpha \choose j}x(t-jh),
\end{align}
where ${\alpha \choose 0}=1$ and for $j\in\N$, $j>0$
\begin{align}
{\alpha \choose j}=\prod_{i=0}^{j-1}\frac{\alpha-i}{i+1}=\frac{\Gamma(\alpha+1)}{\Gamma(\alpha-j+1)j!}.
\end{align}
The \emph{forward-shifted} counterpart of $\Delta^\alpha_h$ is defined as
${_{F}\Delta^\alpha_h} x(t)=\Delta^\alpha_h x(t+h)$.
Now, define
\begin{align}
c^\alpha_j=(-1)^j{\alpha \choose j}= {j-\alpha-1 \choose j},
\end{align}
and notice for all $j\in\N$
that $|c^\alpha_j|\leq \alpha^j/j!$, thus, the sequence
$(c^\alpha_j)_j$ is absolutely summable and, because of the
uniform boundedness of $x$, the series in~\eqref{eq:gl-diff}
converges, therefore, $\Delta^\alpha_h$ is well-defined.
It is worth noticing that for $\alpha\in\N$ it is $c_j^\alpha=0$
for $j\geq \lceil \alpha \rceil $, but this property does not
hold for $\alpha \notin \N$.
As a result, at time $t$ and for non-integer orders
$\alpha$ the whole history of $x$ is needed in order to
estimate $\Delta^\alpha_h x(t)$.
The Gr\"{u}nwald-Letnikov difference operator gives
rise to the Gr\"{u}nwald-Letnikov derivative
of order $\alpha$ which is defined as~\cite[Sec.~{20}]{Sam+93}
\begin{align}
D^\alpha x(t)=\lim_{h\to 0^+}\frac{_{F}\Delta^\alpha_h x(t)}{h^\alpha}
=\lim_{h\to 0^+}\frac{\Delta^\alpha_h x(t)}{h^\alpha},
\end{align}
insofar as both limits exist.
This derivative is then used to describe fractional-order
dynamical systems with state $x:\Re\to\Re^n$ and
input $u:\Re\to\Re^m$ as follows:
\begin{align}\label{eq:frac-sys}
\sum_{i=1}^{l} A_i D^{\alpha_i} x(t)=
\sum_{i=1}^{r} B_i D^{\beta_i} u(t),
\end{align}
where $l,r\in\N$, $A_i$ are $B_i$ are matrices of
opportune dimensions, all $\alpha_i$ and $\beta_i$ are nonnegative,
and by convention $D^0 x(t) = x(t)$ for any $x$.
In an Euler discretisation fashion we approximate the $D^{\alpha}$
in~\eqref{eq:frac-sys} using either $h^{-\alpha} {_{F}}\Delta^{\alpha}_h$
or $h^{-\alpha} \Delta^{\alpha}_h$ for a fixed step size $h$ as in~\cite{OrtCoiTru15}.
\red{In particular, we use ${_{F}}\Delta^{\alpha}_h$ for the derivatives
of the state and $\Delta^{\alpha}_h$ for the input variables}.
We define $x_k=x(kh)$ and $u_k=u(kh)$ for $k\in\mathbb{Z}$ so the
discretisation of~\eqref{eq:frac-sys} becomes
\begin{align}\label{eq:frac-sys-discrete}
\sum_{i=1}^{l}\bar{A}_i \Delta_h^{\alpha_i} x_{k+1}=
\sum_{i=1}^{r}\bar{B}_i \Delta_h^{\beta_i} u_{k},
\end{align}
with $\bar{A}_i=h^{-\alpha_i}A_i$ and
$\bar{B}_i=h^{-\beta_i}B_i$.
The involvement of infinite-dimensional operators in the
system dynamics deem these systems computationally intractable
and call for approximation methods for their simulation
and the design of feedback controllers.
In what follows, we will approximate~\eqref{eq:frac-sys-discrete}
by a finite-dimensional state-space system treating the approximation
as a bounded additive disturbance. We then propose a control
setting which guarantees robust stability properties fo
~\eqref{eq:frac-sys-discrete}.
\subsection{Finite-dimension approximation}\label{sec:foa}
Discrete-time fractional-order dynamical
systems are essentially systems with infinite memory and an
infinite number of state variables. As a result, standard stability theorems and
control design methodologies for finite-dimensional systems cannot be applied directly.
To this end we introduce the following \textit{truncated Gr\"{u}nwald-Letnikov difference operator}
of length $\nu$ given by
\begin{align}
\Delta^\alpha_{h,\nu} x_k =
\sum_{j=0}^{\nu}c_{j}^{\alpha}x_{k-j},
\end{align}
System~\eqref{eq:frac-sys-discrete} is the approximated by the following
system using $\nu\geq 1$
\begin{align}\label{eq:frac-sys-approx}
{\sum_{i=1}^{l}\bar{A}_i\Delta_{h,\nu}^{\alpha_i} x_{k+1}=
\sum_{i=1}^{r}\bar{B}_i\Delta_{h,\nu}^{\beta_i} u_{k}}.
\end{align}
System~\eqref{eq:frac-sys-approx} can be written in state
space format as a linear time-invariant system with a
proper choice of state variables $\tilde{x}_k$ as we shall explain
in this section. In the common case where the right-hand side
of~\eqref{eq:frac-sys-approx} is of the simple form
$Bu_k$, it is straightforward to recast the system in
state-space form. Here, we study the more general case of
equation~\eqref{eq:frac-sys-approx}, which can be written in
the form
\begin{align}\label{eq:qwerty}
\sum_{j=0}^{\nu}\hat{A}_j x_{k-j+1} = \sum_{j=0}^{\nu}\hat{B}_j u_{k-j},
\end{align}
with $\hat{A}_j =\sum_{i=1}^{l}\bar{A}_i c_{j}^{\alpha_i}$
and $\hat{B}_j=\sum_{i=1}^{r}\bar{B}_i c_{j}^{\beta_i}$ for $j\in\N_{[0,\nu]}$.
We hereafter assume that matrix $\hat{A}_0$ is nonsingular.
With this assumption, the discrete-time dynamical system~\eqref{eq:qwerty}
becomes a \textit{normal system}, that is, future states can be determined using
past states in a unique fashion and can be written as
a linear time-invariant system~\cite[Chap.~1]{Duan10}.
Defining $\tilde{A}_j=-\hat{A}_0^{-1}\hat{A}_j$
and $\tilde{B}_j=\hat{A}_0^{-1}\hat{B}_j$,
the dynamic equation~\eqref{eq:qwerty}
becomes
\begin{align}\label{eq:zxcv}
x_{k+1}=\sum_{j=0}^{\nu-1} \tilde{A}_j x_{k-j}
+ \sum_{j=1}^{\nu}\tilde{B}_j u_{k-j} + \tilde{B}_0 u_{k}.
\end{align}
This can be written in state space form with state
variable $\tilde{x}_k=(x_k, x_{k-1},\ldots, x_{k-\nu+1},
u_{k-1},\ldots, u_{k-\nu})'$ as
\begin{align}\label{eq:lti}
\tilde{x}_{k+1}=A\tilde{x}_{k}+Bu_{k}.
\end{align}
System~\eqref{eq:lti} is an ordinary finite-dimensional
LTI system which will be used in the next section to
formulate a model predictive control problem.
Throughout the rest of the paper we assume that
the pair $(A,B)$ is stabilisable.
The truncated difference operator
$\Delta_{h,\nu}^{\alpha}$ introduces some error in the
system dynamics. In particular, the fractional-order difference
operator ${\Delta_h^\alpha}$ can be written as
\begin{align}
\red{{\Delta_{h}^{\alpha}}=\Delta_{h,\nu}^{\alpha}+R^{\alpha}_{h,\nu}},
\end{align}
where $R^{\alpha}_{h,\nu}:\ell^{\infty}_{n}\to\ell^{\infty}_{n}$
is the operator $R^{\alpha}_{h,\nu}(x_k)=\sum_{j=\nu+1}^{\infty}c_{j}^{\alpha}x_{k-j}$.
Let $X$ be a compact convex subset in $\Re^n$
containing $0$ in its interior and at
time $k$ assume that $x_{k-j}\in X$ for all $j\in\N$. Then,
by the assumption that $x_{k-j}\in X$ for all $j\in\N$,
\begin{equation}\label{eq:Rnu}
R^{\alpha}_{h,\nu}(x_k)\in \bigoplus_{j=\nu+1}^{\infty}c_{j}^{\alpha}X.
\end{equation}
For all $\nu\in\N$, the right-hand side of~\eqref{eq:Rnu} is a convex
compact set with the origin in its interior.
Equation~\eqref{eq:frac-sys-discrete} can now be rewritten using the augmented
state variable $\tilde{x}$ (cf.~\eqref{eq:lti})
leading to the following linear uncertain system
\begin{align}\label{eq:uncertain-lti}
\tilde{x}_{k+1}=A\tilde{x}_{k}+Bu_{k}+Gd_k,
\end{align}
where $d_k$ is a (bounded) additive disturbance term (which depends
on $x_{k-\nu-j}$ and $u_{k-\nu-j}$ for $j\in\N$) with $G=\smallmat{I&0&\ldots&0}'$.
Assume that $u_{k-j}\in U$ for $j=1,2,\ldots$ and
$x_{k-j}\in X$ for $j\in\N$, where $X$ and $U$ are
convex compact sets containing $0$ in their interiors.
Then, $d_k$ is bounded in a compact set $D_{\nu}$ given by
\begin{align}\label{eq:D}
D_{\nu}{=}D_{\nu}^{x}\oplus D_{\nu}^{u},
\end{align}
where
\begin{subequations}
\begin{align}
D_{\nu}^{x}&=\bigoplus_{i=1}^{l} -\hat{A}_0^{-1}\bar{A}_i \bigoplus_{j=\nu+1}^{\infty} c_{j}^{\alpha_i} X,\\
D_{\nu}^{u}&=\bigoplus_{i=1}^{r} \hat{A}_0^{-1}\bar{B}_i\bigoplus_{j=\nu+1}^{\infty} c_{j}^{\beta_i}U.
\end{align}
\end{subequations}
Under the prescribed assumptions $D_{\nu}$ is a compact set.
Hereafter, we shall use the notation $A^\ast_i = -\hat{A}_0^{-1}\bar{A}_i$
and $B^\ast_i=\hat{A}_0^{-1}\bar{B}_i$.
Recall that for a balanced set $C\subseteq\Re^n$
and scalars $\lambda_1, \lambda_2$ it is
$\lambda_1 C\oplus \lambda_2 C=(|\lambda_1|+|\lambda_2|)C$.
In case $X$ and $U$ are balanced sets, the above expressions for $D_{\nu}^{x}$
and $D_{\nu}^{u}$
can be simplified. First, for $\nu\in\N$, we define the function
$\Psi_{\nu}:\Re_+\to\Re_+$ as follows
\begin{align}
\Psi_{\nu}(\alpha)=\sum_{j=\nu+1}^{\infty}|c_j^\alpha|.
\end{align}
Then, $D_{\nu}^{x}$ is written as the finite Minkowski sum
\begin{align}\label{eq:Dn-formula}
D_{\nu}^{x} =\bigoplus_{i\in\N_{[1,l]}} A^\ast_i \Psi_{\nu}(\alpha_i) X,
\end{align}
and of course the same simplification applies to $D_{\nu}^{u}$ if $U$
is a balanced set. Notice that the computation of $D_{\nu}^{x}$
by~\eqref{eq:Dn-formula} boils down to determining a finite
Minkowski sum, which is possible when constraints are polytopic~\cite{GriStu93},
while overapproximations exists when they are ellipsoidal~\cite{KurEllips97}.
The size of $D_\nu$ is controlled by the choice of $\nu$;
$D_\nu$ can become arbitrarily small provided that a
sufficiently large $\nu$ is chosen.
Notice also that $D_\nu\to \{0\}$ as $\nu\to\infty$.
In light of~\eqref{eq:uncertain-lti}, the fractional system
can be controlled by standard methods of robust control
such as min-max~\cite{DieBjo04} or tube-based MPC
\cite{RawMay09}; here we use the latter approach. In what follows,
we elaborate on how the tube-based MPC methodology can be
applied for the control of fractional-order systems.
\red{Various integer-order approximation methodologies have
been proposed in the literature such as continued fraction expansions of the
system's transfer function, the approximation methods of Carlson, Matsuda,
Oustaloup, Chareff and more (see~\cite{approx2000review} for an overview).
Methods which are based on the approximation of the system dynamics in
a given frequency range cannot lead to the formulation of an LTI system
with a bounded disturbance as in~\eqref{eq:uncertain-lti} and, as a result,
cannot be used to guarantee stability for constrained systems as we will
present in the next section.}
\section{Model Predictive Control}
\subsection{Tube-based Model Predictive Control}\label{sec:tb-mpc}
Model predictive control is a class of advanced control algorithms
where the control action is calculated at every time instant by solving
a constrained optimisation problem where a \textit{performance index}
is optimised. This performance index is used to choose an optimal sequence of
control actions among the set of such admissible sequences, while
corresponding state sequences are produced using a system model.
The first element of the optimal sequence is applied to the system;
this control scheme defines the receding horizon control approach~\cite{RawMay09}.
When the process model is inaccurate, the modelling error must be taken
into account to guarantee the satisfaction of state constraints and
closed-loop stability properties. Tube-based MPC is a flavour of MPC
which leads to robust closed-loop stability while the accompanying
optimisation problem is computationally tractable (unlike the min-max
version of MPC~\cite{RawMay09}).
Here, we require that the state and input variables are
constrained in the sets $X\subseteq \Re^n$ and
$U\subseteq\Re^m$ respectively, both convex, compact and
contain the origin in their interior. The constraints are written
as follows, this time involving $\tilde{x}$:
\begin{subequations}\label{eq:cstr-sys}
\begin{align}
\tilde{x}_k&\in \tilde{X},\label{eq:cstr-state}\\
u_k&\in U\label{eq:cstr-input},
\end{align}
\end{subequations}
for all $k\in\N$ and where $\tilde{X}=X^{\nu}\times U^{\nu}$,
\textit{i.e.}, $\tilde{x}=(x_k, x_{k-1},\ldots, x_{k-\nu+1},
u_{k-1},\ldots, u_{k-\nu})'\in\tilde{X}$ if and only if $x_{k-i}\in X$ for $i\in\N_{[0,\nu-1]}$
and $u_{k-i}\in U$ for all $i\in \N_{[1,\nu]}$.
Typically, in MPC $\tilde{X}$ and $U$ can be polytopes
or ellipsoids, but for our analysis no particular assumptions on $X$
and $U$ need to be imposed.
The fractional-order system is controlled
by an input $u$ which is computed according to
\begin{align}\label{eq:feedback-policy}
u_k=v_k+Ke_k,
\end{align}
where $v_k$ is a control action computed by the tube-based
MPC controller and $e_k$ is defined as the deviation between
the actual system state and the response of the nominal system,
that is $e_{k}=\tilde{x}_{k}-\tilde{z}_{k}$.
In particular, the nominal dynamics in terms of the
nominal state $\tilde{z}_k$ with input $v_k$ is
\begin{align}\label{eq:sys-nominal}
\tilde{z}_{k}=A\tilde{z}_{k-1}+Bv_{k-1}.
\end{align}
Matrix $K$ in~\eqref{eq:feedback-policy} is chosen so that
the matrix $A_K=A+BK$ is strongly stable. For $k\in \N$ let
\begin{align}
S_k^\nu=\bigoplus_{i=0}^{k} A_K^i G D_{\nu}.
\end{align}
The set $S_{\infty}^\nu=\lim_{k\to\infty} S_k^\nu$, is well-defined (the limit
exists), is compact, and is positive invariant for the deviation dynamics
$e_{k+1}=A_K e_{k}+Gd_{k}$. In what follows, $S_\infty^\nu$
will be assumed to contain the origin in its interior. For the needs
of tube-based MPC, any over-approximation of $S_{\infty}^\nu$
may be used instead~\cite{RakKerKouMay04}.
Having chosen $\tilde{z}_{0}=\tilde{x}_{0}$, it is
$\tilde{x}_k\in \{\tilde{z}_k\}\oplus S_{\infty}^\nu$ for all
$k\in\N$. This implies that constraint~\eqref{eq:cstr-state}
is satisfied if $\tilde{z}_k\in X\ominus S_{\infty}^\nu$ and
constraint~\eqref{eq:cstr-input} is satisfied if $v_k\in U\ominus KS_{\infty}^\nu$.
These constraints will then be involved in the
formulation of the MPC problem which produces the control
actions $v_k=v_k(\tilde{z}_k)$.
The MPC problem amounts to the minimisation of a
performance index $V_N$ along an horizon of future
time instants, known as the \textit{prediction horizon},
given the state of the nominal system $\tilde{z}_{k}$ at time $k$.
Let $N$ be the prediction horizon. We use the notation
$\tilde{z}_{k+i\mid k}$ for the predicted state of the nominal
system at time $k+i$ using feedback information at time $k$.
Let $\mathbf{v}_{k}=\{v_{k+i\mid k}\}_{i\in\N_{[0,N-1]}}$
be a sequence of input values and $\{\tilde{z}_{k+i\mid k}\}_
{i\in\N_{[1,N]}}$
the corresponding predicted states obtained b
~\eqref{eq:sys-nominal}, \textit{i.e.}, it is
\begin{align}
\tilde{z}_{k+i+1\mid k}=A\tilde{z}_{k+i\mid k}+Bv_{k+i\mid k}, \text{ for } i\in\N_{[0,N-1]}
\end{align}
We introduce a performance index
$V_N:\Re^{\bar{n}}{\times} \Re^{mN}{\to} \Re_+$ given the current state of the system
$\tilde{z}_{k\mid k}=\tilde{z}_{k}$
\begin{align}
V_N(\tilde{z}_{k\mid k}, \mathbf{v}_{k}){=}
V_f(\tilde{z}_{k+N\mid k})
{+}\sum_{i=0}^{N-1}\ell(\tilde{z}_{k+i\mid k}, v_{k+i\mid k}),
\end{align}
where $\ell$ and $V_f$ are typically quadratic functions.
We assume that $\ell(z,v)=z'Qz + v'Rv$,
where $Q$ is symmetric, positive semidefinite and $R$ is
symmetric positive definite and $V_f(z) = z'Pz$,
where $P$ is symmetric and positive definite.
The following constrained optimisation problem is then solved:
\begin{align}\label{eq:mpc_optim_prob}
\mathbb{P}_N{:}\ &V_N^\star(\tilde{z}_{k})=\min_{\mathbf{v}_{k}\in
\mathcal{V}_N(\tilde{z}_{k})} V_N(\tilde{z}_{k}, \mathbf{v}_{k}),
\end{align}
where $\mathcal{V}_N(\tilde{z}_{k})$ is the set of all input sequences
$\mathbf{v}_k$ with $v_{k+i\mid k}\in U\ominus KS$ for all $ i{\in}\N_{[0,N-1]}$
so that $\tilde{z}_{k+i\mid k}\in \tilde{X}\ominus S$, for all
$i{\in}\N_{[0,N{-}1]}$ and $\tilde{z}_{k+N\mid k}\in \tilde{X}_f$ given that
$\tilde{z}_{k\mid k}=\tilde{z}_k$,
where $S$ is any over-approximation of $S_{\infty}^\nu$,
\textit{i.e.}, $S\supseteq S_{\infty}^\nu$ and
$\tilde{X}_f\subseteq \tilde{X}$
is the \textit{terminal constraints set}.
In what follows we always assume that $\tilde{X}\ominus S$ and
$U\ominus KS$ are nonempty sets with the origin in their
interior. In regard to the terminal cost function $V_f$ and
the terminal constraints set $\tilde{X}_f$ we assume the following:
\begin{assumption}\label{asum:MPC_std_assumptions}
$V_f$ and $\tilde{X}_f$ satisfy the standard
stabilising conditions in~\cite{May+00} \red{which are
(i) $\tilde{X}_f \subseteq \tilde{X},$ $0\in \tilde{X}_f$, $\tilde{X}_f$ is closed,
(ii) there is a controller $\kappa_f:\tilde{X}_f\to U$ so that
$\tilde{X}_k$ is positively invariant for the nominal system~\eqref{eq:sys-nominal} under $\kappa_f$, i.e.,
$A\tilde{x}+B\kappa_f(\tilde{x})\in \tilde{X}_f$ for all $\tilde{x}\in\tilde{X}_f$, and
(iii) $V_f$ is a local Lyapunov function in $\tilde{X}_f$ for the $\kappa_f$-controlled system.}
\end{assumption}
\begin{remark}\label{rmr:stability-conditions}
Matrix $P$ in $V_f$ is typically
chosen to be the (unique) solution of the
discrete-time algebraic Riccatti equation
$P=(A+BF)'P(A+BF)+Q+F'RF$
with $F=-(B'PB+R)^{-1}B'PA$ and $\tilde{X}_f$ to the maximal
invariant constraint admissible set for the system $\tilde{z}_{k+1}=(A+BF)\tilde{z}_{k}$.
Alternatively, one may choose
$\tilde{X}_f$ to be an ellipsoid of the form $\tilde{X}_f=
\{z:V_f(z)\leq \gamma\}$ and $\gamma>0$ is chosen so that
$\tilde{X}_f\subseteq \tilde{X}$ and $K\tilde{X}_f\subseteq U$; such
a set can be computed according to~\cite[Sec.~{8.4.2}]{BoyVan09}.
The latter is a better choice from a computational point of
view especially in high dimensional spaces although the optimisation
problem becomes a quadratically-constrained quadratic problem.
~\hfill\ensuremath{\diamond}
\end{remark}
The solution of $\mathbb{P}_N$, namely the optimiser
\begin{align}
v^\star(\tilde{z}_k)=\operatorname*{argmin}\limits_{\mathbf{v}_{k}\in
\mathcal{V}_N(\tilde{z}_{k})} V_N(\tilde{z}_{k}, \mathbf{v}_{k}),
\end{align}
defines the control law $\kappa_N(\tilde{z}_k)=v_0^\star(\tilde{z}_k)$
and leads to the closed-loop dynamics
\begin{subequations}\label{eq:composite}
\begin{align}
\tilde{x}_{k+1}&=
A\tilde{x}_k+B\rho(\tilde{z}_k,\tilde{x}_k)+Gd_k,\\
\tilde{z}_{k+1}&=
A\tilde{z}_{k}+B\kappa_N(\tilde{z}_{k})\label{eq:nominal}.
\end{align}
\end{subequations}
Stability properties of the closed-loop system are
hereafter derived and stated with respect to the composite system~\eqref{eq:composite} with
state variable $(\tilde{x}, \tilde{z})$.
\subsection{Stabilising conditions}\label{sec:stabilisation}
In this section we study the stability properties of the
controlled closed-loop system presented previously.
Apart from the well-known stability results in robust MPC,
we prove that, under certain conditions, the controlled trajectories
of the system are asymptotically stable to the origin (see
Theorem~\ref{thm:stability-2}).
The following result, which readily follows
from~\cite[Prop.~3.15]{RawMay09}, states that the system's
state converges towards $S_{\infty}$ exponentially provided
that $S=S_\infty$ is used in the formulation of the MPC
problem.
\begin{theorem}[Rawlings \& Mayne~\cite{RawMay09}]\label{thm:stability-1}
Assume that the MPC control law $\kappa_N$ stabilises
the nominal dynamical system~\eqref{eq:nominal}.
The set $S_{\infty}\times \{0\}$ is exponentially stable for
system~\eqref{eq:composite} with region of attraction
$(Z_N\oplus S_\infty)\times Z_N$, where $Z_N$ is the
domain of $\mathcal{V}_N$, \textit{i.e.}, $Z_N=\{x:
\mathcal{V}_N(x)\neq \varnothing\}$.
\end{theorem}
In addition, the controlled trajectory of the system's
state $x_k$ and input $u_k$ satisfy constraints
\eqref{eq:cstr-sys} at all time instants $k\in\N$.
Notice that $S_\infty$ can become arbitrarily small with
an appropriate choice of $\nu$ and the system's state can
be steered this way very close to the origin, although,
in practice large values of $\nu$ should be avoided to
limit the computation complexity of optimisation problem
$\mathbb{P}_N$. In addition to Theorem~\ref{thm:stability-1},
we are going to prove that the state converges exactly to the
origin and the origin is an asymptotically stable equilibrium point
of the controlled system under certain conditions.
The stability conditions we are about to postulate are
easy to verify and can be used for the design of stabilising model
predictive controllers. Hereafter, we shall assume that there are no
derivatives acting on system's inputs, \textit{i.e.}, $r=1$, $\beta_1=0$.
The main result of this section is stated as follows:
\begin{theorem}[Asymptotic stability]\label{thm:stability-2}
Assume that $X$ is compact and balanced, Assumption~\ref{asum:MPC_std_assumptions}
is satisfied and there is an $\epsilon\in(0,1)$ so that the following condition holds:
\begin{align}\label{eq:asym-stab-condition}
\bigoplus_{j\in\N}A_{K}^j G D&\subseteq \mathcal{B}_{\epsilon}^{\bar{n}},
\end{align}
where $D$ is the set
\begin{align}\label{eq:D2}
D= \bigoplus_{i{\in}\N_{[1,l]}}\Psi_{\nu}(\alpha_i) A^\ast_i \mathcal{B}^n.
\end{align}
Assume also that there is a $\sigma>0$ so that
$S_\infty\subseteq \mathcal{B}_\sigma^{\bar{n}} \subseteq Z_N\oplus S_\infty$.
Then, the origin is an asymptotically stable equilibrium point for~\eqref{eq:composite}.
\end{theorem}
\begin{pf}
The proof can be found in the appendix.
\end{pf}
\begin{remark}
The vector space $\Re^{\bar{n}}$ can be written as the direct sum
of vector spaces $L_1,\ldots,L_\nu$, each of dimension $n$, so that
$\tilde{x}_k\in\Re^{\bar{n}}$ if and only if $x_{k-j+1}\in L_{j}$ for
$j\in\N_{[1,\nu]}$.
Assume that $S_\infty\cap L_i$ has nonempty interior in the topology
of $L_i$. Then, in Theorem~\ref{thm:stability-2} one may drop the requirement that
$S_\infty\subseteq \mathcal{B}_{\sigma}^{\bar{n}}$ by replacing the norm
$\|\cdot\|$ of $\Re^{\bar{n}}$ by the Minkowski functional of $S_\infty$, that is
\begin{align}
\mathrm{p}[S_\infty](\tilde{x})=\inf_{\lambda>0}\{\lambda S_\infty \ni \tilde{x}\}.
\end{align}
The norm-ball $\mathcal{B}_\epsilon^{\bar{n}}$ becomes
$\mathcal{B}_\epsilon^{\bar{n}} = \{x: \mathrm{p}[S_\infty](x)<\epsilon\}$
and the induced matrix norm is modified accordingly, while on
$L_i$ we replace the norm by $\mathrm{p}[S_{\infty}\cap L_i](x)$.
This is based on a useful property of $\mathrm{p}[S_{\infty}\cap L_i](x)$ which
is stated in Appendix~\ref{sec:appPropertiesPS}.~\hfill\ensuremath{\diamond}
\end{remark}
\begin{remark}
Assume that $D$ in Theorem~\ref{thm:stability-2} is a polytope (for example,
the 1-norm or the infinity-norm is used).
Using the results presented in~\cite{RakKerKouMay04}, given a tolerance $t>0$,
there is a $\beta>1$ and an $s\in\N$ so that the polytope
$$F_{\beta,s}=\beta\bigoplus_{i=0}^{s}A_K^i GD$$
be an $t$-outer approximation of
\begin{align}
F=\bigoplus_{i=0}^{\infty}A_K^i GD,
\end{align}
in the sense that $F\subseteq F_{\beta,s} \subseteq F\oplus \mathcal{B}_t^{\bar{n}}$.
Then, the stabilising condition of Theorem~\ref{thm:stability-2} is
satisfied if $F_{\beta,s}\subseteq \mathcal{B}_{\epsilon}^{\bar{n}}$ and this
condition is easier to check computationally.~\hfill\ensuremath{\diamond}
\end{remark}
\begin{remark}
Since $A_K$ is a strictly Hurwitz matrix, there is a finite $a\in\N$
so that $\|A_K^{j}\|<1$ for all $j> a$. Then $F$ can be written as
\begin{align}\label{eq:x454}
F=\bigoplus_{i=0}^{a}A_K^i GD \oplus \bigoplus_{i=0}^{\infty}A_K^{a+1+i} GD,
\end{align}
where the first term is finitely determined and in case $D$ is a polytope, it is
also a polytope. Let $\delta^\ast=\max_{d\in D}\|d\|$ (which is well-defined and finite
because $D$ is compact). The second term of $F$ in~\eqref{eq:x454} can be over-approximated
\begin{align*}
\bigoplus_{i=0}^{\infty}A_K^{a+1+i} GD
&\subseteq \bigoplus_{i=0}^{\infty}A_K^{a+1+i} G\mathcal{B}_{\delta^\ast}
\subseteq \bigoplus_{i=0}^{\infty} \mathcal{B}_{\delta^\ast\|A_K^{a+1}\|^i}\\
&\subseteq \mathcal{B}_{\delta^\ast\sum_{i=0}^{\infty}\|A_K^{a+1}\|^i}
=\mathcal{B}_{\frac{\delta^\ast}{1-\|A_K^{a+1}\|}}.
\end{align*}
This is based on the observation that for a matrix
$B\in\Re^{n\times m}$ it is $B\mathcal{B}^m\subseteq
\mathcal{B}_{\|B\|}^n$, where $\|B\|$ is the operator
norm defined in Section \mbox{\ref{sec:mathematics}}.
As a result we have that for $B_1, B_2\in\Re^{n\times m}$,
it is
$B_1\mathcal{B}^m\oplus B_2\mathcal{B}^m\subseteq \mathcal{B}^n_{\|B_1\|}\oplus
\mathcal{B}^n_{\|B_2\|} \subseteq \mathcal{B}^n_{\|B_1\|+\|B_2\|}$.
If $a$ is adequately large and/or $\delta^\ast$ is adequately small,
it will be $\frac{\delta^\ast}{1-\|A_K^{a+1}\|}<\epsilon<1$ (for some $\epsilon$)
and then we can check the following stability condition
\begin{align}
\bigoplus_{i=0}^{a}A_K^i GD \subseteq \mathcal{B}_{\epsilon-\frac{\delta^\ast}{1-\|A_K^{a+1}\|}},
\end{align}
which entails stabilising condition~\eqref{eq:asym-stab-condition} and is
easier to verify.~\hfill\ensuremath{\diamond}
\end{remark}
\subsection{Computational complexity}
In this section we discuss the computational complexity of the proposed
scheme and give some guidelines for the selection of $\nu$.
By Theorem~\ref{thm:stability-2}, an adequately large value of $\nu$
leads to the satisfaction of the stabilising conditions of the theorem.
Naturally, for a given $\alpha>0$, one would be
interested to know the minimum order of approximation $\nu_\epsilon^\alpha$ for which
$\Psi_{\nu_\epsilon^\alpha}(\alpha)<\epsilon$, where $\epsilon>0$ is a desired threshold.
With $\nu=\nu_\epsilon^\alpha$, the MPC problem one needs to solve is formulated for a
system that has $\nu_\epsilon^\alpha$ as many states as the
original fractional system. Clearly, a parsimonious selection of $\nu$ is of
major importance for a computationally tractable controller design.
The designer needs to choose $\epsilon$ in order to strike a good
balance between performance and computational cost.
Indicatively, for $\epsilon=0.05$ and $\alpha=0.7$ we need $\nu=15$,
whereas for the same $\epsilon$ and $\alpha=1.3$ we need $\nu=4$.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{f1}
\caption{Dependence of $\Psi_{\nu}(\alpha)$ on $\nu$ for various values of $\alpha$.
The smaller $\alpha$ is, the slower the convergence of $\Psi_{\nu}(\alpha)$ becomes.}
\label{fig:Psinu}
\end{figure}
\section{Numerical Example}
We apply the proposed methodology to the
fractional-order system $D^\alpha x(t) = Ax(t) + Bu(t)$
with
\begin{align}\label{eq:sys-xmpl}
A = \begin{bmatrix}
1 & 0.9\\
-0.9 & -0.2
\end{bmatrix}, B = \begin{bmatrix}
0\\
1
\end{bmatrix},
\end{align}
and $x\in\Re^2$, $u\in\Re$ and $\alpha=0.7$.
Matrix $A$ has eigenvalues $0.4\pm 0.678 i$ and the
unactuated open-loop system is unstable.
We discretise the system with sampling period $h=0.1$
and we use $\nu=20$ based on Figure~\ref{fig:Psinu}
so that $\Psi_{\nu}(\alpha)\cong 0.041$ is adequately small (leading
to an adequately small set $D_\nu$).
This way, we derive a discrete-time LTI system
of the form $\tilde{x}_{k}=A\tilde{x}_{k-1}+Bu_{k-1}$
as in Section~\ref{sec:foa}.
The system state and input are subject to the constraints
\begin{subequations}\label{eq:xmpl-constr}
\begin{align}
-\begin{bmatrix}
3\\
3
\end{bmatrix} \leq &x_k \leq
\begin{bmatrix}
3\\
3
\end{bmatrix},\\
-0.5 \leq &u_k \leq 0.5.
\end{align}
\end{subequations}
The terminal cost $V_f$ and the terminal constraints set
$\tilde{X}_f$ were computed according \red{so that
Assumption~\ref{asum:MPC_std_assumptions} is satisfied}. In particular
$\tilde{X}_f$ was chosen to be a sublevel set of $V_f$
as explained in Remark~\ref{rmr:stability-conditions}, that is
$X_f = \{x: V_f(x) \leq \gamma \}$, where $\gamma=0.015$.
The prediction horizon was chosen to be \red{$N=100$}
and the closed-loop state and input trajectories of the
controlled system are presented in Figure~\ref{fig:sim_frac_01}
starting from the initial condition \red{$x_0=(2, 0)$}.
Notice that the imposed constraints~\eqref{eq:xmpl-constr}
are satisfied at all time instants and the control action
saturates at its limit $u=0.5$. A phase portrait of the
controlled system, starting from various initial points, is
shown in Figure~\ref{fig:phase_portait} and as one can see
all trajectories converge to the origin.
In order to demonstrate the effect of $\nu$ on the system's
closed-loop behaviour, in Figure~\ref{fig:simulations}
we present simulations with fixed
prediction horizon $N=100$ and different values of $\nu$
for system~\eqref{eq:sys-xmpl} starting from the initial
state~$x_0=(2, -3)$.
\red{The average computation time for $\nu=20$ (over $150$ random (feasible) initial points $\tilde{x}_k$)
was found to be $38ms$ and the $99\%$-quantile was $43ms$ (maximum observed
runtime: $47.3ms$).
For a larger problem with $\nu=50$, the average runtime was
$108ms$ and the $99\%$-quantile was $195ms$ (max. $206ms$).
The optimisation problem was formulated using
the MATLAB toolbox YALMIP~\cite{YALMIP} and the solver MOSEK (https://www.mosek.com/).
All computations were carried out on an Intel Core i7-4510U, $4\times 2.0GHz$,
$8GB$ RAM 64-bit system running Ubuntu 14.04.
}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{f2}
\caption{Closed-loop simulations of system~\eqref{eq:sys-xmpl}
with the proposed MPC controller with $\nu=20$ and $N=100$.
State (up) and input (down) trajectories.}
\label{fig:sim_frac_01}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{f3}
\caption{Phase portrait of the closed-loop system starting from various different initial points \red{using $\nu=20$ and $N=100$}.}
\label{fig:phase_portait}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{f4}
\caption{Responses for different approximation orders $\nu$ and fixed prediction horizon $N=100$.}
\label{fig:simulations}
\end{figure}
\section{Conclusions and future work directions}
In this paper we proposed a tube-based MPC scheme for fractional
systems which guarantees the satisfaction of state and input
constraints. No assumptions on the fractional orders $\alpha_i$ were imposed
other than that they be nonnegative, so the results presented here are
valid also for non-commensurate systems.
We make use of a linear and finite-dimensional approximation of the
original dynamics and discuss how the order of approximation relates
to the computational complexity and stability properties of the resulting
controlled system.
The proposed control methodology features two important stability
properties: first, it converges exponentially fast to a convex neighbourhood
of the origin and, second, under certain conditions the origin is an
asymptotically stable equilibrium point of the controlled system.
\red{In future work we will consider the discrepancy between the discrete-time
fractional-order system and the original continuous-time system when the MPC control action
is applied by a hold element. Only recently have such problems been solved for
constrained linear time-invariant systems~\cite{SopPatSar13}.}
\begin{ack}
This work was funded by project 11$\Sigma\text{YN}$.10.1152, which is co-financed
by the EU and Greece, Operational Program ``Competitiveness
\& Entrepreneurship'', NSFR 2007-2013 in the context of GSRT National action ``Cooperation''.
\end{ack}
\bibliographystyle{plain}
|
\section{\label{sec1} Introduction}
Diffusion advection processes in turbulent environments represent both experimentally and theoretically important topic of study in the field of fluid motion \cite{Yoshizawa,Biskamp,MoninBook,McComb,Shraiman}. In this respect, the so-called Prandtl number is frequently used to compactly characterize the quantitative properties of flows under the study \cite{Biskamp,MoninBook}. For all admixture types, it is defined as the dimensionless ratio of the coefficient of kinematic viscosity to the corresponding diffusion coefficient of given admixture. For example in the case of thermal diffusivity, the corresponding (scalar) Prandtl number equals to the ratio of kinematic viscosity to the coefficient of molecular diffusivity \cite{MoninBook}. Since both the kinematic viscosity and the diffusion coefficient for given admixture are material and flow specific quantities the resulting Prandtl numbers have always to be specified at distinct conditions required to characterize the flow and are thus often found in property tables alongside other material specific properties \cite{Yoshizawa,Biskamp,Coulson,Chua}.
However, in the high Reynolds number limit the state of fully developed turbulence manifests itself by reaching effective material and flow independent values for both the kinematic viscosity and the corresponding diffusion coefficient. We commonly refer to such effective values as the turbulent viscosity coefficient and turbulent diffusion coefficient \cite{MoninBook,McComb}. Consequently, in fully developed turbulent flows the resulting values of Prandtl numbers are universal for given admixture and do not depend on microscopic nor macroscopic properties of the flow under the consideration. Usually, we refer to them as turbulent Prandtl numbers of given admixture type \cite{Yoshizawa,Biskamp,Chang,VasilevBook}.
In other words, the state of fully developed turbulence allows for studying of advection diffusion processes on a general material and flow unbiased manner \cite{MoninBook,McComb}. Moreover, it is well known that fully developed turbulent systems are well tractable for analytic investigations which would otherwise be difficult or even impossible \cite{VasilevBook,AdzhemyanBook}. Fully developed turbulent flows represent thus theoretically as well as experimentally valuable scenario for analytic studies of how different admixtures are transported within the underlying turbulent environment.
In this respect, several authors have recently analyzed the question of how tensorial nature of admixtures
under the consideration may alter the diffusion advection processes, see for example Refs.\,\cite{Jurcisin2014,Antonov2015,Antonov2015a,Jurcisin2016,Adzhemyan2005} for more details. As a starting point for the present analysis, we discuss briefly Refs.\,\cite{Jurcisin2014,Adzhemyan2005} where the aforementioned turbulent Prandtl number have been used to approach the problem. In Ref.\,\cite{Adzhemyan2005}, turbulent scalar Prandtl number has been investigated in the model of passive advection while in Ref.\,\cite{Jurcisin2014} two other models, namely the so called kinematic MHD model and a passively advected vector field within the $A = 0$ model, have been included into the analysis. As argued by authors of Ref.\,\cite{Jurcisin2014}, introduction of spatial parity violation (helicity) into the turbulent flow represents not only a more realistic physical scenario compared to the corresponding fully symmetric case but it additionally has the advantage of pronouncing different tensorial properties of the model under the study. Thus, based on the helical values of the corresponding Prandtl numbers a comparative analyses is performed in Ref.\,\cite{Jurcisin2014}. As a result, authors of Ref.\,\cite{Jurcisin2014} argue that structure of interactions exerts a more profound impact on diffusion-advection processes than the tensorial nature of the advected field itself. However, only three selected models are analyzed in Ref.\,\cite{Jurcisin2014} and strictly speaking the conclusions made by authors of Ref.\,\cite{Jurcisin2014} are merely hypotheses when extended beyond the range of the three studied models. The reason is that in Ref.\,\cite{Jurcisin2014} interactions could not be varied continuously. Nevertheless, kinematic MHD model and the aforementioned $A = 0$ model represent two special cases of the general $A$ model \cite{Jurcisin2014,Antonov2015,Antonov2015a,Jurcisin2016} with $A$ being a real parameter \cite{Arponen2009}. Thus, we may easily bring the kinematic MHD model and the $A = 0$ model of Ref.\,\cite{Jurcisin2014} onto a same footing by using the framework of the general $A$ model which allows direct description of a spectrum of different interactions by continuous variation of parameter $A$ (for details see Sec.\,\ref{sec2}). For $A = 1$ and $A = 0$ the two physically important cases of kinematic MHD and the $A = 0$ model of passively advected advection are recovered. Additionally, at $A = -1$ another important case of the so called linearized Navier-Stokes equations arises as a special case of the general $A$ model \cite{Arponen2009}. Thus, the general $A$ model represents a tool to unite several distinct but physically important cases into one single model. The advantage of such a generalization lies then in allowing for continuous variation of interaction structures which on the other hand greatly simplifies the analysis of influence of tensorial structures on diffusion-advection processes at least in the case of vector admixtures.
A step towards such an analysis has already been undertaken in Ref.\,\cite{Jurcisin2016}, however only the case of fully symmetric turbulent environment has been considered and consequently only a limited insights have been gained into the problem. The assertions made by authors of Ref.\,\cite{Jurcisin2014} could therefore not be verified in Ref.\,\cite{Jurcisin2016}. It is therefore of high interest to analyze the general $A$ model with broken spatial parity and verify the hypothesis made in Ref.\,\cite{Jurcisin2014}. Moreover, the general $A$ model has also attracted a lot of attention recently from the point of view of their scaling properties, see for example Refs.\,\cite{Antonov2015,Antonov2015a,Antonov2003,Arponen2009}. But up to date only the case of fully symmetric turbulent environment has been analyzed. It is thus of high importance to include helical (violation of spatial parity) effects into the analysis of general $A$ model. For this purpose, it is the scope of the present paper to calculate for the first time the corresponding turbulent Prandtl number for the general $A$ model in fully developed turbulent environments with broken spatial parity. The resulting turbulent Prandtl number becomes then effectively a function of helical effects via the helical parameter $\rho$ (see Sec.\ref{sec2} for the definition) as well as function of the interaction parameter A. At this place, we also note that authors of Ref.\,\cite{Jurcisin2016} varied $A$ only in the range of $-1 \leq A \leq 1$ but according to Ref.\,\cite{Arponen2009} $A$ is in principle not bound to the interval $-1 \leq A \leq 1$. We therefore extend our analysis to the all possible values of $A$ but show later that constraints on $A$ arise artificially within the approach used in the present paper.
To perform the investigations discussed above, we use the well established tools of field renormalization group (RG) technique as presented for example in Refs.\cite{VasilevBook,AdzhemyanBook,Zinn} which has widely been used in the field of fully developed turbulence without admixtures \cite{Adzhemyan1983,Adzhemyan2003,Adzhemyan2003a,Adzhemyan1988,Adzhemyan1996,Adzhemyan2005double,Adzhemyan2006double} as well as for advection diffusion processes of several admixtures including passive scalar admixture \cite{Adzhemyan1983a,Adzhemyan2005,Adzhemyan1998,Adzhemyan2001,Pagani2015,Novikov2003}, magnetic admixtures \cite{Adzhemyan1985,Adzhemyan1987} and also vector admixtures \cite{Jurcisin2014,Antonov2015,Antonov2015a,Jurcisin2016,Arponen2009,Adzhemyan2013,Arponen2010,Novikov2006}. Two loop techniques for calculation of the turbulent Prandtl number within the $A$ model used here are similar to those carried out in Ref.\,\cite{Adzhemyan2005}. The resulting helical values of turbulent Prandtl number are then analyzed to finally investigate the hypothesis raised by authors of Ref.\,\cite{Jurcisin2014}. In this respect, the context of the general $A$ model has also been used to further discuss the validity of two loop results on kinematic MHD obtained in Ref.\,\cite{Jurcisin2014}.
The paper is structured as follows. In Sec.\,\ref{sec2}, the $A$ model of passive advection of vector admixture is defined via stochastic differential equations. The emphasis is laid on the meaning of the parameter $A$ for the structure of interactions. In Sec.\,\ref{sec3}, field theoretic equivalent of stochastic differential equations of the $A$ model is introduced. The UV renormalization of the model is discussed in Sec.\,\ref{sec4} which is then concluded with the calculation of the IR stable fixed point of basic RG equations. Two loop calculation of the helical Prandtl number is presented in Sec.\,\ref{sec5} where also the helical dependence of the turbulent Prandtl number is discussed with special attention given to the influence of tensorial interaction structures on the diffusion advection processes in the $A$ model studied here. Obtained results are then briefly reviewed in Sec.\,\ref{sec6}.
\section{Model $A$ of passive vector advection with spatial parity violation}\label{sec2}
We consider a passive solenoidal vector field $\mathbf{b} \equiv \mathbf{b}(x)$ driven by a helical turbulent environment given by an incompressible velocity field $\mathbf{v} \equiv \mathbf{v}(x)$ where $x \equiv (t, \mathbf{x})$ with $t$ denoting the time variable and $\mathbf x$ the $d$ dimensional spatial position (later $d = 3$ strictly). Apparently, $\mathbf v$ and $\mathbf b$ are divergence free vector fields satisfying $\partial \, . \, \mathbf{b} = \partial \, . \, \mathbf{v} = 0$. Additionally, within a general $A$ model of passive advection the following system of stochastic equations is required:
\begin{eqnarray}
\partial_t {\mathbf b}&=&\nu_0 u_0 \triangle {\mathbf b} -({\mathbf v}\cdot {\mathbf \partial}) {\mathbf b} + A({\mathbf b}\cdot {\mathbf
\partial}) {\mathbf v} -\partial P+ {\mathbf f^{b}}, \label{BB}
\\
\partial_t {\mathbf v}&=&\nu_0 \triangle {\mathbf v} -({\mathbf v}\cdot {\mathbf \partial}) {\mathbf v} -\partial Q + {\mathbf f^v}, \label{vv}
\end{eqnarray}
where $\partial_t \equiv \partial / \partial_t$, $\partial_i \equiv \partial / \partial_{x_i}$, $\Delta \equiv \partial^2$ is the Laplace operator, $\nu_0$ is the bare viscosity coefficient, $u_0$ is the bare reciprocal Prandtl number, $P \equiv P (x)$ and $Q \equiv Q(x)$ represent the pressure fields while the stochastic terms $\mathbf f^v$, $\mathbf f^b$ and the parameter $A$ are discussed later in this section. The subscript $0$ identifies unrenormalized quantities in what follows (see Sec.\,\ref{sec4} for more details).
Let us now briefly review the physical meaning of $A$ in Eq.\,(\ref{BB}). First, we note that Galilean symmetry requires $A$ only to be real with $A \in -1, 0, 1$ attracting most of the interest \cite{Jurcisin2016,Adzhemyan2013,Arponen2009,Arponen2010}. For $A = 1$ the kinematic MHD model is recovered, $A = 0$ leads to passive advection of a vector field in turbulent environments and finally $A = -1$ represents the model of linearized Navier-Stokes equations \cite{Arponen2009}. The parameter $A$ stands in front of the so called stretching term \cite{Adzhemyan2013} and due its continuous nature it represents a measure of specific interactions allowed by Galilean symmetry. Varying $A$ thus allows to investigate a variety of passively advected vector admixtures which only differ in their properties regarding interactions. According to \cite{Arponen2009}, parameter $A$ may take any real values but due to the special cases $A \in \{-1, 0, 1\}$ it is frequently only discussed in the smallest possible continuous interval encompassing all the three models, see for example Ref.\,\cite{Jurcisin2016}. Contrary, we extend the analysis to all physically allowed values of $A$, see Sec.\,\ref{sec5} for more details which allows a straightforward discussion of influence of interactions on advection diffusion processes.
The previously undefined stochastic terms $\mathbf f^v$ and $\mathbf f^b$ introduced in Eqs.\,(1) and (2) represent sources of fluctuations for $\mathbf v$ and $\mathbf b$. For energy injection of $\mathbf b$ we assume transverse Gaussian random noise $\mathbf{f^ b} = \mathbf{f^ b} (x)$ with zero mean via the following correlator:
\begin{equation}
D_{ij}^{b}(x;0)\equiv \langle f^b_i(x)f^b_j(0)\rangle=\delta(t)C_{ij}({\mathbf |{\mathbf x}|}/L), \label{CorelFb}
\end{equation}
where $L$ is an integral scale related to the corresponding stirring of $\mathbf b$ while $C_{ij}$ is required to be finite in the limit $L \rightarrow \infty$ and for $|x| \gg L $ it should rapidly decrease, but remains otherwise unspecified in what follows. Contrary, the transverse random force per unit mass $\mathbf{f^v} = \mathbf{f^v} (x)$ simulates the injection of kinetic energy into the turbulent system on large scales and must suit the description of real infrared (IR) energy pumping. To allow the later application of RG technique we shall assume a specific, power-like form of injection as usual for fully developed turbulence within the RG approach (for more details see Refs.\,\cite{VasilevBook,AdzhemyanBook,Adzhemyan1996}). Nevertheless, although a specific form is used universality of fully developed turbulence ensures that results obtained here may easily be extended to all fully developed turbulent flows. Additionally, it allows easy generalization to environments with broken spatial parity which is performed via tensorial properties of the correlator of $\mathbf{f^ v}$. For this purpose, we prescribe the following pair correlation function with Gaussian statistics:
\begin{eqnarray}
D_{ij}^v(x;0) &\equiv& \langle f^v_i(x) f^v_j(0) \rangle = \nonumber \\ &=& \delta(t)\int \frac{d^d {\mathbf k}}{(2\pi)^d} D_0
k^{4-d-2\varepsilon} R_{ij}({\mathbf k}) e^{i {\mathbf k}\cdot {\mathbf x}}. \label{CorelFv}
\end{eqnarray}
Here, $d$ denotes the spatial dimension of the system, $\mathbf k$ is the wave number, $k$ denotes $|k|$, $D_0 ≡ g_0 \nu_0^3 > 0$ is the positive amplitude with $g_0$ being the coupling constant of the present model related to the characteristic ultraviolet (UV) momentum scale $\Lambda$ by the relation $g_0 \simeq \Lambda^{2\varepsilon}$. The term $R_{ij} (k)$ appearing in Eq.\,(\ref{CorelFv}) encodes the spatial parity violation of the underlying turbulent environment and its detailed structure is discussed separately in the text below. Finally, the parameter $\varepsilon$ is related to the exact form of energy injection at large scales and assumes value of $2$ for physically relevant infrared energy injection. However, as usual in the RG approach to the theory of critical behavior, we treat $\varepsilon$ formally as a small parameter throughout the whole RG calculations and only in the final step its physical value of $2$ is inserted \cite{VasilevBook,Zinn}.
In Eq.\,(\ref{CorelFv}), we encounter typical momentum integrations which lead to two troublesome regions, namely the IR region of low momenta and UV region of high momenta as discussed in detail in Refs.\,\cite{VasilevBook,AdzhemyanBook}. Frequently, these troublesome integration regions are avoided by directly prescribing all relevant micro- and macroscopic properties of the flow. Here, we use the universality of fully developed turbulent flows to avoid unnecessary specifications. Thus, we only demand real IR energy injection of energy via Eq.\,(\ref{CorelFv}) and neglect the exact macroscopic structure of the flow by introducing a sharp IR cut-off $k \geq m$ for integrations over $\mathbf k$ with $L$ assumed to be much bigger than $1/m$. Using sharp cut-off, IR divergences like those in Eq.\,(\ref{CorelFv}) are avoided. As already done for Eq.\,(\ref{CorelFv}), the IR cut-off is understood implicitly in the whole paper and we shall stress out its presence only at the most crucial stages of the calculation. Contrary, UV divergences and their renormalization play central role in calculations presented here.
Finally, let us now turn our attention to the projector $R_{ij}$ in Eq.\,(\ref{CorelFv}) which controls all of the properties of the spatial parity violation in the present model. In the case of fully symmetric isotropic incompressible turbulent environments the projector $R_{ij} (k)$ assumes the usual form of the ordinary transverse projector
\begin{equation}
P_{ij}({\mathbf k})= \delta_{ij} - k_i k_j/k^2, \label{ProjectorP}
\end{equation}
see Ref.\,\cite{VasilevBook} for more details. In the case of helical flows, where spatial parity is violated, we specify Eq.\,(\ref{CorelFv}) in the form of a mixture of a tensor and a pseudotensor. Assuming isotropy of the flow we may divide the projector $R_{ij}$ in Eq.\,(\ref{CorelFv}) into two parts, i.e., $R_{ij} (k) = P_{ij} (k) + H_{ij} (k)$ where $H_{ij} (k)$ also respects the transversality of present fields. The ordinary non-helical transverse projector $P_{ij}$ is thus shifted by a helical contribution $H_{ij} (k)$ given as
\begin{equation}
H_{ij}({\mathbf k})=i \rho \, \varepsilon_{ijl} k_l/k. \label{ProjectorH}
\end{equation}
Here, $\epsilon_{ijl}$ is the Levi-Civita tensor of rank $3$ and the real parameter $\rho$ satisfies $|\rho| \in \left[ 0, 1 \right]$ due to the requirement of positive definiteness of the correlation function. Obviously, $\rho = 0$ corresponds to fully symmetric (non-helical) case whereas $\rho = 1$ means that parity is fully broken. The nonzero helical contribution leads to the presence of nonzero correlations $\langle \mathbf{v} . \it{rot} \, \mathbf{v} \rangle$ in the system.
We finally conclude the section by discussing the structure of interactions in Eqs.\,(\ref{BB}) and (\ref{vv}). Obviously, according to Eq.\,(ref{vv}) admixture field $\mathbf b$ does not disturb evolution of the velocity field $\mathbf v$. In other words, velocity field $\mathbf v$ is completely detached from the influence of admixtures as required by demanding passive advection. Of course, real problems usually involve at least some small amount of mutual interaction between the flow and its admixtures. However, even in the case of active admixtures there exist regimes which correspond to the passive advection problem as seen for example in the case of MHD problem with active magnetic admixture with its so-called kinetic regime controlled by the kinetic fixed point of the RG equations (see, e.g., Ref.\,\cite{Adzhemyan1985}). Such a situation corresponds to the passive advection obtained within the present model when $A = 1$ in Eqs.\,(\ref{BB}) and (\ref{vv}). The present picture of passive advection within the $A$ model represents thus a highly interesting physical scenario.
\section{Field theoretic formulation of the model \label{sec3}}
According to the Martin-Sigia-Rose theorem \cite{Martin}, the system of stochastic differential Eqs.\,(\ref{BB}) and (\ref{vv}) is equivalent to a field theoretic model of the double set of fields $\Phi = \{ v, b, v^{\prime}, b^{\prime} \}$ where unprimed fields correspond to the original fields of Eqs.\,(\ref{BB}) and (\ref{vv}) while primed fields are auxiliary response fields \cite{VasilevBook}. The field theoretic model is then defined via Dominicis-Janssen action functional
\begin{eqnarray}
S(\Phi) &=& \frac{1}{2} \int dt_1\,d^d{\mathbf x}_1\,dt_2\,d^d{\mathbf x}_2 \nonumber \\ && \hspace{-0.5cm} \Large[ v^{\prime}_i(x_1) D^v_{ij}(x_1;x_2) v^{\prime}_j(x_2) + b^{\prime}_i(x_1) D^b_{ij}(x_1;x_2) b^{\prime}_j(x_2) \Large] \nonumber
\\
&+& \int dt\,d^d{\mathbf x} \{ {\mathbf v^{\prime}} [-\partial_t +\nu_0 \triangle - ({\mathbf v} \cdot {\mathbf \partial}) ]{\mathbf v} \nonumber
\\
&& \hspace{-0.3cm} +\, {\mathbf b^{\prime}} [-\partial_t {\mathbf b} + \nu_0 u_0 \triangle {\mathbf b} - ({\mathbf v}\cdot {\mathbf \partial}) {\mathbf b} + A ({\mathbf b}\cdot {\mathbf \partial}) {\mathbf v} ] \}, \nonumber
\\
\label{BareS} \,
\end{eqnarray}
where $x_l=(t_l,{\mathbf x}_l)$ with $l=1,2$, $D_{ij}^b$ and $D_{ij}^v$ are given in Eqs.\,(\ref{CorelFb}) and (\ref{CorelFv}), respectively, and required summations over dummy indices $i, j \in 1, 2, 3$ are implicitly assumed. Auxiliary fields and their original counterparts $\mathbf v$, $\mathbf b$ share the same tensor properties which means that all fields appearing in the present model are transverse. The pressure terms $\partial Q$ and $\partial P$ from Eqs.\, (\ref{BB}) and (\ref{vv}) respectively do not appear in action (\ref{BareS}) because transversality of auxiliary fields $\mathbf{v}^{\prime} (x)$ and $\mathbf{b}^{\prime} (x)$ allows to integrate these out of the action (\ref{BareS}) by using the method of partial integration.
\begin{figure}
\vspace{0.5cm}
\begin{center}\includegraphics[width=5cm]{fig1.pdf}\end{center}
\caption{Graphical representation of the propagators of the model.
\label{fig1}}
\end{figure}
The field theoretic model of Eq.\,(\ref{BareS}) has a form analogous to the corresponding expression of Ref.\,\cite{Jurcisin2016} but includes via $D_{ij}^v$ the more general helical situation which was not considered by authors of Ref.\,\cite{Jurcisin2016}. In the frequency-momentum representation the following set of bare propagators is obtained:
\begin{eqnarray}
\langle b_i^{\prime} b_j \rangle_{0} = \langle b_i b_j^{\prime} \rangle_{0}^* & = & \frac{P_{ij}({\mathbf k})}{i\omega+\nu_{0} u_0 k^{2}},\label{Propagator_Bb}
\\
\langle v_i^{\prime} v_j \rangle_{0} = \langle v_i v_j^{\prime} \rangle_{0}^* & = & \frac{P_{ij}({\mathbf k})}{i\omega+\nu_{0} k^{2}},\label{Propagator_Vv}
\\
\langle b_i b_j\rangle_{0} & = & \frac{C_{ij}({\mathbf k})}{|-i\omega+\nu_{0} u_0 k^{2}|^2},\label{Propagator_bb}
\\
\langle v_i v_j\rangle_{0} & = & \frac{g_0 \nu_0^3 k^{4-d-2\varepsilon}R_{ij}({\mathbf k})}{|-i\omega+\nu_{0} k^{2}|^2},\label{Propagator_vv}
\end{eqnarray}
with helical effects already appearing in the propagator (\ref{Propagator_vv}). Function $C_{ij}(k)$ is the Fourier transform of the function $C_{ij} (r/L)$ which appears in Eq.\,(\ref{CorelFb}), but remains arbitrary in the calculations that follow. Propagators are represented as usual by dashed and full lines, where dashed lines involve velocity type of fields and full lines represent vector admixture type fields. Auxiliary fields are denoted using a slash in the corresponding propagators as shown in Fig.\,\ref{fig1} \cite{VasilevBook}.
Field theoretic formulation of the $A$ model contains also two different triple interaction vertices, namely $b_i^{\prime}(-v_{j}\partial_{j}b_i+ A b_j\partial_j v_i)=b_i^{\prime}v_{j}V_{ijl} b_l$ and $-v_i^{\prime}v_{j}\partial_{j} v_i=v_i^{\prime}v_{j}W_{ijl} v_l/2$. In the momentum-frequency representation, $V_{ijl}=i(k_j\delta_{il}-A k_l \delta_{ij})$ while $W_{ijl}=i(k_l \delta_{ij}+k_j\delta_{il})$. In both cases, momentum $\mathbf k$ is flowing into the vertices via the auxiliary fields ${\mathbf b^{\prime}}$ and ${\mathbf v^{\prime}}$, respectively. In the end, let us also briefly remind that formulation of the stochastic problem given by Eqs.\,(\ref{BB})-(\ref{vv}) through the field theoretic model with the action functional (\ref{BareS}) allows one to use the well-defined field theoretic means, e.g., the RG technique, to analyze the problem \cite{VasilevBook,Collins}.
\begin{figure} [t]
\vspace{0.5cm}
\begin{center}\includegraphics[width=9.cm]{fig2.pdf}\end{center}
\caption{Two interaction vertices of the $A$ model. The $W_{ijk}$ type of vertex involves only velocity type of fields $\mathbf{v}$ and $\mathbf{v^{\prime}}$ with $W_{ijl}=i(k_l \delta_{ij}+k_j\delta_{il})$. The second interaction vertex $V_{ijk}$ is the only basic diagrammatic object of the corresponding Feynman rules for present model which contains $A$ dependent contribution in the form of $V_{ijl}=i(k_j\delta_{il}-A k_l \delta_{ij})$. \label{fig2}}
\end{figure}
\section{Renormalization group analysis} \label{sec4}
The RG analysis performed here requires to determine all relevant UV divergences in the present model. Therefore, we employ the analysis of canonical dimensions which allows to identify all objects (graphs) containing the so called superficial UV divergences as they turn out to be the only relevant divergences left for the subsequent RG analysis in the present paper. For details, see Refs.\,\cite{VasilevBook,AdzhemyanBook,Zinn}.
Since the present $A$ model belongs to the class of the so called two scale models \cite{VasilevBook,AdzhemyanBook,Adzhemyan1996}, an arbitrary quantity $Q$ has a canonical dimension $d_Q = d^k_Q + d^{\omega}_Q$, where $d_Q^k$ corresponds to the canonical dimension of $Q$ connected with the momentum scale and $d^{\omega}_Q$ corresponds to the frequency scale. Our general helical model differs from the simple model studied in Ref.\,\cite{Jurcisin2016} by inclusion of $\rho$ dependent terms which encode helical effects of turbulent environments with broken parity. Therefore, non-helical results of Ref.\,\cite{Jurcisin2016} have to be carefully reexamined for the present model. Nevertheless, in the limit of $\rho \rightarrow 0$, the general helical $A$ model has to give the same results as its non-helical counterpart of Ref.\,\cite{Jurcisin2016}. A straightforward calculation of canonical dimensions in the present model shows that $d^k_{\rho} = d^{\omega}_{\rho} = 0$ while all the other remaining quantities posses canonical dimensions as in Ref.\,\cite{Jurcisin2016}.
In conclusion, analysis of canonical dimensions shows that the helical $A$ model posses dimensionless coupling constant $g_0$ at $\varepsilon = 0$. The present model is thus logarithmic at $\varepsilon = 0$ which means that in the framework of minimal subtraction scheme, as used in what follows, all possible UV divergences are of the form of poles in $\varepsilon$ \cite{Zinn,Collins}. Then, using the general expression for the total canonical dimension of an arbitrary 1-irreducible Green's function $\langle \Phi \ldots \Phi \rangle_{1-ir}$, which plays the role of the formal index of the UV divergence, together with the symmetry properties of the model, one finds that for physical dimension $d=3$, the superficial UV divergences are present only in the 1-irreducible Green's functions $\langle v_i^{\prime} v_j \rangle_{1-ir}$ and$\langle b_i^{\prime} b_j \rangle_{1-ir}$. Thus, all divergences can be removed by counterterms of the forms $\mathbf{v}^{\prime} \Delta \mathbf{v}$ or $\mathbf{v}^{\prime} \Delta \mathbf{v}$ which leads to multiplicative renormalization of the parameters $g_0$, $u_0$, and $\nu_0$ via renormalization constants $Z_i = Z_i (g, u; d, \rho; \varepsilon)$ as
\begin{equation}
\nu_{0}=\nu Z_{\nu}, \quad g_{0}=g\mu^{2\varepsilon}Z_{g}, \quad u_{0}=u Z_{u},\label{RenormVGU}
\end{equation}
where the dimensionless parameters $g,u$, and $\nu$ are the renormalized counterparts of the corresponding bare ones and $\mu$ is the renormalization mass required for dimensional regularization as used in the present paper. Quantities $Z_i = Z_i (g, u; d, \rho; \varepsilon)$ contain poles in $\varepsilon$.
However, there exist one additional problem when passing from the non-helical to the general parity broken $A$ model. Strictly speaking, the above conclusions are completely true only in the non-helical case. In the general case ($0 < |\rho| \leq 1$), linear divergences in the form of $\mathbf{b^{\prime}} . rot \, \mathbf{b}$ appear in the 1-irreducible Green’s function $\langle b^{\prime}_i b_j \rangle_{1-ir}$, see Ref.\,\cite{Adzhemyan1987} for more details. Removing them multiplicatively, corresponding linear terms would have to be introduced into the action functional. On the other hand, such new terms would lead to the instability causing the exponential growth in time of the response function $\langle b^{\prime}_i b_j \rangle$. A correct treatment inherently requires a genuine interplay between the underlying helical velocity field and its admixtures which is beyond the scope of passive advection $A$ model. Therefore, we shall leave the problem of the linear divergences untouched in the present paper and concentrate only on the problem of the existence and stability of the IR scaling regime, which can be studied without considering the linear divergences as already done for similar problem for example in Ref.\,\cite{Jurcisin2014}. However, we stress out that the full problem can only be solved when $A$ model with active admixtures is considered which should be the next logical in continuing the present analysis to more complicated systems.
Bearing the problem of linear $\rho$ divergences in mind, we continue the RG analysis by writing the renormalized action functional as
\begin{eqnarray}
S(\Phi) &=& \frac{1}{2} \int dt_1\,d^d{\mathbf x}_1\,dt_2\,d^d{\mathbf x}_2 \nonumber \\ && \hspace{-0.5cm} \Large[ v^{\prime}_i(x_1) D^v_{ij}(x_1;x_2) v^{\prime}_j(x_2) + b^{\prime}_i(x_1) D^b_{ij}(x_1;x_2) b^{\prime}_j(x_2) \Large] \nonumber
\\
&+& \int dt\,d^d{\mathbf x} \{ {\mathbf v^{\prime}} [-\partial_t +\nu Z_1 \triangle - ({\mathbf v} \cdot {\mathbf \partial}) ]{\mathbf v} \nonumber \\ && \hspace{-0.3cm} +\, {\mathbf b^{\prime}} [-\partial_t {\mathbf b} + \nu u Z_2 \triangle {\mathbf b} - ({\mathbf v}\cdot {\mathbf \partial}) {\mathbf b} + A ({\mathbf b}\cdot {\mathbf \partial}) {\mathbf v} ] \}, \nonumber
\\
\label{BareSr}
\end{eqnarray}
with $Z_1$ and $Z_2$ being the renormalization constants connected with the previously defined renormalization constants $Z_i = Z_i (g, u; d, \rho; \varepsilon)$ with $i \in \nu, g, \mu$ via equations
\begin{equation}
Z_{\nu}=Z_{1},\quad Z_{g}=Z_{1}^{-3},\quad Z_{u}=Z_2
Z_{1}^{-1}.\label{Renorm12}
\end{equation}
Each of the renormalization constants $Z_ 1$ and $Z_2$ corresponds to a different class of Feynman diagrams (as discussed below) but they share an analogous structure within the MS scheme: the $n$-th order of perturbation theory corresponds to $n$-th power of $g$ with the corresponding expansion coefficient containing a pole in $\varepsilon$ of multiplicity $n$ and less, i. e.:
\begin{eqnarray}
Z_{1}(g;d,\rho;\varepsilon) &=& 1+\sum_{n=1}^{\infty} g^n \sum_{j=1}^n \frac{z^{(1)}_{nj}(d,\rho)}{\varepsilon^j}, \label{Z1expansion}
\\
Z_{2}(g,u;d,\rho;\varepsilon) &=& 1+\sum_{n=1}^{\infty} g^n \sum_{j=1}^n \frac{z^{(2)}_{nj}(u,d,\rho)}{\varepsilon^j}, \label{Z2expansion}
\end{eqnarray}
where we defined $\varepsilon$ independent terms $z^{(1)}_{nj}(d,\rho)$ and $z^{(2)}_{nj}(u,d,\rho)$ and explicitly divided them by corresponding poles over $\varepsilon$. Using the last expressions with renormalized variables inserted leads to divergence free 1-irreducible Green's functions $\langle v_i^{\prime} v_j \rangle_{1-ir}$ and $\langle b_i^{\prime} b_j \rangle_{1-ir}$. Moreover, 1-irreducible Green's functions $\langle v_i^{\prime} v_j \rangle_{1-ir}$ and $\langle b_i^{\prime} b_j \rangle_{1-ir}$ are associated with the corresponding self-energy operators $\Sigma^{v^{\prime}v}$ and $\Sigma^{b^{\prime} b}$ by the Dyson equations which in frequency-momentum representation read
\begin{eqnarray}
\langle v_i^{\prime}v_j \rangle_{1-ir}&=&[ \, \, \, i\omega \, \, \, -\nu_0 p^2 + \Sigma^{v^{\prime}v}(\omega, p)] P_{ij}({\mathbf p}),\label{Dyson1}
\\
\langle b_i^{\prime}b_j \rangle_{1-ir}&=&[i\omega-\nu_0 u_0 p^2+ \Sigma^{b^{\prime}b}(\omega, p)]P_{ij}({\mathbf p}).\label{Dyson2}
\end{eqnarray}
Thus, substitution of $e_{0}= e \mu^{d_e} Z_{e}$ for $e=\{g,u,\nu \}$ is required to lead to UV convergent Eqs.\,(\ref{Dyson1}) and (\ref{Dyson2}) which in turn determine the renormalization constants $Z_1$ and $Z_2$ up to an UV finite contribution. However, by choosing the minimal subtraction (MS) scheme in what follows we require all renormalization constants have the form of 1 + \textit{poles in $\varepsilon$}. In the end, one gets explicit expressions for coefficients $z_{nj}^{(i)}$, $i = 1, 2$ in Eqs.\,(\ref{Z1expansion}) and (\ref{Z2expansion}) in the corresponding order of the perturbation theory. As explained earlier, only logarithmic divergences are considered within the general $A$ model of passive advection and possible linear divergences in $\rho$ remain untreated.
The aim of the present paper consists of deriving two-loop perturbative results for the $A$ model with helical effects included via proper definition of Eq.\,(\ref{CorelFv}). Since in the limit $\rho \rightarrow 0$ the less general non-helical model of Ref.\,\cite{Jurcisin2016} is recovered, all non-helical results of Ref.\,\cite{Jurcisin2016} have to be reproduced here. Moreover, all quantities depending exclusively on velocity field $\mathbf v$ follow only from stochastic Navier-Stokes equation (\ref{vv}) and the correlator (\ref{CorelFv}). In Refs.\,\cite{Jurcisin2014,Adzhemyan1988}, exactly the same conditions have been imposed on velocity type of fields $\mathbf{v}$ and $\mathbf{v^{\prime}}$ in two loop calculations of the given model. Consequently, the corresponding quantities depending exclusively on velocity type of fields in the present paper model have to equal those obtained in Refs.\,\cite{Jurcisin2014,Adzhemyan1988}. Taking together, $Z_1$ in the present model must be the same as in Ref.\,\cite{Adzhemyan1988} while non-helical values of $Z_2$ in the generalized helical $A$ model must reproduce results of Ref.\,\cite{Jurcisin2016}. Thus, before generalizing the approach of Refs.\,\cite{Jurcisin2016,Adzhemyan2005} to the more general $A$ model with helical contributions, we review results of Refs.\,\cite{Jurcisin2014,Jurcisin2016,Adzhemyan2005} which are relevant for the present paper.
Let us start with coefficients related to the underlying turbulent environment given by $\mathbf v$ which comprise the renormalization coefficient $Z_1$. As stated above, the present model and the model under the study in Refs.\,\cite{Jurcisin2014,Adzhemyan1988} have the same renormalization constant $Z_1$. Its one-loop expansion coefficient $z_{11}^{(1)}$ therefore reads
\begin{equation}
z^{(1)}_{11}=-\frac{S_d}{(2\pi)^d}\frac{(d-1)}{8(d+2)}, \label{z1_11}
\end{equation}
where $S_d$ is the surface area of the $d$-dimensional unit sphere defined as $S_d \equiv 2 \pi^{d/2} / \Gamma(d/2)$ with $\Gamma(x)$ being the standard Euler's Gamma function. Thus, no helical contributions at one-loop level emerge for quantities involving only velocity type of fields $\mathbf{v}$ and $\mathbf{v^{\prime}}$. The two loop order coefficient $z^{(1)}_{22}$ is in Ref.\,\cite{Adzhemyan1988} shown to satisfy
\begin{equation}
z^{(1)}_{22}=-\left(z^{(1)}_{11}\right)^2, \label{z1_22}
\end{equation}
Consequently, $z^{(1)}_{22}$ is actually also $\rho$ independent. Thus, only the remaining coefficient $z^{(1)}_{21}$ contains helical contributions to $Z_1$ . Nevertheless, the corresponding expression from Ref.\,\cite{Adzhemyan2005} is rather huge and we shall not reprint it here.
\begin{figure}
\vspace{0.5cm}
\begin{center}\includegraphics[width=8.5cm]{fig3.pdf}\end{center}
\caption{One-loop and two-loop diagrams that contribute to the self-energy operator $\Sigma^{b^{\prime}b} (\omega, p)$ in Eq.\, (\ref{Dyson2}).
\label{fig3}}
\end{figure}
Let us now reexamine the calculations of $Z_2$ done by authors of Ref.\,\cite{Jurcisin2016} with special attention given to the extension of the procedure to the more general helical $A$ model of passive advection as considered here. For this purpose we shall analyze the structure of the self-energy operator $\Sigma^{b^{\prime} b}$ in the Dyson equation (\ref{Dyson2}). In the two loop order, $\Sigma^{b^{\prime} b}$ equals the sum of singular parts of nine one-irreducible Feynman diagrams as shown in Fig.\,{\ref{fig3}}. Using the notation of Ref.\,\cite{Jurcisin2016} for the sake of easier comparison, we write down the two-loop approximation of $\Sigma^{b^{\prime} b}$ as
\begin{equation}
\Sigma^{b^{\prime} b} = \Gamma^{(1)} + \Gamma^{(2)} = \Gamma^{(1)} + \sum\limits_{l=1}^{8} s_l \Gamma^{(2)}_l \label{Dyson2structure}
\end{equation}
where $\Gamma(1)$ represents the single one-loop diagram shown in Fig.\,\ref{fig3} and $\Gamma^{(2)}$ represents the sum of eight two-loop diagrams shown in Fig.\,\ref{fig3}. Terms $s_l$, $l = 1, \ldots, 8$ denote the corresponding symmetry factors which equal $1$ for all diagrams except of the fourth with $s_4 = 1/2$.
The single one loop diagram of Fig.\,\ref{fig3} apparently does not include the propagator $\langle v_i v_j \rangle_0$ which is the only diagrammatic object that contains helical contributions. The corresponding coefficient $z_{11}^{(2)}$ that follows from the $\Gamma^{(1)}$ contribution is thus actually also $\rho$ independent. Since all non-helical quantities in the present helical $A$ model must reproduce the corresponding values of Ref.\,\cite{Jurcisin2016} the following $z_{11}^{(2)}$ expansion coefficient must be obtained (as verified also by direct calculation):
\begin{eqnarray}
z^{(2)}_{11} &=& -\frac{S_d}{(2\pi)^d} \nonumber
\\
&\times&
\frac{ (d^2-3) (u+1) + A \left[ d + u (d-2) \right] + A^2 (1 + 3u) }{4d(d+2)u(u+1)^2 } \nonumber
\\ \label{z2_11}
\end{eqnarray}
which in the case of $A = -1$, a special case focused later on in the paper, simplifies to
\begin{equation}
z^{(2)}_{11}=-\frac{S_d}{(2\pi)^d}\frac{(2u-1) +d(d-1)(u+1) }{4 d (d+2) u (u+1)^2}.
\label{z2_11a}
\end{equation}
Let us now analyze the contributions to $\Gamma^{(2)}$ which determine $z^{(2)}_{22} (d, \rho)$ and $z^{(2)}_{21} (d, \rho)$. As already stated, there are eight two loop diagrams contributing to $\Gamma^{(2)}$. After a quick inspection we notice that each of the diagrams contains two $ \langle v_i v_j \rangle_0$ propagators which are linearly dependent on helicity parameter $\rho$. Thus, all two loop diagrams can depend only quadratically on $\rho$ (linear dependencies are not relevant for present calculations and are dropped systematically). Thus, using notation equivalent to that of Ref.\,\cite{Jurcisin2016} we may write the divergent part of $\Gamma^{(2)}$ in the following form:
\begin{eqnarray}
\Gamma^{(2)}&=&\frac{g^2\nu\,p^2\,S_d}{16(2\pi)^{2d}} \left(\frac{\mu}{m}\right)^{4\varepsilon} \frac{1}{\varepsilon}
\nonumber
\\
&\times& \Bigg\{ \frac{S_d}{\varepsilon} C^{\rho} + B^{(0)} + \rho^2 \delta_{3d} B^{(\rho)} \Bigg\}, \label{Gamma2structure}
\end{eqnarray}
where $C^{\rho}$, $B^{(0)}$ and $B^{(\rho)}$ are for now on undetermined. We note that $d, g, p, \mu, u, m$ dependent factors in Eq.\,(\ref{Gamma2structure}) could principally by absorbed into $C^{\rho}$, $B^{(0)}$ and $B^{(\rho)}$, but in order to comply with the notation of Ref.\,\cite{Jurcisin2016} the specific form of Eq.\,(\ref{Gamma2structure}) is used. Since by definition, $B^{(0)}$ encodes the non-helical contributions of the corresponding diagrams we notice that it must yield the same result as obtained in Ref.\,\cite{Jurcisin2016}. However, $B^{(0)}$ was not explicitly introduced in Ref.\,\cite{Jurcisin2016} but it may easily be expressed via following equation:
\begin{eqnarray}
B^{(0)} &=& S_{d-1} \int_0^1 d x\,\, (1-x^2)^{(d-1)/2} \,\, B, \label{Bzero}
\end{eqnarray}
where variable $x$ denotes the cosine of the angle between two independent loop momenta $\mathbf k$ and $\mathbf q$ of the two-loop diagrams, i.e., $x ={\mathbf k}.{\mathbf q}/|k|/|q|$ and $B$ is a function explicitly introduced by authors of Ref.\,\cite{Jurcisin2016}. Nevertheless, $B$ is a complicated function of $u$ and $A$ as shown in Appendix of Ref.\,\cite{Jurcisin2016} and shall not be reprinted here. We merely notice that within the scope of the present calculations we have determined $B^{(0)}$ directly by methods discussed later in connection with helical contributions in the present model. We state in advance that the special non-helical values of Ref.\,\cite{Jurcisin2016}, expressed via Eq.\,(\ref{Bzero}) above, have been confirmed to hold within the present general helical $A$ model. On the other hand, the expression $C^{\rho}$ is directly related to the second order pole coefficient of $Z_2$, namely to $z_{11}^{(2)} (d, \rho)$. Although, we denoted this contribution with superscript $\rho$, in reality it must be independent of helical contributions when divergences linear in $\rho$ are left untouched as done in the present paper. The reason for vanishing of the possible $\rho$ dependence lies in the one-loop order of the present generalized $A$ model which is completely free of any helical effects. Consequently, second order $\varepsilon$ pole contributions to $\Gamma^{(2)}$ have to remain also $\rho$ independent. Particularly, it means that superscript $\rho$ in $C^{\rho}$ may be dropped, i. e., $C^{\rho} \equiv C$. Because $\rho$ dependencies are not present in $C^{\rho} \equiv C$ it must equal to Eq.(32) of Ref.\,\cite{Jurcisin2016} yielding thus the corresponding $z_{22}^{(2)} (d, \rho)$ actually also $\rho$ independent:
\begin{equation}
z^{(2)}_{22} (d,\rho) = z^{(2)}_{22} (d) = -\frac{S_d^2}{(2\pi)^{2d}} \frac{C}{16u}. \label{z2_22}
\end{equation}
The coefficient $C^{\rho} \equiv C$ may be calculated directly. At this place, we only review its form and postpone the details of calculation for later on. In accordance to Ref.\,\cite{Jurcisin2016} and to calculations performed within the scope of the present article, $C^{\rho} \equiv C$ is a polynomial of fourth order in $A$ while corresponding coefficients are complicated rational functions of $d$ and $u$ and shall not be reprinted here, for details see Eq.\,(32) in Ref.\,\cite{Jurcisin2016}.
Taking together, in Eqs.\,(\ref{z1_11})-(\ref{z2_22}) we have briefly discussed results common among the present model and models of Refs.\,\cite{Jurcisin2014,Jurcisin2016}. Passing to our generalized helical $A$ model requires now an explicit calculation of helical contributions to $\Gamma^{(2)}$ . Now, we stress out that although $B^{(\rho)}$ is calculated with explicit $d$ dependence, the helical contributions make only sense for $d = 3$ as already stated several times and made explicit by insertion of Kronecker delta $\delta_{d3}$ into the Eq.\,(\ref{Gamma2structure}).
However, before going further, let us now explain the general character of $A$ dependencies in expressions $C^{\rho} \equiv C$, $B^{(0)}$ and $B^{(\rho)}$ without considering the details of the corresponding calculations. According to Fig.\,\ref{fig3} and Eqs.(\ref{Dyson2structure}) and (\ref{Gamma2structure}), all of the discussed expressions are connected with diagrams $\Gamma^{(1)}$ or $\Gamma^{(2)}_l$ with $l = 1, \ldots 8$. Noting now that parameter $A$ appears only in the $V_{ijl}$ type vertex as a linear function we may gain direct insights into the structure of $A$ dependencies of given diagrams. To this end, imagine now a diagram with only two vertices of $V_{ijl}$ type. Since each of the vertices contains only a linear function of $A$ when necessary summations on dummy field indices are performed we get an overall dependence which may include the most a quadratical term in $A$ as a result of two linear terms in $A$ being multiplied together. In other words, the resulting diagram may therefore be only a polynomial in $A$ of order $2$ the most. The same reasoning extends also to the case when four $V_{ijl}$ type vertices appear simultaneously in given diagram. Here, the resulting polynomial must be of order four in $A$. Of coarse, since $V_{ijl}$ type vertices are of tensorial nature, summation over field indices in given diagram may lower the actual order of the polynomials in $A$ while some polynomial coefficients may also vanish completely. However, under any circumstances higher powers of $A$ may not emerge in the graphs. Using now the previous conclusions, diagrams $\Gamma_l$ with $l = 1, \ldots 8$ contain two or four $V_{ijl}$ type vertices and their sum $\Gamma^{(2)}$ must consequently be a polynomial in $A$ of the order $4$ the most. Subsequently, since $C^{\rho} \equiv C$ is proportional to the second order pole in $\varepsilon$ of $\Gamma^{(2)}$ it must also be a polynomial of the order $4$ the most. Parameters $B^{(0)}$ and $B^{(\rho)}$ are proportional to the corresponding parts of $\Gamma^{(2)}$ and must therefore also be polynomials in $A$ with order $4$ the most.
Although previous discussions determine the structure of the diagrams, only direct calculation may give us the needed coefficients of the resulting polynomials in $A$. Thus, we have to perform the calculation of the coefficients $z_{21}^{(2)} (u,d, \rho)$ and $z_{22}^{(2)} (d, \rho)$ directly. As already seen, $z_{21}^{(2)} (u,d, \rho)$ has to comply with Eq.\,(\ref{z2_21nohelicity}) in the limit $\rho \rightarrow 0$. On the other hand, since all helical properties of the generalized helical $A$ model are encoded by the term $B^{(\rho)}$ and linear $\rho$ divergences are left out in the passive advection within the $A$ model we already note that $z_{21}^{(2)} (u,d, \rho)$ contains a quadratic term in $\rho$ as the only $\rho$ dependent part. However, to correctly determine the exact term proportional to $\rho^2$ we are required to calculate $B^{(\rho)}$. For this purpose, we use the Dyson equation (\ref{Dyson2}), the relation (\ref{Dyson2structure}), and the structure of $\Gamma^{(2)}$ as given by Eq.\,(\ref{Gamma2structure}). In the end, $z_{21}^{(2)}(u,d,\rho)$ is found as (once again notation of Ref.\,\cite{Jurcisin2016} is used)
\begin{equation}
z_{21}^{(2)} (u,d, \rho) = \frac{S_d S_{d-1}}{16 u (2\pi)^{2d}} \left( B^{(0)} + \rho^2 \delta_{d3} B^{(\rho)} \right),
\label{z2_21helical}
\end{equation}
where $B^{(0)}$ and $B^{(\rho)}$ are defined via Eqs.\,(\ref{Gamma2structure}) and (\ref{Bzero}), respectively. According to Eq.\,(\ref{z2_21helical}), $B^{(\rho)}$ is given by eight two loop diagrams of Fig.\,\ref{fig3} which have a graphical representation equal to that of Refs.\,\cite{Jurcisin2014,Jurcisin2016,Adzhemyan2005} but are inherently different because of helical effects included via the propagator $\langle v_i v_j \rangle_0$. In close analogy to Eq.\,(\ref{Bzero}) we write $B^{(\rho)}$ as
\begin{eqnarray}
B^{(\rho)} &=& S_{d-1} \int_0^1 d x\,\, (1-x^2)^{(d-1)/2} \,\, \sum_{l=1}^8 s_l B^{(\rho)}_{l}, \label{Brho}
\end{eqnarray}
and define thus $B^{(\rho)}_{l}$ to be helical contributions from the corresponding parts of $\Gamma_|^{(2)}$ diagrams. Thus, as already discussed, when the limit $\rho \rightarrow 0$ is imposed on Eq.\,(\ref{z2_21helical}) the resulting value gives the $B^{(0)}$ coefficient which then in turn complies with its corresponding counterpart of Ref.\,\cite{Jurcisin2016}. On the other hand, for $\rho \neq 0$ the eight two-loop graphs contain nonzero terms which then via $B^{(\rho)}$ encode all of the helical effects investigated here. In other words, result of Ref.\,\cite{Jurcisin2016} are only a special case of the present calculations when appropriate limits are taken while for $0 < |\rho| \leq 1$ the corresponding expressions are completely unknown and require to be calculated here. For this purpose, for diagrams $\Gamma^{(2)}_l$ with $l = 2, \ldots 8$, we utilize the derivative technique outlined in Ref.\,\cite{Adzhemyan2005} whose prerequisites are fulfilled for selected diagrams with $l = 2, \ldots 8$. However, in the case of diagram $\Gamma^{(2)}_1$, only its non-helical value, a special case of the model considered here, can by evaluated using the derivative technique of Ref.\,\cite{Adzhemyan2005}. Therefore, the well established techniques outlined for example in Ref.\,\cite{VasilevBook} are used for the remaining graph $\Gamma^{(2)}_1$. Nevertheless, calculations for all graphs are quite straightforward, however they result in complicated lengthy expressions and we present them in the Appendix of the present paper.
In the end, we have to reexamine the influence of helicity on the properties of the IR scaling regime and its stability. First of all, since fields $\mathbf v$, $\mathbf v^{\prime}$, $\mathbf b$, and $\mathbf b^{\prime}$ are not renormalized the following simple relation is satisfied:
\begin{figure}
\vspace{0.5cm}
\begin{center}\includegraphics[width=8cm]{fig4.pdf}\end{center}
\caption{ (Color online) Dependence of one loop inverse turbulent Prandtl number $u^{(1)}_{*}$ on parameter $A$ in region $-2 \leq A \leq 2$. Note that for $A = -1$ one obtains $u_{*}^{(1)} = 1$. Apparently, one loop values of $u_{*}^{(1)}$ are always positive ($u_{*}^{(1)} \rightarrow \infty$ for $A \rightarrow \pm \infty$) and therefore physical for all arbitrary real $A$.
\label{fig4}}
\end{figure}
\begin{equation}
W^R(g,u,\nu,\mu,\cdots)=W(g_0,u_0,\nu_0,\cdots), \label{RenormW}
\end{equation}
It states that renormalized connected correlation functions $W^R=\langle \Phi \dots \Phi \rangle^R$ differ from their unrenormalized counterparts $W=\langle \Phi \dots \Phi \rangle$ only by the choice of variables (renormalized or unrenormalized) and in the corresponding perturbation expansion (in $g$ or $g_0$), where dots stand for other arguments which are untouched by renormalization, e.g., the helicity parameter or coordinates \cite{VasilevBook,AdzhemyanBook,Collins}. This however means that unrenormalized correlation functions are independent of the scale-setting parameter $\mu$ of dimensional regularization. Thus, applying the differential operator $\mu \partial_{\mu}$ at fixed unrenormalized parameters on both sides of Eq.\,(\ref{RenormW}) gives the basic differential RG equation of the following form \cite{VasilevBook,AdzhemyanBook}:
\begin{equation}
[\mu \partial_{\mu} + \beta_g\partial_g + \beta_u\partial_u-\gamma_{\nu} \nu
\partial_{\nu}] W^R(g,u,\nu,\mu,\cdots)=0, \label{BasicRG}
\end{equation}
where the so-called RG functions (the $\beta$ and $\gamma$ functions) are given as follows:
\begin{eqnarray}
\beta_g &\equiv& \mu \partial_{\mu}g=g(-2\varepsilon + 3\gamma_1), \label{BetaG}
\\
\beta_u &\equiv& \mu \partial_{\mu}u=u(\gamma_1-\gamma_2), \label{BetaU}
\\
\gamma_{i} &\equiv& \mu \partial_{\mu} \ln Z_i,\quad i=1,2,\label{GammaS}
\end{eqnarray}
and are based on relations among the renormalization constants (\ref{Renorm12}) together with explicit expressions of $Z_1$ and $Z_2$ given by (\ref{Z1expansion}) and (\ref{Z2expansion}), respectively. To obtain the IR asymptotic behavior of the correlation functions deep inside of the inertial interval we need to identify the coordinates ($g^∗$, $u^∗$ ) of the corresponding IR stable fixed point where $\beta_g$ and $\beta_u$ vanish, i. e.:
\begin{equation}
\beta_g(g_{*})=0, \quad \beta_u(g_{*},u_{*})=0, \label{BetaZero}
\end{equation}
where $g_{*}\neq 0$ and $u_{*}\neq 0$ in two loop approximation are required to have the form
\begin{eqnarray}
g_{*}&=&g_{*}^{(1)} \varepsilon + g_{*}^{(2)} \varepsilon^2 + O(\varepsilon^3), \label{Gexpansion} \\
u_{*}&=&u_{*}^{(1)}+u_{*}^{(2)} \varepsilon + O(\varepsilon^2). \label{Uexpansion}
\end{eqnarray}
It may be verified by direct calculation that at non-trivial fixed points the following expressions hold:
\begin{eqnarray}
g_{*}^{(1)}&=&\frac{(2\pi)^d}{S_d}\frac{8(d+2)}{3(d-1)}, \label{Gstar1}
\\
g_{*}^{(2)}&=&\frac{(2\pi)^d}{S_d}\frac{8(d+2)}{3(d-1)} \lambda, \label{Gstar2}
\\
u_{*}^{(1)}&=&\frac{1}{3a_2}\left( -2a_2 - \frac{\sqrt[3]{2} b_1}{\sqrt[3]{b_2 + b_3}} + \frac{ \sqrt[3]{b_2 + b_3}}{\sqrt[3]{2}} \right), \label{Ustar1}
\\
u_{*}^{(2)}&=&\frac{2(d+2)}{d [1+2 u^{(1)}_{*}]} \Biggl[\lambda - \frac{128 (d+2)^2}{3(d-1)^2} \mathcal{B}(u^{(1)}_{*}) \Biggr], \label{Ustar2}
\end{eqnarray}
where $\lambda$ is related to the coefficient $z_{21}^{(1)}$ in Eq.\,(\ref{Z1expansion}) as
\begin{equation}
\lambda = \frac{2}{3} \frac{(2\pi)^{2d}}{S_d^2}\left(\frac{8(d+2)}{d-1}\right)^2 z_{21}^{(1)}. \label{Lambda}
\end{equation}
Coefficient $\mathcal{B}(u^{(1)}_{*})$ is discussed in the text below. Let us now give the explicit expressions for
$a_i$ with $i \in 0, 1, 2$ and $b_i$ with $i \in 1, 2, 3$. They read:
\begin{eqnarray}
b_1 &=& a_2 \left( 3a_1 - 4a_2\right) \label{b1}
\\
b_2 &=& a_2^2 \left(-27 a_0 + 18 a_1 -16 a_2 \right) \label{b2}
\\
b_3 &=& \sqrt{4 b_1^3 + b^2_2} \label{b3}
\\
a_0 &=& 2 \left[ d^2 -3 + A(A+d)\right] \label{a0}
\\
a_1 &=& 6 (1-A^2) - 2 A (d-2) - d(d+1) \label{a1}
\\
a_2 &=& d(d-1) \label{a2}
\end{eqnarray}
The value of the coefficient $a_1$ differs from that presented in Ref.\,\cite{Jurcisin2016} where most probably a typesetting error occurred since all further results of Ref.\,\cite{Jurcisin2016} agree with corresponding results of the present helical $A$ model when the limit $\rho \rightarrow 0$ is taken. Moreover, $a_1$ presented in Ref.\,\cite{Jurcisin2016} takes the same form as the current one when the (probably misplaced) brackets are corrected.
As already mentioned, one loop results given by Eqs.\,(\ref{Gstar1}) and (\ref{Ustar1}) are free of helical contributions. Further, $g^{(2)}_{*}$ depends exclusively on the properties of the underlying velocity field which in turn means that it is common within a class of models with passively advected admixtures as discussed for example in Ref.\,\cite{Jurcisin2014}. In more detail, $g^{(2)}_{*}$ is completely determined by $\lambda$ from Eq.\,(\ref{Lambda}) and takes exactly the same value as the corresponding quantity in Refs.\,\cite{Jurcisin2016}. However, $u^{(2)}_{*}$ is model specific and known only for special choices of $A \in 0, 1 $, see Ref.\,\cite{Jurcisin2014} for more details. Here, it is expected to contain helical contributions via the quantity $\mathcal{B}(u^{(1)}_{*})$ which in turn is completely given by the coefficient $z_{21}^{(2)}$ in Eq.\,(\ref{z2_21helical}) and it obtains the following value at $u = u^{(1)}_{*}$ :
\begin{eqnarray}
\mathcal{B}(u^{(1)}_{*},\rho ) = \frac{(2\pi)^{2d}}{S^2_d} \, z^{(2)}_{21} (u_{*}^{(1)},\rho) \label{Bstar}
\end{eqnarray}
We retained the $d$ dependencies for notation purpose. However as already mentioned, only spatial dimension $d = 3$ is physically meaningful when helical effects are considered. The IR behavior of the fixed point is determined by the matrix of the first derivatives which is given as
\begin{equation}
\Omega_{ij} = \left(
\begin{array}{cc}
\partial \beta_g / \partial g & \partial \beta_g / \partial u \\
\partial \beta_u / \partial g & \partial \beta_u / \partial u \\
\end{array}
\right) \label{IRmatrix}
\end{equation}
and is evaluated for given ($g_{*}$, $u_{*}$). The present matrix has a triangular form since $\beta_g$ is independent of $u$ and thus $\partial \beta_g/ \partial u =0$. Thus, diagonal elements $\partial \beta_g / \partial g$ and $\partial \beta_u / \partial u$ correspond directly to the eigenvalues of the present matrix. Subsequently, using numerical analysis one can show that real parts of diagonal elements are positive for all values of $A$ in vicinity of $\epsilon =0$. Furthermore, we have aslo shown that including spatial parity violation shifts the values of the present matrix even further to positive values. In the end, we stress out the well known fact that $\beta$ functions of the present model are exactly given even at the one-loop order since all higher order terms cancel mutually which means that the anomalous dimensions $\gamma_1 = \gamma_2$ equal exactly $2\varepsilon/3$ at the IR stable fixed point.
\section{Helicity and the turbulent Prandtl number} \label{sec5}
As discussed in the text above, all one loop contributions to the renormalization constants $Z_1$ and $Z_2$ are free of helical contributions even when turbulent environments with broken spatial parity are considered explicitly \cite{Jurcisin2016,Adzhemyan2005}. In the previous section, we have therefore determined two loop values of renormalization constants $Z_1$ and $Z_2$ which in fact do manifest helical effects for both renormalization constants. Additionally, stable non-trivial IR fixed point is shown to exists for given $g^*$ and $u^*$ in two loop order of calculation. Therefore, one may expect (as already seen for example in Refs.\cite{Jurcisin2014,Jurcisin2016}) that two loop order is sufficient to capture the leading order helical contributions to the required turbulent Prandtl number which then of coarse correspond to the two loop order of given perturbative theory. We prove this assertion in the subsequent text by explicit determination of the corresponding values of turbulent Prandtl number for a range of values of the continuous parameter $A$. However, we show explicitly that some regions of $A$ have to be omitted when spatial parity violation is weak enough.
Two-loop calculation presented here is to a large extent based on Ref.\,\cite{Adzhemyan2005} where the turbulent Prandtl number in the simple model of passive advection of a scalar field has been calculated for completely symmetrical turbulent environment. As shown for example in Ref.\,\cite{Jurcisin2014}, the approach of Ref.\,\cite{Adzhemyan2005} may successfully by used also in helical environments. Moreover, although tensorial properties of passively advected fields considered in Refs.\,\cite{Jurcisin2014,Adzhemyan2005,Jurcisin2016} are to a large extent different from those considered in the present model, one further analogy is observed when the corresponding correlators of stochastic pumping are reviewed. Therefore, the actual calculations of two loop Prandtl number performed here are closely analogous to those of Refs.\,\cite{Jurcisin2014,Adzhemyan2005} and the resulting two loop expression for the Prandtl number is analogous to Eq.\,(33) of Ref.\,\cite{Adzhemyan2005}. Moreover, due to the passive nature of the admixtures considered here and in Refs.\,\cite{Jurcisin2014,Jurcisin2016,Adzhemyan2005}, we note that all (partial) results which depend not on the admixture field $\mathbf b$ have to be identical in Refs.\,\cite{Jurcisin2014,Jurcisin2016,Adzhemyan2005} as well as in the present model. In explicit, properties of the helical environments with given admixture type which fully correspond to the model given by differential Eqs.\,(\ref{BB}) and (\ref{vv}) are in two loop order of the perturbation theory of the corresponding field theoretic model completely encoded by the Feynman graphs of Fig.\,\ref{fig3}.
Taking together, although present calculations are analogous to that of Refs.\,\cite{Jurcisin2014,Adzhemyan2005}, all quantities inherently connected with given admixtures and their interactions have to be reexamined here. We also stress out that formula (33) of Ref.\,\cite{Adzhemyan2005} holds inside of the inertial interval and does not depend on the renormalization scheme. The details of the calculations are outlined in Ref.\,\cite{Adzhemyan2005} and we omit them consequently. The resulting two loop expression for Prandtl number is obtained as
\begin{eqnarray}
u_{eff}&=& u^{(1)}_{*} \Biggl(1 + \varepsilon \Biggl\{ \frac{1+u^{(1)}_{*}}{1+2 u^{(1)}_{*}} \Biggl[\lambda - \frac{128 (d+2)^2}{3(d-1)^2} \mathcal{B}(u^{(1)}_{*}) \Biggr] \nonumber
\\
&& + \frac{(2\pi)^d}{S_d}\frac{8 (d+2)}{3(d-1)}\left[a_{v}-a_{b}(u^{(1)}_{*})\right]\Biggl\} \Biggr), \label{InversePrandtl}
\end{eqnarray}
where $\varepsilon$ and dimension $d$ are taken to their physical values of $\varepsilon = 2$ and $d = 3$, the one loop value of turbulent Prandtl number $u_{*}^{(1)}$ has already been given in Eq.\,(\ref{Ustar1}), $\mathcal{B}(u^{(1)}_{*})$ is given in Eq.\,(\ref{Bstar}) and $\lambda$ in Eq.\,(\ref{Lambda}). The following numerical value corresponds to $\lambda$ in $d = 3$ as considered here for helical environments:
\begin{equation}
\lambda = -1.0994 - 0.0556 \times 10^{-3} \rho^2 \label{LambdaNumerical}
\end{equation}
which is the same as in Ref.\,\cite{Jurcisin2014} since $\lambda$ in independent of the admixture type for passive advection. The remaining parameters $a_v$ and $a_b$ which enter Eq.\,(\ref{InversePrandtl}) are discussed in the text below. Let us first notice that $a_v$ and $a_b$ represent the finite parts of one-loop diagrams with two external velocity type fields $\mathbf v$, $\mathbf v^{\prime}$ and two admixture type fields $\mathbf b$, $\mathbf b^{\prime}$ respectively. Since turbulent velocity environments here and in Ref.\,\cite{Adzhemyan2005} are the same, the coefficient $a_v$ must be also the same and we shall not reproduce its analytic form here. In $d = 3$, it can however be easily evaluated numerically as
\begin{equation}
a_v = -0.047718/(2\pi^2). \label{AvNumerical}
\end{equation}
Contrary, $a_b$ is model specific and is given by the finite part of one-loop one irreducible diagram $\Gamma^{(1)}$ making it thus also $\rho$ independent. As already discussed, the present generalized helical $A$ model and the less general model introduced in Ref.\,\cite{Jurcisin2016} have all one-loop quantities including $a_b$ identical due to helical effects beeing pronounced first in the two-loop order. Since $a_b$ plays a crucial role in two-loop calculation of inverse turbulent Prandtl number that follows show it explicitly in the present paper. After straightforward calculation discussed for example in Ref.\,\cite{Adzhemyan2005}, one obtains $a_b$ in the same form as authors of Ref.\,\cite{Jurcisin2016}. It reads:
\begin{widetext}
\begin{eqnarray}
a_b(u) &=&-\frac{S_{d-1}}{2 u (d-1) (2\pi)^d}\int_0^{\infty} dk \int_{-1}^1 dx\, (1-x^2)^{\frac{d-1}{2}} \nonumber
\\
&\times& \left\{
\frac{k \left[ k^3 x A (1-A) + k^2 ( x^2 (1-A^2)+A+d-2) +2 k x (d-1)+d-1 \right]}{ (k^2+2 k x +1) \left[ (1+u)k^2+2 u k x + u \right]} \right. \nonumber
\\
&-& \left. \frac{ \theta(k-1) \left[ k A (1-A)(1+u)x +A^2(1+3u)x^2 +A (1+u-2(1+2u)x^2) +(1+u)(x^2+d-2) \right] }{k(1+u)^2} \right\} \label{Ab}
\end{eqnarray}
\end{widetext}
\begin{figure*}
\vspace{0.5cm}
\begin{center}\includegraphics[width=15cm]{fig5.pdf}\end{center}
\caption{ (Color online) Dependence of $u^{(0)}$ and $u^{(\rho)}$ on parameter $A$ shown in regions $-2 \leq A \leq 3$ and $-2 \leq A \leq 2$ respectively. Quantity $u^{(0)}$ corresponds to non-helical value of inverse Prandtl number, while $u^{(\rho)}$ represents helical contribution to the inverse Prandtl number. Points represent numerical values obtained from Eq.\,(\ref{Split}).
\label{fig5}}
\end{figure*}
\noindent with $\theta(k-1)$ being the usual Heaviside step function with $k-1$ as argument. The expression (\ref{Ab}) is be easily obtained by direct calculation of the single one loop diagram shown in Fig.\,\ref{fig3}. One further difference manifested even at one loop order lies in the already calculated value of $u_{*}^{(1)}$. Due to the tensorial interaction structures in the present model it obtains the form of Eq.\,(\ref{Ustar1}) which is of coarse different from the corresponding value obtained in Ref.\,\cite{Adzhemyan2005}. Using the expression (\ref{InversePrandtl}) with all necessary coefficients now known due to Eqs.\,(\ref{Ustar1}), (\ref{Lambda}), (\ref{Bstar}), (\ref{AvNumerical}) and (\ref{Ab}), we may proceed to the actual calculation of the Prandtl number. As already discussed, using RG techniques in theory of critical behavior requires to substitute $\varepsilon = 2$ in the final expressions as thoroughly discussed for example in Refs.\,\cite{VasilevBook,McComb}. The spatial dimension is set to $d=3$ as required by the nature of the helical problem. Inserting all necessary quantities into the Eq.\,(\ref{InversePrandtl}) we obtain its values for arbitrary $A$. In other words, we get the inverse turbulent Prandtl number $u_{eff}$ as a function of $A$ which we indicate here explicitly by denoting the corresponding values as $u_{eff} (A)$.
Since the corresponding Eqs.\,(\ref{InversePrandtl}), (\ref{Ustar1}), (\ref{Lambda}), (\ref{Bstar}), (\ref{AvNumerical}) and (\ref{Ab}) are all known in analytic form also the resulting turbulent inverse Prandtl number has an analytic form. Nevertheless, due to the complicated analytic structure of the coefficient $z^{(2)}_{12}$, the expression in Eq.\,({\ref{Bstar}}) is a complicated analytic function of model variables. Thus, the resulting analytic expression for the inverse Prandtl number is also lengthy and complicated. Consequently, we shall not show it here explicitly (all necessary coefficients for it's calculation are discussed either in the main body of the article or in its Appendix) but instead we split $u_{eff} (A)$ into its non-helical part $u^{(0)} (A)$ and its corresponding helical contribution $\rho^2 u^{(\rho)} (A)$ in the following way:
\begin{equation}
u_{eff} (A) = u^{(0)} (A) + \rho^2 u^{(\rho)} (A). \label{Split}
\end{equation}
Note that both $u^{(0)} (A)$ and $u^{(\rho)} (A)$ are defined to be independent of $\rho$ but $u^{(\rho)} (A)$ is the coefficient which stands in front of the helical contribution in Eq.\,(\ref{Split}) and encodes thus all helical effect of the present model. As before, both $u^{(0)} (A)$ and $u^{(\rho)} (A)$ are quite complicated analytic functions of model parameters and we therefore present them here via their graphical representation given in Fig.\,\ref{fig5} which is sufficient for the interpretation of the obtained results. Moreover, the corresponding numerical values are given in Tab.\,\ref{tab1} for few selected values of parameter $A$. In Fig.\,\ref{fig5}, we plot $u^{(0)} (A)$ in the region of $-2 \leq A \leq 3$ as it contains all zero points of the present function. Due to the same reasoning, $u^{(\rho)} (A)$ is plotted in a smaller region of $-2 \leq A \leq 2$. The actual turbulent Prandtl number $Pr_A$ is then given as the inverse of $u_{eff} (A)$. In explicit:
\begin{equation}
Pr_A = \frac{1}{u^{(0)} (A) + \rho^2 \, u^{(\rho)} (A)}. \label{Prandtl}
\end{equation}
However, in the immediately following text we shall rather use the corresponding values of the inverse turbulent Prandtl number as they better suit our next discussion. Afterwards, we discuss turbulent Prandtl numbers in helical $A$ model for selected values of $A$.
Let us now therefore consider non-helical part $u^{(0)} (A)$ of the inverse turbulent Prandtl. First, we note that in the range $-1 \leq A \leq 1$ non-helical values of the function $u^{(0)} (A)$ obtained here are in complete agreement with those obtained by authors of Ref.\,\cite{Jurcisin2016} for a simpler non-helical case (see Fig.\,(8) in Ref.\,\cite{Jurcisin2016} and the corresponding analytic expressions in the Appendix of the same reference). However, in Ref.\,\cite{Jurcisin2016} only the region $-1 \leq A \leq 1$ is investigated and thus the important zero points of function $u^{(0)} (A)$ have not been discussed in any way. However, problems which arise at zero points of $u^{(0)} (A)$ clearly manifest the limits of perturbative two-loop approach as used here. Physically, when $u^{(0)} (A)=0$ the effective value of the corresponding diffusion coefficient for given $A$ should be infinitesimally small which is of coarse non-physical. Nevertheless, zero points of the function $u^{(0)} (A)$ are present and located at $A = -1.723$ and $A = 2.800$ (numerical values rounded on the last digit). Consequently, by approaching the zero points of $u^{(0)}$, turbulent Prandtl numbers would obtain infinitely large values. Additionally, according to Fig.\,\ref{fig5}, inverse turbulent Prantdl number would be negative in regions $A < -1.723$ and $A > 2.800$. The effective diffusion coefficients would in such cases obtain non-physical values which clearly must be avoided.
Thus, for non-helical turbulent environments constraints $-1.723 \leq A \leq 2.800$ must be imposed on values of $A$ in the two loop order of perturbation theory. Additionally, values of $A$ close to zero points of $u^{(0)} (A)$ should also be considered only with extreme caution as the resulting turbulent Prandtl numbers tend to $+\infty$ at the border of the allowed interval. On the other hand, such a problem did apparently not occur for the corresponding one loop values as clearly demonstrated in Fig.\,\ref{fig4} in the present paper. It is therefore clear, that constraints for non-helical environments arise only in connection with the two loop order calculation used here and are therefore inherently given by the structure of perturbation theory of the $A$ model. In other words, such constraints are not inherent to values of $A$ outside of the usually studied region $-1 \leq A \leq 1$ and represent only an artifact of the perturbative approach. Such a conclusion is supported also by the special case of $A = -1$ discussed later in more detail. For now on, we stress out that all previous conclusions are completely true only in non-helical environments. Bearing in mind the constraints on $A$ in non-helical case, we also notice that $u^{(0)} (A)$ has a maximum at $A=0.7128$ (rounded on last presented number) and is quite well stable in the range of approximately $-0.5<A<1.5$ which in connection to the results on one-loop order values presented for example in Fig.\,\ref{fig4} also explains the remarkable stability of models with $A=0$ and $A= 1$ against the order of perturbation theory as already noticed in Ref.\,\cite{Jurcisin2016}. Qualitatively similar picture holds also when helical contributions are considered as discussed below.
Let us now finally turn our attention to $u^{(\rho)}$ which encodes the much needed helical contributions of our generalized helical $A$ model. Its graphical representation is given in Fig.\,\ref{fig5} and is directly connected with the inverse Prandtl number $u_{eff} (A)$ by Eq.\,(\ref{InversePrandtl}). Consequently, the sign of $u^{(\rho)}$ determines the character of helical dependence of $u_{eff} (A)$. In explicit, for positive (negative) values of $u^{(\rho)} (A)$ the corresponding inverse turbulent Prandtl number will be a monotonically growing (descending) function of helicity parameter $\rho$. The zero points of $u^{(\rho)} (A)$ turn out therefore to represent very important special cases of the general $A$ model. Their location is easily determined numerically based on the previous analysis with resulting values being $-1.516$, $-1.000$, $0.325$ and $0.912$ (numbers rounded at the last presented digit).
Furthermore, inserting values of functions $u^{(0)}(A)$ and $u^{(\rho)}(A)$ into Eq.\,(\ref{InversePrandtl}) one may easily calculate the inverse turbulent Prandtl number $u_{eff}(A)$ as a function of $A$ for selected values of $\rho$. The resulting values are presented in Fig.\,\ref{fig6} and show highly interesting behavior. In non-helical case, the resulting turbulent Prandtl numbers have been shown to obtain unphysical values in restricted intervals $A < -1.723$ and $A > 2.800$. However, $u^{(\rho)}(A)$ is according to Fig.\,\ref{fig5} in both restricted intervals not only positive but it also evidently satisfies $u^{(\rho)}(A)>|u^{(0)}(A)|$. Therefore, when exceeding some critical value of helicity parameter $\rho$ for given value of $A$ the corresponding inverse turbulent Prandtl number $u_{eff}=u^{(0)} +\rho^2 u^{(\rho)}$ must get positive. In other words, when parity violation is strong enough, the resulting inverse turbulent Prandtl number obtains always positive values. Thus, introducing parity violation into the turbulent system improves perturbative series for the present model as shown explicitly in Fig.\,\ref{fig6}. In this respect, we also notice that increasing $\rho$ from $0$ up to $\rho\approx0.5$ enlarges the region of physically allowed values of $A$. However, the allowed region of $A$ grows according to Fig.\,\ref{fig6} infinitely when helicity parameter $\rho$ is increased further. Strikingly, it is not required to reach the maximum possible violation of parity ($|\rho|=1$) to remove the constraints on $A$. Contrary, by exceeding a critical value of $\rho= 0.749$ (rounded on the last presented digit) we remove any constraints on $A$ completely. In other words, beyond the critical value of $\rho =0.749$ all inverse turbulent Prandtl numbers are positive and thus physical. Consequently, exceeding the threshold of $\rho=0.749$ stabilizes the diffusion advection processes in the general $A$ model to a large extent. Calculations within the two loop order of the corresponding perturbative theory are then well defined which even further our hypothesis regarding the artificiality of constraints imposed on values of $A$ in non-helical environments. The interplay between the interaction parameter $A$ and parameter $\rho$ describing the amount of spatial parity violation is thus proven to be highly non-trivial.
\begin{figure}[t]
\begin{center}\includegraphics[width=7cm]{fig6.pdf}\end{center}
\caption{ (Color online) Inverse turbulent Prandtl numbers $u_{eff}$ as function of $A$ shown in the range of $-2 \leq A \leq 3$ for selected values of helical parameter $\rho$. Shown are values of $u_{eff}$ for $\rho=0$ (black), $\rho=0.4$ (blue), $\rho=0.5$ (magenta), $\rho=0.7$ (red) and $\rho=1$ (orange). \label{fig6} }
\end{figure}
Additionally, in Fig.\,\ref{fig6} we may identify values of $A$ for which the helical dependence of inverse turbulent Prandtl number is relatively small. Such regions are all connected with the regions of negative values of $u^{(\rho)}$ and correspond therefore to the union of interval $-1.516 \leq A \leq -1.000$ with $0.325 \leq A \leq 0.912$. Interestingly, we notice that two special cases $A=-1$ and $A=1$ lie either directly in such regions ($A=-1$ case) or are located in a close vicinity of these ($A=1$ case). First, let us discuss the case of linearized helical NS equations with $A = -1$ which up to date has not been investigated in any way. According to the performed numerical analysis of Eq.\,(50), $u^{(\rho)}$ is less than $10^{-8}$ at $A = -1$ which in limits of accuracy means that $u^{(\rho)}$ is actually equivalent to zero and consequently $A = -1$ corresponds directly to the zero point of $u^{(\rho)} (A)$. We stress out that this is not just a trivial influence of vanishing of all helical terms in two-loop diagrams $\Gamma_l^{(2)}$ with $l=1, \ldots, 8$. In fact, separately each diagram contains corresponding helical terms which however mutually cancel each other when all diagrams are summed up together as required in deriving of $\Sigma^{b^{\prime}b}$. As a consequence, at $A = -1$ the properties of the flow are completely independent of spatial parity violation of the underlying fully developed turbulent velocity flow. Moreover, this result is most probably independent of perturbation order as suggested by $u^{(0)}$ being exactly one (within the accuracy of the present numerical analysis) at both the first and the second order of the corresponding perturbation order. A similar hypothesis has already been stated by authors of Refs.\,\cite{Jurcisin2016} for non-helical values. Here, we however demonstrate that such a behavior persist even in helical environments.
In this respect, it is also worth to mention that for $A = 1.038$ (value obtained numerically and rounded on the last presented digit) one and two loop values of non-helical inverse Prandtl number do also coincide, a result which was not observed in Ref.\,\cite{Jurcisin2016} due to constraining the analysis only at $-1 \leq A \leq 1$. However, unlike for the $A = -1$ case, helical effects are present quite significantly for the $A = 1.038$ case (as later discussed more closely, the difference between the non-zero value of the turbulent Prandtl number and its minimal value at $|\rho|=1$ is around $7 \%$.) which means that the model of linearized Navier-Stokes equations corresponding to the $A = -1$ case in the present model has unique features. The remaining three zero points of $u^{(\rho)}(A)$, namely $A \in \{ -1.516, 0.325, 0.912 \}$, do not show the same behavior. Instead, their one and two loop values differ significantly. In other words, although the remaining three zero points of $u^{(\rho)} (A)$ also lead to models stable against helical effects in two loop order, there is no indication that higher order of perturbation theory preserve location of the zero points for the analog of $u^{(\rho)}(A)$ calculated in higher orders.
\begin{table*}[t]
\begin{tabular}{ |c |c |c |c |c |c |c |c |c |c |c |c | }
\hline
$A$ & $-2$ & $-1.5$ &$-1.0$ & $-0.5$ & $0$ & $+0.5$ & $+1.0$ & $+1.5$ & $+2.0$ & $+2.5$ & $+3.0$
\\ \hline
$u^{(0)} (A)$ & $-0.4663$ & $+0.3726$ & $+1.000$ & $+1.2705$ & $+1.3685$ & $+1.4436$ & $+1.4205$ & $+1.2145$ & $+0.8343$ & $+0.3339$ & $-0.2355$
\\ \hline
$u^{(\rho)}(A)$ & $+1.0503$ & $-0.0163$ & $-0.000$ & $+0.3587$ & $+0.2376$ & $-0.0854$ & $+0.0623$ & $+0.9444$ & $+2.4408$ & $ +4.3269$ & $+6.4228$
\\ \hline
\end{tabular}
\caption{Turbulent Prandtl number for the present helical model is given as $Pr_A = 1 / (u^{(0)} (A) + \rho^2 \, u^{(\rho)} (A) )$ with numerical values of $u^{(0)} (A)$ and $u^{(\rho)} (A)$ given for selected values of $A$ and rounded on the last presented digit. Values at $A=-2$, $A=2.5$ and $A=3$ demonstrate physical constraints that must be imposed on values of $A$ in two loop calculations as performed here. In fact, for non-helical case only values satisfying $-1.723 < A < 2.800$ are considered in the present paper. Let us however note that for sufficiently large values of $\rho$, namely $\rho < 0.749$ all turbulent Prandtl number obtain positive and thus physically meaningful values. \label{tab1}}
\end{table*}
On the other hand, the equality of one and two loop results for $A = 1.038$ explains another up to date not well understood result of Ref.\,\cite{Jurcisin2014}. Here, the authors have observed that kinematic MHD model corresponding to $A=1$ of the present model is remarkably stable against one- and two-loop order corrections. Using however the previous result we easily explain this as a consequence of $A=1$ case lying in the proximity of $A = 1.038$ where one- and two-loop order values are identical. Such a situation is of coarse true only in the present two-loop order of the calculation. In higher orders of the perturbation theory, the corresponding polynomials over $A$ which occur in diagrams of Fig.\,\ref{fig3} are of higher orders and consequently the intersection between higher order analog of $u^{(0)} (A)$ and one-loop order result $u^{(1)}_{*}$ may dramatically shift to new values. Additionally, contrary to the $A=-1$ case, there is no evidence from helical values that the location of $A = 1.038$ would be fixed in higher order loop calculations. Thus, the relatively small contribution of the two-loop order corrections to the inverse Prandtl number of the kinematic MHD model should be clearly attributed to the present two-loop order of calculation. Additionally to this, $A=1$ case lies close to the border of the interval $A>0.912$ with $A=0.912$ being value where inverse turbulent Prandtl number is independent of any helical effects. Since the function $u^{(\rho)} (A)$ is continuous it consequently causes the helical effects for all models in vicinity of the $A=0,912$ case to be relatively well stable against helical effects. Such an effect is now clearly manifested also in the kinematic MHD model corresponding to $A=1$ case of the present model. Here, the corresponding inverse turbulent magnetic Prandtl number changes less than $5\%$ of its original non-helical value as observed for example in Ref.\cite{Jurcisin2014}.
Contrary to the previous two special cases discussed above, another physically important model corresponding to the $A = 0$ case of the present model lies deep in the interval of positive values of $u^{(\rho)} (A)$ and is thus located far away form points $A=0.325$ and $A=0.912$ where the function $u^{(\rho)} (A)$ has its zero points located. On the other hand, the $A=0$ model as studied for example in Ref.\,\cite{Jurcisin2014} lies relatively closely to the local maximum of the function $u^{(\rho)} (A)$ on the interval of $0.325 \leq A \leq 0.912$. Consequently, helical effects in the $A=0$ model are pronounced far greatly (almost by maximum possible amount in the interval of positive values of $u^{(\rho)}(A)$) as seen for example on the almost $20 \%$ change of the inverse turbulent Prandtl number in he helical environments. In other words, function $u^{(\rho)} (A)$ represents an easy tool to asses the importance of helical effects for given values of $A$ in the present model and explains previously unidentified context between the $A=0$ and $A=1$ models.
Finally, let us discuss the obtained values of helical turbulent Prandtl numbers which follow from Eq.\,(\ref{InversePrandtl}) and the functions $u^{(0)}$ and $u^{(\rho)}$ which appear therein. For selected parameters $A$, we show their corresponding numerical values in Tab.\,\ref{tab1} while their graphical representation is given in Fig.\ref{fig7} for the three physically important models of $A \in \{ -1, 0, 1\}$. As before, function $u^{(\rho)}(A)$ encodes the behavior of turbulent Prandtl numbers in respect to $\rho$ for all physically admissible values of $A$ which as shown before clearly depend also on the helicity parameter $\rho$. For the case of turbulent Prandtl numbers it has according to the Eq.\,(\ref{Prandtl}) the following meaning: For $u^{(\rho)}(A)>0$ turbulent Prandtl number is a decreasing function of $\rho$, for $u^{(\rho)}(A)<0$ turbulent Prandtl number is an increasing function of $\rho$ and finally for $u^{(\rho)}(A)=0$ turbulent Prandtl number is independent of $\rho$. This means that turbulent Prandtl number do increase with helicity parameter $\rho$ only for values of $A$ satisfying $-1.516 < A < -1$ and $0.325 < A< 0.912$. Excluded the zero points of $u^{(\rho)}(A)$, the remaining values of $A$ lead to monotonically decreasing helical turbulent Prandtl numbers as already seen in Ref.\,\cite{Jurcisin2014} for the special cases $A=0$ and $A=1$. While for $A=0$ model helical effects are pronounced more effectively due to reasons discussed above, the turbulent Prandtl number for $A=1$ model corresponding to kinematic MHD model is less sensitive to helical effects due to its above discussed proximity to the $A=0.912$ case. We also stress out, that there are no restrictions on $A$ when the threshold of $\rho \approx 0.749$ is exceeded. Thus, corresponding helical dependences of turbulent Prandtl numbers may for $\rho > 0.749$ also be considered. Consequently, we see that not the internal vectorial nature of the admixture itself but their interactions with the underlying turbulent field $\mathbf v$, as described by the parameter $A$, are crucial for developing different patterns in regard to helical effects and their influence on diffusion advection processes.
\begin{figure}[t]
\begin{center}\includegraphics[width=7cm]{fig7.pdf}\end{center}
\caption{ (Color online) Helical dependence of turbulent Prandtl numbers $Pr_{t,A}$ for three physically important models with $A \in \{ -1,0,1\}$ shown in the range of $\rho \leq 1$. Presented curves correspond to values $A=-1$ (blue), $A=0$ (magenta) and $A=1$ (red). \label{fig7} }
\end{figure}
Taking together, we have shown that the impact of the interactions as given via the parameter value of $A$ has a highly non-trivial impact on diffusion-advection processes when helical turbulent environments are considered. The resulting dependencies are truly complicated functions of $A$ and lead to non-trivial effects in connection with helicity parameter $\rho$. Therefore, instead of the tensorial nature of the admixture itself we have clearly identified the tensorial structure of interactions to be a more dominant factor which effectively alters advection diffusion process in fully developed turbulent environments. Thus, assertions made by authors of Ref.\,\cite{Jurcisin2014} must partially be revided at least for the case of vector admixtures advected passively in turbulent environments and the greater than expected impact of interactions on the actual advection diffusion processes must be recognized. Additionally, we once again stress out that present calculations clearly demonstrate that helical effects excert stabilizing effect on diffusion advection processes.
\section{Conclusion} \label{sec6}
Using the field theoretic renormalization group technique in the two-loop approximation, we have obtained analytic expressions for turbulent Prandtl number within the general $A$ model of passively advected vector impurity. Compared to Ref.\,\cite{Jurcisin2016} a more realistic scenario with effects of broken spatial parity has been considered by defining appropriate correlators of stochastic driving forces. Technically, the presence of broken spatial parity is described by helicity parameter $\rho$ ranging from $|\rho| = 0$ (no parity breaking) to $|\rho|=1$ (highest possible violation of spatial parity). Since our general helical $A$ model encompasses the less general non helical model of Ref.\,\cite{Jurcisin2016} we have been able to recover the results of Ref.\,\cite{Jurcisin2016} within the present calculations. However, the parameter $a_1$, in Eq, (46) has been shown to differ from the corresponding non helical value of Ref.\,\cite{Jurcisin2016}. However since further results show no differences and only the parameter a 1 from Ref.\,\cite{Jurcisin2016} is clearly not reproducing the well established results of Refs.\,\cite{Jurcisin2016}, we attribute the difference merely to a typographic error made by authors of Ref.\,\cite{Jurcisin2016}.
Furthermore, additionally to helical effects we extended our study of the $A$ model to arbitrary real values of $A$ as suggested by Ref.\,\cite{Arponen2009} whereas in Ref.\,\cite{Jurcisin2016}, only the interval $-1 \leq A \leq$ 1 is considered. Nevertheless, although one loop values of physical quantities have all been shown to obtain meaningful values when passing to the two loop order we noticed negative values of turbulent Prandtl numbers for $A < -1.723$ and $A > 2.800$ (numbers rounded at the last presented digit) in non-helical case. Furthermore, we show that helical effects effectively enhance stability in the present model and lift of the restrictions imposed on $A$ when a critical threshold of $\rho \approx 0.749$ (rounded on the last presented digit) is exceeded. This points towards the conclusion that restrictions of $A$ to interval $-1.723 \leq A \leq 2.800$ are most probably only an artifact of two loop order perturbative calculations. Furthermore, in Sec.\,\ref{sec5} we have shown that Feynman diagrams corresponding to the $n$-th order of perturbation theory will generally have a form of polynomials in $A$ with the highest possible power of $A$ being $2n$. We therefore expect that higher orders of loop calculations shift or let even completely vanish all the zero points of inverse turbulent Prandtl number. Such a behavior has already been observed in one loop order for the analogous quantity $u^{(1)}_{*}$. Thus, it would be of high interest to go beyond the limits of two loop order, however such an analysis is technically demanding and beyond the scope of the present paper. Nevertheless, two loop order values obtained deep in the interval of $-1.723 \leq A \leq 2.800$ are clearly free of any problems which means that physically most interesting cases of $A \in \{-1, 0, 1\}$ can be safely considered at least in the two loop order of the perturbation theory. Thus, all restrictions on the values of $A$ should be considered as an artifact of the perturbative approach used in the present work.
For the case of the model of linearized Navier-Stokes equations ($A = -1$) we have obtained helical values of turbulent Prandtl number equal $1$ regardless of the presence of helical effects. It is therefore natural to expect that also higher orders of perturbation theory may preserve the same, a hypothesis already stated by authors of Ref.\,\cite{Jurcisin2016}. This adds another argument in favor of hypothesis that problems with range of physically admissible values of $A$ could be resolved completely in higher orders. Physically, the resulting values demonstrate remarkable stability of $A = -1$ case against helical effects.
Effectively, the $A=1$ case corresponding to kinematic MHD model has been shown to have some similarities to $A=-1$ model with regard to its helical properties. Varying $A$ continuously allowed us to show that high stability of $A=1$ model is not due the vectorial nature of the admixture but due to its interactions given by $A=1$. Since it lies in the proximity of the $A=0.912$ case, where helical effects are not present in two-loop order, it must consequently be effectively less sensitive to helical effects. Contrary $A=0$ model is shown to lie far from values of $A$ where helical effects are not present. Consequently it shows significant dependence of turbulent Prandtl number on $\rho$.
Taking together, the case of $A = 1$ corresponding to the kinematic MHD turbulence, the case $A = 0$ model of passive vector admixture and the model of linearized Navier-Stokes equations have been brought into the context of the more general $A$ model. The interactions encoded by values of $A$ result in various patterns of behavior of turbulent Prandtl numbers. Thus, in regions of $1.516 \leq A \leq 1.000$ and $0.325 \leq A \leq 0.912$ (all numbers rounded on the last digit) the corresponding two loop turbulent Prandtl number are monotonically growing with $\rho$. Moreover, for values of $A \in \{ -1.516; -1; 0.325; 0.912\}$ turbulent Prandtl numbers are independent of $\rho$ but as previously discussed only $A=-1$ case is believed to retain this property also in higher order loop calculations. Finally, the remaining values of $A$ which belong to physically admissible region posses monotonically decreasing turbulent Prandtl numbers when $\rho$ is increased. We thus conclude that varying the interactions by changing the values of $A$ has a more profound effect on advection diffusion processes than the tensorial character of the admixture itself which significantly refines the conclusions made by authors of Ref.\,\cite{Jurcisin2016}.
\begin{acknowledgments}
The work was supported by VEGA Grant No.\,$1/0222/13$ of the Ministry of Education, Science, Research and Sport of the Slovak Republic. The authors gratefully acknowledge the hospitality of the Bogoliubov Laboratory of Theoretical Physics of the Joint Institute for Nuclear Research, Dubna, Russian Federation. P.Z. likes to express his gratitude to M.\,Dan\v{c}o and M.\,Jur\v{c}i\v{s}in for fruitful discussions which helped to carry out investigations presented here.
\end{acknowledgments}
|
\section{Introduction}
We consider a family of exact infinite-energy solutions of two three-dimensional (3D) fluid models with a damping term, the incompressible Euler equations
\begin{equation}
\label{dampede}
\begin{cases}
&\bfu_t+\bfu\cdot\nabla \bfu+\alpha \bfu=-\nabla p,
\\
&\diver\bfu=0,
\end{cases}
\end{equation}
and the inviscid Boussinesq system
\begin{equation}
\label{dampedb}
\begin{cases}
&\bfu_t+\bfu\cdot\nabla \bfu+\alpha \bfu=-\nabla p+\theta\mathbf{e}_3,
\\
&\theta_t+\bfu\cdot\nabla\theta=0,
\\
&\diver\bfu=0.
\end{cases}
\end{equation}
In \eqref{dampede}-\eqref{dampedb}, $\bfu$ represents the fluid velocity, $p$ is the scalar pressure, $\theta$ is the scalar temperature in the context of thermal convection or the density in the modeling of geophysical fluids, $\mathbf{e}_3=(0,0,1)^T$, and $\alpha\in\mathbb{R}^+$ is a real parameter. For $\alpha=0$, \eqref{dampede} reduces to the standard 3D Euler equations describing the motion of an ideal, incompressible homogeneous fluid, while \eqref{dampedb} becomes the standard 3D inviscid Boussinesq system modeling large scale atmospheric dynamics and oceanic flows \cite{gill, majda1, pedlosky}. If $\theta\equiv0$, \eqref{dampedb} reduces to \eqref{dampede}. When $\alpha\bfu$ in \eqref{dampede} is replaced by the diffusion term $-\nu\Delta\bfu$, we obtain the classical 3D Navier-Stokes equations. The global (in time) regularity problem for the aforementioned 3D models are long-standing open problems in mathematical fluid dynamics. See Constantin \cite{constantin1} for a history
and survey of results on the 3D Euler regularity problem and Fefferman \cite{fefferman} for a more precise account of the Navier-Stokes regularity problem. In general, the main obstacle in obtaining global existence of smooth solutions of these 3D models for general initial data is controlling nonlinear growth due to vortex stretching \cite{hou22, hou1}. To gain insight into this challenging problem, many researchers have turned their efforts to the 2D viscous Boussinesq equations
\begin{equation}
\label{b}
\begin{cases}
&\bfu_t+\bfu\cdot\nabla \bfu=-\nabla p+\nu\Delta\bfu+\theta\mathbf{e}_2,
\\
&\theta_t+\bfu\cdot\nabla\theta=\kappa\Delta\theta,
\\
&\diver\bfu=0.
\end{cases}
\end{equation}
System \eqref{b} can be shown to be formally identical to the 3D Euler or Navier-Stokes equations for axisymmetric swirling flows and retains key features of the 3D models such as the vortex stretching mechanism (see, e.g., \cite{majdabertozzi}). The global regularity issue for \eqref{b} has been settled in the affirmative under various degrees of viscosity and dissipation: with full viscosity $\nu>0$ and $\kappa>0$, partial viscosity $\nu>0$ and $\kappa=0$, or $\nu=0$ and $\kappa>0$, for anisotropic models \cite{adhikaria01, chae07, hou0, lai}, and with fractional Laplacian dissipation (see \cite{yang} and references therein). In contrast, the question of global regularity for \eqref{b} in the inviscid case $\nu=\kappa=0$ remains open; it is not apparent how to control vortex stretching when there is no dissipation ($\nu=0$) and no thermal diffusion ($\kappa=0$). Using a somewhat different approach, Adhikaria et al. \cite{adhikaria} replaced $\nu\Delta\bfu$ and $\kappa\Delta\theta$ in \eqref{b} with damping terms $-\alpha\bfu$ and $-\beta\theta$ for $\alpha>0$ and $\beta>0$ real parameters. Although the resulting damping effects are insufficient to control vortex stretching for general initial data, the authors showed that a local (in time) solution will persist globally in time if the initial data is small enough in some homogeneous Besov space.
The aim of this work is to examine how damping affects the global regularity of a particular class of infinite-energy solutions of \eqref{dampede} and \eqref{dampedb} which, in the absence of damping ($\alpha=0$), blowup in finite-time from smooth initial data. More particularly, the fluid velocity and temperature considered here have the form
\begin{equation}
\label{ansatz}
\bfu(\bfx,z,t)=(u(\bfx,t),v(\bfx,t),z\gamma(\bfx,t)),\qquad \theta(\bfx,z,t)=z\rho(\bfx,t)
\end{equation}
for $(\bfx,z)=(x,y,z)$. Our spatial domain will be the semi-bounded 3D channel
\begin{equation}
\label{domain}
\Pi\equiv\{(\bfx,z)\in Q\times\mathbb{R}\}
\end{equation}
of rectangular periodic cross-section $Q\equiv[0,1]^2$ with $\bfu$ and $\theta$ both periodic in the $x$ and $y$ variables with period one. Note that the unbounded geometry of \eqref{domain} in the vertical direction endows the fluid under consideration with, at best, locally finite kinetic energy.
Solutions of the 3D incompressible Euler equations
\begin{equation}
\label{e}
\begin{cases}
&\bfu_t+\bfu\cdot\nabla \bfu=-\nabla p
\\
&\diver\bfu=0
\end{cases}
\end{equation}
of the form \eqref{ansatz}i) were proposed by Gibbon et al. \cite{gibbon1}, and then shown to blowup in finite time numerically by Ohkitani and Gibbon \cite{ohkitani}, and analytically by Constantin \cite{constantin2}. See also \cite{childress, stuart, sarria1, sarria2} for blowup results of other similar infinite-energy solutions of the Euler and Boussinesq equations in two and three dimensions. For convenience of the reader, we now summarize Constantin's blowup result.
The set-up used by Constantin is effectively the same as the one considered here. Imposing a velocity field of the form \eqref{ansatz}i) on \eqref{e} subject to periodicity in the $x$ and $y$ variables of period one and with spatial domain \eqref{domain}, it follows that the vertical component of the velocity field, $z\gamma$, satisfies the vertical component of the momentum equation \eqref{e}i) if the mean-zero function $\gamma=-\left(u_x+v_y\right)$ solves the nonlocal two-dimensional equation
\begin{equation}
\label{undampedgammaeqn}
\gamma_t+\bfu'\cdot\nabla\gamma=-\gamma^2+2\int_{Q}{\gamma^2\,d\mathbf{x}}
\end{equation}
with $\bfu'=(u,v)$. For $(\bfa,t)\equiv(a_1,a_2,t)$, $\bfa\in Q$, Constantin constructed the solution formula
\begin{equation}
\label{undampedsln}
\begin{split}
\gamma(\bfY(\bfa,t),t)=-\frac{\tau'(t)}{\tau(t)}\left\{\frac{1}{1+\gamma_0(\bfa)\tau(t)}-\left(\int_{Q}{\frac{d\bfa}{1+\gamma_0(\bfa)\tau(t)}}\right)^{-1}\int_{Q}{\frac{d\bfa}{(1+\gamma_0(\bfa)\tau(t))^2}}\right\}
\end{split}
\end{equation}
for $\gamma(\bfa,0)=\gamma_0(\bfa)$, $\tau$ satisfying the initial value problem (IVP)
\begin{equation}
\label{undampedtau}
\tau'(t)=\left(\int_Q{\frac{d\bfa}{1+\gamma_0(\bfa)\tau(t)}}\right)^{-2},\qquad \tau(0)=0,
\end{equation}
and $\bfa\to\mathbf{Y}(\bfa,t)$ the 2D flow-map defined by
\begin{equation*}
\label{undampedflowmap}
\frac{d\mathbf{Y}}{dt}=\bfu'(\mathbf{Y}(\bfa,t),t),\qquad \mathbf{Y}(\bfa,0)=\bfa.
\end{equation*}
By comparing the blowup rates of the time integrals in \eqref{undampedsln} against the local term, Constantin proved the following blowup result for a large class of smooth initial data $\gamma_0$.
\begin{theorem}[\cite{constantin2}]
\label{blowupundamped}
Consider the initial boundary value problem for \eqref{undampedgammaeqn} with smooth mean-zero initial data $\gamma_0$ and periodic boundary conditions. Suppose $\gamma_0$ attains a negative minimum $m_0$ at a finite number of locations $\bfa_0\in Q$ and, near these locations, $\gamma_0$ has non-vanishing second-order derivatives. Set $\tau^*=-\frac{1}{m_0}$ and let
\begin{equation}
\label{undampedeulertime}
t^E(\tau)=\int_0^{\tau}\left(\int_{Q}{\frac{d\bfa}{1+\gamma_0(\bfa)\mu}}\right)^2d\mu.
\end{equation}
Then there exists a finite time $T^E>0$, given by
\begin{equation*}
\label{undampedeulerblowup}
T^E\equiv\lim_{\tau\nearrow\tau^*}t^E(\tau),
\end{equation*}
such that both the maximum and minimum values of $\gamma$ diverge to positive and respectively negative infinity as $t\nearrow T^E$.
\end{theorem}
The outline for the remainder of the paper is as follows. In Section \ref{prelim} we introduce the damped two-dimensional equations \eqref{dampedeqnb}-\eqref{dampedeqne} and summarize the main results of the paper. Then in Section \ref{solution} we derive the solution formulae \eqref{gammab}-\eqref{tauivpe}, which we use in Section \ref{proofs} to prove the Theorems.
\section{Preliminaries}
\label{prelim}
\subsection{The Damped Two-dimensional Equations}\hfill
As stated in the previous section, we are interested in the global regularity of solutions of \eqref{dampede} and \eqref{dampedb} of the form \eqref{ansatz} subject to periodic boundary conditions in the $x$ and $y$ variables (period one), and with spatial domain \eqref{domain}. First note that, from incompressibility and periodicity, $\gamma=-\left(u_x+v_y\right)$ satisfies the mean-zero condition
\begin{equation}
\label{zeromean}
\int_Q{\gamma(\bfx,t)\,d\bfx}=0.
\end{equation}
Then imposing the ansatz \eqref{ansatz} on the damped Boussinesq system \eqref{dampedb}, it is easy to check that the vertical component of the velocity field, $z\gamma$, and the scalar temperature, $z\rho$, satisfy the vertical component of \eqref{dampedb}i) and equation \eqref{dampedb}ii) if $\gamma$ and $\rho$ solve the nonlocal 2D system
\begin{equation}
\label{dampedeqnb}
\begin{cases}
\gamma_t+\bfu'\cdot\nabla\gamma=\rho-\gamma^2-\alpha\gamma+I(t),\quad &\bfx\in Q,\quad t>0,
\\
\rho_t+\bfu'\cdot\nabla\rho=-\gamma\rho,\qquad &\bfx\in Q,\quad t>0
\end{cases}
\end{equation}
with $\bfu'=(u,v)$, $\alpha>0$ a real parameter, and
\begin{equation}
\label{nonlocalb}
I(t)=2\int_Q{\gamma^2(\bfx,t)\,d\bfx}-\int_Q\rho(\bfx,t)\,d\bfx.
\end{equation}
For $\rho\equiv0$, \eqref{dampedeqnb}-\eqref{nonlocalb} reduces to
\begin{equation}
\label{dampedeqne}
\begin{cases}
\gamma_t+\bfu'\cdot\nabla\gamma=-\gamma^2-\alpha\gamma+I(t),\quad &\bfx\in Q,\quad t>0,
\\
I(t)=2\int_Q{\gamma^2(\bfx,t)\,d\bfx},
\end{cases}
\end{equation}
which is just the associated 2D equation obtained from the vertical component of the damped 3D Euler system \eqref{dampede}.
For simplicity we will refer to equations \eqref{undampedgammaeqn} and \eqref{dampedeqne}, and the system \eqref{dampedeqnb}-\eqref{nonlocalb}, as \emph{the undamped Euler equation}, \emph{the damped Euler equation}, and \emph{the damped Boussinesq system}, respectively.
Before summarizing the main results of the paper, we define some notation that will be helpful in differentiating among solutions of the various equations under consideration.
\subsection{Notation}\hfill
For $\alpha>0$, we denote by $(\gamma^{B}_{\alpha},\rho_{\alpha})$ and $\gamma^{E}_{\alpha}$ the solution of the damped Boussinesq system \eqref{dampedeqnb}-\eqref{nonlocalb} and the damped Euler equation \eqref{dampedeqne}, respectively. The undamped ($\alpha=0$) counterparts of $(\gamma^{B}_{\alpha},\rho_{\alpha})$ and $\gamma^E_{\alpha}$ will be denoted by dropping the subscript, i.e., $(\gamma^{B},\rho)$ and respectively $\gamma^E$, with the latter given (along characteristics) by formula \eqref{undampedsln}. Other notation will be introduced in later sections in a similar manner. Lastly, by $C$ we mean a generic positive constant that may change in value from line to line.
\subsection{Summary of Results}\hfill
For a smooth initial condition $\gamma_0$ satisfying the conditions of Theorem \ref{blowupundamped}, we determine in Theorem \ref{dampedeulerthm} below ``how much'' damping is required for the solution $\gamma^E_{\alpha}$ of the damped Euler equation \eqref{dampedeqne} to persist globally in time, or alternatively, for the finite-time blowup of the solution $\gamma^E$ of the undamped Euler equation \eqref{undampedgammaeqn} to be suppressed.
\begin{theorem}
\label{dampedeulerthm}
Consider the damped Euler equation \eqref{dampedeqne} with smooth mean-zero initial data $\gamma_0$ and periodic boundary conditions. Suppose $\gamma_0$ satisfies the conditions of Theorem \ref{blowupundamped}, so the solution $\gamma^E$ of the undamped Euler equation \eqref{undampedgammaeqn} blows up at a finite time $T^E$. Then for $\alpha\geq1/T^E$, the solution $\gamma^E_{\alpha}$ of the damped Euler equation \eqref{dampedeqne} exists globally in time. More particularly, for $\alpha=1/T^E$, $\gamma^E_{\alpha}$ converges to a non-trivial steady state as $t\to+\infty$, whereas, for $\alpha>1/T^E$, convergence is to the trivial steady state. In contrast, if $0<\alpha<1/T^E$, then there exists a finite time $T_{\alpha}^E>T^E$ such that the maximum and minimum values of $\gamma^E_{\alpha}$ diverge to positive and respectively negative infinity as $t\nearrow T_{\alpha}^E$. More particularly, let $\tau^*=-1/m_0$ for $m_0$ the negative minimum of $\gamma_0$ attained at a finite number of points in $Q$, and set $t_{\alpha}^E(\tau)=-\frac{1}{\alpha}\ln\left|1-\alpha t^E(\tau)\right|$ for $t^E(\tau)$ the undamped Euler time variable \eqref{undampedeulertime}. Then for $0<\alpha<1/T^E$, the finite blowup time is $T_{\alpha}^E\equiv\lim_{\tau\nearrow\tau^*}t_{\alpha}^E(\tau)$.
\end{theorem}
Before summarizing our next result, we note that Gibbon and Ohkitani \cite{gibbon3} established a regularity criterion of BKM \cite{BKM} type whereby a solution $\gamma^E$ of \eqref{undampedgammaeqn} blows up at a finite time $T^E$ if and only if
\begin{equation}
\label{bkm}
\int_0^{T^E}{\|\gamma^E(\bfx,s)\|_{\infty}}ds=+\infty
\end{equation}
for $\|\cdot\|_{\infty}$ the supremum norm. Let $(\bfu, p)$ be a solution of the damped Euler system \eqref{dampede} for $\bfu$ as in \eqref{ansatz}i). Since the vertical component $\omega=v_x-u_y$ of the vorticity $\boldsymbol{\omega}=\curl\bfu$ satisfies
\begin{equation*}
\label{dampedvorticityeqn}
\omega_t+\bfu'\cdot\grad\omega=(\gamma_{\alpha}^E-\alpha)\omega,
\end{equation*}
we may refer to $\gamma_{\alpha}^E$ as the vorticity stretching rate. In the spirit of \eqref{bkm}, a similar BKM-type criterion can be established for a solution of the damped Euler equation \eqref{dampedeqne} where the blowup
time $T_{\alpha}^E$ is defined as the smallest time at which
$$\int_0^{T_{\alpha}^E}{\|\gamma_{\alpha}^E(\bfx,s)\|_{\infty}}ds=+\infty.$$
Additionally, the argument used to prove Theorem 1.1 in \cite{sarria2}, together with Theorem \ref{dampedboussinesqthm} below, shows that the same regularity criterion\footnote[1]{With $\gamma_{\alpha}^E$ replaced by $\gamma_{\alpha}^B$} is true for a solution of the damped Boussinesq system \eqref{dampedeqnb}-\eqref{nonlocalb}. Part one of our next Theorem shows that for smooth $\gamma_0$ satisfying the conditions of Theorem \ref{blowupundamped} and $\rho_0\geq0$, the existence of a finite blowup time $T^E$ for the undamped Euler equation \eqref{undampedgammaeqn} leads to finite-time blowup of the damped Boussinesq system \eqref{dampedeqnb}-\eqref{nonlocalb} if the damping coefficient satisfies $0<\alpha<1/T^E$ and there is at least one point $\bfa_1\in Q$ such that $\gamma_0(\bfa_1)=m_0$ and $\rho_0(\bfa_1)=0$. In particular, we find that the time integral of $\gamma_{\alpha}^B$ diverges to \emph{negative infinity} for $\bfa=\bfa_1$, in agreement with the aforementioned BKM-type criterion. The second part of Theorem \ref{dampedboussinesqthm} shows that finite-time blowup is not restricted to nonnegative $\rho_0$, although the blowup mechanism we uncover for nonpositive $\rho_0$ is of a different nature and the singularity is effectively one dimensional. Briefly, we show the existence of smooth $\gamma_0$ attaining its negative minimum $m_0$ at infinitely many points $\bfa'\in Q$, and $\rho_0\leq0$ satisfying $\rho_0(\bfa')\neq0$, such that for $T^B>0$ the finite blowup time of the associated undamped Boussinesq system (i.e. \eqref{dampedeqnb}-\eqref{nonlocalb} with $\alpha=0$), if $0<\alpha<1/T^B$, then the time integral of $\gamma_{\alpha}^B$ diverges to \emph{positive infinity} for $\bfa\neq\bfa'$.
\begin{theorem}
\label{dampedboussinesqthm}
Consider the damped Bousinesq system \eqref{dampedeqnb}-\eqref{nonlocalb} with periodic boundary conditions.
\begin{enumerate}
\item Suppose $\gamma_0$ satisfies the conditions of Theorem \ref{blowupundamped}, so the solution $\gamma^E$ of the undamped Euler equation \eqref{undampedgammaeqn} blows up at a finite time $T^E$. Further, let $0<\alpha<1/T^E$, so that by Theorem \ref{dampedeulerthm} the solution $\gamma_{\alpha}^E$ of the damped Euler equation \eqref{dampedeqne} blows up at a finite time $T_{\alpha}^E$. Assume $\gamma_0$ attains its negative minimum $m_0$ at finitely many points $\bfa_i\in Q$, $1\leq i\leq n$, and $\rho_0\geq0$ is smooth with $\rho_0(\bfa_j)=0$ for some $1\leq j\leq n$. Then there exists a finite time $T_{\alpha}^B$, satisfying $0<T_{\alpha}^B<T_{\alpha}^E$, such that
$$J(\bfa_j,t)= \operatorname{exp}\,\left\{-\int_0^t{\gamma_{\alpha}^B(\bfX(\bfa_j,s),s)ds}\right\}\to+\infty$$
as $t\nearrow T_{\alpha}^B$ for $J=\det\left\{\frac{\partial\bfX}{\partial\bfa}\right\}$ the Jacobian of the 2D flow-map (see \eqref{dampedflowmap}).
\vspace{0.05in}
\item There exist smooth mean-zero initial data $\gamma_0$ attaining its negative minimum $m_0$ at infinitely many points $\bfa'\in Q$, and smooth $\rho_0\leq0$ with $\rho_0(\bfa')\neq0$, such that for $T^B>0$ the finite blowup time of the associated undamped Boussinesq system, if $0<\alpha<1/T^B$ and $\bfa\neq\bfa'$, then
$$J(\bfa,t)= \operatorname{exp}\,\left\{-\int_0^t{\gamma_{\alpha}^B(\bfX(\bfa,s),s)ds}\right\}\to0$$
as $t\nearrow T_{\alpha}^B=-\frac{1}{\alpha}\ln\left(1-\alpha T^B\right)>T^B$.
\end{enumerate}
\end{theorem}
\section{Solution along characteristics}
\label{solution}
In this section we derive the representation formula \eqref{gammab}-\eqref{tauivpb} for the general solution $(\gamma_{\alpha}^B(\bfx,t),\rho_{\alpha}(\bfx,t))$ (along characteristics) of the damped Boussinesq system \eqref{dampedeqnb}-\eqref{nonlocalb}. Our approach is a natural generalization to higher dimensions of the arguments in \cite{sarria1, sarria2}, and is somewhat more direct than the argument used in \cite{constantin2} to derive \eqref{undampedsln}-\eqref{undampedtau}.
For $(\bfa,t)\equiv(a_1,a_2,t)$, $\bfa\in Q$, and $\bfu'=(u,v)$, define the 2D flow-map $\bfa\to\bfX(\bfa,t)$ as the solution of the IVP
\begin{equation}
\label{dampedflowmap}
\frac{d\mathbf{X}}{dt}=\bfu'(\bfX(\bfa,t),t),\qquad \mathbf{X}(\bfa,0)=\bfa.
\end{equation}
Integrating equation \eqref{dampedeqnb}ii) along the flow-map yields
\begin{equation}
\label{eq1}
\rho_{\alpha}(\bfX(\bfa,t),t)=\rho_0(\bfa)\operatorname{exp}\,\left\{-\int_0^t{\gamma_{\alpha}^B(\bfX(\bfa,s),s)\,ds}\right\}.
\end{equation}
Note that
\begin{equation}
\label{rho}
\rho_{\alpha}(\bfX(\bfa,t),t)=\rho_0(\bfa)J(\bfa,t)
\end{equation}
for $J(\bfa,t)=\det\left\{\frac{\partial\bfX}{\partial\bfa}\right\}$ the Jacobian determinant of $\bfX$ satisfying
\begin{equation}
\label{jac}
J(\bfa,t)=\text{exp}\,\left\{-\int_0^t{\gamma_{\alpha}^B(\bfX(\bfa,s),s)\,ds}\right\}.
\end{equation}
Formula \eqref{rho} follows directly from \eqref{eq1} and the fact that the Jacobian satisfies
\begin{equation}
\label{jacid}
J_t(\bfa,t)=-J(\bfa,t)\gamma_{\alpha}^B(\bfX(\bfa,t),t),\qquad J(\bfa,0)\equiv1.
\end{equation}
Next we use the above formulas to derive a second-order linear ODE for $\xi=J^{-1}$.
From \eqref{dampedeqnb}i), \eqref{rho}, and \eqref{jacid}, it follows that
\begin{equation}
\label{eq2}
\begin{split}
\frac{\partial}{\partial t}(\gamma_{\alpha}^B(\bfX(\bfa,t),t))=\rho_0(\bfa)J(\bfa,t)-\left(\frac{J_t(\bfa,t)}{J(\bfa,t)}\right)^2+\alpha\frac{J_t(\bfa,t)}{J(\bfa,t)}+I(t).
\end{split}
\end{equation}
Then differentiating \eqref{jacid} with respect to $t$, and using \eqref{jacid} and \eqref{eq2} on the resulting expression leads to
\begin{equation}
\label{eq3}
\begin{split}
J_{tt}&=\left(2\frac{J_t^2}{J^2}-\rho_0J-\alpha\frac{J_t}{J}-I(t)\right)J.
\end{split}
\end{equation}
Setting $\xi=J^{-1}$ in \eqref{eq3} yields, after some rearranging,
\begin{equation}
\label{ode}
\xi_{tt}(\bfa,t)+\alpha\xi_t(\bfa,t)-I(t)\xi(\bfa,t)=\rho_0(\bfa),
\end{equation}
a second-order linear ODE parametrized by $\bfa\in Q$. Fix $\bfa\in Q$ and set $f(t)=\xi(\bfa,t)$. Then by \eqref{jacid} and \eqref{ode}, $f$ solves the IVP
\begin{equation}
\label{ode2}
f''(t)+\alpha f'(t)-I(t)f(t)=\rho_0,\qquad f(0)=1,\,\,\, f'(0)=\gamma_0(\bfa)
\end{equation}
for ${}^{'}\equiv\frac{d}{dt}$. To find a general solution of \eqref{ode2} we follow a standard variation of parameters argument. Consider the associated homogeneous equation
\begin{equation}
\label{homo}
f_h''(t)+\alpha f_h'(t)-I(t)f_h(t)=0.
\end{equation}
Let $\phi_1(t)$ and $\phi_2(t)$ be two linearly independent solutions of \eqref{homo} satisfying $\phi_1(0)=\phi_2'(0)=1$ and $\phi_1'(0)=\phi_2(0)=0$. Then by reduction of order, the general solution of \eqref{homo} takes the form $f_h(t)=\phi_1(t)(c_1(\bfa)+c_2(\bfa)\tau_{\alpha}(t))$ for
\begin{equation*}
\label{tau}
\tau_{\alpha}(t)=\int_0^t{\frac{e^{-\alpha s}}{\phi_1^2(s)}\,ds}.
\end{equation*}
Now consider a particular solution $f_p(t)=v_1(\bfa,t)\phi_1(t)+v_2(\bfa,t)\phi_2(t)$ of \eqref{ode2} for $\phi_2(t)=\tau_{\alpha}(t)\phi_1(t)$ and $v_1$ and $v_2$ to be determined. After some standard computations, we can write the general solution of \eqref{ode2} as
\begin{equation}
\label{particular}
\xi(\bfa,t)=\phi_1(t)\left[\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)\right]
\end{equation}
for
\begin{equation}
\label{psi}
\begin{split}
\Psi(\bfa,t)=1+\gamma_0(\bfa)\tau_{\alpha}(t)
\end{split}
\end{equation}
and
\begin{equation}
\label{sigma}
\begin{split}
\sigma(t)&=\int_0^t{e^{\alpha s}\tau_{\alpha}(s)\phi_1(s)ds}-\tau_{\alpha}(t)\int_0^t{e^{\alpha s}\phi_1(s)ds}
\\
&=
-\int_0^t\int_s^t{\frac{\phi_1(s)}{\phi_1^2(z)}e^{\alpha(s-z)}dzds}.
\end{split}
\end{equation}
The Jacobian is now obtained from \eqref{particular} and $\xi=J^{-1}$ as
\begin{equation}
\label{dampedjacobianb0}
J(\bfa,t)=\frac{\phi_1^{-1}(t)}{\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)}.
\end{equation}
To find $\phi_1$ note that for fixed $\bfa\in Q$ and $\mathbf{c}\in\mathbb{Z}^2$, the IVP
\begin{equation}
\label{flow0}
\mathbf{Z}'(t)=\bfu'(\mathbf{Z}(t),t),\qquad \mathbf{Z}(0)=\bfa+\mathbf{c}
\end{equation}
has a unique solution as long as $\bfu'=(u,v)$ stays smooth. Then by periodicity of $\bfu'$ and \eqref{dampedflowmap}, $\mathbf{Z}_1(t)=\bfX(\bfa,t)+\mathbf{c}$ and $\mathbf{Z}_2(t)=\bfX(\bfa+\mathbf{c},t)$ both solve \eqref{flow0} with the same initial condition. Thus $\bfX(\bfa,t)+\mathbf{c}=\bfX(\bfa+\mathbf{c},t)$, which implies that
\begin{equation}
\label{mean1}
\int_Q{J(\bfa,t)d\bfa}\equiv1.
\end{equation}
Integrating \eqref{dampedjacobianb0} now yields
\begin{equation}
\label{phib}
\phi_1(t)=\int_Q{\frac{d\bfa}{\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)}}.
\end{equation}
For $i\in\mathbb{Z}^+$, set
\begin{equation*}
\label{kgamma}
\mathcal{K}_i(\bfa,t)=\frac{\gamma_0(\bfa)}{(\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t))^i}\,,\qquad
\mathcal{\bar K}_i(t)=\int_Q{\frac{\gamma_0(\bfa)}{(\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t))^i}d\bfa}
\end{equation*}
and
\begin{equation*}
\label{krho}
\mathcal{L}_i(\bfa,t)=\frac{\rho_0(\bfa)}{(\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t))^i}\,,\qquad
\mathcal{\bar L}_i(t)=\int_Q{\frac{\rho_0(\bfa)}{(\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t))^i}d\bfa}.
\end{equation*}
Then using \eqref{rho}, \eqref{jacid}, \eqref{dampedjacobianb0} and \eqref{phib} we obtain, after a lengthy but straight-forward computation, the solution formula
\begin{equation}
\label{gammab}
\begin{split}
\gamma^B_{\alpha}(\bfX(\bfa,t),t)=\tau_{\alpha}'(t)\left(\mathcal{K}_1(\bfa,t)-\frac{\mathcal{\bar K}_2(t)}{\phi_1(t)}\right)+
\left(\mathcal{L}_1(\bfa,t)-\frac{\mathcal{\bar L}_2(t)}{\phi_1(t)}\right)\int_0^t{e^{\alpha s}\phi_1(s)ds}
\end{split}
\end{equation}
for $\tau_{\alpha}(t)$ satisfying
\begin{equation}
\label{tauivpb}
\tau_{\alpha}'(t)=\frac{e^{-\alpha t}}{\phi_1^{2}(t)},\qquad \tau_{\alpha}(0)=0.
\end{equation}
Moreover, the Jacobian \eqref{dampedjacobianb0} becomes
\begin{equation}
\label{dampedjacobianb}
J(\bfa,t)=\frac{1}{\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)}\left(\int_Q{\frac{d\bfa}{\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)}}\right)^{-1}.
\end{equation}
Since \eqref{rho} implies that if $\rho_0\equiv0$, then $\rho_{\alpha}\equiv0$ for as long as the solution is defined, setting $\rho_0\equiv0$ in \eqref{gammab}-\eqref{tauivpb} yields, after some rearranging, the general solution of the damped Euler equation \eqref{dampedeqne} as
\begin{equation}
\label{gammae}
\begin{split}
\gamma^E_{\alpha}(\bfX(\bfa,t),t)=-\frac{\tau_{\alpha}'(t)}{\tau_{\alpha}(t)}\left\{\frac{1}{1+\gamma_0(\bfa)\tau_{\alpha}(t)}-\left(\int_Q{\frac{d\bfa}{1+\gamma_0(\bfa)\tau_{\alpha}(t)}}\right)^{-1}\int_Q{\frac{d\bfa}{(1+\gamma_0(\bfa)\tau_{\alpha}(t))^2}}\right\}
\end{split}
\end{equation}
for
\begin{equation}
\label{tauivpe}
\tau_{\alpha}'(t)=e^{-\alpha t}\left(\int_Q{\frac{d\bfa}{1+\gamma_0(\bfa)\tau_{\alpha}(t)}}\right)^{-2},\qquad \tau_{\alpha}(0)=0.
\end{equation}
\section{Proof of the Main Theorems}
\label{proofs}
The blowup result in Theorem \ref{blowupundamped} for the solution $\gamma^E$ of the undamped Euler equation \eqref{undampedgammaeqn} is established by estimating blowup rates for the integral terms in \eqref{undampedsln} under the assumption that the smooth initial data $\gamma_0$ behaves quadratically near the points where its minimum is attained \cite{constantin2}. Since we are interested in how damping can arrest this blowup, we consider the same class of initial data. In particular, this means that the blowup rates derived in \cite{constantin2} for the integral terms also hold here; however, and for convenience of the reader, we outline how to obtain these estimates in the proof of Theorem \ref{dampedeulerthm} below.
\begin{proof}[\textbf{Proof of Theorem \ref{dampedeulerthm}}]
Suppose the mean-zero initial data $\gamma_0(\bfa)$ is smooth and attains its negative minimum $m_0$ at a finite number of locations $\bfa_0\in Q$. Then the spatial term
\begin{equation}
\label{spatial}
\frac{1}{1+\gamma_0(\bfa)\tau_{\alpha}(t)}
\end{equation}
in \eqref{gammae} diverges to positive infinity when $\bfa=\bfa_0$ as $\tau_{\alpha}$ approaches $$\tau^*=-\frac{1}{m_0},$$
and remains finite and positive for all $0\leq\tau\leq\tau^*$ and $\bfa\neq\bfa_0$. Suppose $\gamma_0$ has nonzero second-order partials near $\bfa_0$, so that locally,
\begin{equation*}
\label{approx}
m_0+\frac{1}{2}\lambda_2|\bfa-\bfa_0|^2\leq \gamma_0(\bfa)\leq
m_0+\frac{1}{2}\lambda_1|\bfa-\bfa_0|^2
\end{equation*}
for $0\leq|\bfa-\bfa_0|\leq r$, $r>0$ small, and $\lambda_1>\lambda_2>0$ the eigenvalues of the Hessian matrix of $\gamma_0$ at $\bfa_0$.\footnote[2]{If $\lambda_1=\lambda_2=\lambda$, then near $\bfa_0$, $\gamma_0\sim m_0+\frac{1}{2}\lambda|\bfa-\bfa_0|^2$ and the blowup rates \eqref{int1}- \eqref{int2} still hold.} This implies that
\begin{equation}
\label{approx2}
C_1\ln\left(1+\frac{\lambda_1 r^2}{2\epsilon}\right)\leq
\int_{\mathcal{D}}\frac{d\bfa}{\epsilon+\gamma_0(\bfa)-m_0}
\leq
C_2\ln\left(1+\frac{\lambda_2 r^2}{2\epsilon}\right)
\end{equation}
for $\epsilon>0$ small, $C_i=\frac{2\pi}{\lambda_i}>0$, $i=1, 2$, and $\mathcal{D}$ the disk centered at $\bfa_0$ of radius $r$. Setting $\epsilon=\frac{1}{\tau_{\alpha}}+m_0$ in \eqref{approx2}, it follows that
\begin{equation}
\label{int1}
\int_{Q}\frac{d\bfa}{1+\gamma_0(\bfa)\tau_{\alpha}}\sim -C\ln(\tau^*-\tau_{\alpha})
\end{equation}
for $\tau^*-\tau_{\alpha}>0$ small. By a similar argument,
\begin{equation}
\label{int2}
\int_{Q}\frac{d\bfa}{(1+\gamma_0(\bfa)\tau_{\alpha})^2}\sim\frac{C}{\tau^*-\tau_{\alpha}}.
\end{equation}
Next we need the corresponding behavior of the exponential term
\begin{equation}
\label{exp}
e^{-\alpha t}=\operatorname{exp}\left\{-\alpha t_{\alpha}^E(\tau_{\alpha})\right\}
\end{equation}
in \eqref{gammae}-\eqref{tauivpe} as $\tau_{\alpha}\nearrow \tau^*$. In \eqref{exp} we have introduced the notation $t= t_{\alpha}^E$ to differentiate the time variable in \eqref{gammae}-\eqref{tauivpe} from that in \eqref{gammab}-\eqref{tauivpb}. For $\alpha>0$, we will refer to $t_{\alpha}^E$ and $t_{\alpha}^B$ as \emph{the damped Euler time variable} and respectively \emph{the damped Boussinesq time variable}.
Since $\gamma_0$ satisfies the conditions of Theorem \ref{blowupundamped}, the limit $T^E\equiv\lim_{\tau_{\alpha}\to\tau^*}t^E(\tau_{\alpha})$, for
\begin{equation}
\label{undampedeulertime2}
t^E(\tau_{\alpha})=\int_0^{\tau_{\alpha}}\left(\int_{Q}{\frac{d\bfa}{1+\gamma_0(\bfa)\mu}}\right)^2d\mu
\end{equation}
the associated \emph{undamped Euler time variable} \eqref{undampedeulertime}, is positive and finite, and represents the blowup time for $\gamma^E$ in Theorem \ref{blowupundamped}. That $T^E$ is finite follows from \eqref{int1} and \eqref{undampedeulertime2} which imply, for $\tau^*-\tau_{\alpha}>0$ small, the asymptotic relation
\begin{equation}
\label{asymp}
T^E-t^E\sim (\tau^*-\tau_{\alpha})(\ln(\tau^*-\tau_{\alpha})-1)^2
\end{equation}
whose right-hand side vanishes as $\tau_{\alpha}\nearrow\tau^*$.
Since \eqref{tauivpe} implies that the damped Euler time variable satisfies
\begin{equation}
\label{time2}
t_{\alpha}^E(\tau_{\alpha})=-\frac{1}{\alpha}\ln\left|1-\alpha t^E(\tau_{\alpha})\right|,
\end{equation}
it follows that the behavior of the exponential \eqref{exp} is determined by the limit
\begin{equation}
\label{limit}
T_{\alpha}^E\equiv\lim_{\tau_{\alpha}\nearrow\tau^*}t_{\alpha}^E(\tau_{\alpha}),
\end{equation}
which in turn depends on whether $0<\alpha<1/T^E$ or $\alpha\geq1/T^E$.
\subsection*{Case 1 - Finite-time blowup for $0<\alpha<1/T^E$.}\hfill
For $0<\alpha<1/T^E$, the argument of the logarithm in \eqref{time2} satisfies $0<1-\alpha t^E(\tau_{\alpha})\leq1$ for all $0\leq\tau_{\alpha}\leq\tau^*$. This implies that the limits
\begin{equation*}
\label{case1time}
\lim_{\tau_{\alpha}\nearrow\tau^*}t_{\alpha}^E(\tau_{\alpha})=T_{\alpha}^E,\quad\qquad\lim_{\tau_{\alpha}\nearrow\tau^*}\operatorname{exp}\left\{-\alpha t_{\alpha}^E(\tau_{\alpha})\right\}=\operatorname{exp}\left\{-\alpha T_{\alpha}^E\right\}
\end{equation*}
are both positive and finite. Using estimates \eqref{int1} and \eqref{int2} on \eqref{gammae}, it follows that for $\bfa=\bfa_0$ the spatial term dominates and $\gamma_{\alpha}^E$ diverges to negative infinity,
$$\gamma_{\alpha}^E(\bfX(\bfa_0,t),t)\sim-\frac{C}{(\tau^*-\tau_{\alpha})\ln^2(\tau^*-\tau_{\alpha})}\to-\infty$$
as $t\nearrow T_{\alpha}^E$. If instead $\bfa\neq\bfa_0$, the second term in the bracket of \eqref{gammae} now dominates and the blowup is to positive infinity,
$$\gamma_{\alpha}^E(\bfX(\bfa,t),t)\sim-\frac{C}{(\tau^*-\tau_{\alpha})\ln^3(\tau^*-\tau_{\alpha})}\to+\infty.$$
From \eqref{time2} note that the blowup time $T_{\alpha}^E$ approaches the undamped blowup time $T^E$ as the damping coefficient vanishes. Further, for $\alpha>0$ and $\tau_{\alpha}\geq0$, we have that $t_{\alpha}^E(\tau_{\alpha})\geq t^E(\tau_{\alpha})$ with equality only at $\tau_{\alpha}=0$. Thus $T_{\alpha}^E>T^{E}$.
\subsection*{Case 2 - Convergence to a nontrivial steady state for $\alpha=1/T^E$.}\hfill
For $\alpha=1/T^E$,
\begin{equation}
\label{case2time}
\lim_{\tau_{\alpha}\nearrow\tau^*}t_{\alpha}^E(\tau_{\alpha})=-T^E\lim_{\tau_{\alpha}\nearrow\tau^*}\left|1-\frac{t^E(\tau_{\alpha})}{T^E}\right|=+\infty
\end{equation}
and the exponential \eqref{exp} vanishes,
$$\lim_{\tau_{\alpha}\nearrow\tau^*}\operatorname{exp}\left\{-\alpha t_{\alpha}^E(\tau_{\alpha})\right\}=0.$$
To determine how fast we use \eqref{asymp} and \eqref{time2} to obtain
\begin{equation}
\label{exp2}
\operatorname{exp}\left\{-\alpha t_{\alpha}^E(\tau_{\alpha})\right\}
=\frac{1}{T^E}(T^E-t^E(\tau_{\alpha}))\sim C(\tau^*-\tau_{\alpha})(\ln(\tau^*-\tau_{\alpha})-1)^2
\end{equation}
for $\tau^*-\tau_{\alpha}>0$ small. Using \eqref{int1}-\eqref{int2} and \eqref{case2time}-\eqref{exp2} on \eqref{gammae}, we see that $\gamma_{\alpha}^E$ converges to a mean-zero nontrivial steady state as $t\to+\infty$,
\begin{equation*}
\label{ae}
\lim_{t\to+\infty}\gamma_{\alpha}^{E}(\bfX(\bfa,t),t)=
\begin{cases}
-C,\qquad & \bfa=\bfa_0,
\\
0,\qquad & \bfa\neq\bfa_0.
\end{cases}
\end{equation*}
\subsection*{Case 3 - Convergence to the trivial steady state for $\alpha>1/T^E$.}\hfill
For $\alpha>1/T^E$, \eqref{time2} implies the existence of $0<\tau_1<\tau^*$ such that $t^E(\tau_{\alpha})\nearrow\frac{1}{\alpha}<T^E$ as $\tau_{\alpha}\nearrow\tau_1$. Then $t_{\alpha}^E(\tau_{\alpha})\to+\infty$ as $\tau_{\alpha}\nearrow\tau_1$, and
$$\lim_{\tau_{\alpha}\nearrow\tau_1}\operatorname{exp}\left\{-\alpha t_{\alpha}^E(\tau_{\alpha})\right\}=0.$$
Since the spatial and integral terms in \eqref{gammae} stay positive and finite for $0\leq\tau_{\alpha}\leq\tau_1$, it follows that $\gamma_{\alpha}^E(\bfx,t)\to0$ as $t\to+\infty$ for all $\bfx\in Q$.
\end{proof}
Before proving Theorem \ref{dampedboussinesqthm} we establish the following Lemma.
\begin{lemma}
\label{lemma}
Let $\rho_0(\bfx)\geq0$ and set $\tau^*=-1/m_0$. If $0\leq\tau_{\alpha}(t)<\tau^*$ for all $t\in\Xi\equiv[0,T)$, $0<T\leq+\infty$, then $0<\phi_1(t)<+\infty$ on $\Xi$.
\end{lemma}
\begin{proof}
Suppose $\rho_0(\bfa)\geq0$ and $0\leq\tau_{\alpha}(t)<\tau_*$ for all $t\in\Xi\equiv[0,T)$, $0<T\leq+\infty$. The latter and \eqref{psi} imply that $\Psi(\bfa,t)=1+\gamma_0(\bfa)\tau_{\alpha}(t)>0$ for all $(\bfa,t)\in Q\times\Xi$. Since $\phi_1(t)$ satisfies $\phi_1(0)>0$, there exists, by continuity, a positive $t_1\in\Xi$ such that $0<\phi_1(t)<+\infty$ for all $t\in[0,t_1)\subset\Xi$. Note that $\phi_1(t)$ cannot diverge to positive infinity at $t_1$. Indeed, suppose
\begin{equation}
\label{phiinf}
\lim_{t\nearrow t_1}\phi_1(t)=+\infty.
\end{equation}
Since $\rho_0(\bfa)\geq0$, $\Psi(\bfa,t)>0$ for all $(\bfa,t)\in Q\times[0,t_1]$, and $\sigma(t)\leq0$ on $[0,t_1)$ (by \eqref{sigma}ii)),
$$\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)\geq\Psi(\bfa,t)>0$$
for all $(\bfa,t)\in Q\times[0,t_1)$, and thus
\begin{equation}
\label{upsidedown}
\phi_1^{-1}(t)\geq\left(\int_Q{\frac{d\bfa}{\Psi(\bfa,t)}}\right)^{-1}>0
\end{equation}
for $t\in[0,t_1)$. From \eqref{phiinf} and \eqref{upsidedown} it follows that
$$\lim_{t\nearrow t_1}\int_Q{\frac{d\bfa}{\Psi(\bfa,t)}}=+\infty,$$
contradicting $\Psi(\bfa,t)>0$ for all $(\bfa,t)\in Q\times\Xi$.
Now suppose there exists $t_2\in\Xi$ such that $\phi_1(t)$ vanishes as $t\nearrow t_2$, namely,
\begin{equation}
\label{phi0}
\lim_{t\nearrow t_2}\int_Q{\frac{d\bfa}{\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)}}=0.
\end{equation}
Since $\rho_0(\bfa)\geq0$ and $0<\phi_1(t)<+\infty$ on $[0,t_2)\subset\Xi$, $\sigma(t)\leq0$ is continuous on $[0,t_2)$. Then boundedness of $\tau_{\alpha}$ on $[0,t_2]$ and $\Psi(\bfa,t)-\rho_0(\bfa)\sigma(t)>0$ for all $(\bfa,t)\in Q\times[0,t_2)$ imply that, for \eqref{phi0} to hold, $\sigma(t)\to-\infty$ as $t\nearrow t_2$. From this and \eqref{sigma}i) it follows that
\begin{equation*}
\label{a}
\lim_{t\nearrow t_2}\left\{\tau_{\alpha}(t)\int_0^t{e^{\alpha s}\phi_1(s)ds}\right\}=+\infty,
\end{equation*}
or since $0\leq\phi_1<+\infty$ on $[0,t_2]$, $\tau_{\alpha}(t)\to+\infty$ as $t\nearrow t_2$, contradicting $0\leq\tau_{\alpha}<\tau^*$ on $\Xi$.
\end{proof}
\begin{proof}[\textbf{Proof of Theorem \ref{dampedboussinesqthm}}]
Suppose $\gamma_0(\bfx)$ satisfies the conditions in Theorem \ref{blowupundamped} and $\rho_0(\bfx)\geq0$ for all $\bfx\in Q$. Further, and without loss of generality, assume there is one location $\bfa_1\in Q$ such that $\rho_0(\bfa_1)=0$ and $\gamma_0(\bfa_1)=m_0$. Set $\tau_*=-1/m_0$. Then Lemma \ref{lemma} and formulas \eqref{sigma} and \eqref{dampedjacobianb} imply that, as long as $0\leq\tau_{\alpha}<\tau_*$,
\begin{equation}
\label{jacest1}
J(\bfa,t)\geq\frac{1}{1+\gamma_0(\bfa)\tau_{\alpha}(t)-\rho_0(\bfa)\sigma(t)}\left(\int_Q{\frac{d\bfa}{1+\gamma_0(\bfa)\tau_{\alpha}(t)}}\right)^{-1}
\end{equation}
for all $\bfa\in Q$. Setting $\bfa=\bfa_1$ on \eqref{jacest1} and using \eqref{int1}, it follows that
\begin{equation*}
\label{jacest2}
J(\bfa_1,t)\geq\frac{1}{1+m_0\tau_{\alpha}(t)}\left(\int_Q{\frac{d\bfa}{1+\gamma_0(\bfa)\tau_{\alpha}(t)}}\right)^{-1}\sim-\frac{C}{(\tau^*-\tau_{\alpha})\ln(\tau^*-\tau_{\alpha})}
\end{equation*}
for $\tau^*-\tau_{\alpha}>0$ small, and therefore
\begin{equation*}
\label{jacest4}
J(\bfa_1,t)\to+\infty
\end{equation*}
as $\tau_{\alpha}\nearrow\tau^*$. Next, for $0<\alpha<1/T^E$ and $T^E>0$ the finite blowup time of the undamped Euler equation \eqref{undampedgammaeqn}, we prove the existence of a finite $T_{\alpha}^B>0$ such that $t\nearrow T_{\alpha}^B$ as $\tau_{\alpha}\nearrow\tau^*$. From \eqref{phib}, \eqref{tauivpb} and Lemma \ref{lemma},
\begin{equation}
\label{timeb1}
e^{-\alpha t}dt\leq\left(\int_Q{\frac{d\bfa}{1+\gamma_0(\bfa)\tau_{\alpha}}}\right)^2d\tau_{\alpha}
\end{equation}
for $0\leq\tau_{\alpha}<\tau^*$. Denote the damped Boussinesq time variable $t=t_{\alpha}^B(\tau_{\alpha})$. Then integrating \eqref{timeb1} between 0 and $ t_{\alpha}^B$ yields the upper-bound
\begin{equation}
\label{timeb2}
t_{\alpha}^B(\tau_{\alpha})\leq-\frac{1}{\alpha}\ln\left|1-\alpha t^E(\tau_{\alpha})\right|=t_{\alpha}^E(\tau_{\alpha})
\end{equation}
for $t_{\alpha}^B$ in terms of the damped Euler time variable $t_{\alpha}^E$, which in turn depends on the undamped Euler time variable $t^E$ in \eqref{undampedeulertime} and the damping coefficient $\alpha>0$. Since $\gamma_0$ satisfies the conditions of Theorem \ref{blowupundamped}, the limit $T^E\equiv\lim_{\tau_{\alpha}\nearrow\tau^*}t^E(\tau_{\alpha})$ is positive and finite. Suppose $0<\alpha<1/T^E$. Then Theorem \ref{dampedeulerthm} implies that $T_{\alpha}^E\equiv\lim_{\tau_{\alpha}\nearrow\tau^*}t_{\alpha}^E(\tau_{\alpha})$ is also positive and finite. Thus letting $\tau_{\alpha}\nearrow\tau^*$ in \eqref{timeb2}, we see that the blowup time $T_{\alpha}^B>0$ is finite and satisfies $$T_{\alpha}^B\equiv\lim_{\tau_{\alpha}\nearrow\tau^*}t_{\alpha}^B(\tau_{\alpha})\leq T_{\alpha}^E.$$
This finishes the proof of the first part of the Theorem. For the second part we adapt an argument used in \cite{childress, sarria2} to construct blowup and respectively global infinite-energy solutions of the 2D Euler and inviscid Boussinesq equations.
Let $\rho_0(\bfx)=-\sin^2(2\pi x)$. Then we look for a solution of \eqref{ode} satisfying $\xi(\bfx,0)\equiv1$ and $\xi_t(\bfx,0)=\gamma_0(\bfx)$. Suppose $\mu_1(t)$ and $\mu_2(t)$ solve
\begin{equation}
\label{mueqn}
\mu_1''+\alpha\mu_1'-I(t)\mu_1=0,\qquad \mu_2''+\alpha\mu_2'-I(t)\mu_2=1
\end{equation}
with $\mu_1(0)=\mu_1'(0)=1$ and $\mu_2(0)=0$, $\mu_2'(0)\neq0$. Then
\begin{equation}
\label{mu}
\mu(\bfx,t)=\mu_1(t)+\rho_0(\bfx)\mu_2(t)
\end{equation}
satisfies
$$\mu_{tt}+\alpha\mu_t-I(t)\mu=\rho_0,\qquad \mu(\bfx,0)=1.$$
Note that $\mu_2'(0)\neq0$ must be such that $\gamma_0(\bfx)=\mu_t(\bfx,0)=1+\rho_0(\bfx)\mu_2'(0)$
has mean zero over $Q$, as required by \eqref{zeromean}. Now, since $J$ satisfies \eqref{mean1} it follows that
$$1=\int_Q{\frac{d\bfx}{\mu_1-\sin^2(2\pi x)\mu_2}},$$
which yields the relation
\begin{equation}
\label{relation}
\mu_2=\mu_1-\frac{1}{\mu_1}.
\end{equation}
Using \eqref{relation} we see that $\gamma_0(\bfx)=\cos(4\pi x)$, which satisfies the mean-zero condition \eqref{zeromean}. Next, using \eqref{relation} on \eqref{mueqn}ii) to eliminate the nonlocal term $I(t)$ in \eqref{mueqn}i) gives, after some simplification,
$$\frac{d}{dt}\left(\frac{\mu_1'}{\mu_1}\right)+\alpha\frac{\mu_1'}{\mu_1}=\frac{\mu_1}{2}.$$
Dividing both sides by $\mu_1$, differentiating the resulting equation, and then setting $N(t)=\mu_1'/\mu_1$ now leads to
\begin{equation}
\label{i}
\frac{d}{dt}\left(N'-\frac{1}{2}N^2+\alpha N\right)=\alpha N^2.
\end{equation}
Since $N(0)=1$ and $N'(0)=\frac{1}{2}-\alpha$, we integrate \eqref{i} and use a standard Gronwall-type argument to obtain
\begin{equation}
\label{z0}
z'\geq\frac{1}{2}e^{-\alpha t}z^2,\qquad z(0)=1
\end{equation}
for $z(t)=e^{\alpha t}N(t)$. Set
\begin{equation}
\label{timeneg}
T_{\alpha}^B=-\frac{1}{\alpha}\ln\left(1-2\alpha\right)
\end{equation}
for $0<\alpha<1/2$. Note that $T_{\alpha}^B$ is positive and finite. Then solving \eqref{z0} for $t\in[0,T_{\alpha}^B)$ gives
$$z(t)\geq\frac{2\alpha}{e^{-\alpha t}-(1-2\alpha)},$$
or since $z(t)=e^{\alpha t}N(t)=e^{\alpha t}\frac{d}{dt}\left(\ln\mu_1(t)\right)$,
\begin{equation}
\label{mu1}
\mu_1(t)\geq\frac{4\alpha^2}{\left(e^{-\alpha t}-(1-2\alpha)\right)^2}.
\end{equation}
From \eqref{mu1} it follows that $\mu_1\to+\infty$ as $t\nearrow T_{\alpha}^B$. Then by \eqref{mu} and \eqref{relation},
$$\frac{1}{J(\bfx,t)}\geq\cos^2(2\pi x)\mu_1(t)+\frac{\sin^2(2\pi x)}{\mu_1(t)},$$
which implies that $J(\bfx,t)\to0$ as $t\nearrow T_{\alpha}^B$ for $\bfx\in\mathcal{B}\equiv\{(x,y)\in Q\,|\, x\notin\{1/4,3/4\}\}$. Note that $\mathcal{B}$ are precisely the points where $\gamma_0(\bfx)=\cos(4\pi x)$ does not equal its negative minimum.
Lastly, by setting $\alpha=0$ in \eqref{i} it is easy to see that the Jacobian of the associated undamped Boussinesq system vanishes as $t\nearrow 2$, which agrees with $T^B_{\alpha}\to2$ in \eqref{timeneg} as $\alpha\to0$. Thus we may write $0<\alpha<1/2$ as $0<\alpha<1/T^B$ for $T^B=2$ the blowup time for the solution of the undamped Boussinesq system with the same initial data.
\end{proof}
Note that the blowup result in part 2 of Theorem \ref{dampedboussinesqthm} is effectively one dimensional in the sense that the initial data depends only on one of the two coordinate variables. Currently we do not know if this particular blowup is suppressed for $\alpha\geq1/T^B$, or if a solution of the damped Boussinesq system \eqref{dampedeqnb}-\eqref{nonlocalb} persist globally in time for $\alpha\geq1/T^E$ when the initial data satisfies the conditions of part one of Theorem \ref{dampedboussinesqthm} \emph{and} is such that the associated undamped solution blows up in finite time.
\section{Acknowledgments}
This work was partially supported by the Div III \& P Research Funding Committee at Williams College.
\makeatletter \renewcommand{\@biblabel}[1]{\hfill#1.}\makeatother
|
\section{Introduction}
Understanding the nature of nonlinear interactions is at the core of the
problem of turbulence. It has been known for quite some time that the
transfer of energy between scales in a turbulent flow involves groups
of three spatial modes, namely a {\it triad} $({\bf k}, {\bf p}, {\bf
q})$ such that ${\bf k} = {\bf p} + {\bf q}$. The concept was first
introduced by \cite{Kraichnan58}, where he showed that triadic
interactions are conservative, and later used this representation to
formulate his Direct Interaction Approximation theory. Later,
\cite{Lee75} analysed triads according to geometric arguments. The
subsequent growth of computing power allowed analysis of triadic
interactions in direct numerical simulations, and \cite{Domaradzki90a}
showed that while energy transfer in the inertial range is mostly
local in wave number space, nonlocal triads (i.e., triads involving
modes with disparate wave numbers) can have large amplitudes (although
they are less numerous, and thus the flux is dominated by the local
triads, see \citet{Eyink09,Aluie09}). An important result was obtained
by \cite{Waleffe92}, who decomposed the velocity field in terms of
helical modes, and also analysed locality aspects of the nonlinear
interactions. This allowed him to identify which triads, when
isolated, contribute to a direct energy transfer (energy going from
large to small scales), and which to an inverse transfer (energy going
from small to large scales).
The concept of triadic interactions remains relevant to present
date. Nonlinear interactions were analysed by
\cite{Mininni06,Mininni08}, by studying local and nonlocal triads, and
the shell-to-shell energy transfer functions in wavenumber
space. Recently, \cite{Cheung14} formulated an exact representation of
the nonlinear triads using a combination matrix, and were able use it
along with minimal assumptions to obtain the Kolmogorov spectrum from
the Navier-Stokes equation. The helical decomposition of
\cite{Waleffe92} was also used recently to build ``decimated''
versions of the Navier-Stokes equation \citep{Biferale13}, where the
nonlinear terms in the equation are split into contributions from
each kind of triad, and which can then be turned on or off to see how
they affect the energy cascade. In a similar way, \cite{Moffatt14}
analysed the effect of triad truncation on the velocity and vorticity
field of the Euler equation. Finally, decimation models in which
wave-wave-wave interactions and wave-vortex-wave interactions were
differentiated have been used to study rotating stratified
turbulence \citep{Remmel10,Hernandez-duenas14}.
In flows with restitutive forces and in which waves can be present
(e.g., in rotating and/or stratified flows, or in magnetohydrodynamics),
an important concept arises which is that of resonants triads. These are
triads $({\bf k},{\bf p},{\bf q})$ which also satisfy the resonant
condition $\omega ({\bf k}) = \omega ({\bf p}) +\omega ({\bf q})$, with
$\omega ({\bf k})$ being the dispersion relation of the waves. If a flow
is dominated by rapidly varying waves, non-resonant interactions should,
in principle, die out in front of resonant ones, thus leaving the bulk
of the nonlinear energy transfer to the resonant triads. This has been
exploited in theories of weak turbulence (i.e., in systems in which the
flow is completely given by a superposition of interacting dispersive
waves), as done for interfacial waves in fluids or for waves in plasmas
\citep{Zakharov,Nazarenko} with varying degrees of success
\citep{Newell11}. Experimental evidence of such resonant wave
interactions has been found, e.g., in capillary wave turbulence
\citep{Aubourg15,Aubourg16} and in gravity-capillary waves
\citep{Haudin16}.
The Coriolis force in rotating flows gives rise to inertial waves which
in experiments and in simulations coexist with eddies
\citep{Staplehurst08,Bokhoven08}. As a result, although weak turbulence
theories can give some insight into the energy transfer mechanisms
\citep{Galtier03,Nazarenko11}, more general formulations of wave
turbulence in the strong regime are needed to describe the flow
\citep{Cambon89,Cambon97}. The first attempts to study resonant triads
in these flows were carried out by \cite{Newell69}, who studied how
these triads become the preferred energy transfer mechanism and its
implication for the formation of planetary zonal flows. Extensions of
Rapid Distortion Theory and of the Eddy-Damped Quasi-Normal Markovian
closure to rotating turbulent flows rely heavily on resonant
interactions, and can correctly capture the development of anisotropy in
rotating turbulence \citep{Cambon89,Cambon97,Bellet06}. Moreover, this
approach is useful to understand how the flow becomes quasi-two
dimensional, with energy in three-dimensional modes being transferred
preferentially towards modes with smaller vertical wavenumber through a
subset of the resonant triads. Within the framework of the helical
decomposition, \cite{Waleffe93} also considered the resonant triads, and
pointed out that a parametric instability may be the mechanism behind
the preferential transfer of energy towards quasi-two dimensional modes:
the resonant condition $\omega ({\bf k}) = \omega({\bf p}) + \omega
({\bf q})$ is more easily satisfied by modes with small vertical
wavenumber, thus being preferred by the nonlinear coupling. This
tendency in the energy transfer towards quasi-two dimensionalisation has
been confirmed both in numerical simulations \citep{Sen12,Horne13} and
in laboratory experiments \citep{Campagne15}. However, the parametric
instability mechanism of \citet{Waleffe93} is valid for isolated
triads; in a real turbulent flow, in which each triad is coupled to a
miriad of other triads, it is unclear whether this is the actual
mechanism responsible for the quasi-two dimensionalisation.
Moreover, wave turbulence theories are valid when the wave period is
much shorter than the eddy turnover time. As a result, many of these
arguments fail when the Rossby number is moderate, or when the
vertical wavenumber approaches zero (as there are no waves in these
modes), and thus they cannot predict whether energy is transferred
into pure two-dimensional modes. Also, wave turbulence theories are
inhomogeneous in scale space, and even for small Rossby numbers there
can exist a sufficiently small scale such that the time scales of the
eddies and of the waves become of the same order thus violating its
hypothesis \citep{Pouquet10,Mininni12}. In recent years, several
efforts were made to detect inertial waves in rotating turbulence,
quantify their energy, and identify their role in the anisotropic
transfer of energy. Some relied on the fact that bounded domains have
resonant frequencies which can be spotted in a temporal spectrum
\citep{Bewley07,Rieutord12,Lamriben11}. Others analysed temporal
decorrelation functions to determine at which scales wave action was
predominant \citep{Favier10,Clark14}. \cite{Campagne15} identified the
presence of inertial waves by analysing the two-point spatial
correlation of the time transformed velocity fields obtained from PIV
measurements. Also, the space and time resolved energy spectrum
was calculated both numerically \citep{Clark14,Clark15b} and
experimentally \citep{Yarom14,Campagne15}. While all this evidence
points to a strong presence of waves in rotating flows, and thus of
resonant interactions, studies of the contribution of resonant and
near-resonant triads in rotating turbulence are scarce.
Recently, experimental evidence of three-wave resonant interactions has
been found in a rotating flow \citep{Bordes12}. In numerical
simulations, \cite{Chen05} compared rotating flows computed in grids of
$128^3$ spatial points with simulations of the Navier-Stokes equation in
two dimensions, and concluded that resonant triads play a more dominant
role as rotation is increased, while they also raised concerns on the
validity of wave turbulence arguments for the long time dynamics of the
flow. Following \cite{Waleffe92}, \cite{Smith05} considered numerical
simulations of truncated systems in which only some interactions were
preserved, to identify which triads were responsible for the development
of anisotropy. The authors concluded that near-resonant interactions
were needed to reproduce the quasi-two dimensionalisation of the flow,
while non-resonant triads reduce this anisotropic transfer. More
recently, \cite{Alexakis15} analysed a large numerical dataset of
rotating flows and concluded that the dynamics of the quasi-two
dimensional component of the flow can only be correctly captured if
near-resonant and non-resonant interactions are taken into account.
Also recently, \cite{Gallet15} showed that two-dimensional flows are
preferred solutions of rotating flows for small enough Rossby number,
indicating that a description of the energy transfer solely in terms
of resonant triads has limitations even in the limit of very strong
rotation. The role of near-resonant interactions is also
important to understand the limit of infinite domains, see
\citet{Cambon04,Chen05,Bourouiba08} for discussions. However,
direct measurements of resonant interactions in turbulent flows are
hard to find, owing primarily to the massive amount of data that needs
to be extracted and analysed from either experiments or numerical
simulations.
The aim of this paper is to directly quantify how different triads
contribute to the energy transfer in rotating turbulence. It is worth
mentioning that a similar analysis was performed recently on
experimental data of gravity-capillary waves measured on the surface of
a liquid \citep{Aubourg16}, where interactions are also between
three waves. The analysis, based on phenomenological arguments,
allowed direct identification of resonant interactions. Here we
develop a theoretical formalism for three-waves interactions in a
rotating flow that allows explicit derivation of third order
correlation functions between modes. We do this by deriving a
{\it contribution function}, a function that measures the
contribution of each triad to the total energy transfer as a function
of the wavenumber and frequency, and a {\it normalised contribution
function} that measures the characteristic timescale at which an
interaction takes place. Both allow the measurement of how relevant
and how well tuned (i.e., how resonant) a given triad is. We then use
these tools to analyse results from direct numerical simulations of
rotating turbulence. The formalism can be extended to other systems
with three or more wave interactions.
We start in Sec.~\ref{theoback} with a brief explanation of the nature
of nonlinear interactions in the Navier-Stokes equations, followed by
a description of our numerical simulations. Then, in
Sec.~\ref{explicacion} we derive the aforementioned contribution
function, which we then use to analyse the data from numerical
simulations of rotating turbulence in Sec.~\ref{numresults}. Finally,
in Sec.~\ref{conclusions} we present our conclusions.
\section{Resonant triads}
\label{theoback}
\subsection{Nonlinear interactions in Navier-Stokes}
In a rotating frame, the Navier-Stokes equations for an incompressible
fluid with velocity ${\bf u}$ and under the action of a mechanical
forcing ${\bf F}$ read
\begin{gather}
\frac{\partial {\bf u}}{\partial t} = - ({\bf u} \cdot {\bm \nabla})
{\bf u}
- 2 \Omega \hat{z} \times {\bf u}
- {\bm \nabla} {\cal P} + \nu \nabla^2 {\bf u} + {\bf F} ,
\label{momentum}
\\
{\bm \nabla} \cdot {\bf u} =0 ,
\label{incompressible}
\end{gather}
where ${\cal P}$ is the total pressure (including the centrifugal
force, and normalised by the uniform fluid mass density), $\hat{z}$ is
parallel to the rotation axis, $\Omega$ is the rotation frequency, and
$\nu$ is the kinematic viscosity. The Reynolds number, defined as
usual as $\textrm{Re} = UL/\nu$ (where $U$ is the r.m.s.~velocity and
$L$ is the energy injection scale) quantifies the strength of the
nonlinear term against viscous damping.
Using the incompressibility condition given by
Eq.~(\ref{incompressible}), and for ${\bf F}=0$, the Fourier transform
of Eq.~\eqref{momentum} can be written as
\begin{equation}
\label{transformed}
\left( \frac{\partial}{\partial t}
- \nu k^2 + 2 \Omega \mathbb{P}_{\bf k} \hat{z} \times \right) {\bf u_k}
= -i \mathbb{P}_{\bf k} \sum_{{\bf p}+{\bf q}={\bf k}} ({\bf u_p}
\cdot {\bf q}) {\bf u_q} ,
\end{equation}
where $[\mathbb{P}_{\bf k}]_{ij} = \delta_{ij} - k_i k_j /k^2$
is the projector operator, which projects in the direction
perpendicular to ${\bf k}$ to enforce incompressibility. All
terms on the l.h.s.~of the equation are linear in ${\bf u_k}$ and do
not couple modes with different ${\bf k}$. So, after transforming
the nonlinear term in Eq.~\eqref{momentum}, the resulting
convolution on the r.h.s.~of this equation tells us that only modes
${\bf p}$ and ${\bf q}$ in triads satisfying
${\bf k}={\bf q} + {\bf p}$ can give or receive energy from the mode
with wave vector ${\bf k}$. Note this is not a unique property
of the Navier-Stokes equation but of any nonlinear equation with
quadratic nonlinearities.
\subsection{Rotating flows and relevant time scales}
In the presence of rotation (or of other restitutive forces), waves
can be excited with well defined frequencies for each wave vector,
given by the dispersion relation of the waves $\omega({\bf k})$. For a
rotating flow, the dispersion relation of inertial waves is
\begin{equation}
\omega({\bf k}) = \pm \frac{2 \Omega k_z}{k}.
\label{dispersion}
\end{equation}
We can then define the Rossby number as
\begin{equation}
\textrm{Ro} = \frac{U}{2L \Omega} ,
\end{equation}
which measures the ratio of the rotation period to the turnover time
of the large-scale eddies. As a result, for small Rossby number we can
expect waves to be faster than eddies, at least for a range of
scales. Indeed, in the absence of forcing and viscous effects,
Eqs.~\eqref{momentum} and \eqref{incompressible} have waves with
dispersion relation \eqref{dispersion} as exact solutions.
We can thus define several relevant time scales, as in a turbulent
flow one can identify different characteristic times for each
possible interaction. The sweeping of the small scale eddies
(of size $\sim 1/k$) by the large scale flow is described by the
{\it sweeping time}
\citep{Chen89}
\begin{equation}
\tau_S \sim \frac{1}{U k} .
\label{sweeping}
\end{equation}
Note that sweeping does not result in a transfer of energy
across scales. Sweepping corresponds to the advection of the small
scale eddies by a large scale flow, and can act even in the absence
of a mean flow (i.e., just a random evolution of the modes at large
scales can result in {\it random sweeping}). The advection of the
small scale eddies in real space corresponds to a rotation of the
Fourier modes by $e^{iUkt}$. The interaction of similar sized
eddies, which result in nonlinear transfer of energy, is described
by the {\it nonlinear time scale},
\begin{equation}
\tau_{NL} \sim \frac{1}{k\sqrt{k E(k)}} ,
\label{nonlintime}
\end{equation}
where $E(k)$ is the energy spectrum of the flow. Finally, the time
scale for the interaction of waves modes is expected to be proportional
to the wave period, i.e.,
\begin{equation}
\tau_\omega \sim \frac{1}{|\omega({\bf k})|} .
\label{wavetime}
\end{equation}
While in homogenous and isotropic turbulence the dominant Eulerian
time is the sweeping time \citep{Chen89}, when waves are present the
dominant time scale can be either the sweeping time, the nonlinear
time, or the wave period depending on which is the fastest at a given
scale \citep{Favier10,Servidio11,Clark14,Clark15b}. As a result, these
time scales imply that depending on the shape of the energy
spectrum $E(k)$, the approximation that waves are faster than the
eddies for small enough $\textrm{Ro}$ may break down for sufficiently
large wave numbers, or for a subset of the Fourier modes (e.g., for
modes with small vertical wavenumber). Below we present a more
detailed estimation of which is the dominant time scale for each mode
in our numerical simulations.
\subsection{Resonant interactions}
For modes for which the waves are much faster than the eddies, we can
assume wave dynamics dominate the evolution, while the eddies
contribute to a slow modulation of the amplitude of the waves. Thus,
we can write ${\bf u_k} = {\bf U_k} e^{i \omega_{\bf k} t}$. In
practice, this approximation is done after decomposing the modes
${\bf u_k}$ into the helical eigenstates ${\bf h}_\pm ({\bf k})$ of
the linearised Eq.~\eqref{momentum},
${\bf u_k} = a_+({\bf k}) {\bf h}_+ ({\bf k}) +
a_-({\bf k}) {\bf h}_- ({\bf k})$,
where the subindices $+$ and $-$ correspond to the two possible
polarisation of the waves, see e.g., \cite{Waleffe93}. However, for
the purpose of the following discussion it is better to work in terms
of ${\bf u_k}$, as those modes are more easily accessed in numerical
simulations.
Replacing in Eq.~\eqref{transformed} we obtain
\begin{equation}
\left( \frac{\partial}{\partial t} - \nu k^2 +
2 \Omega \mathbb{P}_{\bf k} \hat{z} \times \right) {\bf U_k}
= -i \mathbb{P}_{\bf k} \sum_{{\bf p}+{\bf q}={\bf k}} ({\bf U_p}
\cdot {\bf q}) {\bf U_q}
e^{-i [\omega (\bf k) - \omega ({\bf p}) - \omega ({\bf q})]t}.
\end{equation}
Integrating over several periods of the waves, the nonlinear term can
give a non-negligible energy transfer only if triads are resonant,
i.e., if $\omega ({\bf k}) = \omega ({\bf p}) + \omega ({\bf q})$. In
practice, near-resonant triads with
\begin{equation}
\gamma_r ({\bf k}, {\bf p}, {\bf q}) =
\frac{
\min\{|\omega ({\bf k}) \pm \omega ({\bf p}) \pm \omega
({\bf q})|\}}
{2 \Omega} =
\cal{O}(\textrm{Ro}) ,
\end{equation}
are also expected to be relevant \citep[see, e.g.,][]{Alexakis15}. We
will call $\gamma_r$ the resonance factor, as it measures how close to
resonance a given triad is in the framework of wave turbulence
theory. The minimum and the plus-minus signs added in the latter
equation are due to the fact that our modes ${\bf u_k}$ mix both
polarisations of the inertial waves.
\section{Contribution of nonlinear triads to the energy transfer and
to the eddy decorrelation}
\label{explicacion}
In the traditional picture of turbulence, energy is transferred
towards smaller scales as the eddy gets sufficiently deformed (and
thus, decorrelated in time) by the interaction with other eddies. As
we are interested in understanding the role of the waves in the energy
transfer, we need an expression for the contribution of each triad to
the decorrelation of individual modes (and thus, to the distribution
of energy per wavenumber). To do this we define
${\bf u_k} = {\bf u_k}(t)$ and ${\bf u'}_{\bf k} = {\bf u_k} (t')$
with $t'=t-\tau$, and multiply Eq.~\eqref{transformed} by
${\bf u'}^*_{\bf k}$. After averaging over the time $t'$ and assuming
the system is in a turbulent steady state, we obtain
\begin{equation}
\left( \frac{\partial}{\partial t} - \nu k^2 \right)
\left< {\bf u'}^*_{\bf k} \cdot {\bf u_k} \right>_{t'}
+ 2 \Omega \left< {\bf u'}^*_{\bf k} \cdot
( \hat{z} \times {\bf u_k} ) \right>_{t'} =
-i \sum_{{\bf p}+{\bf q}={\bf k}} \left< {\bf u'}^*_{\bf k}
\cdot ({\bf u_p} \cdot {\bf q}) {\bf u_q} \right>_{t'} ,
\label{firststep}
\end{equation}
where the projector $\mathbb{P}_{\bf k}$ was dropped as the dot
product with ${\bf u'}^*_{\bf k}$ ensures only components of the
terms perpendicular to ${\bf k}$ survive (as
${\bf k} \cdot {\bf u'}^*_{\bf k} = 0 $ from the incompressibility
condition). In Eq.~(\ref{firststep}) we also assume that the complex
conjugate is added (this must be assumed in all the following
equations).
The viscous term on the l.h.s.~of Eq.~(\ref{firststep}) is just
responsible for damping of the correlations in a viscous time scale
which grows as the Reynolds number. Thus, for large Reynolds numbers
its effect can be neglected in comparison to the wave and the
nonlinear time scales. After assuming the system is in a turbulent
steady state, we can then rewrite this equation in terms of
functions that depend only on the time lag $\tau$ as
\begin{equation}
\frac{\partial}{\partial \tau} \Gamma_{\bf k}(\tau)
+ 2\Omega \Delta_{\bf k}(\tau) =
-i \sum_{{\bf p}+{\bf q}={\bf k}} \Theta({\bf k},{\bf q},{\bf p},\tau)
\label{eqgamt}
\end{equation}
where
\begin{equation}
\Gamma_{\bf k} (\tau) = \langle {\bf u}^*_{\bf k}(t') \cdot {\bf
u_k}(t'+\tau) \rangle_{t'}
\end{equation}
and
\begin{eqnarray}
\Delta_{\bf k} (\tau) &=& \langle {\bf u}^*_{\bf k}(t') \cdot
\left[ \hat{z} \times {\bf u_k}(t'+\tau) \right] \rangle_{t'} \\
{} &=&
\langle u_y^*({\bf k}, t') u_x({\bf k},t'+\tau) \rangle_{t'}
- \langle u_x^*({\bf k},t') u_y({\bf k},t'+\tau) \rangle_{t'}
\end{eqnarray}
are time correlation functions for the mode ${\bf k}$ (with $\Gamma$
the usual time correlation function used in isotropic and homogeneous
turbulence), and the third-order time correlation is
\begin{equation}
\Theta ({\bf k},{\bf p},{\bf q},\tau) = \langle {\bf u}^*_{\bf k} (t') \cdot
\left[ {\bf u_p}(t'+\tau) \cdot {\bf q} \right] {\bf u_q}(t'+\tau)
\rangle_{t'} .
\label{thetadef}
\end{equation}
The function $\Delta_{\bf k}$ is zero for $\tau = 0$, and can be
removed from these equations for all time lags if correlation
functions are written for the amplitudes of the helical eigenstates
${\bf h_\pm}({\bf k})$. Moreover, even in terms of the Fourier modes
of the velocity ${\bf u}_{\bf k}$, after adding the complex conjugate
and assuming the system is in a turbulent steady state (i.e., that the
statistical properties of the signals are homogeneous in time), this
function can be neglected.
In a turbulent flow, the correlation function $\Gamma_{\bf k}$ is thus
expected to decrease to $1/e$ of its value at $\tau=0$ on a
timescale that may be either $\tau_S$, $\tau_{NL}$, or
$\tau_\omega$. This decorrelation results from the interaction with
all triads, with the contribution from each triad measured by the
triple correlation $\Theta ({\bf k},{\bf q},{\bf p},\tau)$. Thus,
computation of this function should allow identification of the
dominant interactions responsible for the energy cascade discriminated
by time scale. Note also that for $\tau =0$, $\Theta$ reduces to the
usual transfer function $T({\bf k},{\bf p},{\bf q})$ that measures the
strength of each individual triad
\citep{Kraichnan58,Domaradzki90a,Waleffe92,Mininni11b}.
As the Fourier transform of the correlation function is proportional
to the power spectrum, we have
\begin{equation}
\widehat{\Gamma_{\bf k} (\tau)} = 2 E({\bf k},\omega) ,
\end{equation}
and as from the property of derivatives of Fourier transformed
functions
\begin{equation}
\widehat{\frac{\partial}{\partial \tau} \Gamma_{\bf k}(\tau)} =
-2 i \omega E({\bf k},\omega) ,
\end{equation}
we thus arrive to
\begin{equation}
\label{ekwtheta}
2 \omega E({\bf k},\omega) = \sum_{{\bf p}+{\bf q}={\bf k}}
\widehat{\Theta} ({\bf k},{\bf q},{\bf p},\omega).
\end{equation}
Note that $\widehat{\Theta}$ quantifies how much each triad
$({\bf k}, {\bf p}, {\bf q})$ and each frequency $\omega$ contribute
to the space and time (four-dimensional) energy spectrum. Also, how
well tuned $\widehat{\Theta}$ is around a given $\omega({\bf k})$
can be used to identify how close to resonance a triad actually is. We
will thus call $\widehat{\Theta}$ the {\it contribution
function}.
We can gain further insight into the meaning of $\widehat{\Theta}$ by
studying the case of a fluid in which only waves are present. In this
particular case, we can write
${\bf u}_{\bf k} = {\bf U}_{\bf k} e^{i \omega_{\bf k} t}$, and we can
neglect any slow dependence of ${\bf U}_{\bf k}$ in time. Bearing
aside normalisation factors for simplicity, we have
\begin{eqnarray}
\widehat{\Theta} ({\bf k},{\bf q},{\bf p},\omega)
&=& \int_{-\infty}^{\infty} e^{i \omega \tau}
\left< {\bf u}^*_{\bf k} (t') \cdot \left[ {\bf
u_p}(t'+\tau) \cdot {\bf q} \right] {\bf
u_q}(t'+\tau) \right>_{t'} \mathrm{d} \tau
\nonumber \\
&=& \int_{-\infty}^{\infty} e^{i (\omega +\omega_{\bf p} +
\omega_{\bf q}) \tau} \left< {\bf U}^*_{\bf k}
\cdot \left( {\bf U_p} \cdot {\bf q}\right) {\bf
U_q} \, e^{-i (\omega_{\bf k} -\omega_{\bf p} -
\omega_{\bf q}) t'}\right>_{t'} \mathrm{d} \tau
\nonumber \\
&=& {\bf U}^*_{\bf k} \cdot ({\bf U_p} \cdot {\bf q}) {\bf U_q}
\, \delta(\omega-\omega_{\bf k}) .
\label{eq:delta}
\end{eqnarray}
So in this case $\widehat{\Theta}$ only contributes to the frequency
$\omega = \omega({\bf k})$, and thus only resonant triads contribute
to $E({\bf k},\omega)$. In practice $\widehat{\Theta}$ will not always
be sharply peaked around $\omega({\bf k})$, as shown below. The width
of the peak can therefore be used to quantify how resonant a triad is.
It is much easier, both conceptually and practically, to work with a
symmetrised $\widehat{\Theta}$, namely
\begin{equation}
\widehat{\Theta}^S ({\bf k},{\bf q},\omega) = \frac{1}{2} \left[
\widehat{\Theta} ({\bf k},{\bf q},{\bf p}={\bf k}-{\bf q},\omega) +
\widehat{\Theta} ({\bf k},{\bf p}={\bf k}-{\bf q},{\bf q},\omega)
\right].
\end{equation}
From here on after every mention of $\Theta$ and its Fourier transform
will be in this symmetrised form. The superscript $S$ shall therefore
be dropped.
\section{Numerical results}
\label{numresults}
\begin{figure}
\centering
\includegraphics[scale=0.5]{fig1.png}
\caption{Contour levels of the energy spectrum as a function of
parallel and perpendicular wave numbers, in the simulation of
rotating turbulence with $\Omega=8$. Note the anisotropy of the
spectrum, with most of the energy accumulating near modes with
$k_\parallel \approx 0$. The white region corresponds to modes
with $\tau_\omega<\tau_s<\tau_{NL}$ (i.e., modes dominated by
the waves), the light grey region to modes with
$\tau_s<\tau_\omega<\tau_{NL}$, and the dark grey region to
modes with $\tau_s<\tau_{NL}<\tau_{\omega}$ (see text for
details). Grey regions thus correspond to modes for which
sweeping gives the fastest time scale (with dark grey indicating
``slow'' modes). The location of four modes relevant for the
analysis are marked in the figure: the ``wave'' modes
${\bf k}=(0,0,8)$ (marked with a blue triangle) and
${\bf k}=(0,5,5)$ (marked with a red star), a ``slow''
(two-dimensional) mode in the dark grey region
${\bf k}=(0,30,0)$ (marked with a cyan circle), and the mode
${\bf k}=(0,0,45)$ in the light grey region (marked with a green
square).}
\label{ekk}
\end{figure}
\subsection{Numerical simulations}
The code GHOST \citep{Gomez05,Mininni11} is used to solve
Eqs.~\eqref{momentum} and \eqref{incompressible} using a parallel
pseudo-spectral method with a second order Runge-Kutta scheme for the
time evolution. The $2/3$-rule is used for dealiasing. As will be seen
below, computation of the contribution function requires high cadence
I/O in time, and a significant amount of storage (note spatial
information needs to be saved with twice the frequency of the fastest
waves in the system). As a result, only simulations with moderate
resolution can be performed. Here we present two simulations using
grids of $N^3=512^3$ points in a three-dimensional periodic box.
Both simulations are identical except for the value of $\Omega$. In
one of the simulations $\Omega=4$, while in the other $\Omega=8$. The
simulations were started from the fluid at rest, and energy was
injected via a mechanical forcing. We chose a Taylor-Green forcing of
the form
\begin{eqnarray}
{\bf F} &=& F_0 \left[ \sin(k_{\textrm{TG},x} x)
\cos(k_{\textrm{TG},y} y)
\cos(k_{\textrm{TG},x} z) \hat{x} \right.
\nonumber \\
{} && \left. -\cos(k_{\textrm{TG},x} x)
\sin(k_{\textrm{TG},y} y)
\cos(k_{\textrm{TG},z} z) \hat{y} \right] ,
\end{eqnarray}
with $F_0 = 0.277$, ${\bf k}_\textrm{TG} = (1,1,1)$ (which results in
$L=2\pi/k_\textrm{TG}=2\pi/\sqrt{3}$), and $\nu=6.5\times10^{-4}$ in
dimensionless units (for unit velocity and a box of length $2\pi$).
While other forcings will presumably produce similar results,
Taylor-Green forcing was chosen because it has been reported to
result in a larger amplitude of wave modes when compared with
random-in-time isotropic forcing \citep{Clark14}. The system was
let to reach a turbulent steady state with $U\approx0.9$, which
translates to a Reynolds number of approximately 5000, and a Rossby
number of $0.03$ for $\Omega=4$ and of $0.015$ for $\Omega=8$.
Once in t he turbulent steady state the simulations were allowed to run
for over 12 large scale turn over times, the time span over which the
following analysis was carried on.
\subsection{Energy spectrum and decorrelation times}
\begin{figure}
\centering
\includegraphics[scale=0.6]{fig2.pdf}
\caption{Partial reconstruction of
$-{\partial \Gamma_{{\bf k}}}/{\partial \tau}$ via the sum over
${\bf q}$ of the third-order correlator
$\Theta({\bf k}, {\bf q}, \tau)$, from Eq.~\eqref{eqgamt}
(computation of the real part by adding the complex conjugate
is implied). The two panels show, from top to bottom, the
results for ${\bf k}=(0,30,0)$ and for ${\bf k}=(0,0,8)$. The
time lag $\tau$ is normalised by the dominant time scale
($\tau_s$ in the top panel, and $\tau_\omega$ in the bottom
panel); the different timescales in these units are also
shown as a reference by the horizontal bars in each figure.
We recover the expected behaviour for the correlation
functions, ${\bf k}=(0,0,8)$ (which is a fast mode) gets
locked to the wave period, while ${\bf k}=(0,30,0)$ (which is
slow) evolves in the sweeping time scale. The integral of
these functions gives the correlation function, which decays
rapidly on the dominant time scale.}
\label{thetat}
\end{figure}
\begin{figure}
\centering
\includegraphics[scale=0.6]{fig3.pdf}
\caption{$\vert \widehat{\Theta}({\bf k}=(0,0,8),{\bf
q},\omega)\vert$ as a function of $\omega$, for a fixed
fast mode ${\bf k}$, and for different values of ${\bf
q}$. Interactions with other wave modes (blue and red
curves) show a well tuned spectrum centred around the wave
frequency of the mode ${\bf k}$, indicating interactions are
close to resonance. On the other hand, interactions with
eddy modes (green and cyan curves) display a wider
spectrum. All subsequent analyses of the contribution
function focus on its maximum amplitude and
on how well tuned each triad is (i.e., on the width of the
peak around the maximum).}
\label{thetaw}
\end{figure}
Before proceeding to the analysis of the contribution function, we
first discuss some general properties of the simulations. In
Fig.~\ref{ekk} we show contour levels of the axisymmetric energy
spectrum for the simulation with $\Omega=8$, as a
function of the perpendicular and parallel wave numbers (with the
perpendicular and parallel directions defined with respect to the axis
of rotation). Whilst in isotropic turbulence one would expect these
contours to be circular, the effect of rotation in these flows impose
a clear anisotropy, with a preferred accumulation of energy in modes
with $k_\parallel \approx 0$ as predicted by
\cite{Cambon89}, \cite{Cambon97}, and \cite{Waleffe92}. Moreover, a
significant fraction of the energy is in modes with $k_\parallel = 0$,
for which resonant interactions cannot account for.
Three different regions are shaded in Fig.~\ref{ekk}. The white region
corresponds to modes with $\tau_\omega<\tau_s<\tau_{NL}$. This is the
region of ``fast'' (or ``wave'') modes, for which the period of the
waves is the fastest time scale. The grey region corresponds to modes
with $\tau_s<\tau_\omega<\tau_{NL}$. Although these modes are often
considered to be ``fast'', as shown in \cite{Clark14} these modes are
decorrelated in a timescale which is of the order of the sweeping
time. In other words, in the Eulerian frame, the dominant time scale
for these modes is given by the sweeping, which is the shortest
available time. Finally, in the dark grey region the modes have
$\tau_s<\tau_{NL}<\tau_{\omega}$. This is the region of ``slow''
modes for which the eddies are faster than the waves.
The three shaded regions are shown only as a reference. To
compute the value of $\tau_{NL}$ at each wave number using
Eq.~(\ref{nonlintime}), an estimation of $E(k)$ is needed. For
simplicity, instead of using the measured spectrum, we use the
phenomenological expression for non-helical rotating turbulence
$E(k)\sim \epsilon^{1/2} \Omega^{1/2} k^{-2}$
\citep{Zhou95,Muller07,Mininni12}. In \citet{Clark14} it was shown,
from direct computation of the decorrelation times, that this choice
results in a good estimation of the dominant time scale for modes
laying in the inertial range (i.e., at small and intermediate wave
numbers). At large wavenumbers, where the spectrum drops
exponentially as a result of viscous damping, $\tau_{NL}$ departs
from this estimation. However, we will not consider modes in the
viscous range, for which also the viscous damping time can be
relevant.
The ordering of the time scales described above has implications for
the behaviour of the time correlation function
$\Gamma_{\bf k}(\tau)$. In the white region of Fig.~\ref{ekk},
$\Gamma_{\bf k}(\tau)$ is expected to decay to $1/e$ of its value for
$\tau=0$ in one wave period of the mode with wave vector ${\bf k}$
\citep{Favier10,Clark14}. This time scale is what we define as the
decorrelation time: after this time, the mode ${\bf k}$ has
significantly decorrelated from its previous state. As already
mentioned, in the two grey regions the function
$\Gamma_{\bf k}(\tau)$ decays to $1/e$ of its value for $\tau=0$ in a
time equal to $\tau_s$ \citep{Clark14}. We will thus consider modes in
these three regions to compute the third order correlators
$\Theta({\bf k}, {\bf p}, {\bf q}, \tau)$ and
$\hat{\Theta}({\bf k}, {\bf p}, {\bf q}, \omega)$. In particular, in
Fig.~\ref{ekk} we indicate two fast modes ${\bf k}=(0,5,5)$ and
$(0,0,8)$, a ``swept'' mode $(0,0,45)$, and a slow mode
$(0,30,0)$. These modes will be used in several examples below.
\begin{figure}
\centering
\includegraphics[width=6.5cm]{fig4a.pdf}
\includegraphics[width=6.5cm]{fig4b.pdf}
\caption{Intensity (as a function of ${\bf q}$) of
the maximum of the contribution function,
$\max_\omega \{\vert \widehat{\Theta} ({\bf k},{\bf q},\omega)
\vert\}$. In each panel, ${\bf k}$ is fixed, and the maximum of
$|\widehat{\Theta}|$ is plotted for all available triads by
varying ${\bf q}$. Two ${\bf k}$ modes are considered,
{\it a)} ${\bf k}=(0,0,8)$ and {\it b)} $(0,5,5)$, both
dominated by waves. The black dots in the centre indicate the
modes ${\bf q}=\pm{\bf k}$. Two prominent features arise. One is
the effect of the anisotropy of the flow, as triads with larger
amplitudes are distributed along horizontal bands (i.e.,
coupling the ${\bf k}$ modes with modes with smaller
vertical wave numbers). The other is the defect along the line
${\bf q} = \alpha {\bf k}$, as collinear modes do not contribute
to the triads in an incompressible fluid.}
\label{peak_contribution}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=6.5cm]{fig5a.pdf}
\includegraphics[width=6.5cm]{fig5b.pdf}
\caption{Intensity of the peak values of the
normalised contribution function for each triad, given by
$\max_\omega \{\vert \widehat{\Theta} ({\bf k},{\bf q},\omega)
\vert\} /E(q_y,q_z)$. In each panel, ${\bf k}$ is
fixed for two modes dominated by waves: {\it a)}
${\bf k}=(0,0,8)$, and {\it b)} ${\bf k} = (0,5,5)$. The
dashed lines represent the modes with
$\tau_\omega = \tau_{NL}$. Modes with $\tau_\omega<\tau_{NL}$
(those above the upper dashed line) have higher frequencies (i.e.,
shorter time scales), and are thus preferred. Note however there
is a non-negligible leakage towards slow modes ${\bf q}$ with
$\tau_\omega \gtrsim\tau_{NL}$ (i.e., modes slightly below the
dashed upper curve).}
\label{peak_frequency}
\end{figure}
\subsection{Analysis of third-order time correlators}
Equation \eqref{eqgamt} indicates that the nonlinear interaction with
all triads gives rise to the time decorrelation of the mode at a given
${\bf k}$. In other words, the apparently random contribution of
all nonlinear triads results in the deformation of the structure to
the point that the mode decorrelates with itself, and thus transfers
its energy to other modes in the allowed triads. As a result,
$\Gamma_{\bf k}(\tau)$ decreases for short increments $\tau$, and then
fluctuates around zero. $-\partial \Gamma_{\bf k}(\tau)/\partial \tau$
should then start from zero for $\tau=0$, increase to a maximum, and
then fluctuate with the dominant time scale of the mode. Figure
\ref{thetat} shows a partial reconstruction of
$-{\partial \Gamma_{{\bf k}}}/{\partial \tau}$ by computing a partial
sum over ${\bf q}$ of the $\Theta({\bf k}, {\bf q}, \tau)$ function
for ${\bf k}=(0,0,8)$ and for ${\bf k}=(0,30,0)$. The partial
reconstruction is done using Eqs.~\eqref{eqgamt} and
\eqref{thetadef}, i.e., we sum $\Theta$ over the subset of ${\bf p}$
and ${\bf q}$ modes available for the analysis and that satisfy the
relation ${\bf p} + {\bf q} = {\bf k}$. Due to the large amount of
data required for this computation, we only consider modes in the
$(k_x=0,k_y,k_z)$ plane, and therefore we only sum over the triads
with ${\bf p} \cdot \hat{x} = {\bf q} \cdot \hat{x}=0$,
resulting in the partial reconstruction mentioned
above. Nonetheless, this suffices to get the expected behaviour for
the time derivative of the decorrelation functions. For the mode
${\bf k}=(0,0,8)$, which is a fast mode, the time derivative gets
locked to the wave period, while for ${\bf k}=(0,30,0)$, which is a
slow mode, the dominant time scale is the sweeping time. Here and in
the following, except when duly noted, all results shown are for the
$\Omega=8$ simulation.
\begin{figure}
\centering
\includegraphics[width=6.5cm]{fig6a.pdf}
\includegraphics[width=6.5cm]{fig6b.pdf}
\caption{Close up of the geometric distribution of the peak value
of the normalised contribution function for each triad, given by
$\max_\omega \{\vert \widehat{\Theta} ({\bf k},{\bf q},\omega)
\vert\} /E(q_y,q_z)$. As in Fig.~\ref{peak_frequency}, in each
panel ${\bf k}$ is fixed to
consider two modes dominated by waves: {\it a)}
${\bf k}=(0,0,8)$, and {\it b)} ${\bf k} = (0,5,5)$. The dashed
curves represent the modes with $\tau_\omega = \tau_{NL}$, and
the circles represent the near-resonant modes (according to the
theoretical prediction) with $\gamma_r<0.1$. In good
agreement with wave turbulence theories, normalised resonant
triads have large amplitudes, but some near-resonant triads are
also strong. In panel ${\it b)}$, some of these near-resonant triads
have non-negligible coupling with slow modes (see the circles
near $q_y \approx 10$ and $q_z \approx 0$) allowing for energy
transfer towards these modes.}
\label{zoom}
\end{figure}
\subsection{Contribution functions}
We now present our analysis of the behaviour of the contribution
function $\Theta$. Figure \ref{thetaw} shows the value of
$\vert \widehat{\Theta}({\bf k}=(0,0,8),{\bf q},\omega)\vert$ as a
function of $\omega$, for four different values of ${\bf q}$ (i.e.,
for four different triads). All of them peak at
$\omega_0 = \omega({\bf k})$, which is the wave frequency of the mode
${\bf k}$. This was checked for other values of ${\bf k}$ as well, and
a peak in the corresponding wave frequency was observed in all
cases except for the modes with $k_\parallel \approx 0$ (i.e., the
slow modes), for which no discernible peak is present. Moreover, and
in spite of the presence of a peak for modes with $k_\parallel > 0$,
it is worth noting that the width of the peak depends strongly on the
nature of the other modes in the triad. While interactions with other
wave modes have most of the power in the peak and are well tuned
(i.e., the peak is relatively narrow), interactions with slow modes
can have large amplitudes but with a broad spectrum.
This gives a direct way to identify not only the strength of a given
triad, but also to measure how resonant the triad is, as more resonant
triads are expected to result in a sharper spectrum of
$ \widehat{\Theta}({\bf k},{\bf q},\omega)$ per virtue of
Eq.~(\ref{eq:delta}). Therefore, to simplify the analysis, we can
focus on a few modes ${\bf k}$, explore all available values of
${\bf q}$ on a triad with ${\bf k}$, and look only at the the maximum
value of $\widehat{\Theta}$ (for all $\omega$) and on the relative
width of the maximum (i.e., on how well tuned the interaction is
around $\omega_0$).
In Fig.~\ref{peak_contribution} we show
$\max_\omega \{\vert\widehat{\Theta} ({\bf k},{\bf q},\omega) \vert\}$
for two modes ${\bf k}=(0,0,8)$ and $(0,5,5)$, as a function of all
possible values of ${\bf q}$ in the $(0,q_y,q_z)$ plane. The
anisotropic nature of rotating turbulence makes a stellar apparition
here, as the distribution of values is clearly influenced by it. The
result indicates that triads which are elongated along the horizontal
direction have larger amplitudes, which is compatible with the
prediction that energy tends to go towards the slow modes (with
$q_z\approx 0$) as discussed in \cite{Waleffe92}. Indeed, the triads
with larger amplitudes are located in a horizontal band within $-k_z
\lesssim q_z \lesssim k_z$. There are also strong triads that couple
the ${\bf k}$ mode with modes with larger vertical wavenumber (i.e.,
triads in the horizontal bands $k_z \lesssim |q_z| \lesssim 2 k_z$)
which are compatible with an anisotropic transfer of a fraction of
the energy towards larger wave numbers (i.e., smaller
scales). Finally, collinear modes (i.e., modes with ${\bf q} = \alpha
{\bf k}$) make no contribution to the triads as a result of the
incompressibility of the fluid.
\begin{figure}
\centering
\includegraphics[width=6.5cm]{fig7a.pdf}
\includegraphics[width=6.5cm]{fig7b.pdf}
\caption{Peak values of the normalised contribution function for
each triad, given by
$\max_\omega \{\vert \widehat{\Theta} ({\bf k},{\bf q},\omega)
\vert\} /E(q_y,q_z)$. The panels correspond to two fixed values
of ${\bf k}$: {\it a)} a small-scale ``fast'' mode
${\bf k}=(0,0,45)$, and {\it b)} a small-scale ``slow'' mode
${\bf k} = (0,30,0)$. The dashed curves
represent the modes with $\tau_\omega = \tau_{NL}$. The role
played by resonant and near-resonant interactions in these cases
is less clear.}
\label{peak_frequency_nowave}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=6.5cm]{fig8.pdf}
\caption{Peak values of the normalised contribution function for
each triad,
$\max_\omega \{\vert \widehat{\Theta} ({\bf k}=(0,0,8),{\bf
q},\omega) \vert\}/E(q_y,q_z)$, for the simulation
with weaker rotation (i.e., larger Rossby
number). Results are similar to the case with stronger rotation,
although the contrast in intensity between triads involving
fast and slow waves is less clear.}
\label{peak_frequency_rossby}
\end{figure}
\subsection{Normalised contribution functions}
Having said this, it can be argued that the strongest triads correspond
to modes with $q_z \approx 0$ only as a result of the anisotropic
energy spectrum shown in Fig.~\ref{ekk}: the modes with small vertical
wavenumber have most of the energy, and as a result triads involving
those modes will have larger amplitudes. Therefor, we can normalise the
triads by the energy of the {\bf q} mode in the triad, i.e., we can
compute
$\max_\omega\{|\hat{\Theta}({\bf k},{\bf q}, \omega)|\}/E({\bf q}).$ If
this is done, from Eqs.~(\ref{eqgamt}) and (\ref{ekwtheta}) the
normalised contribution function has units of inverse time (i.e., of
frequency). This time can thus be interpreted as the time scale of the
energy transfer mechanism, as it is often done in turbulence
theories.
We now turn to the analysis of these normalised contribution
functions. In Fig.~\ref{peak_frequency} we show the peak value of the
normalised functions for ${\bf k}=(0,0,8)$ and $(0,5,5)$. Two dashed
curves indicate the modes for which $\tau_\omega = \tau_{NL}$, i.e.,
modes with the eddy turnover time equal to the period of the
waves. Modes respectively above and below the upper and lower curves
have $\tau_\omega < \tau_{NL}$, and are dominated by the waves. Modes
between the two dashed curves have $\tau_\omega > \tau_{NL}$, and are
slow modes dominated by the eddies.
Anisotropic effects are still evident in Fig.~\ref{peak_frequency}
after the normalisation, but the role played in the triads by the
waves starts to become more clear. Although wave modes have
less energy, after normalisation it becomes evident that triadic
interactions of the wave modes ${\bf k}=(0,0,8)$ and $(0,5,5)$ with
other wave modes are relatively stronger than interactions with
slow modes. If the normalised contribution function is interpreted as
an inverse transfer time, it then implies that the transfer between
triads involving waves is faster than triads involving slow modes, and
thus should be preferred for the interaction. This is true even for
modes ${\bf q}$ with $\tau_s<\tau_\omega<\tau_{NL}$, i.e., for modes
with sweeping time faster than the wave period (but with the waves
still dominating over the eddies).
Large amplitudes (or, equivalently, shorter transfer times) can be
seen in Fig.~\ref{peak_frequency} in the vicinity of
${\bf q} \approx {\bf k}$. Although at first glance it would appear
that this is due to local (in Fourier space) interactions being the
most prominent, closer inspection reveals that it is also due to the
effect of resonances. In Fig.~\ref{zoom} we show a close up of the
normalised contribution functions in Fig.~\ref{peak_frequency} for
small values of $|{\bf q}|$. Circles mark the modes that satisfy the
theoretical near-resonant condition up to a value of
$\gamma_r=0.1$. It is evident that many of the strongest triads
correspond to resonant or near-resonant triads. This is in very good
agreement with wave turbulence theories of rotating turbulence, which
predict that resonant triads should dominate the coupling between
modes \citep{Newell69}. However, this also
explains how the system transfers energy towards slow modes, which are
inaccessible in weak wave turbulence approximations. The data in
Fig.~\ref{zoom} indicates that not only resonant triads are relevant,
but that near-resonant triads play an equally important role (at least
for the case of a periodic flow). Indeed, large amplitudes can be seen
around the circles in Fig.~\ref{zoom} in a region that is even
broader than the fan corresponding to the condition
$\gamma_r=0.1$. Close observation of Fig.~\ref{zoom} {\it a)} and
{\it b)} (as well as the observation of other modes ${\bf k}$ not
shown here) gives rise to the following picture: For ${\bf k} =
(0,0,8)$, resonant and near-resonant interactions couple this mode
with some modes with $q_z/q < k_z/k$, thus allowing an
energy exchange between these modes. Energy can thus be transferred
towards modes with smaller vertical wavenumber, in agreement with the
arguments in \cite{Cambon89} and \cite{Waleffe93}. For
${\bf k} = (0,5,5)$, the process is repeated, but now some
near-resonant interactions allow for a coupling (and thus a transfer)
with slow modes. This is compatible with observations in
\cite{Smith05}, \cite{Alexakis15}, and \cite{Gallet15}. Moreover, in
Fig.~\ref{zoom} {\it b)} a non-negligible coupling with slow modes can
be observed even for non-resonant triads (see the region between the
two-dashed curves with $q_y>0$), indicating that as energy approaches
the slow modes the role of non-resonant interactions may also become
more relevant.
\begin{figure}
\centering
\includegraphics[width=6.5cm]{fig9a.pdf}
\includegraphics[width=6.5cm]{fig9b.pdf}
\caption{Inverse relative bandwidth (quality factor) of the peak in
each contribution function, as a function of ${\bf q}$, and for
fixed ${\bf k} =(0,0,8)$, {\it a)} for the simulation with $Ro
\approx 0.03$, and {\it b)} for the simulation with $Ro \approx
0.015$. Larger quality factors correspond to sharper bandwidths
of the triads relative to their central frequency, and thus to more
resonant interactions (i.e., the factor quantifies how well tuned a
triad is). Circles indicate modes which satisfy the theoretical
near-resonance condition with $\gamma_r<0.1$. A good agreement
is observed between the theoretical condition and the quality
factor of the contribution function, specially for the flow with
smaller Rossby number. Note however that relatively large
quality factors are observed for branches which are wider than
the condition $\gamma_r<0.1$, indicating again the important
role played by near-resonant interactions.}
\label{tuning}
\end{figure}
\subsection{Comparison with small-scale and slow modes}
To gain more certainty on the effect of the waves in the triadic
interactions, we compare now the previous results with the normalised
contribution function for two modes: a small-scale mode with
${\bf k}=(0,0,45)$, which is dominated by sweeping (but with the wave
period faster than the eddy turnover time), and a small-scale slow
mode with ${\bf k}=(0,30,0)$ which has zero wave frequency. Here, by
small-scale, we refer to modes with wave numbers significantly larger
than the forced wave number. The resulting normalised
contribution functions are shown in
Fig.~\ref{peak_frequency_nowave}. In both cases, the division
given by the curve with $\tau_\omega = \tau_{NL}$ is less evident, and a
superposition of the modes expected
to be resonant or near-resonant (not shown)
indicates no clear correlation between the strength of the triad and
the theoretical resonant or near-resonant condition. For the mode
${\bf k}=(0,0,45)$, the normalised contribution function indicates
that coupling is stronger for modes with $q_z \approx k_z$, an effect
associated with the anisotropy of the flow, while the coupling with slow
modes is negligible. The slow mode ${\bf k}=(0,0,45)$ shows more
interesting features. The mode seems to be more strongly coupled with
other slow modes ${\bf q}$ in the vicinity of ${\bf k}$, and with
modes with large $q_z$ (of the order of $|{\bf k}|$, compatible with
local triadic interactions, although these interactions are
non-resonant).
\subsection{Effect of the Rossby number}
We can also compare the results in the two simulations with different
Rossby number, to quantify the effect of changing the rotation
frequency in the intensity of resonant and near-resonant
triads. As an illustration, Fig.~\ref{peak_frequency_rossby} shows the
geometric distribution of the peak values of the normalised
contribution function for the mode ${\bf k}=(0,0,8)$ in the simulation
with larger Rossby number. The same result as in
Fig.~\ref{peak_frequency} is obtained, but the contrast between modes
above and below the $\tau_\omega=\tau_{NL}$ curve is less
marked. Also, the region of modes dominated by eddies (i.e., of
slow modes) increases as expected, and the boundary between the
two regions indicated by the change in intensity of the triads moves
accordingly. This confirms that the changes in intensity in
Fig.~\ref{peak_frequency} indeed separate triads involving slow and
fast modes, and is also consistent with the behaviour expected in a
rotating flow as the Rossby number is varied.
\subsection{Direct measurement of the resonance level of each triad}
One of the most important implications of the contribution function is
that it allows a direct measurement of how well tuned a triad is,
i.e., of how resonant the interaction between three modes is. This was
already discussed in the context of Fig.~\ref{thetaw}, where we showed
that some triads display a narrower peak around the wave frequency
than others. But now we can put this observation on firmer grounds.
As it follows from Eq.~(\ref{eq:delta}), for a perfectly resonant
triad the contribution function should be a delta distribution
centred around $\omega_0=\omega({\bf k})$. Near-resonant and
non-resonant interactions broaden the peak. This broadening can be
measured using the quality factor
\begin{equation}
Q = \frac{\omega_0}{\Delta \omega} .
\end{equation}
In other words, the $Q$ factor is the inverse relative width of the
peak of $\widehat{\Theta}({\bf k},{\bf q},\omega)$ as a function
of the frequency. We estimate $\Delta \omega$ by calculating the
width of the peak in the spectrum (see Fig.~\ref{thetaw}) at half
the amplitude of the maximum value. In the theory of resonators, the
$Q$ factor is often interpreted as the ratio of the energy stored to
the energy lost in a system. In our case, the larger the $Q$ factor
the more resonant the triad is, and the less energy of the mode
${\bf k}$ is lost (i.e., given) to non-resonant modes.
Figure \ref{tuning} shows this quality factor for each contribution
function for fixed ${\bf k}=(0,0,8)$ and for all possible values of
${\bf q} = (0,q_y,q_z)$, for the two simulations with different Rossby
numbers. Superimposed to the quality factor, circles mark the modes that
satisfy the theoretical near-resonant condition up to a value of
$\gamma_r=0.1$. In particular for the simulation with smaller Rossby
number, triads with large quality factors (i.e., well tuned triads)
more or less coincide with triads with small $\gamma_r$, specially for
the branches in the upper-left and lower-right quadrants on the figure
on the right in Fig.~\ref{tuning}. In other words, the quality factor
defined above gives a good measure of how resonant a triad
is. Moreover, three features in Fig.~\ref{tuning} are worth
emphasising: First, the $Q$ factor has maximum values $\approx 6$
which is of the order of, although a bit smaller than, what is
often found in electrical or mechanical resonators. In other words,
even resonant triads display broadband peaks in the frequency
spectrum. Second, the area covered by triads with the largest values
of $Q$ is relatively larger than the area corresponding to the
circles with $\gamma_r<0.1$, confirming the importance of interactions
that are even marginally near-resonant. The scaling of this
behaviour with Rossby number and domain size can be important for
theories of rotating turbulence in infinite domains, in which the
behavior of the resonant and slow manifolds, as well as the
relevance of near-resonances, are unclear
\citep{Cambon04,Chen05,Bourouiba08}. Third, there are some modes
with relatively large $Q$ values (compared with the mean amplitude of
$Q$ for all modes) that do not correspond to resonant or near-resonant
modes in wave turbulence theory. Most notably, some of these modes are
modes with small $|{\bf q}|$ and lying in the region of the slow modes.
\section{Conclusions}
\label{conclusions}
One of the central problems in turbulence theory involves the
understanding of how modes interact non-linearly, specially in systems
with restitutive forces for which eddies and waves can coexist, and
for which resonances can strongly affect the nonlinear triadic
interactions. In these problems, a direct investigation of how each
triad of modes contributes to the overall dynamics is quite cumbersome
and complicated. To tackle this problem we have derived a
contribution function that characterises the spatio-temporal behaviour
of each triad, and more importantly, their contribution to the energy
transfer and to the spatio-temporal spectrum of the turbulent
flow.
We used this function to study the case of rotating turbulence, in
which eddies coexist with inertial waves, and where triadic resonant
interactions are expected to be dominant \citep{Newell69},
transferring their energy preferentially towards modes with small vertical
wave numbers \citep{Cambon89,Waleffe93}. However, this picture fails to
explain how energy continues to be transferred anisotropically to
``slow'' two-dimensional modes, as the wave turbulence approximation
breaks down in the vicinity of those modes. Previous results in
simulations at low resolution or in truncated systems
\citep{Chen05,Smith05} indicate that near-resonant triads can be
responsible for this latter transfer, but it is still unclear whether
these interactions remain to be relevant as the turbulence level is
increased, or as the Rossby number is decreased. Some recent results
suggest this to be the case \citep{Alexakis15,Gallet15}.
We computed the contribution function for a large number of triads in
two simulations of rotating turbulence at spatial resolutions of
$512^3$ grid points. As the contribution function is a third-order
time-correlation function for each Fourier triad, this requires a
massive analysis of spatio-temporal data. The main results show
or confirm that:
(1) For ``wave'' modes with $\tau_\omega<\tau_s<\tau_{NL}$ (i.e.,
modes for which the wave period is faster than the sweeping time and
the eddy turnover time), the coupling between triads is strongly
anisotropic. Triads which are elongated along the horizontal direction
have larger amplitudes, a result which is compatible with the
prediction that energy tends to go towards modes with smaller parallel
wave numbers. This result is also in agreement with the
proposed mechanism of parametric instability, which was obtained for
isolated triads \citep{Waleffe93}, while the data analysis presented
here considers the system with all possible couplings between the
triads.
(2) After normalising the triads by the energy in one of modes, it is
found that the transfer between triads involving wave modes is faster
than the transfer between triads that couple the wave mode with a slow
mode, and thus the former can be expected to be the preferred ones for
interactions, also in agreement with predictions from wave turbulence
theory.
(3) However, near-resonant and non-resonant interactions are
non-negligible, and couple the wave modes to slow modes, thus allowing
for energy transfer into that region of spectral space.
(4) The contribution function is peaked around the frequency of each
mode, and thus can be used to define a quality factor $Q$ that
measures how resonant a triad is. While resonant triads are compatible
with relatively larger values of the $Q$ factor, the analysis shows that
some marginally near-resonant and non-resonant triads also display
tuning with the wave frequency and are such that can couple fast and
slow modes. Further studies of this result can be importat for
theories of rotating turbulence in infinite domains, in which the
nature and coupling of the modes in the slow and in the resonant
manifolds with the rest of the modes is a matter of debate.
(6) For modes for which $\tau_\omega$ is larger than $\tau_s$ or
$\tau_{NL}$, the relevance of resonant and near-resonant triads
decreases rapidly.
(7) Finally, varying the Rossby number qualitatively preserves these
results, at least in the short range of values considered here.
These results are in agreement with major theoretical predictions for
the behaviour of nonlinear interactions in rotating turbulence, as
mentioned above, and can shed some light on the recent results
concerning the behaviour or two-dimensional modes for very small
Rossby numbers. In this context, an obvious shortcoming of the present
study is the lack of a parametric study of the behaviour of the triads
for even smaller Rossby numbers, or for larger Reynolds numbers. The
need to properly resolve in time the fastest waves and to store with
high time cadence the data, to then perform the spatio-temporal
analysis of each triad, precludes for the moment studies with faster
rotation or with larger spatial resolution. However, we believe that
the results presented here can be useful to quantitatively assess the
relevance of resonant, near-resonant, and non-resonant triads at
moderate Rossby numbers. Also, the formalism presented here can be
extended to analyse other systems in which resonant interactions
are also believed to play a central role (see, e.g., the recent
studies by \citet{Aubourg15,Aubourg16}).
\begin{acknowledgements}
The authors acknowledge support from Grant Nos. PIP
11220090100825, UBACYT 20020110200359, and PICT 2011-1529.
PCdL thanks Luca Biferale for fruitful discussions and comments.
\end{acknowledgements}
\bibliographystyle{jfm}
|
\section{Introduction}\label{s1}
Given a real \emph{base} $q>1$ and a finite \emph{alphabet} $A\subset\mathbb R$ (having at least two elements), by an \emph{expansion} of a real number $x$ we mean a sequence $c=(c_i)\subset A$ satisfying the equality
\begin{equation*}
\pi_q(c)=\pi_q(c_i):=\sum_{i=1}^{\infty}\frac{c_i}{q^i}=x.
\end{equation*}
We denote by $A^{\infty}$ the set of all sequences $c=(c_i)\subset A$, by $U_{A,q}$ the set of numbers $x$ having a unique expansion, and by $U'_{A,q}\subset A^{\infty}$ the set of the corresponding expansions.
The structure of the \emph{univoque set} $U_{A,q}$ is well known for the regular alphabets $A=\set{0,1,\ldots,m}$, $m=1,2,\ldots ;$ see, e.g., the reviews \cite{DevKom2016}, \cite{Kom2011} and their references.
For the general case a number of basic results have been given in \cite{Ped2005}.
Writing $a:=\min A$ and $b:=\max A$, it is clear that the lexicographically smallest and greatest sequences $a^{\infty}$ and $b^{\infty}$ belong to $U'_{A,q}$ for all $q>1$.
Here and in the sequel we emply the notation of symbolic dynamics.
For example, we denote by $a^{\infty}$ the constant sequence $a,a,\ldots,$ by $(ab)^{\infty}$ the periodic sequence $a,b,a,b,\ldots,$ by $\set{ab,abb}^{\infty}$ the set of sequences formed by the blocks $A_1,A_2,\ldots,$ where each block is equal to one of the words $ab$ or $abb$, and by $d^*\set{ab,abb}^{\infty}$ the union of the sequences of the form $c$, $dc$, $ddc$\ldots with some sequence $c\in\set{ab,abb}^{\infty}$.
For the regular alphabets $A=\set{0,1,\ldots,m}$ the univoque sets $U'_{A,q}$ are increasing with $q$.
In the general case the following holds:
\begin{proposition}\label{p11}
Fix an alphabet $A$.
\begin{enumerate}[\upshape (i)]
\item If $q>1$ is sufficiently close to one, then $U'_{A,q}=\set{a^{\infty},b^{\infty}}$.
\item If
\begin{equation*}
q>1+\frac{a_J-a_1}{\min_{j>1} \set{a_j-a_{j-1}}},
\end{equation*}
then $U'_{A,q}=A^{\infty}$.
\item If
\begin{equation}\label{11}
1<q\le R_A:=1+\frac{a_J-a_1}{\max_{j>1} \set{a_j-a_{j-1}}},
\end{equation}
and $p>q$, then $U'_{A,q}\subset U'_{A,p}$.
\end{enumerate}
\end{proposition}
It follows from the proposition that there exist two critical bases $p_A$ and $r_A$ such that $1<p_A\le r_A$, and
\begin{align*}
q\in (1,p_A)&\Longrightarrow U'_{A,q}\quad\text{is finite;}\\
q\in (p_A,r_A)&\Longrightarrow U'_{A,q}\quad\text{is countably infinite;}\\
q\in (r_A,\infty)&\Longrightarrow U'_{A,q}\quad\text{is uncountable.}
\end{align*}
It was observed without proof by Erd\H os \cite{Erd1996} that either $U'_{A,q}$ is countable, or it has the power of continuum: this property holds without assuming the continuum hypothesis.
A proof was given by Baker \cite{Baker2015}; see also \cite[Theorem 2.3.1, p. 22]{DevKom2016}.
\begin{example}\label{e12}
For $A=\set{0,1}$ we have $p_A=\varphi\approx 1.61803$ (the \emph{Golden Ratio}) and $r_A=q'\approx 1.78723$ (the \emph{Komornik--Loreti constant}, see \cite{KomLor1998}).
The first result was proved by Dar\'oczy et al. \cite{DarJarKat1986}, \cite{DarKat1993} (see also \cite{ErdJooKom1990} and \cite{SidVer1998}), while the second was established by Glendinning and Sidorov \cite{GleSid2001}.
More precisely,
\begin{align*}
q\in (1,\varphi]&\Longrightarrow U'_{A,q}=\set{0^{\infty},1^{\infty}};\\
q\in (\varphi,q')&\Longrightarrow U'_{A,q}\quad\text{is countably infinite;}\\
q\in [q',\infty)&\Longrightarrow U'_{A,q}\quad\text{has the power of continuum;}\\
q\in (2,\infty)&\Longleftrightarrow U'_{A,q}=\set{0,1}^{\infty}.
\end{align*}
For example, $\set{0^*(10)^{\infty}, 1^*(01)^{\infty}}\subset U'_{A,q}$ for all $q>\varphi$.
See also de Vries \cite{Dev2009} for stronger set-theoretical results.
Since the critical bases are invariant for non-constant affine transformations of the alphabet, these results remain valid for all two-letter alphabets.
\end{example}
\begin{example}\label{e13}
For more general regular alphabets $A_m=\set{0,1,\ldots, m}$, $m=2,3,\ldots,$ the critical bases $p_{A_m}$ have been determined by Baker \cite{Baker2014}: they are integer if $m$ is even, and quadratic irrational numbers if $m$ is odd.
The bases $r_{A_m}$ have been determined in \cite[p. 425]{DevKom2009} for $m=2$ and by Kong, Li and Dekking \cite{KongLiDekking2010} for $m\ge 3$. Generalizing the just mentioned theorem of Glendinning and Sidorov, they have proved that the bases $r_{A_m}$ coincide with the generalized Komornik--Loreti constants introduced in \cite{KomLor2002}, and hence they are transcendental numbers.
For example, we have $r_{A_2}\approx 2.53595.$
\end{example}
As an example we recall some results concerning the \emph{generalized golden ratios} $p_A$ for all three-letter alphabets.
By an affine transformation it suffices to consider the alphabets $\set{0,1,m}$ with $m\in [2,\infty)$.
Properties (i)-(viii) below have been obtained in \cite{KomLaiPed2010} (see Theorem 1.1, the proof of Lemma 5.3 and Remark 5.12); (ix) is due to Lai \cite{Lai2011}.
\begin{theorem}[KLP]\label{t12}
We consider the alphabets $\set{0,1,m}$ with $m\in [2,\infty)$, and we write $p_m$ instead of $p_{\set{0,1,m}}$ for brevity.
\begin{enumerate}[\upshape (i)]
\item The function $m\mapsto p_m$ is continuous.
\item We have $2\le p_m\le P_m:=1+\sqrt{m/(m-1)}$ for all $m$.
\item $p_m=2\Longleftrightarrow m\in\set{2,4,8,\ldots}$.
\item $C:=\set{m\ :\ p_m=1+\sqrt{m/(m-1)}}$ is a Cantor set.
\item In each connected component\footnote{The indices $d$ run over a set of sequences closely related to the Sturmian sequences in symbolic dynamics.} $(m_d,M_d)$ of $(2,\infty)\setminus C$ there exists a point $\mu_d$ such that $p_m$ is decreasing in $[m_d,\mu_d]$ and increasing in $[\mu_d,M_d]$.
Moreover, $p_m$ is given by two explicit formulas in $(m_d,\mu_d]$ and $[\mu_d,M_d)$, respectively.
\item If $q\in (1,p_m)$, then $U'_{A,q}=\set{0^{\infty},m^{\infty}}$.
Furthermore, $U'_{A,p_m}=\set{0^{\infty},m^{\infty}}$ if and only if $m\in [m_d,M_d]$ where $(m_d,M_d)$ is a connected component of $(2,\infty)\setminus C$.
\item If $q>p_m$, then $\set{m^*\delta'}\subset U'_{A,q}$ with some sequence $\delta'\ne m^{\infty}$ depending on $m$, to be described later.
\item If $q\in (1,P_m)$, then each element of $U'_{m,q}\setminus\set{0^{\infty}}$ has the form $0^*c$ with some $c\in\set{1,m}^{\infty}$.
\item If $q\in (1,P_m)$, then $U'_{m,q}$ is countable.
\end{enumerate}
\end{theorem}
It follows from these results that
\begin{equation}\label{12}
2\le p_m\le P_m\le r_m
\end{equation}
for all $m\ge 2$.
An interesting open problem is the determination of the critical bases $r_A$ for all three-letter alphabets.
The fact that $r_A$ is transcendental for the simplest such alphabet $\set{0,1,2}$ indicates the difficulty of this problem.
A more tractable problem is suggested by property (viii) in Theorem \ref{t12} above: this implies that for $q>p_m$ not only $U'_{m,q}$, but already $U'_{m,q}\cap \set{1,m}^{\infty}$ is infinite.
Motivated by this example we may investigate the size of $U'_{A,q}\cap B^{\infty}$ instead of $U'_{A,q}$, with any given subset $B$ of the alphabet $A$.
We have always $U'_{A,q}\cap B^{\infty}\subset U'_{B,q}$, but the converse inclusion may fail.
To prove the first assertion we observe that if a sequence $(c_i)\in B^{\infty}$ does not belong to $U'_{B,q}$, then there is another sequence $(d_i)\in B^{\infty}$ satisfying $\pi_q(d_i)=\pi_q(c_i)$.
Since $B\subset A$, this shows that $(c_i)$ does not belong to $U'_{A,q}$ either.
The second assertion follows from the following counterexample:
\begin{example}\label{e15}
Let $A=\set{0,1,2}$, $B=\set{0,1}$ and $q=2$.
Then the constant sequence $1^{\infty}$ belongs to $B^{\infty}=\set{0,1}^{\infty}$, but it does not belong to $U'_{A,q}$ because
\begin{equation*}
\pi_2(1^{\infty})=\pi_2(20^{\infty}).
\end{equation*}
\end{example}
Proposition \ref{p11} also implies the existence of $p_{A,B}, r_{A,B}\in (1,\infty]$ satisfying $p_{A,B}\le r_{A,B}$ and such that
\begin{align*}
q\in (1,p_{A,B})&\Longrightarrow U'_{A,q}\cap B^{\infty}\quad\text{is finite;}\\
q\in (p_{A,B},r_{A,B})&\Longrightarrow U'_{A,q}\cap B^{\infty}\quad\text{is countably infinite;}\\
q\in (r_{A,B},\infty)&\Longrightarrow U'_{A,q}\cap B^{\infty}\quad\text{is uncountable.}
\end{align*}
If $B$ is empty or has a unique element, then the set $U'_{A,q}\cap B^{\infty}$ has at most one element, so that $p_{A,B}=r_{A,B}=\infty$.
Otherwise the problem is non-trivial.
For ternary alphabets the above mentioned property may be expressed by the equality
\begin{equation*}
p_{\set{0,1,m},\set{1,m}}=p_{\set{0,1,m}}(=p_m).
\end{equation*}
Motivated by this we focus on the determination of
\begin{equation*}
r_m:=r_{\set{0,1,m},\set{1,m}},
\end{equation*}
i.e., we investigate only unique expansions not containing the zero digit.
For brevity we write henceforth
\begin{equation*}
V'_{m,q}:=U'_{\set{0,1,m},q}\cap \set{1,m}^{\infty}.
\end{equation*}
We have thus
\begin{align*}
q\in (1,p_m)&\Longrightarrow V'_{m,q}=\set{m^{\infty}}\quad\text{is finite;}\\
q\in (p_m,r_m)&\Longrightarrow V'_{m,q}\quad\text{is countably infinite;}\\
q\in (r_m,\infty)&\Longrightarrow V'_{m,q}\quad\text{is uncountable.}
\end{align*}
Now we are ready to state our results on ternary alphabets.
First we will complete the inequalities \eqref{12}:
\begin{proposition}\label{p13}
We have
\begin{equation}\label{13}
r_m\le R_m:=1+\frac{m}{m-1}
\end{equation}
for all $m\in [2,\infty)$.
\end{proposition}
Observe that $R_m$ is equal to the right side expression of \eqref{11} for the alphabet $\set{0,1,m}$.
Then we will determine the critical base $r_m$ for $m$ belonging to some special intervals.
We start with the first connected component (see property (v) in Theorem \ref{t12})
\begin{equation*}
[\mu_d,M_d)=[2,1+\alpha)
\end{equation*}
of $[2,\infty)\setminus C$, where $\alpha=1.32472\ldots$ denotes the first Pisot number.
This corresponds to the smallest ``admissible sequence'' $d=0^{\infty}$, as explained below, in Subsection \ref{ss91} of the Appendix.
We recall (see \eqref{92} below) that $p_m=m$ for all $m\in [2,1+\alpha]$.
\begin{proposition}\label{p14}\mbox{}
If $m\in [2,1+\alpha]=[2,2.32472\ldots]$, then $r_m$ is the unique positive solution $q$ of the equation
\begin{equation*}
\pi_q(m1^{\infty})=m-1.
\end{equation*}
Moreover, $V'_{m,r_m}$ is still countable.
\end{proposition}
See Figure \ref{f11}.
\begin{figure}
\includegraphics[scale=0.8]{f-P13.pdf}
\caption{Proposition \ref{p14}: $p_m$, $P_m$, $r_m$ and $R_m$ in $[2,2.32472]$}\label{f11}
\end{figure}
\begin{remark}
A direct computation (see \eqref{94} below) shows that
\begin{equation*}
r_m=\frac{2m-1+\sqrt{4m-3}}{2m-2}
\end{equation*}
in the component $[2,1+\alpha]$.
In particular, $r_2=1+\varphi\approx 2.61803$.
\end{remark}
Next we consider the connected component
\begin{equation*}
(m_d,M_d)=(2.80194\ldots ,4.54646\ldots)
\end{equation*}
of $(2,\infty)\setminus C$; see Figure \ref{f12}.
This corresponds to the smallest ``admissible sequence'' $d=(10)^{\infty}$, as explained in Subsection \ref{ss93} of the Appendix, where the formula of $p_m=\max\set{p_m',p_m''}$ is also given.
\begin{figure}
\includegraphics[scale=0.8]{f-P14}
\caption{Proposition \ref{p15}: $p''_m$, $p'_m$, $P_m$ and $R_m$ in $[2.80194,4.54646]$}\label{f12}
\end{figure}
We will determine $r_m$ for $m$ belonging to three subintervals
\begin{equation*}
[m_d,m_1],\quad [m_2,m_3],\quad [m_4,M_d]
\end{equation*}
of $[m_d,M_d]$, with
\begin{equation*}
m_1\approx 2.91286,\quad m_2\approx 2.992,\quad m_3\approx 3.10204,\quad m_4\approx 3.30278.
\end{equation*}
The precise definitions of these numbers will be given during the proof of the following proposition.
\begin{proposition}\label{p15}\mbox{}
\begin{enumerate}[\upshape (i)]
\item If $m\in [m_d,m_1]\approx [2.80194,2.91286]$, then $r_m$ is the unique positive solution of the equation
\begin{equation*}
\pi_q(m^{\infty}-(1m)^{\infty})=1.
\end{equation*}
\item If $m\in [m_2,m_3]\approx [2.992,3.10204]$, then $r_m$ is the unique positive solution of the equation
\begin{equation*}
\pi_q\left(mm1(m11m)^{\infty}\right)=m-1.
\end{equation*}
\item If $m\in [m_4,M_d]\approx [3.30278,4.54646]$, then $r_m$ is the unique positive solution of the equation
\begin{equation*}
\pi_q(m(m1)^{\infty})=m-1.
\end{equation*}
\end{enumerate}
\end{proposition}
\noindent See Figures \ref{f13}, \ref{f14}, \ref{f15} for the three separate cases and Figure \ref{f16} for a global picture.
Part of Figure \ref{f16} is shown in greater detail in Figure \ref{f17} to understand the situation concerning the curves $r_m$ of the middle and right intervals.
\begin{figure}
\includegraphics[scale=0.8]{f-P14-1}
\caption{Proposition \ref{p15} (i): $p_m$, $P_m$, $r_m$, $R_m$ and an auxiliary curve in $[2.80194,2.91286]$}\label{f13}
\end{figure}
\begin{figure}
\includegraphics[scale=0.8]{f-P14-2}
\caption{Proposition \ref{p15} (ii): $p_m$, $P_m$, $r_m$, $R_m$ and an auxiliary curve in $[2.992,3.10214]$}\label{f14}
\remove{P14-2.pdf=fig2-20160309.pdf}
\remove{Before we had the file f54.pdf labeled as f54.}
\end{figure}
\begin{figure}
\includegraphics[scale=0.8]{f-P14-3}
\caption{Proposition \ref{p15} (iii): $p''_m$, $p'_m$, $P_m$, $r_m$, $R_m$ and an auxiliary curve in $[3.30278,4.54646]$}\label{f15}
\end{figure}
\begin{figure}
\includegraphics[scale=0.8]{f-P14-global}
\caption{Proposition \ref{p15}: global picture in $[3.30278,4.54646]$}\label{f16}
\end{figure}
\begin{figure}
\includegraphics[scale=0.8]{f-P14-23}
\caption{Proposition \ref{p15}: the mutual positions of the curves defining $r_m$ in the middle and right zones}\label{f17}
\end{figure}
\begin{remarks}\mbox{}
\begin{enumerate}[\upshape (i)]
\item A straightforward computation shows that $r_m$ is the unique solution $q>2$ of the polynomial equations
\begin{align*}
&(m-1)q=q^2-1,\\
&(m-1)(q^6-2q^5+q^4-q^3-q^2+2q-1)=q^5+q^3,\\
&(m-1)(q^3-q^2-2q+1)=q^2+q
\end{align*}
in the three cases, respectively.
In particular,
\begin{equation*}
r_m=\frac{m-1+\sqrt{(m-1)^2+4}}{2}
\end{equation*}
in the first case (see \eqref{97}).
\item While $p_m$ is given by two formulas in $[m_d,M_d]$ (see property (v) in Theorem \ref{t12}), $r_m$ is given by three different formulas in different subintervals of $[m_d,M_d]$.
This indicates that the determination of $r_m$ might be more difficult than that of $p_m$.
\end{enumerate}
\end{remarks}
In the following sections we prove Propositions \ref{p11}--\ref{p15}.
All proofs are mathematically complete, except that of Propositions \ref{p15} (ii).
For the latter we admit some intermediate results obtained by symbolic computations and computer simulations; otherwise the proof would become too long.
In order to facilitate the reading, many technical computations are collected in an Appendix at the end of the paper.
\section{Proof of Proposition \ref{p11}}\label{s2}
Let us write $A=\set{a_1<\cdots<a_J}$, $J\ge 2$.
We recall\footnote{See, e.g., the proof of \cite[Theorem 2.2]{KomLaiPed2010}.} the following description of $U'_{A,q}$:
\begin{proposition}\label{p21}
Let $(c_i)\in A^{\infty}$ and $q>1$.
\begin{enumerate}[\upshape (i)]
\item If $(c_i)\in A^{\infty}$ satisfies for some $q>1$ the conditions
\begin{align}
&\sum_{i=1}^\infty \frac{c_{n+i}-a_1}{q^i} < a_{j+1}-a_j \qtq{whenever} c_n=a_j<a_J\label{21}
\intertext{and}
&\sum_{i=1}^\infty \frac{a_J-c_{n+i}}{q^i} < a_j-a_{j-1} \qtq{whenever} c_n=a_j>a_1,\label{22}
\end{align}
then $(c_i)\in U'_{A,q}$.
\item Conversely, if
\begin{equation}\label{23}
1<q\le 1+\frac{a_J-a_1}{\max_{j>1} \set{a_j-a_{j-1}}} (\le J)
\end{equation}
and $(c_i)\in U'_{A,q}$, then the inequalities \eqref{21} and \eqref{22} are satisfied.
\end{enumerate}
\end{proposition}
\begin{proof}[Proof of Proposition \ref{p11}]
\mbox{}
\begin{enumerate}[\upshape (i)]
\item This is proved in \cite[Corollary 2.7]{KomLaiPed2010}.
\item If
\begin{equation*}
q>1+\frac{a_J-a_1}{\min_{j>1} \set{a_j-a_{j-1}}},
\end{equation*}
then every sequence $(c_i)\in A^{\infty}$ satisfies \eqref{21} and \eqref{22}, so that $U'_{A,q}=A^{\infty}$.
\item We apply Proposition \ref{p21}.
Since $q$ satisfies \eqref{23} by our assumption, $U'_{A,q}$ is characterized by the conditions \eqref{21} and \eqref{22}.
Since $p>q$, both conditions remain valid by changing $q$ to $p$, and therefore $U'_{A,q}\subset U'_{A,p}$.
\end{enumerate}
\end{proof}
\begin{example}\label{e22}
We illustrate the usefulness of Proposition \ref{p21} by reproving some assertions on the alphabet $A=\set{0,1}$, mentioned above.
It follows from this proposition that in a base $q\in (1,2]$ a sequence $(c_i)\in\set{0,1}^{\infty}$ belongs to $U'_{A,q}$ if and only if
\begin{align}
&\sum_{i=1}^\infty \frac{c_{n+i}}{q^i} < 1 \qtq{whenever} c_n=0\label{24}
\intertext{and}
&\sum_{i=1}^\infty \frac{1-c_{n+i}}{q^i} < 1 \qtq{whenever} c_n=1.\label{25}
\end{align}
If $q\in (1,\varphi]$, then $q^{-1}+q^{-2}\ge 1$, so that both words $011$ and $100$ are forbidden in every sequence $(c_i)\in U'_{A,q}$.
This leaves only the possibilities: $0^{\infty}$, $1^{\infty}$, $0^*(10)^{\infty}$ and $1^*(01)^{\infty}$.
Since
\begin{equation*}
q^{-1}+q^{-3}+q^{-5}+\cdots\ge 1,
\end{equation*}
none of the sequences $0^*(10)^{\infty}$, $1^*(01)^{\infty}$ satisfies \eqref{24} and \eqref{25} either, so that $U'_{A,q}=\set{0^{\infty},1^{\infty}}$.
If $q\in (\varphi,\infty)$, then all sequences $0^*(10)^{\infty}$, $1^*(01)^{\infty}$ satisfy \eqref{24} and \eqref{25} and hence belong to $U'_{A,q}$.
Therefore $p_A=\varphi$.
On the other hand, if $q>2$, then
\begin{equation*}
q^{-1}+q^{-3}+q^{-5}+\cdots<1,
\end{equation*}
so that \eqref{24} and \eqref{25} are satisfied for all sequences $(c_i)$ of zeros and ones.
Therefore $U'_{A,q}=A^{\infty}$.
\end{example}
\section{Proof of Proposition \ref{p13}}\label{s3}
We will show that if $q\ge 1+m/(m-1)$ and $k$ is a sufficiently large integer, then
\begin{equation*}
\set{m^{\ell}1\quad :\quad \ell\ge k}^{\infty}\subset V'_{m,q},
\end{equation*}
and hence $V'_{m,q}$ has the power of continuum.
Applying Proposition \ref{p21} (i) for this special case, it suffices to show that each sequence
\begin{equation*}
(c_i)\in \set{m^{\ell}1\quad :\quad \ell\ge k}^{\infty}
\end{equation*}
satifies the following four conditions:
\begin{align*}
&\sum_{i=1}^\infty \frac{c_{n+i}}{q^i} < 1 \qtq{whenever} c_n=0;\\
&\sum_{i=1}^\infty \frac{c_{n+i}}{q^i} < m-1 \qtq{whenever} c_n=1;\\
&\sum_{i=1}^\infty \frac{m-c_{n+i}}{q^i} < 1 \qtq{whenever} c_n=1;\\
&\sum_{i=1}^\infty \frac{m-c_{n+i}}{q^i} < m-1 \qtq{whenever} c_n=m.
\end{align*}
The first condition is trivially satisfied because there is no zero digit.
The second condition follows by using our assumption on $q$:
\begin{equation*}
\sum_{i=1}^\infty \frac{c_{n+i}}{q^i}<\sum_{i=1}^\infty \frac{m}{q^i}=\frac{m}{q-1}\le m-1.
\end{equation*}
For the proof of the fourth condition first we observe that
\begin{equation*}
\sum_{i=1}^\infty \frac{m-c_{n+i}}{q^i}\le \pi_q\left(((m-1)0^k)^\infty\right)=(m-1)\frac{q^k}{q^{k+1}-1}
\end{equation*}
if $c_n=m$.
We conclude by observing that our assumption $q\ge 1+m/(m-1)$ implies that $q>2$, and then $q^k/(q^{k+1}-1)<1$ for all $k\ge 0$.
Finally, we have
\begin{equation*}
\sum_{i=1}^\infty \frac{m-c_{n+i}}{q^i}\le \pi_q\left((0^k(m-1))^\infty\right)=\frac{m-1}{q^{k+1}-1}
\end{equation*}
if $c_n=1$.
Since $q>1$, the third condition follows by choosing a sufficiently large integer $k$ satisfying $q^{k+1}>m$.
\section{Preliminary lemmas on ternary alphabets}\label{s4}
We establish three elementary lemmas on ternary alphabets that will be frequently used in the sequel.
\begin{lemma}\label{l41}
Let $q>2$ and $(a_i), (b_i)\in\set{1,m}^{\infty}$.
If $(a_i)<(b_i)$, then $\pi_q(a_i)<\pi_q(b_i)$.
\end{lemma}
\begin{proof}
If $n$ is the first index such that $a_n<b_n$, then
\begin{equation*}
\pi_q(b_i)-\pi_q(a_i)
\ge \frac{m-1}{q^n}-\sum_{i=n+1}^{\infty}\frac{m-1}{q^i}
=\frac{m-1}{q^n}\left(1-\frac{1}{q-1} \right) >0.\qedhere
\end{equation*}
\end{proof}
\begin{lemma}\label{l42}\mbox{}
\begin{enumerate}[\upshape (i)]
\item If $(c_i)\in\set{1,m}^{\infty}$ satisfies for some $q>2$ the conditions
\begin{equation}\label{41}
\pi_q(c_{n+i})<m-1
\qtq{and}
\pi_q(m-c_{n+i})<1
\end{equation}
whenever $c_n=1$, then $(c_i)\in V'_{m,q}$.
\item Conversely, if $q\in(2,R_m]$ and $(c_i)\in V'_{m,q}$, then the inequalities \eqref{41} are satisfied.
\end{enumerate}
\end{lemma}
\begin{proof}\mbox{}
\begin{enumerate}[\upshape (i)]
\item Proposition \ref{p21} (i) contains two other conditions: $\pi_q(c_{n+i})<1$ whenever $c_n=0$, and $\pi_q(m-c_{n+i})<m-1$ whenever $c_n=m$.
The first condition may be omitted because there are no zero digits.
Since $q>2$, the second condition is automatically satisfied, even if $c_n\ne m$:
\begin{equation*}
\pi_q(m-c_{n+i})\le \pi_q((m-1)^{\infty})=\frac{m-1}{q-1}<m-1
\end{equation*}
for all $n$.
\item This is a special case of Proposition \ref{p21} (ii).
\end{enumerate}
\end{proof}
Henceforth a finite block $a_0\cdots a_k$ of digits is called \emph{forbidden} if it cannot occur in any sequence $(c_i)\in V'_{m,q}$.
\begin{lemma}\label{l43}
Let $q\in(2,R_m]$ and $a_1\cdots a_k\in\set{1,m}^k$.
If
\begin{equation*}
\pi_q(a_1\cdots a_k1^{\infty})\ge m-1
\qtq{or of}
\pi_q(a_1\cdots a_km^{\infty})\le \frac{m}{q-1}-1,
\end{equation*}
then the block $1a_1\cdots a_k$ is forbidden.
\end{lemma}
\begin{proof}
If $(c_i)\in V'_{m,q}$ and $c_mc_{m+1}\cdots c_{m+k}=1a_1\cdots a_k$ for some $m$, then applying the preceding two lemmas we obtain
\begin{align*}
\pi_q(a_1\cdots a_k1^{\infty})
&=\pi_q(c_{m+1}\cdots c_{m+k}1^{\infty})\\
&\le \pi_q(c_{m+1}\cdots c_{m+k}c_{m+k+1}\cdots)<m-1
\end{align*}
and
\begin{align*}
\pi_q(a_1\cdots a_km^{\infty})
&=\pi_q(c_{m+1}\cdots c_{m+k}m^{\infty})\\
&\ge \pi_q(c_{m+1}\cdots c_{m+k}c_{m+k+1}\cdots)>\frac{m}{q-1}-1.\qedhere
\end{align*}
\end{proof}
\section{Proof of Proposition \ref{p14}}\label{s5}
We recall that in this section we have $d=0^{\infty}$, $\mu_d=2$ and $M_d\approx 2.32472$.
Henceforth we will write $f\sim g$ if $f$ and $g$ have the same sign.
We need a lemma:
\begin{lemma}\label{l51}
For each $m\in [2,M_d]$ there exists a number $r(m)>1$ such that
\begin{equation}\label{51}
\pi_q(m1^{\infty}) -(m-1)\sim r(m)-q.
\end{equation}
Furthermore,
\begin{equation*}
P_m\le r(m)<R_m,
\end{equation*}
and the equality $r(m)=P_m$ holds only if $m=M_d$.
\end{lemma}
\begin{proof}
The first assertion follows by observing that $\pi_q(m1^{\infty})$ is a continuous decreasing function of $q\in(0,\infty)$, and
\begin{equation*}
\lim_{q\to 1}\pi_q(m1^{\infty})=\infty,\quad \lim_{q\to \infty}\pi_q(m1^{\infty})=0.
\end{equation*}
The second assertion follows from \eqref{94} and the formulas \eqref{95}, \eqref{96} of the Appendix.
\end{proof}
\begin{proof}[Proof of Proposition \ref{p14}]\mbox{}
\emph{First step.} If $1<q\le r(m)$, then
\begin{equation*}
\pi_q(m1^{\infty})\ge m-1.
\end{equation*}
Since $2<P_m\le r(m)<R_m$ by \eqref{51}, by Proposition \ref{p11} (iii) it is sufficient to consider the case $2<q\le r(m)$.
Then we may apply Lemma \ref{l43} to infer that $1m$ is a forbidden block.
Hence $V'_{m,q}\subset\set{m^{\infty},m^*1^{\infty}}$, and thus $V'_{m,q}$ is countable.
\medskip
\emph{Second step.}
If $q>r(m)$, then $V'_{m,q}$ has the power of continuum.
Indeed, our assumption on $q$ implies that $\pi_q(m1^{\infty})< m-1$.
There exists therefore a positive integer $k$ satisfying
\begin{equation}\label{52}
\pi_q((m1^k)^{\infty})<m-1.
\end{equation}
We complete the proof by showing that each sequence
\begin{equation*}
(c_i)\in\set{m1^{\ell}\ :\ \ell=k,k+1,\ldots}^{\infty}
\end{equation*}
belongs to $V'_{m,q}$.
Since $q>P_m\ge 2$, it is sufficient to check the conditions of Lemma \ref{l42}.
In fact, the conditions \eqref{41} are satisfied for all $n$: we have
\begin{equation*}
\pi_q(c_{n+i})\le \pi_q((m1^k)^{\infty})<m-1
\end{equation*}
by the lexicographic inequality $(c_{n+i})\le (m1^k)^{\infty}$, Lemma \ref{l41} and \eqref{52}, and
\begin{equation*}
\pi_q(m-c_{n+i})\le \pi_q((m-1)^{\infty})=\frac{m-1}{q-1}<1
\end{equation*}
because $c_{n+i}\ge 1$ for all $n$, and $q>m$.
\medskip
It follows from the preceding steps that $r_m=r(m)$.
\end{proof}
\section{Proof of Proposition \ref{p15} (i)}\label{s6}
We recall that in this section $d=(10)^{\infty}$ and $m_d\approx 2.80194$.
First we define $m_1$:
\begin{lemma}\label{l61}\mbox{}
\begin{enumerate}[\upshape (i)]
\item The system of equations
\begin{equation*}
\pi_q(m^{\infty}-(1m)^{\infty})=1\qtq{and} \pi_q(mm1^{\infty})=m-1
\end{equation*}
has a unique solution $(m_1,q_1)$ with $q_1>2$.
We have
\begin{equation*}
m_1\approx 2.91286\qtq{and} q_1\approx 2.34018.
\end{equation*}
\item For each $m\in [m_d,m_1]$ there exists a number $r(m)>1$ such that
\begin{align*}
1<q<r(m)&\Longrightarrow \pi_q(m^{\infty}-(1m)^{\infty})>1\qtq{and} \pi_q(mm1^{\infty})>m-1,\\
q>r(m)&\Longrightarrow \pi_q(m^{\infty}-(1m)^{\infty})<1.
\end{align*}
\item We have
\begin{equation}\label{61}
P_m\le r(m)<R_m,
\end{equation}
and the equality $r(m)=P_m$ holds only if $m=m_d$.
\end{enumerate}
\end{lemma}
\noindent See Figure \ref{f13}.
\begin{proof}\mbox{}
\begin{enumerate}[\upshape (i)]
\item Since
\begin{equation*}
\pi_q(m^{\infty}-(1m)^{\infty})=\pi_q\left(((m-1)0)^{\infty} \right) =\frac{(m-1)q}{q^2-1}
\end{equation*}
and
\begin{equation*}
\pi_q(mm1^{\infty})=\frac{m-1}{q}+\frac{m-1}{q^2}+\frac{1}{q-1},
\end{equation*}
the system of equations is equivalent to
\begin{equation*}
m-1=\frac{q^2-1}{q}\qtq{and} m-1=\frac{q^2}{(q-1)(q^2-q-1)}.
\end{equation*}
Eliminating $m$ we obtain the equation
\begin{equation*}
q^2(q-1)(q^2-q-3)=1.
\end{equation*}
The left hand side is increasing in $[2,\infty)$ and changes sign in $(2,3)$, hence it has a unique solution $q_1>2$, satisfying $q_1\in (2,3)$.
Then
\begin{equation*}
m_1=1+\frac{q_1^2-1}{q_1}=1+q_1-\frac{1}{q_1}\in \left(1+2-\frac{1}{2},1+3-\frac{1}{3}\right)
\end{equation*}
because the function $q\mapsto 1+q-1/q$ is increasing for $q>0$.
The numerical values are obtained by evaluating the root $q_1$ of the above polynomial equation.
\item The function
\begin{equation*}
q\mapsto \pi_q(m^{\infty}-(1m)^{\infty})=(m-1)\pi_q\left((10)^{\infty} \right)
\end{equation*}
is continuous and decreasing in $(1,\infty)$, and
\begin{equation*}
\lim_{q\to 1}\pi_q(m^{\infty}-(1m)^{\infty})=\infty,\quad \lim_{q\to \infty}\pi_q(m^{\infty}-(1m)^{\infty})=0.
\end{equation*}
Therefore there exists a unique number $r(m)\in (1,\infty)$ satisfying
\begin{equation*}
\pi_q(m^{\infty}-(1m)^{\infty})-1\sim r(m)-q.
\end{equation*}
Since the function $q\mapsto \pi_q(mm1^{\infty})$ is also decreasing in $q$, it remains to show that
\begin{equation*}
\pi_{r(m)}(mm1^{\infty})\ge m-1.
\end{equation*}
Denoting by $q=f(m)$ and $q=g(m)$ the solutions of the equations
\begin{equation*}
\pi_q(m^{\infty}-(1m)^{\infty})=1\qtq{and} \pi_q(mm1^{\infty})=m-1,
\end{equation*}
respectively, it suffices to show by monotonicity that $f\le g$ in $[m_d,m_1]$.
Since $f(m_1)=g(m_1)$ by (i), it suffices to prove that $f$ is increasing and $g$ is decreasing in $[m_d,m_1]$.
The increasingness of $f$ follows from the explicit formula
\begin{equation*}
f(m)=\frac{m-1+\sqrt{(m-1)^2+4}}{2}
\end{equation*}
(see \eqref{912}).
Furthermore, since
\begin{equation*}
\pi_q(mm1^{\infty})=m-1\Longleftrightarrow \frac{1}{m-1}=q-2+q^{-2}
\end{equation*}
(see \eqref{915}) and since the right hand side has a positive derivative
\begin{equation*}
(q-2+q^{-2})'=1-2q^{-3}>0
\end{equation*}
for all $q>2$, $m$ is a decreasing function of $g(m)$, and hence its inverse function $g$ is also decreasing.
\item This follows from \eqref{912}, \eqref{913} and \eqref{914} in the Appendix.
\end{enumerate}
\end{proof}
\begin{proof}[Proof of Proposition \ref{p15} (i)]\mbox{}
\emph{First step.} If $1<q\le r(m)$, then $V'_{m,q}$ is countable.
Indeed, it follows from our assumptions that
\begin{equation*}
\pi_q(mm1^{\infty})\ge m-1
\qtq{and}
\pi_q\left(m^{\infty}-(1m)^{\infty}\right)\ge 1.
\end{equation*}
We claim that the blocks $1mm$ and $11$ are forbidden.
As in the proof of Proposition \ref{p14}, we may assume that $2<q\le r(m)$.
If $(c_i)\in\set{1,m}^{\infty}$ contains a block $c_nc_{n+1}c_{n+2}=1mm$, then
\begin{equation*}
\pi_q(c_{n+i})\ge \pi_q(mm1^{\infty})\ge m-1
\end{equation*}
by our assumption, and hence $(c_i)\notin V'_{m,q}$ by \eqref{61} and Lemma \ref{l42}.
Next assume that $(c_i)\in\set{1,m}^{\infty}$ does not contain any block $1mm$, but it contains a block $c_nc_{n+1}=11$.
Then using also the preceding observation,
\begin{equation*}
\pi_q(m-c_{n+i})\ge \pi_q\left( m^{\infty}-(1m)^{\infty}\right)\ge 1
\end{equation*}
by hypothesis, and therefore $(c_i)\notin V'_{m,q}$ again.
It follows that $V'_{m,q}\subset\set{m^{\infty}, m^*(1m)^{\infty}}$, and hence it is countable.
\medskip
\emph{Second step.} If $q>r(m)$, then $V'_{m,q}$ has the power of continuum.
Indeed, it follows from our assumption that $\pi_q(m^{\infty}-(1m)^{\infty})<1$.
There exists therefore a (sufficiently large) positive integer $k$ satisfying
\begin{equation*}
\pi_q\left(m^{\infty}-\left((1m)^k1 \right)^{\infty}\right)<1.
\end{equation*}
We complete the proof by showing the inclusion
\begin{equation*}
\set{(1m)^{\ell}1\ :\ \ell\ge k}^{\infty}\subset V'_{m,q}.
\end{equation*}
We have to check that each sequence $(c_i)$ in the left hand side set satisfies the conditions of Lemma \ref{l42}.
For this we use the sequence $\delta'=(m1)^{\infty}$, introduced at the beginning of Section 4 in \cite{KomLaiPed2010}; see also \cite[Lemma 3.9]{KomLaiPed2010}.
If $c_n=1$, then \cite[Lemma 5.11]{KomLaiPed2010} and the relation $q>p_m$ imply $\delta'\in V'_{m,q}$ and hence
\begin{equation*}
\pi_q(c_{n+i})<\pi_q\left((m1)^{\infty}\right)=\pi_q(\delta')<m-1,
\end{equation*}
while
\begin{equation*}
\pi_q(m-c_{n+i})\le \pi_q\left(m-\left((1m)^k1 \right)^{\infty}\right)<1
\end{equation*}
by the choice of $k$.
\medskip
It follows from the preceding steps that $r_m=r(m)$.
\end{proof}
\section{Proof of Proposition \ref{p15} (ii)}\label{s7}
In this section we still have $d=(10)^{\infty}$, $m_d\approx 2.80194$ and $M_d\approx 4.54646]$.
We consider the subinterval $[m_2,m_3]:=[2.992,3.10214]$ of $(m_d,M_d)$.
\begin{lemma}\label{l71}
For each $m\in [m_2,m_3]$ there exists a number $r(m)>1$ such that
\begin{align*}
1<q\le r(m)&\Longrightarrow \pi_q\left(mm1(m11m)^{\infty}\right)\ge m-1,\\
q>r(m)&\Longrightarrow \pi_q\left(mm1(m11m)^{\infty}\right)<m-1\quad\text{and}\\
&\hspace*{19mm}\pi_q(m^{\infty}-(1mm1)^{\infty})<1.
\end{align*}
Furthermore, $P_m<r(m)<R_m$.
\end{lemma}
\noindent See Figure \ref{f14}.
\begin{proof}
The existence of $r(m)>1$ satisfying
\begin{align*}
&\Longrightarrow \pi_q\left(mm1(m11m)^{\infty}\right)>m-1\qtq{if}1<q<r(m),\\
&\Longrightarrow \pi_q\left(mm1(m11m)^{\infty}\right)<m-1\qtq{if}q>r(m)
\end{align*}
follows by monotonicity as in the proof of Lemma \ref{l51}.
The inequality
\begin{equation*}
\pi_q(m^{\infty}-(1mm1)^{\infty})<1\qtq{if}q>r(m)
\end{equation*}
follows from the choice of $m_3$, see Figure \ref{f14}.
The same figure also shows the inequalities $P_m<r(m)<R_m$.
\end{proof}
\begin{proof}[Proof of Proposition \ref{p15} (ii)]
Fix $m\in [m_2,m_3]$ arbitrarily.
A numerical computation suggests that for each fixed $m\in [m_2,m_3]$, if $q$ is sufficiently small, then all elements of $V'_{m,q}$ belong to the set
\begin{equation*}
\set{1m(11mm)^k\ :\ k=1,2,\ldots}^{\infty}.
\end{equation*}
In this set the words $111$, $1mmm$, $11m11$ do not occur, and we may find similarly forbidden words of length $5,6,\ldots .$
As we show below, a careful study of the forbidden words of length $\le 7$ allows us to determine the critical base $r_m$.
\medskip
\emph{First step.}
If $2<q\le r(m)$, then the application of Lemma \ref{l43} shows that the following seven words are forbidden:
\remove{For $m=3$ and $q=2.37034$}
\begin{align*}
&111,\\
&1mmm,\\
&11m11,\\
&11m1m1,\\
&1mm1mm,\\
&11m1mm1,\\
&1mm1m1m.
\end{align*}
Using the theory of finite automata (see Figure \ref{f71}) we infer from this that each sequence of $V'_{q,m}$ appears in the following list:
\begin{align*}
&m^{\infty},\\
&m^*(1m)^{\infty},\\
&m^*(1m)^*11mm1\set{1mm1,m11mm1}^{\infty},\\
&m^*(1m)^*1mm1\set{1mm1,m11mm1}^{\infty}.
\end{align*}
\begin{figure}
\definecolor{white}{RGB}{255,255,255}
\centering
\scalebox{0.9}{\begin{tikzpicture}[y=0.80pt, x=0.8pt,yscale=-1, inner sep=0pt, outer sep=0pt,
draw=black,fill=black,line join=miter,line cap=rect,miter limit=10.00,line width=0.800pt]
\begin{scope}[shift={(56.0,-188.0)},draw=white,fill=white]
\path[fill,rounded corners=0.0000cm] (-56.0000,188.0000) rectangle
(403.0000,398.0000);
\end{scope}
\begin{scope}[cm={{1.0,0.0,0.0,1.0,(56.0,-188.0)}},line cap=butt,miter limit=1.45,line width=1.600pt]
\path[draw] (15.0000,367.7871) circle (0.4233cm);
\path[draw] (116.9032,367.7871) circle (0.4233cm);
\path[draw] (233.9916,367.7871) circle (0.4233cm);
\path[draw] (116.9032,224.7135) circle (0.4233cm);
\path[draw] (233.9916,224.7135) circle (0.4233cm);
\path[draw] (372.3973,224.7135) circle (0.4233cm);
\path[draw] (233.9916,288.2503) circle (0.4233cm);
\path[draw] (372.3973,367.7871) circle (0.4233cm);
\path[draw] (303.1945,367.7871) circle (0.4233cm);
\path[draw] (22.5000,354.7851) -- (22.5000,346.2861) -- (23.1250,342.5677) --
(25.0000,339.9118) -- (28.1250,338.3183) -- (32.5000,337.7871) --
(35.0000,337.7871) -- (39.3750,338.4121) -- (42.5000,340.2871) --
(44.3750,343.4121) -- (45.0000,347.7871) -- (45.0000,350.2871) --
(44.0003,357.7871) -- (42.7506,359.6621) -- (41.0010,360.2871) --
(37.0020,360.2871);
\path[fill] (28.0020,360.2871) -- (41.5020,365.9121) -- (38.1270,360.2871) --
(41.5020,354.6621) -- cycle;
\path[fill,line width=0.800pt] (28.1484,330.2558) node[above right] (text42)
{$m$};
\path[draw] (30.0000,367.7871) -- (92.9033,367.7871);
\path[fill] (101.9033,367.7871) -- (88.4033,362.1621) -- (91.7783,367.7871) --
(88.4033,373.4121) -- cycle;
\path[fill,line width=0.800pt] (62.1577,360.2558) node[above right] (text48)
{$1$};
\path[draw] (131.9032,367.7871) -- (209.9916,367.7871);
\path[fill] (218.9916,367.7871) -- (205.4916,362.1621) -- (208.8666,367.7871) --
(205.4916,373.4121) -- cycle;
\path[fill,line width=0.800pt] (171.6535,360.2558) node[above right] (text54)
{$1$};
\path[draw] (110.2712,354.3441) -- (90.6507,304.3098) -- (88.6049,294.9252) --
(89.7687,285.3909) -- (92.7313,275.1091) -- (99.5192,256.3433) --
(104.4660,245.0733);
\path[fill] (108.0833,236.8322) -- (97.5067,246.9330) -- (104.0138,246.1034) --
(107.8080,251.4546) -- cycle;
\path[fill,line width=0.800pt] (76.7827,323.8662) node[above right] (text60)
{$m$};
\path[draw] (122.8588,238.4951) -- (136.4920,259.1536) -- (140.6230,267.9134) --
(142.0000,277.5000) -- (142.0000,296.0000) -- (141.0791,305.8242) --
(138.3164,315.2968) -- (135.1836,323.2032) -- (129.0570,339.2021) --
(126.6139,345.9042);
\path[fill] (123.5317,354.3600) -- (133.4399,343.6028) -- (126.9992,344.8473) --
(122.8702,339.7499) -- cycle;
\path[fill,line width=0.800pt] (147,291.2852) node[above right] (text66) {$1$};
\path[draw] (131.9032,224.7135) -- (209.9916,224.7135);
\path[fill] (218.9916,224.7135) -- (205.4916,219.0885) -- (208.8666,224.7135) --
(205.4916,230.3385) -- cycle;
\path[fill,line width=0.800pt] (169.8459,217.1823) node[above right] (text72)
{$m$};
\path[draw] (248.9916,224.7135) -- (348.3973,224.7135);
\path[fill] (357.3973,224.7135) -- (343.8973,219.0885) -- (347.2723,224.7135) --
(343.8973,230.3385) -- cycle;
\path[fill,line width=0.800pt] (299.4005,217.1823) node[above right] (text78)
{$1$};
\path[draw] (361.9680,235.4945) -- (250.6785,350.5374);
\path[fill] (244.4209,357.0061) -- (257.8501,351.2141) -- (251.4607,349.7289) --
(249.7644,343.3922) -- cycle;
\path[fill,line width=0.800pt] (299.4005,282.7297) node[above right] (text84)
{$1$};
\path[draw] (233.9916,352.7871) -- (233.9916,312.2503);
\path[fill] (233.9916,303.2503) -- (228.3666,316.7503) -- (233.9916,313.3753) --
(239.6166,316.7503) -- cycle;
\path[fill,line width=0.800pt] (217.7885,332.5538) node[above right] (text90)
{$m$};
\path[draw] (233.9916,273.2503) -- (233.9916,248.7135);
\path[fill] (233.9916,239.7135) -- (228.3666,253.2135) -- (233.9916,249.8385) --
(239.6166,253.2135) -- cycle;
\path[fill,line width=0.800pt] (217.7885,261.0171) node[above right] (text96)
{$m$};
\path[draw] (372.3973,239.7135) -- (372.3973,343.7871);
\path[fill] (372.3973,352.7871) -- (378.0223,339.2871) -- (372.3973,342.6621) --
(366.7723,339.2871) -- cycle;
\path[fill,line width=0.800pt] (356.1942,300.7855) node[above right] (text102)
{$m$};
\path[draw] (357.3973,367.7871) -- (327.1945,367.7871);
\path[fill] (318.1945,367.7871) -- (331.6945,373.4121) -- (328.3195,367.7871) --
(331.6945,362.1621) -- cycle;
\path[fill,line width=0.800pt] (334.002,360.2558) node[above right] (text108)
{$1$};
\path[draw] (288.1945,367.7871) -- (257.9916,367.7871);
\path[fill] (248.9916,367.7871) -- (262.4916,373.4121) -- (259.1166,367.7871) --
(262.4916,362.1621) -- cycle;
\path[fill,line width=0.800pt] (264.7991,360.2558) node[above right] (text114)
{$1$};
\path[draw] (-14.1508,342.5950) -- (-3.1587,352.0944);
\path[fill] (3.6508,357.9791) -- (-2.8855,344.8960) -- (-4.0099,351.3588) --
(-10.2414,353.4079) -- cycle;
\end{scope}
\end{tikzpicture}}
\caption{Automaton for the seven forbidden words}\label{f71}
\end{figure}
\begin{figure}
\centering
\scalebox{0.9}{
\definecolor{white}{RGB}{255,255,255}
\begin{tikzpicture}[y=0.80pt, x=0.8pt,yscale=-1, inner sep=0pt, outer sep=0pt,
draw=black,fill=black,line join=miter,line cap=rect,miter limit=10.00,line width=0.800pt]
\begin{scope}[shift={(56.0,-189.0)},draw=white,fill=white]
\path[fill,rounded corners=0.0000cm] (-56.0000,189.0000) rectangle
(403.0000,398.0000);
\end{scope}
\begin{scope}[cm={{1.0,0.0,0.0,1.0,(56.0,-189.0)}},line cap=butt,miter limit=1.45,line width=1.600pt]
\path[draw] (15.0000,367.7871) circle (0.4233cm);
\path[draw] (116.9032,367.7871) circle (0.4233cm);
\path[draw] (233.9916,367.7871) circle (0.4233cm);
\path[draw] (116.9032,224.7135) circle (0.4233cm);
\path[draw] (233.9916,224.7135) circle (0.4233cm);
\path[draw] (372.3973,224.7135) circle (0.4233cm);
\path[draw] (233.9916,288.2503) circle (0.4233cm);
\path[draw] (22.5000,354.7851) -- (22.5000,346.2861) -- (23.1250,342.5677) --
(25.0000,339.9118) -- (28.1250,338.3183) -- (32.5000,337.7871) --
(35.0000,337.7871) -- (39.3750,338.4121) -- (42.5000,340.2871) --
(44.3750,343.4121) -- (45.0000,347.7871) -- (45.0000,350.2871) --
(44.0003,357.7871) -- (42.7506,359.6621) -- (41.0010,360.2871) --
(37.0020,360.2871);
\path[fill] (28.0020,360.2871) -- (41.5020,365.9121) -- (38.1270,360.2871) --
(41.5020,354.6621) -- cycle;
\path[fill,line width=0.800pt] (28.1484,331.2558) node[above right] (text38)
{$m$};
\path[draw] (30.0000,367.7871) -- (92.9033,367.7871);
\path[fill] (101.9033,367.7871) -- (88.4033,362.1621) -- (91.7783,367.7871) --
(88.4033,373.4121) -- cycle;
\path[fill,line width=0.800pt] (62.1577,361.2558) node[above right] (text44)
{$1$};
\path[draw] (131.9032,367.7871) -- (209.9916,367.7871);
\path[fill] (218.9916,367.7871) -- (205.4916,362.1621) -- (208.8666,367.7871) --
(205.4916,373.4121) -- cycle;
\path[fill,line width=0.800pt] (171.6535,361.2558) node[above right] (text50)
{$1$};
\path[draw] (110.2712,354.3441) -- (90.6507,304.3098) -- (88.6049,294.9252) --
(89.7687,285.3909) -- (92.7313,275.1091) -- (99.5192,256.3433) --
(104.4660,245.0733);
\path[fill] (108.0833,236.8322) -- (97.5067,246.9330) -- (104.0138,246.1034) --
(107.8080,251.4546) -- cycle;
\path[fill,line width=0.800pt] (77.7827,323.8662) node[above right] (text56)
{$m$};
\path[draw] (122.8588,238.4951) -- (136.4920,259.1536) -- (140.6230,267.9134) --
(142.0000,277.5000) -- (142.0000,296.0000) -- (141.0791,305.8242) --
(138.3164,315.2968) -- (135.1836,323.2032) -- (129.0570,339.2021) --
(126.6139,345.9042);
\path[fill] (123.5317,354.3600) -- (133.4399,343.6028) -- (126.9992,344.8473) --
(122.8702,339.7499) -- cycle;
\path[fill,line width=0.800pt] (146,291.2852) node[above right] (text62) {$1$};
\path[draw] (131.9032,224.7135) -- (209.9916,224.7135);
\path[fill] (218.9916,224.7135) -- (205.4916,219.0885) -- (208.8666,224.7135) --
(205.4916,230.3385) -- cycle;
\path[fill,line width=0.800pt] (169.8459,218.1823) node[above right] (text68)
{$m$};
\path[draw] (248.9916,224.7135) -- (348.3973,224.7135);
\path[fill] (357.3973,224.7135) -- (343.8973,219.0885) -- (347.2723,224.7135) --
(343.8973,230.3385) -- cycle;
\path[fill,line width=0.800pt] (299.4005,218.1823) node[above right] (text74)
{$1$};
\path[draw] (361.9680,235.4945) -- (250.6785,350.5374);
\path[fill] (244.4209,357.0061) -- (257.8501,351.2141) -- (251.4607,349.7289) --
(249.7644,343.3922) -- cycle;
\path[fill,line width=0.800pt] (299.4005,283.7297) node[above right] (text80)
{$1$};
\path[draw] (233.9916,352.7871) -- (233.9916,312.2503);
\path[fill] (233.9916,303.2503) -- (228.3666,316.7503) -- (233.9916,313.3753) --
(239.6166,316.7503) -- cycle;
\path[fill,line width=0.800pt] (218.7885,332.5538) node[above right] (text86)
{$m$};
\path[draw] (233.9916,273.2503) -- (233.9916,248.7135);
\path[fill] (233.9916,239.7135) -- (228.3666,253.2135) -- (233.9916,249.8385) --
(239.6166,253.2135) -- cycle;
\path[fill,line width=0.800pt] (218.7885,261.0171) node[above right] (text92)
{$m$};
\path[draw] (-14.1508,342.5950) -- (-3.1587,352.0944);
\path[fill] (3.6508,357.9791) -- (-2.8855,344.8960) -- (-4.0099,351.3588) --
(-10.2414,353.4079) -- cycle;
\end{scope}
\end{tikzpicture}}
\caption{Automaton for the eight forbidden words}\label{f72}
\end{figure}
We claim that the block $(1mm1)(m11mm1)$ is also forbidden.
Assume on the contrary that a sequence $(c_i)\in V'_{q,m}$ contains such a block.
Then we infer from the above list that
\begin{equation*}
(c_{n-1+i})\in (1mm1)(m11mm1)\set{1mm1,m11mm1}^{\infty}
\end{equation*}
for some $n\ge 1$.
Therefore $c_n=1$ and
\begin{equation*}
(c_{n+i})\ge (mm1)(m11mm1)(1mm1)^{\infty}=(mm1)(m11m)^{\infty}.
\end{equation*}
Applying Lemma \ref{l41} and the hypothesis $q\le r(m)$ we get
\begin{equation*}
\pi_q(c_{n+i})=\pi_q\left(mm1(m11m)^{\infty}\right)\ge \pi_{r_m}\left(mm1(m11m)^{\infty}\right)=m-1,
\end{equation*}
contradicting the first condition of Lemma \ref{l42}.
This contradiction proves our claim.
Since $(1mm1)(m11mm1)$ is forbidden, the above list corresponding to seven forbidden words may be further reduced: for $q\in (1,r(m)]$ each sequence $(c_i)\in V'_{q,m}$ belongs to the following list:
\begin{align*}
&m^{\infty},\\
&m^*(1m)^{\infty},\\
&m^*(1m)^*1(1mm1)^{\infty},\\
&m^*(1m)^*(1mm1)^{\infty}.
\end{align*}
(See also the corresponding automaton in Figure \ref{f72}.)
Hence $V'_{q,m}$ is countable.
As in the previous proofs, the conclusion remains valid for all $1<q\le r(m)$.
\medskip
\emph{Second step.}
We show that $V'_{q,m}$ has the power of continuum if $q>r(m)$.
By Lemma \ref{l71} there exists a sufficiently large integer $k$ such that
\begin{equation*}
\pi_q\left(mm(1m(11mm)^k)^{\infty}\right)< m-1.
\end{equation*}
We complete the proof by showing that
\begin{equation*}
\set{1m(11mm)^k,1m(11mm)^{k+1}}^{\infty}\subset V'_{q,m}.
\end{equation*}
It suffices to check that if $(c_i)$ is a sequence in the left hand side set and $c_n=1$, then the conditions of Lemma \ref{l42} are satisfied.
Since $c_n=1$, we have
\begin{equation*}
(1mm1)^{\infty}<(c_{n+i})\le mm(1m(11mm)^k)^{\infty}.
\end{equation*}
Applying Lemma \ref{l41} the conditions of Lemma \ref{l42} follow:
\begin{equation*}
\pi_q(c_{n+i})\le \pi_q\left(mm(1m(11mm)^k)^{\infty}\right)< m-1
\end{equation*}
and
\begin{equation*}
\pi_q(c_{n+i})> \pi_q\left((1mm1)^{\infty}\right)>\frac{m}{q-1}-1.
\end{equation*}
\medskip
We conclude from the preceding steps that $r_m=r(m)$.
\end{proof}
\section{Proof of Proposition \ref{p15} (iii)}\label{s8}
As in the preceding two sections, we have $d=(10)^{\infty}$, $m_d\approx 2.80194$ and $M_d\approx 4.54646$.
We start by defining $m_4$:
\begin{lemma}\label{l81}\mbox{}
\begin{enumerate}[\upshape (i)]
\item The system of equations
\begin{equation*}
\pi_q(m(m1)^{\infty})=m-1\qtq{and} q=m-1
\end{equation*}
has a unique solution $(m_4,q_4)$ with $q_4>1$.
We have explicitly
\begin{equation*}
m_4=\frac{3+\sqrt{13}}{2}\approx 3.30278\qtq{and} q_4=\frac{1+\sqrt{13}}{2}\approx 2.30278.
\end{equation*}
\item For each $m\in [m_4,M_d]$ there exists a number $r(m)>1$ such that
\begin{align*}
1<q<r(m)&\Longrightarrow \pi_q(m(m1)^{\infty})\ge m-1\qtq{and} q\le m-1,\\
q>r(m)&\Longrightarrow \pi_q(m(m1)^{\infty})<m-1.
\end{align*}
\item We have
\begin{equation}\label{81}
P_m\le r(m)<R_m,
\end{equation}
and the equality $r(m)=P_m$ holds only if $m=M_d$.
\end{enumerate}
\end{lemma}
\noindent See Figure \ref{f15}.
\begin{proof}\mbox{}
\begin{enumerate}[\upshape (i)]
\item See \eqref{916} below.
\item The function $q\mapsto \pi_q(m(m1)^{\infty})$ is continuous and decreasing in $q\in (1,\infty)$.
Since
\begin{equation*}
\lim_{q\to 1}\pi_q(m(m1)^{\infty})=\infty
\qtq{and}
\lim_{q\to \infty}\pi_q(m(m1)^{\infty})=0,
\end{equation*}
there exists a number $r(m)>1$ satisfying the relation
\begin{equation*}
\pi_q(m(m1)^{\infty})-(m-1)\sim r(m)-q.
\end{equation*}
It remains to show that $r(m)\le m-1$.
For $m=m_4$ we have equality by (i).
For $m>m_4$ we have (see \eqref{916})
\begin{equation*}
\pi_{m-1}(m(m1)^{\infty})<m-1\qtq{and} \pi_{r(m)}(m(m1)^{\infty})=m-1,
\end{equation*}
so that $r(m)<m-1$ by the decreasingness of the function $q\mapsto \pi_q(m(m1)^{\infty})$.
\item This follows from the definition of $r(m)$, \eqref{917} and \eqref{918}.
\end{enumerate}
\end{proof}
\begin{proof}[Proof of Proposition \ref{p15} (iii)]\mbox{}
\emph{First step.} If $1<q\le r(m)$, then $V'_{m,q}$ is countable.
Indeed, it follows from our assumptions that
\begin{equation*}
\pi_q(m(m1)^{\infty})\ge m-1
\qtq{and}
q\le m-1.
\end{equation*}
We claim that the blocks $1mm$ and $11$ are forbidden.
As usual, we may assume that $2<q\le r(m)$.
If $(c_i)\in\set{1,m}^{\infty}$ contains a block $c_nc_{n+1}=11$, then
\begin{equation*}
\pi_q(m-(c_{n+i}))\ge \frac{m-1}{q}\ge 1
\end{equation*}
and hence $(c_i)\notin V'_q$ by Lemma \ref{l42}.
If $(c_i)\in\set{1,m}^{\infty}$ does not contain any block $11$, but contains a block
\begin{equation*}
c_nc_{n+1}c_{n+2}=1mm,
\end{equation*}
then
\begin{equation*}
\pi_q(c_{n+i})\ge \pi_q\left( m(m1)^{\infty}\right)\ge m-1
\end{equation*}
by our assumptions, and hence $(c_i)\notin V'_q$ again by Lemma \ref{l42}.
It follows that $V'_{m,q}\subset\set{m^{\infty}, m^*(1m)^{\infty}}$, and hence it is countable.
\medskip
\emph{Second step.} If $q>r(m)$, then $V'_{m,q}$ has the power of continuum.
Indeed, since $\pi_q(m(m1)^{\infty})< m-1$ by our assumption, we may fix a large integer $k$ satisfying
\begin{equation*}
\pi_q\left( \left( m(m1)^k\right)^{\infty}\right)<m-1.
\end{equation*}
We claim that
\begin{equation*}
\set{m(m1)^{\ell}\ :\ \ell\ge k}^{\infty}\subset V'_{m,q}.
\end{equation*}
For the proof we take an arbitrary sequence $(c_i)$ from the left hand side set.
We have to check the conditions of Lemma \ref{l42}.
If $c_n=1$, then
\begin{equation*}
\pi_q(c_{n+i})\le \pi_q\left( \left( m(m1)^k\right)^{\infty}\right)<m-1
\end{equation*}
by the choice of $k$, and
\begin{equation*}
\pi_q(m-(c_{n+i}))<\pi_q\left( m-(m1)^{\infty}\right)=\pi_q(m-\delta')<1
\end{equation*}
because $q>p_m$ and therefore $\delta'\in V'_{m,q}$ by property (vii) in Theorem \ref{t12}.
\medskip
It follows from the preceding steps that $r_m=r(m)$.
\end{proof}
\section{Appendix}\label{s9}
We assume throughout this section that $m\ge 2$ and $q>1$.
As before, we will write $f\sim g$ if $f$ and $g$ have the same sign.
An elementary computation shows that the definitions
\begin{equation*}
P_m:=1+\sqrt{\frac{m}{m-1}}
\qtq{and}
R_m:=1+\frac{m}{m-1}
\end{equation*}
are equivalent to the relations
\begin{equation}\label{91}
(m-1)P_m(P_m-2)=1
\qtq{and}
(m-1)(R_m-2)=1.
\end{equation}
They will be often used below without explicit reference.
The first two subsections are related to the proof of Proposition \ref{p14}, the following four subsections to that of Proposition \ref{p15}.
Accordingly, the notations $m_d$, $M_d$ refer to the corresponding admissible sequences $0^{\infty}$ or $(10)^{\infty}$.
\subsection{The sequence $1^{\infty}$}\label{ss91}
We recall from \cite{KomLaiPed2010} that the first connected component of $[2,\infty)\setminus C$ is the interval $[2,M_d)$, associated with the smallest admissible sequence $d=0^{\infty}$, where $M_d=2.32472\ldots$ is the unique positive solution of the equation
\begin{equation*}
\pi_{P_m}(m^{\infty}-1^{\infty})=1,
\end{equation*}
and that $p_m=p_m''$ for each $m\in [2,m_d]$, where $p_m''$ is the unique positive solution of the equation
\begin{equation*}
\pi_{p_m''}(m^{\infty}-1^{\infty})=1.
\end{equation*}
It follows from the identity
\begin{equation*}
\pi_q(m^{\infty}-1^{\infty})-1=\frac{m-1}{q-1}-1=\frac{m-q}{q-1}
\end{equation*}
that
\begin{equation}\label{92}
p_m=m\qtq{for all} m\in [2,m_d]=[2,2.32472\ldots],
\end{equation}
and that
\begin{equation}\label{93}
\pi_{P_m}(m^{\infty}-1^{\infty})-1\sim m-P_m,
\end{equation}
whence $M_d=2.32472\ldots$ is the unique solution of the equation $P_m=m$.
Since
\begin{align*}
m=P_m
&\Longleftrightarrow m-1=\sqrt{\frac{m}{m-1}}\\
&\Longleftrightarrow (m-1)^3=m\\
&\Longleftrightarrow (m-1)^3-(m-1)-1=0,
\end{align*}
we conclude that $M_d=1+\alpha$ where $\alpha=1.32472\ldots$ denotes the unique positive root of the polynomial $x^3-x-1$ (this is the first Pisot number).
\subsection{The sequence $m1^{\infty}$}\label{ss92}
We have
\begin{align*}
\pi_q(m1^{\infty})-(m-1)
&=\frac{m-1}{q}+\frac{1}{q-1}-(m-1)\\
&=\frac{q-(m-1)(q-1)^2}{q(q-1)}.
\end{align*}
Hence
\begin{equation}\label{94}
\pi_q(m1^{\infty})=(m-1)\Longleftrightarrow q=r(m):=\frac{2m-1+\sqrt{4m-3}}{2m-2}.
\end{equation}
Furthermore, using \eqref{91} we obtain
\begin{align*}
\pi_{R_m}(m1^{\infty})-(m-1)
&\sim R_m-(m-1)(R_m-1)^2\\
&\sim R_m(R_m-2)-(R_m-1)^2\\
&=-1
\end{align*}
and
\begin{align*}
\pi_{P_m}(m1^{\infty})-(m-1)
&\sim P_m-(m-1)(P_m-1)^2\\
&\sim P_m^2(P_m-2)-(P_m-1)^2\\
&=P_m^3-3P_m^2+2P_m-1\\
&=(P_m-1)^3-(P_m-1)-1.
\end{align*}
Recalling that $x^3-x-1=0$ for $x=\alpha$ and observing that $x^3-x-1<0$ for $x\in (0,\alpha)$, we conclude from the last two relations that
\begin{equation}\label{95}
\pi_{R_m}(m1^{\infty})<(m-1)\qtq{for all} m\in [2,\infty)
\end{equation}
and (since the function $m\mapsto P_m$ is decreasing)
\begin{equation}\label{96}
\pi_{P_m}(m1^{\infty})-(m-1)\sim 1+\alpha-m=M_d-m.
\end{equation}
\subsection{The sequence $(m1)^{\infty}$}\label{ss93}
We recall from \cite{KomLaiPed2010} that the connected component $(m_d,M_d)=(2.80194\ldots,4.54646\ldots)$ of $[2,\infty)\setminus C$, associated with the admissible sequence $d=\set{(10)^{\infty}}$, is defined by the equations
\begin{equation*}
\pi_{P_{m_d}}((m_d1)^{\infty})=m-1\qtq{and} \pi_{P_{M_d}}(M_d^{\infty}-(M_d1)^{\infty})=1,
\end{equation*}
and that $p_m=\max\set{p_m',p_m''}$ for all $m\in [m_d,M_d]$, where $p_m'$ and $p_m''$ are uniquely defined by the equations
\begin{equation*}
\pi_{p_m'}((m1)^{\infty})=m-1\qtq{and} \pi_{p_m''}(m^{\infty}-(m1)^{\infty})=1.
\end{equation*}
We have
\begin{align*}
\pi_q((m1)^{\infty})-(m-1)
&=\frac{1}{q-1}+(m-1)\pi_q((10)^{\infty})-(m-1)\\
&=\frac{1}{q-1}+\frac{(m-1)q}{q^2-1}-(m-1)\\
&=\frac{(q+1)-(m-1)(q^2-q-1)}{q^2-1}.
\end{align*}
Hence
\begin{multline}\label{97}
\pi_q((m1)^{\infty})=(m-1)\Longleftrightarrow\\
q=r(m):=\frac{m-1+\sqrt{(m-1)^2+4}}{2}.
\end{multline}
Furthermore, using \eqref{91} we obtain
\begin{align*}
\pi_{P_m}((m1)^{\infty})-(m-1)
&\sim (P_m+1)-(m-1)(P_m^2-P_m-1)\\
&\sim (P_m+1)P_m(P_m-2)-(P_m^2-P_m-1).
\end{align*}
Hence
\begin{equation}\label{98}
\pi_{P_m}((m1)^{\infty})-(m-1)\sim P_m^3-2P_m^2-P_m+1
\end{equation}
after simplification, and, using the decreasingness of $m\mapsto P_m$,
\begin{equation}\label{99}
P_m^3-2P_m^2-P_m+1\sim m_d-m.
\end{equation}
Finally, it follows from the identity
\begin{align*}
\pi_q(m^{\infty}-(m1)^{\infty})-1
&=(m-1)\pi_q((01)^{\infty})-1\\
&=\frac{m-1}{q^2-1}-1
\end{align*}
that
\begin{align*}
\pi_{P_m}(m^{\infty}-(m1)^{\infty})-1
&=\frac{m-1}{P_m^2-1}-1\\
&=\frac{1}{P_m(P_m-2)(P_m^2-1)}-1\\
&\sim 1-P_m(P_m-2)(P_m^2-1).
\end{align*}
Hence
\begin{equation}\label{910}
\pi_{P_m}(m^{\infty}-(m1)^{\infty})-1\sim -P_m^4+2P_m^3+P_m^2-2P_m+1
\end{equation}
after simplification, and, since the right hand side tends to the positive limit $1$ as $m\to \infty$ and $P_m\to 2$,
\begin{equation}\label{911}
-P_m^4+2P_m^3+P_m^2-2P_m+1\sim m-M_d.
\end{equation}
\subsection{The sequence $(1m)^{\infty}$}\label{ss94}
We have
\begin{align*}
\pi_q(m^{\infty}-(1m)^{\infty})-1
&=(m-1)\pi_q((10)^{\infty})-1\\
&=\frac{(m-1)q}{q^2-1}-1\\
&=\frac{(m-1)q+(1-q^2)}{q^2-1}.
\end{align*}
Hence
\begin{multline}\label{912}
\pi_q(m^{\infty}-(1m)^{\infty})=1\Longleftrightarrow \\
q=r(m):=\frac{m-1+\sqrt{(m-1)^2+4}}{2}.
\end{multline}
Furthermore, using \eqref{91},
\begin{align*}
\pi_{R_m}(m^{\infty}-(1m)^{\infty})-1
&\sim (m-1)R_m+(1-R_m^2)\\
&\sim R_m+(R_m-2)(1-R_m^2)\\
&=-R_m^3+2R_m^2+2R_m-2
\end{align*}
and
\begin{align*}
\pi_{P_m}(m^{\infty}-(1m)^{\infty})-1
&\sim (m-1)P_m+(1-P_m^2)\\
&\sim 1+(P_m-2)(1-P_m^2)\\
&=-P_m^3+2P_m^2+P_m-1.
\end{align*}
If $m\in [2.80194\ldots,2.91286\ldots]$, then $R_m\in [2.5,3)$, and therefore
\begin{align*}
-R_m^3+2R_m^2+2R_m-2
&=(R_m+1)\left( 1.25-(R_m-1.5)^2\right) -1\\
&\le \frac{R_m+1}{4}-1<0.
\end{align*}
We conclude that
\begin{multline}\label{913}
\pi_{R_m}(m^{\infty}-(1m)^{\infty})<1\qtq{for all} \\
m\in [2.80194\ldots,2.91286\ldots].
\end{multline}
Furthermore, using \eqref{99} we obtain that
\begin{equation}\label{914}
\pi_{P_m}(m^{\infty}-(1m)^{\infty})\ge 1\qtq{for all} m\ge 2.80194\ldots,
\end{equation}
with equality only if $m=m_d=2.80194\ldots .$
\subsection{The sequence $mm1^{\infty}$}\label{ss95}
We have
\begin{align*}
\pi_q(mm1^{\infty})-(m-1)
&=\frac{m-1}{q}+\frac{m-1}{q^2}+\frac{1}{q-1}-(m-1)\\
&=\frac{q^2-(m-1)(q-1)(q^2-q-1)}{q^2(q-1)}\\
&=\frac{q^2-(m-1)(q^3-2q^2+1)}{q^2(q-1)},
\end{align*}
implying
\begin{equation}\label{915}
\pi_q(mm1^{\infty})-(m-1)\sim 1-(m-1)(q-2+q^{-2}).
\end{equation}
\subsection{The sequence $m(m1)^{\infty}$}\label{ss96}
We have
\begin{align*}
\pi_q(m(m1)^{\infty})-(m-1)
&=\frac{m-1}{q}+\frac{1}{q-1}+\frac{m-1}{q^2-1}-(m-1)\\
&=\frac{q(q+1)-(m-1)(q^3-q^2-2q+1)}{q(q^2-1)}.
\end{align*}
We evaluate this expression for $q=m-1$, $R_m$ and $P_m$.
For $q=m-1$ we get
\begin{align*}
\pi_{m-1}(m(m1)^{\infty})-(m-1)
&=\frac{q(q+1)-q(q^3-q^2-2q+1)}{q(q^2-1)}\\
&=\frac{-q(q^2-q-3)}{q^2-1}.
\end{align*}
Denoting by $q_4:=(1+\sqrt{13})/2$ the unique positive root of the numerator and setting $m_4:=1+q_4$, it follows that
\begin{equation}\label{916}
\pi_{m-1}(m(m1)^{\infty})-(m-1)\sim m_4-m.
\end{equation}
Next we remark that
\begin{align*}
\pi_{R_m}(m(m1)^{\infty})&-(m-1)\\
&\sim R_m(R_m+1)-(m-1)(R_m^3-R_m^2-2R_m+1)\\
&\sim R_m(R_m+1)(R_m-2)-(R_m^3-R_m^2-2R_m+1)\\
&=-1,
\end{align*}
so that
\begin{equation}\label{917}
\pi_{R_m}(m(m1)^{\infty})<m-1.
\end{equation}
Finally,
\begin{align*}
\pi_{P_m}(m(m1)^{\infty})&-(m-1)\\
&\sim P_m(P_m+1)-(m-1)(P_m^3-P_m^2-2P_m+1)\\
&\sim P_m^2(P_m+1)(P_m-2)-(P_m^3-P_m^2-2P_m+1),
\end{align*}
yielding
\begin{equation*}
\pi_{P_m}(m(m1)^{\infty})-(m-1)\sim P_m^4-2P_m^3-P_m^2+2P_m-1.
\end{equation*}
Using \eqref{911} we conclude that
\begin{equation}\label{918}
\pi_{P_m}(m(m1)^{\infty})-(m-1)\ge 0\qtq{if} m\in [2,M_d],
\end{equation}
with equality only if $m=M_d$.
|
\section{Background}
The model describes an outbreak epidemic with host-vector population dynamics in a geographical region. The epidemic outbreak begins at time 0 in a small sub-region in which the epidemic disease is not yet present. The goal of the model is to aid understanding of how the introduction of a very small number of cases in a specific location in the geographic region will result in a dissipation or a sustained and growing epidemic. The focus of the study is on the importance of spatial effects in these possible outcomes. If the equations of the model do not depend on spatial considerations, then a corresponding system of ordinary differential equations can be analyzed for their asymptotic behavior (Appendix).
The geographical region is denoted by $\Omega \subset R^2$. The background population of uninfected hosts in
$\Omega$ has geographic density $H_u(x,y)$, which is unchanging in time in the demographic and epidemic context of the outbreak. Thus, the model is viewed as applicable to an early phase of the epidemic, during which the epidemic does not alter the basic geographic and demographic population structure of hosts, and the susceptible host population is not altered significantly by immunity to re-infection.
\subsection{Compartments of the Model}
The model consists of the following compartments:
\vspace{0.1in}
\begin{itemize}
\item[ ] The density of infected hosts $H_i(t,x,y)$ at time $t$ at $(x,y) \in \Omega$, with initial condition $H_{i0}(x,y)$.
\item[ ] The density of uninfected vectors $V_u(t,x,y)$ at time $t$ at $(x,y) \in \Omega$, with initial condition
$V_{u0}(x,y)$.
\item[ ] The density of infected vectors $V_i(t,x,y)$ at time $t$ at $(x,y) \in \Omega$, with initial condition $V_{i0}(x,y)$.
\end{itemize}
The initial state $H_{i0}(x,y)$ consists of a relatively small number of infected hosts located at time $0$ in a small sub-region of $\Omega$.
This input corresponds to an arrival of infected hosts from outside $\Omega$. This input is assumed to have a threshold level, which may include multiple cases produced from arriving cases. The background uninfected mosquito population has an initial state $V_{u0}(x,y)$, which decreases as the infected vector population increases. The initial population of infected vectors $V_{i0}(x,y)$ in $\Omega$ is assumed proportional to $H_{i0}(x,y)$.
\subsection{Equations of the Model}
The equations of the model in the case that transmission from vectors to hosts is year-round are
\begin{eqnarray}
\label{ZIKAmodel}
\frac{\partial}{\partial t} H_i(t,x,y) &=& \nabla \cdot \delta_1(x,y) \nabla H_i(t,x,y)
- \lambda(x,y) \, H_i(t,x,y) \label{Eq-1} \\
&+& \sigma_1(x,y) \, H_u(x,y) \, V_i(t,x,y), \nonumber \\
\frac{\partial}{\partial t} V_u (t,x,y) &=& \nabla \cdot \delta_2(x,y) \nabla V_u(t,x,y)
- \sigma_2(x,y) V_u(t,x,y) H_i(t,x,y) \label{Eq-2} \\
&+& \beta(x,y) \bigg( V_u(t,x,y) + V_i(t,x,y) \bigg) \nonumber \\
&-& \mu(x,y) \bigg( V_u(t,x,y) + V_i(t,x,y) \bigg) V_u(t,x,y), \nonumber \\
\frac{\partial}{\partial t} V_i (t,x,y) &=& \nabla \cdot \delta_2(x,y) \nabla V_i(t,x,y)
+ \sigma_2(x,y) V_u(t,x,y) H_i(t,x,y) \label{Eq-3} \\
& & \quad \quad - \mu(x,y) \bigg( V_u(t,x,y) + V_i(t,x,y) \bigg) V_i(t,x,y). \nonumber
\end{eqnarray}
\subsection{Well-posedness of the Model}
{\bf Theorem.}
Let $\Omega$ be a bounded domain in $R^2$ with smooth boundary $\partial \Omega$ such that $\Omega$ lies locally on one side of $\partial\Omega$.
Let $\beta$, $\mu$, $\lambda$, $\sigma_1$, $\sigma_2, \delta_1, \delta_2
\, \in C_+^0(\overline{\Omega})$, and let
$H_{u}, I_0, V_{u0} ,V_{i0} \in C_+^1(\overline{\Omega})$.
There exists a unique global classical solution $\{H_i(t),V_u(t),V_i(t)\} \in C_{+}^{1}(\overline{\Omega}), \, t \geq 0$, to (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}), satisfying boundary conditions
$$
\frac{\partial}{\partial \eta} H_i(t,x,y) = 0, \,
\frac{\partial}{\partial \eta} V_u(t,x,y) = 0, \,
\frac{\partial}{\partial \eta} V_i(t,x,y) = 0, \, (x,y) \in \partial \Omega, \, t >0
$$
and initial conditions
$$
H_i(0,x,y) = H_{i0}(x,y), \, V_u(0,x,y) = V_{u0}(x,y), \, V_i(0,x,y) = V_{i0}(x,y), \, (x,y) \in \Omega.
$$
{\bf Proof.}
We first observe that a unique classical solution $\{H_i(t),V_u(t),V_i(t)\}$ exists in $C^{1}(\overline{\Omega})$
on a maximal interval of existence $[0,T_{max})$ (\cite{Martin}, \cite{Pazy}, \cite{Smoller}).
Standard arguments (\cite{Smoller}) guarantee that $\{H_i(t),V_u(t),V_i(t)\}$ remain nonnegative for $t \in [0,T_{max})$.
Moreover, the classical solution can be globally defined if we can establish uniform \textit{a priori} bounds.
Set $M(t,x,y) = V_u(t,x,y) + V_i(t,x,y)$ and add equations (\ref{Eq-2}) and (\ref{Eq-3}) to obtain
\begin{eqnarray}
\frac{\partial}{\partial t} M(t,x,y)
&=& \nabla \cdot \delta_2(x,y) \nabla M(t,x,y) \label{Eq-4} \\
&+& \, \, \beta(x,y) M(t,x,y)\, \, - \, \, \mu(x,y) M(t,x,y)^2. \nonumber
\end{eqnarray}
Theorem 1 in \cite{Fitz-Langlais} guarantees the existence of a unique global classical solution
$M(t) \in C_+^1(\overline{\Omega})$ to Equation (\ref{Eq-4}) satisfying
$$
\frac{\partial}{\partial \eta} M(t,x,y) = 0, \, (x,y) \in \partial \Omega, \, t \geq 0, \, M(0,x,y) = V_{u0}(x,y) + V_{i0}(x,y), \, (x,y) \in \Omega.
$$
Further, in \cite{Fitz-Langlais} it is proved that there exists $\overline{M} \in C_+^0(\overline{\Omega})$, $\overline{M} \neq 0$, such that $\lim_{t \rightarrow \infty}M(t) = \overline{M} \in C_+^0(\overline{\Omega})$. We note that the disease free equilibrium of (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) is $(0,\overline{M},0)$. From \cite{Fitz-Langlais} there exists $N_1 > 0$ such that
$max_{t \geq 0} \| M(t) \|_{C_+^0(\overline{\Omega})} < N_1$, which implies $\| V_i(t) \|_{C_+^0(\overline{\Omega})}, \,
\| V_u(t) \|_{C_+^0(\overline{\Omega})} < N_1$. Then, since $\lambda > 0$ in (\ref{Eq-1}), there exists $N_2 > 0$ such that $\| H_i(t) \|_{C_+^0(\overline{\Omega})} < N_2$. Consequently, the solution exists globally on $[0,\infty)$.
\section{The Spatially Structured Basic Reproduction Number}
Define the basic reproduction number of the model (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) as
$$R_0(x,y) = \frac{\sigma_1(x,y) \sigma_2(x,y) H_u(x,y)}{\lambda(x,y) \mu(x,y)}.$$
$R_0(x,y)$ is interpreted as the average number of new cases generated by a single case at a given location $(x,y)$
in $\Omega$. Our motivation for this definition is the basic reproduction number $R_0$ of the spatially independent model (Appendix). Simulations of the spatially dependent model show that for certain parameterizations of equations (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}), the solutions have the following behavior:
(1) If $R_0(x,y) < 1$ everywhere in $\Omega$, then the populations of both infected hosts and infected vectors extinguish, and the populations converge to the disease free equilibrium.
(2) If $R_0(x,y) > 1$ in some sub-region $\Omega_0 \subset \Omega$, then the populations of both infected hosts and infected vectors converge to an endemic equilibrium independently of the initial conditions, and even if the average value of $R_0(x,y)$ in all of $\Omega$ is $ < 1$.
\section{The 2015 Zika Outbreak in Rio de Janeiro Municipality}
We apply a version of the model (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) to the 2015 Zika outbreak in Rio de Janeiro Municipality in Brazil. Because disease transmission in the Municipality is seasonal, equations (\ref{Eq-2}) and (\ref{Eq-3}) must be modified to account for seasonality. We assume that there is a mosquito population breeding term
$\beta(t,x,y)$, depending on time. We also assume that there is an on-going mosquito loss term $\mu_1(x,y)$, corresponding to the average mosquito life-span $1 / \mu_1(x,y)$, independent of the carrying capacity loss term $\mu(x,y)$. The modified equations are
\begin{eqnarray}
\label{ZIKAmodelModified}
\frac{\partial}{\partial t} V_u (t,x,y) &=& \nabla \cdot \delta_2(x,y) \nabla V_u(t,x,y)
- \sigma_2(x,y) V_u(t,x,y) H_i(t,x,y) \label{Eq-2mod} \\
&+& \beta(t,x,y) \bigg( V_u(t,x,y) + V_i(t,x,y) \bigg) - \mu_1(x,y) V_u(t,x,y) \nonumber \\
& & \quad - \mu(x,y) \bigg( V_u(t,x,y) + V_i(t,x,y) \bigg) V_u(t,x,y), \nonumber \\
\frac{\partial}{\partial t} V_i (t,x,y) &=& \nabla \cdot \delta_2(x,y) \nabla V_i(t,x,y)
+ \sigma_2(x,y) V_u(t,x,y) H_i(t,x,y) \label{Eq-3mod} \\
& & - \mu(x,y) \bigg( V_u(t,x,y) + V_i(t,x,y) \bigg) V_i(t,x,y) \nonumber - \mu_1(x,y) V_i(t,x,y). \nonumber
\end{eqnarray}
The host population are the people in the Municipality, which in 2016 is approximately 6,000,000, in a geographical region of approximately 1,200 square kilometers (Source: \textit{Instituto Brasileiro de Geografia e Estatistica}). The vector population is the female \textit{Aedes aegypti} mosquito. The Municipality comprises 33 sub-districts, with population densities ranging from 1,000 to 50,000 inhabitants per square kilometer (Figure 1).
A small number of cases were recorded in the Municipality into the summer of 2015, with the highest number of cases in the eastern region of the Municipality (\cite{Brasil}, \cite{Honorio}).
The Brazilian Health Ministry (\cite{Boletim13},\cite{Boletim16}) reported that Rio de Janeiro State (population approximately 16,000,000) registered a count of 25,930 cumulative cases from January 1, 2016 to April 1, 2016 (with incidence of 156.7 cases per 100,000 inhabitants), and 32,312 cumulative cases by April 23, 2016 (with incidence of 195.2 cases per 100,000 inhabitants). The Ministry (\cite{Boletim17},\cite{Boletim18}) reported no new cases in the State from April 24, 2016 to May 7, 2016.
In \cite{Bastos} the weekly case data for Rio de Janeiro Municipality is given from November 1, 2015 through April 10, 2016, during which time the reporting of cases became mandatory. The period can be viewed as the 2015-2016 seasonal mosquito transmission period of the epidemic in the Municipality.
\subsection{Parameterization of the Rio de Janeiro Model}
We simulate this case data for the Rio de Janeiro Municipality with the following parameterization: The time units are weeks. The spatial units are kilometers and $\Omega = (-25,25) \times (-12,12)$.
The average length of the infectious period of infected people is approximately 1 to 2 weeks and we set
$\lambda(x,y) = 1.0$ (\cite{CDC}). The average lifespan of female \textit{Aedes aegypti} mosquitoes is approximately two weeks in an urban environment (\cite{Brady},\cite{Kucharsky},\cite{Otero}), and we set $\mu_1(x,y) =0.5$.
The total uninfected host population is $6,000,000$, with geographical density function $H_u(x,y) = 50.0 + 10^2 \, (1.0 + \sin(0.02 \pi x) \, \cos(0.03 \pi y))$ (Figure 2A), which corresponds approximately to the population density distribution in Figure 1.
Set $\mu_0 = 0.0001$ and the density dependent mosquito loss function
$\mu(x,y) = \mu_0 (1.0 +100 \, gauss(20.0,30.0,x) \times gauss(0.0,30.0,y))$ (Figure 2B), which corresponds to higher levels of mosquito control in the eastern region of the Municipality, where the population density is highest.
Here $gauss(m,sd,x)$ is the probability density function in $x$ of the normal distribution function with mean $m$ and standard deviation $sd$.
Set the transmission parameters $\sigma_1(x,y) = 0.0000051$, $\sigma_2(x,y) =0.78$ (we assume that individual mosquitoes bite multiple people, people receive multiple bites, and the probability of infection of mosquitoes is much higher than the probability of infection of people). Set $\delta_1 = 0.2$, which is a simplified estimate of human dispersal in urban settings. Set $\delta_2 = 0.2$, which is consistent with an estimated adult mosquito dispersal of $30 - 50 \, m$ per day (\cite{Otero}).
The time dependent mosquito breeding function is $\beta(t,x,y) =
\beta_0 \, emg(t,\bar{\mu},\bar{\sigma},\bar{\lambda)}$, where $\beta_0 = 300.0$ and $emg$ is the shifted exponentially modified gaussian
$$emg(t,\bar{\mu},\bar{\sigma},\bar{\lambda)}) = \frac{\bar{\lambda}}{2} Exp\bigg(\frac{\bar{\lambda}}{2}(2 \bar{\mu} + \bar{\lambda} \bar{\sigma}^2 - 2 \, t)\bigg)
Erfc\bigg(\frac{1}{\sqrt{2} \, \bar{\sigma}} (\bar{\lambda} \bar{\sigma}^2 + \bar{\mu} -t)\bigg)$$
Here $Erfc$ is the complementary error function. The parameters are $\bar{\mu} = -2.0$, $\bar{\sigma} = 5.0$,
$\bar{\lambda} = 0.2$.
The graph of the seasonal mosquito breeding function $\beta$ is given is Figure 3 ($\beta$ is independent of $x$ and $y$).
\subsection{Initialization of the Rio de Janeiro Model}
The outbreak begins at time 0 on November 1, 2015 in a northeastern location of the Municipality with high population density. The total number of infected cases at time $0$ is $10$, with spatial distribution $H_i(0,x,y) = 10.0 \, gauss(15.0,1.0,x) \times gauss(0.0,1.0,y)$. At time $0$ the total number of uninfected mosquitoes is $120,000$, distributed uniformly throughout the Municipality. The total number of infected mosquitoes at time $0$ is $100$, with $V_i(0,x,y) = 10.0 H_i(0,x,y)$.
\subsection{Simulation of the Rio de Janeiro Model}
Example 1. The simulation of the model (\ref{Eq-1}), (\ref{Eq-2mod}), (\ref{Eq-3mod}) over the time period November 1, 2015 to May 21, 2016 is graphed in Figures 4, and 5. The simulation agrees qualitatively with the weekly reported case data for Rio de Janeiro Municipality in \cite{Bastos} (Figure 4). The spatial distributions of infected people expand from a very small number of initial cases in a small sub-region of the Municipality, and disperses throughout the eastern region of the Municipality (Figure 5). The mosquito population rises rapidly and reaches carrying capacity at approximately 14 million in earlier 2016. During much of mosquito season, the ratio of mosquitoes to people is approximately $2$ to $1$, which agrees with the ratio in (\cite{Manore}).
Example 2. We repeat the simulation with the only change the location of the initial infected cases. We take $H_i(0,x,y) = 10.0 \, gauss(10.0,1.0,x) \times gauss(-5.0,1.0,y)$ (Figure 6). The infected population again expands from the initial location and disperses throughout the eastern region of the Municipality, but at approximately one-half the number of infected cases as in Example 1. The reason is that the density of susceptible people is lower in this initial location than the initial location in Example 1.
\section{Conclusions}
The model (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) describes criss-cross vector-host transmission dynamics of an epidemic outbreak in a geographical region $\Omega$.
The outbreak occurs with a small number of infected hosts in a small sub-region of the much larger geographical region $\Omega$. The diffusion terms describe the on-going average spatial movement of vectors and hosts in the geographical region. The focus of the model is to describe the geographical spread from an initial localized immigration into the region, in terms of the epidemiological properties of the outbreak vector-host transmission dynamics.
\subsection{Summary of the Outcomes of the Outbreak Model}
We prove that the partial differential equations model (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) is mathematically well-posed, and compare its properties to an analogous ordinary differential equations model in the spatially independent case (Appendix). The outcomes of the model depend on the spatially distributed basic reproduction number
$R_0(x,y)$.
If $R_0(x,y) < 1$ everywhere in $\Omega$, then the epidemic will extinguish.
If $R_0(x,y) > 1$ in some sub-region of $\Omega$, then the epidemic will spread and converge to an endemic equilibrium throughout all of $\Omega$, independently of the location of the sub-region.
\subsection{Summary of the Model Applied to the Zika Outbreak in Rio de Janeiro}
The model (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) is modified to allow seasonality of the vector population, and applied to the 2015-2016 Zika outbreak in Rio de Janeiro Municipality. A simulation of the model provides qualitative agreement with the case data reported by the Brazilian Ministry of Health in \cite{Bastos}. The model simulation suggests that the Zika epidemic in Rio de Janeiro Municipality, will rise each season from initial locations with very small numbers of infected people, and spread throughout the Municipality. The evolution of the epidemic depends on the initial location of infected cases. In general, the evolution of the epidemic depends on the parameterization and initialization of the model, and can be limited only by reduction of the disease parameters throughout the entire region. If the number of actual infected cases is significantly higher than the number of reported cases, then the parameterization must be adjusted accordingly.
\section*{Appendix}
The equations (\ref{Eq-1}),(\ref{Eq-2}),(\ref{Eq-3}) without spatial dependence are
\begin{eqnarray}
\label{ZIKA-ODEmodel}
\quad \quad \frac{d}{d t} H_i(t)&=& - \lambda H_i(t)
+ \sigma_1 \, V_i(t) \, H_u \label{Eq-7X} \\
\frac{d}{d t} V_u (t) &=&
\beta ( V_u(t) + V_i(t) ) -\sigma_2 V_u(t) H_i(t)
- \mu ( V_u(t) + V_i(t) ) V_u(t) \label{Eq-8X} \\
\frac{d}{d t} V_i (t) &=&
\sigma_2 V_u(t) H_i(t)
- \mu ( V_u(t) + V_i(t) ) V_i(t) \label{Eq-9X}
\end{eqnarray}
with initial conditions $H_i(0) = H_{i0}$, $V_u(0) = V_{u0}$, $V_i(0) = V_{i0}$.
Set $R_0 = H_u \sigma_1 \sigma_2 / \lambda \mu$. The behavior of solutions of equations (\ref{Eq-7X}),(\ref{Eq-8X}), (\ref{Eq-9X}) can be classified as follows:
{\bf Proposition}
If $R_0 < 1$, then the only steady states of (\ref{Eq-7X}),(\ref{Eq-8X}),(\ref{Eq-9X}) in $R_+^3$ are $ss_0 = (0,0,0)$, which is unstable in $R_+^3$, and $ss_1 = (0,\beta/\mu,0)$, which is proportional to $\beta$ and locally exponentially asymptotically stable in $R_+^3$.
If $R_0 < 1$, $H_i(0) > 0$, and $V_i(0) = 0$, then $(H_i(t),V_u(t),V_i(t))$ converges to $(0,\overline{M},0)$.
If $R_0 > 1$, then $ss_0$ and $ss_1$ are unstable in $R_+^3$
and there is another steady state in $R_+^3$,
$$ss_2 = \bigg( \frac{\beta (H_u \sigma_1 \sigma_2 -\lambda \mu)}{\lambda \mu \sigma_2},
\frac{\beta \lambda}{H_u \sigma_1 \sigma_2},
\frac{\beta (H_u \sigma_1 \sigma_2 -\lambda \mu)}{H_u \mu \sigma_1 \sigma_2} \bigg)$$
$$= \bigg( \frac{\beta(R_0 -1)}{\sigma_2},
\frac{\beta}{R_0 \mu} ,
\frac{\lambda \beta (R_0 - 1)}{ H_u \sigma_1 \sigma_2} \bigg).$$
which is locally exponentially asymptotically stable in $R_+^3$.
{\bf Proof.}
It can be verified that the steady states of (\ref{Eq-7X}),(\ref{Eq-8X}),(\ref{Eq-9X}) in $R_+^3$ are $ss_0, ss_1$, and $ss_2$. The Jacobian of (\ref{Eq-7X}),(\ref{Eq-8X}),(\ref{Eq-9X}) at $ss_0$ is
$$J(0,0,0) = \left[ \begin{array}{ccc}
-\lambda & 0 & H_u \sigma_1 \\
0 & \beta & \beta \\
0 & 0 & 0 \end{array} \right]$$
with eigenvalues $\{-\lambda,\beta,0\}$, which means that $(0,0,0)$ is unstable.
Let $M(t) = V_i(t) + V_u(t)$. Equations (\ref{Eq-8X}) and (\ref{Eq-9X}) imply $M^{\prime}(t) = \beta M(t) - \mu M(t)^2$, $\lim_{t \rightarrow \infty}M(t) = \overline{M} = \beta / \mu$.
If $H_i(0) > 0$ and $V_i(0) = 0$, then (\ref{Eq-7X}) implies $H_i^{\prime}(0) < 0$. Assume there is a smallest positive time $t^{\ast}$ such that $H_i^{\prime}(t^{\ast}) = 0$. Then (\ref{Eq-7X}) implies $H_i(t^{\ast}) = (\sigma_1 H_u / \lambda) V_i(t^{\ast})$. If $R_0 < 1$, then (\ref{Eq-9X}) implies
$$
V_i^{\prime}(t^{\ast}) = (\sigma_1 \sigma_2 H_u / \lambda) V_i(t^{\ast}) ( M(t^{\ast}) - V_i(t^{\ast}) )
- \mu V_i(t^{\ast}) M(t^{\ast}) < - (\sigma_1 \sigma_2 H_u / \lambda) V_i(t^{\ast})^2 < 0.
$$
Then (\ref{Eq-7X}) implies
$
H_i^{\prime \prime}(t^{\ast}) = - \lambda H_i^{\prime}(t^{\ast}) + \sigma_1 H_u V_i^{\prime}(t^{\ast}) < 0,
$
which implies $H_i(t)$ is strictly decreasing at $t^{\ast}$, yielding a contradiction. Thus, $H_i(t)$ is strictly decreasing for all $t \geq 0$. Let $H_{i,\infty} = \lim_{t \rightarrow \infty } H_i(t) \geq 0$. Assume $H_{i,\infty} >0 $.
Then (\ref{Eq-9X}) implies $lim_{t \rightarrow \infty } V_i(t) = \lambda H_{i.\infty} / \sigma_1 H_u > 0$. Equation (\ref{Eq-8X}) then implies
$\lim_{t \rightarrow \infty}V_u(t) = \beta \overline{M} / (\sigma_2 H_{i,\infty} + \mu \overline{M})$. Then
$(H_{i,\infty},\beta \overline{M} / (\sigma_2 H_{i,\infty} + \mu \overline{M}),\lambda H_{i,\infty} / (\sigma_1 H_u))$ is a steady state of
(\ref{Eq-7X}),(\ref{Eq-8X}),(\ref{Eq-9X}). If $R_0 < 1$, then $H_{i,\infty} =0$, yielding a contradiction. Thus, $H_{i,\infty} = 0$.
The eigenvalues of the Jacobian of (\ref{Eq-7X}),(\ref{Eq-8X}),(\ref{Eq-9X}) at $ss_1$
$$J(0,\beta / \mu,0) = \left[ \begin{array}{ccc}
-\lambda & 0 & H_u \sigma_2 \\
- \beta \sigma_1 / \mu & - \beta &0 \\
\beta \sigma_1 / \mu & 0 & - \beta \end{array} \right]$$
\vspace{0.1in}
\noindent
are
$$\{- \beta,
\frac{ - \beta - \lambda - \sqrt{(\beta - \lambda)^2 + 4 R_0 \beta \lambda}}{2},
\frac{ - \beta - \lambda + \sqrt{(\beta - \lambda)^2 + 4 R_0 \beta \lambda}}{2}\}.$$
Thus, $J(0,\beta / \mu,0)$ is unstable if $R_0 > 1$ and locally exponentially asymptotically stable if $R_0 < 1$.
The Jacobian of (\ref{Eq-7X}),(\ref{Eq-8X}),(\ref{Eq-9X}) at $ss_2$ is
$$\left[ \begin{array}{ccc}
- \lambda & 0 & H_u \sigma_1 \\
- \frac{\beta \lambda}{H_u \sigma_1} & \beta(1- \frac{\lambda \mu}{H_u \sigma_1 \sigma_1}
- \frac{H_u \sigma_1 \sigma_2}{\lambda \mu}) & \beta (1 - \frac{\lambda \mu}{H_u \sigma_1 \sigma_2}) \\
\frac{\beta \lambda}{H_u \sigma_1} & \beta(- 2 + \frac{\lambda \mu}{H_u \sigma_1 \sigma_2}
+\frac{H_u \sigma_1 \sigma_2}{\lambda \mu}) & \beta (- 2 + \frac{\lambda \mu}{H_u \sigma_1 \sigma_2}) \end{array} \right]
$$
$$ = \left[ \begin{array}{ccc}
- \lambda & 0 & \frac{R_0 \lambda \mu}{\sigma_2} \\
- \frac{\beta \sigma_2}{R_0 \mu} & \beta(1- \frac{1}{R_0}
- R_0) & \beta (1 - \frac{1}{R_0}) \\
\frac{\beta \sigma_2}{R_0 \mu} & \beta(- 2 + \frac{1}{R_0}
+ R_0) & \beta (- 2 + \frac{1}{R_0}) \end{array} \right].
$$
\vspace{0.1in}
\noindent
with eigenvalues
$$\{- \beta,
\frac{ - R_0 \beta - \lambda - \sqrt{(R_0 \beta - \lambda)^2 + 4 \beta \lambda}}{2},
\frac{ - R_0 \beta - \lambda + \sqrt{(R_0 \beta - \lambda)^2 + 4 R_0 \beta \lambda}}{2}\}.
$$
Since
$
- (R_0 \beta + \lambda)^2 + (R_0 \beta - \lambda)^2 + 4 \beta \lambda = - 4 (R_0 -1) \beta \lambda < 0
$
if $R_0 >1$, the eigenvalues of the Jacobian at $ss_2$ are strictly negative if $R_0 >1$, which means that $ss_2$ is locally exponentially asymptotically stable if $R_0 >1$.
\section*{Authors contributions}
All authors conceived and developed the study. All authors read and approved the final manuscript.
|
\section{Introduction and overview}
\label{intro}
In the study of partial differential equations (PDEs),
conserved integrals and local continuity equations have many important uses.
They yield fundamental conserved quantities and constants of motion,
which along with symmetries are an intrinsic coordinate-free aspect of the structure
of a PDE system.
They also yield potentials and nonlocally-related systems.
They provide conserved norms and estimates,
which are central to the analysis of solutions.
They detect if a PDE system admits an invertible transformation into a target class of PDE systems
(\eg/, nonlinear to linear, or linear variable coefficient to constant coefficient).
They typically indicate if a PDE system has integrability structure.
They allow checking the accuracy of numerical solution methods
and also give rise to good discretizations (\eg/, conserving energy or momentum).
For a dynamical PDE system in one spatial dimension,
a local continuity equation is a total divergence expression
\begin{equation}
D_t T + D_x X =0
\end{equation}
vanishing on the solution space of the system,
where $T$ is a conserved density and $X$ is a spatial flux.
(Here $D_t$ and $D_x$ are total derivatives with respect to time and space coordinates.)
Every local continuity equation physically represents a conservation law
for the quantity $T$.
The conservation law can be formulated by integrating the local continuity equation over any spatial domain $\Omega\subseteq\mathbb{R}$,
yielding
\begin{equation}\label{conslaw-onedim}
\frac{d}{dt}\int_\Omega T dx = -X\Big|_{\partial\Omega} .
\end{equation}
This shows that the rate of change of the integral of the conserved density $T$
on the domain $\Omega$ is balanced by the net outward flux
through the domain endpoints $\partial\Omega$.
In two and three spatial dimensions,
local continuity equations have the more general total divergence form
\begin{equation}\label{continuityeqn}
D_t T + {\rm Div\,}\vec X=0 .
\end{equation}
The corresponding physical conservation law is given by
\begin{equation}\label{conslaw-multidim}
\frac{d}{dt}\int_\Omega T dV = -\oint_{\partial\Omega}\vec X\cdot\vec\nu dA
\end{equation}
where $\Omega$ is a spatial domain
and $\vec\nu$ is the outward unit normal of the domain boundary.
This conservation law shows that the net outward flux of $\vec X$ integrated over $\partial\Omega$
balances the rate of change of the integral of the conserved density $T$ on $\Omega$.
Another type of conservation law in two and three spatial dimensions
can be formulated on the boundary of a spatial domain $\Omega$,
\begin{equation}\label{conslaw-boundary}
\frac{d}{dt}\oint_{\partial\Omega} \vec T\cdot\vec\nu dA = 0
\end{equation}
holding on the solution space of a PDE system.
This boundary conservation law corresponds to a local continuity equation \eqref{continuityeqn}
in which the conserved density is a total spatial divergence, $T={\rm Div\,}\vec T$,
and the flux is a total spatial curl, $\vec X = {\rm Div\,}\mathbf\Gamma$,
where $\mathbf\Gamma$ is an antisymmetric tensor.
Its physical meaning is that the net flux of $\vec T$ over $\partial\Omega$
is a constant of the motion for the PDE system.
When hydrodynamical PDE systems for fluid/gas flow are considered,
a more physically useful formulation of conservation laws is given by
considering moving spatial domains $\Omega(t)$,
or moving spatial boundaries $\partial\Omega(t)$,
that are transported by the flow of the fluid/gas.
For a moving domain,
a physical conservation law has the form
\begin{equation}\label{conslaw-movingdomain}
\frac{d}{dt}\int_{\Omega(t)} T dV = -\oint_{\partial\Omega(t)}(\vec X-T\vec u)\cdot\vec\nu dA
\end{equation}
where $\vec u$ is the fluid/gas velocity,
and $\vec X-T\vec u=\vec{\mathcal X}$ is the moving flux.
The local continuity equation \eqref{continuityeqn}
is then equivalent to a transport equation
\begin{equation}\label{transporteqn}
(D_t +\vec u\cdot D_x)T = -(\nabla\cdot\vec u)T -{\rm Div\,}\vec{\mathcal X}
\end{equation}
for the conserved density $T$,
with $D_t +\vec u\cdot D_x$ being the material (advective) derivative,
and $\nabla\cdot\vec u$ being the expansion or contraction factor of
an infinitesimal moving volume of the fluid/gas.
If the net moving flux over the domain boundary $\partial\Omega(t)$ vanishes,
then the integral of the conserved density $T$ on the moving domain $\Omega(t)$
is a constant of motion.
For a moving boundary,
the physical form of a conservation law is given by
\begin{equation}\label{conslaw-movingboundary}
\frac{d}{dt}\oint_{\partial\Omega(t)}\vec T\cdot\vec\nu dA =0
\end{equation}
which shows the net flux of $\vec T$ integrated over $\partial\Omega(t)$
is a constant of motion.
In the corresponding transport equation \eqref{transporteqn},
the conserved density is a total spatial divergence,
$T={\rm Div\,}\vec T$,
and the moving flux is a total spatial curl,
$\vec{\mathcal X} = {\rm Div\,}\mathbf\Gamma$,
where $\mathbf\Gamma$ is an antisymmetric tensor.
A related type of conservation law in two and three spatial dimensions
arises from the total spatial divergence of a flux vector that is not a total spatial curl,
\begin{equation}
{\rm Div\,}\vec X=0,
\quad
\vec X \neq {\rm Div\,}\mathbf\Gamma
\end{equation}
holding on the solution space of a PDE system.
This yields a physical conservation law
on any spatial domain $\Omega$ enclosed by
an inner boundary $\partial_-\Omega$ and an outer boundary $\partial_+\Omega$.
The conservation law shows that the net outward flux across each boundary
is the same,
\begin{equation}\label{div-conslaw}
\oint_{\partial_-\Omega}\vec X\cdot\vec\nu_- dA = \oint_{\partial_+\Omega}\vec X\cdot\vec\nu_+ dA
\end{equation}
where $\vec\nu_\mp$ is the outward unit normal of the respective boundaries.
The most well-known method \cite{Olv,1stbook,2ndbook}
for finding conservation laws is Noether's theorem,
which is applicable only to PDE systems that possess a variational principle.
Noether's theorem shows that the infinitesimal symmetries of the variational principle
yield conserved integrals \eqref{conslaw-onedim}, \eqref{conslaw-multidim}, \eqref{div-conslaw} of the PDE system,
including conserved boundary integrals \eqref{conslaw-boundary}
when the PDE system satisfies a differential identity.
In the case of PDE systems that possess a generalized Cauchy-Kovalevskaya form \cite{Mar-Alo,Olv},
all conserved integrals \eqref{conslaw-onedim}, \eqref{conslaw-multidim}, \eqref{div-conslaw} arise from Noether's theorem.
This direct connection between conserved integrals and symmetries
is especially useful because,
typically, the symmetries of a given PDE system have a direct physical meaning
related to basic properties of the system,
while, computationally,
all infinitesimal symmetries of a given PDE system can be found in a systematic way by solving a linear system of determining equations.
Over the past few decades,
a modern formulation of Noether's theorem has been developed
in which the components of a variational symmetry are expressed
as the components of a multiplier whose product with a given variational PDE system
yields a total divergence that reduces on the space of solutions of the PDE system
to a local continuity equation.
The main advantage of this reformulation is that
multipliers can be sought for any given PDE system,
regardless of whether it possesses a variational principle or not.
In general, multipliers are simply the natural PDE counterpart of
integration factors for ordinary differential equations \cite{1stbook},
and for any given PDE system,
a linear system of determining equations can be formulated \cite{Olv,2ndbook}
to yield all multipliers.
As a consequence, local continuity equations can be derived
without any restriction required on the nature of the PDE system.
Moreover, for PDE systems that possess a generalized Cauchy-Kovalevskaya form,
all conserved integrals \eqref{conslaw-onedim}, \eqref{conslaw-multidim}, \eqref{div-conslaw} arise from multipliers \cite{Mar-Alo,Olv}.
A review of the history of Noether's theorem and of the multiplier method for finding conservation laws can be found in \Ref{Kos-Sch}.
In recent years,
the multiplier method has been cast into the form of a generalization of Noether's theorem
which is applicable to PDE systems without a variational principle.
The generalization \cite{AncBlu97,AncBlu02a,AncBlu02b,Anc03}
is based on the structure of the determining system for multipliers,
which turns out to be an augmented, adjoint version of the determining equations for infinitesimal symmetries.
In particular,
multipliers can be viewed as an adjoint generalization of variational symmetries,
and most significantly, the determining system for multipliers
can be solved by the use of the same standard procedure that is used
for solving the determining equations for symmetries \cite{Olv,1stbook,2ndbook}.
Moreover, the physical conservation law determined by a multiplier
can be constructed directly from the multiplier and the given PDE system
by various integration methods \cite{Olv,AncBlu02a,AncBlu02b,Wol02b,Anc03,2ndbook}.
In this modern generalization,
the problem of finding all conservation laws for a given PDE system
thereby becomes a kind of adjoint of the problem of finding all infinitesimal symmetries.
As a consequence, for any PDE system,
there is no need to use special methods or anstazes
(\eg/, \cite{Mor,KarMah,Ibr07,Ibr11,PooHer})
for determining its conservation laws,
just as there is no necessity to use special methods or ansatzes for finding its symmetries.
The present paper is intended to review and extend these recent developments,
with an emphasis on applications to PDE systems arising in physical models.
The most natural mathematical framework
for understanding the methods and the results
is variational calculus in jet spaces \cite{Olv}.
This framework will be given a concrete formulation,
which is useful both for formulating general statements and for doing calculations
for specific PDE systems.
As a starting point, in \secref{examples},
a wide range of examples of local conservation laws and conserved integrals
are presented,
covering dynamical systems that model convection, diffusion, wave propagation, fluid flow, gas dynamics and plasma dynamics,
as well as non-dynamical (static equilibrium) systems.
In \secref{prelim},
for general PDE systems,
the standard formulations of local conservation laws, conserved integrals,
and symmetries,
as well as some other preliminaries, are stated.
Additionally,
local versus global aspects of conservation laws are discussed
and are related to the distinction between trivial and non-trivial conservation laws
and their physical meaning.
This discussion clears up some confusion in the existing literature.
In \secref{tools},
some basic modern tools from variational calculus are reviewed
using a concrete self-contained approach.
These tools are employed in \secref{characteristiceqns}
to derive the determining equations for multipliers and symmetries,
based on a characteristic form for conservation laws and symmetry generators.
An important technical step in this derivation is the introduction of
a coordinatization for the solution space of PDE systems in jet space,
which involves expressing a given PDE system in a solved form for a set of leading derivatives,
after the system is closed by appending all integrability conditions (if any).
This coordinatization is applicable to all PDE systems of physical interest,
including systems that possess differential identities.
It is used to show that the characteristic form for trivial conservation laws is given by
trivial multipliers which vanish on the solution space of a PDE system
when the system has no differential identities.
This directly leads to an explicit one-to-one correspondence
between non-trivial conservation laws and non-trivial multipliers,
taking into account the natural equivalence freedoms in conservation laws and multipliers.
An explicit generalization of this correspondence is established
in the case when a PDE system possesses a differential identity (or set of identities).
The generalization involves considering gauge multipliers \cite{AncPoh}
that arise from a conservation law connected with the differential identity.
These new results significantly broaden the explicit correspondence
between non-trivial conservation laws and non-trivial multipliers
previously obtained \cite{Mar-Alo,Olv,AncBlu02a,AncBlu02b}
only by requiring PDE systems to have a generalized Cauchy-Kovalevskaya form
(which restricts a system from possessing any differential identities).
Furthermore, as another result,
a large class of multipliers that captures physically important conserved integrals
such as mass, momentum, energy, angular momentum
is identified for general PDE systems
by examining the numerous examples of conservation laws presented earlier.
In \secref{Noether},
the variational calculus tools are used to state Noether's theorem in a modern form
for variational PDE systems,
along with the determining equations for variational symmetries.
The generalization of Noether's theorem in modern form to non-variational PDE systems
is explained in \secref{mainresults}.
First, the determining equations for multipliers are shown to be
an augmented, adjoint counterpart of the determining equations for symmetries.
More precisely, the multiplier determining system has a natural division into two subsystems \cite{AncBlu97,AncBlu02a,AncBlu02b}.
One subsystem is the adjoint of the symmetry determining system,
whose solutions can be viewed as adjoint-symmetries (also known as cosymmetries).
The remaining subsystem comprises equations that are necessary and sufficient for
an adjoint-symmetry to be a multiplier,
analogously to the conditions required for an infinitesimal symmetry to be a variational symmetry in the case of a variational PDE system.
Next, the role of a Lagrangian in constructing a conserved integral from a symmetry of a variational principle
is replaced for non-variational PDE systems by several different constructions:
an explicit integral formula, an explicit algebraic scaling formula,
and a system of determining equations,
all of which use only a multiplier and the PDE system itself.
The scaling formula is based on dimensional analysis
and generalizes a formula previously derived only for PDE systems that admit a scaling symmetry \cite{Anc03}.
These main results cover both the case of PDE systems without differential identities
and the case of PDE systems with differential identities.
It is emphasized that this general method for explicitly deriving the conservation laws of PDE systems
reproduces the content of Noether's theorem whenever a PDE system has a variational principle.
(For comparison,
an abstract, cohomological approach to determining conservation laws of PDE systems
can be found in \Ref{KraVin,Ver,Hen}.)
Some concluding remarks,
including discussion of the geometrical meaning of adjoint-symmetries and multipliers,
are provided in \secref{remarks}.
Several running examples will be used to illustrate
the main ideas and the main results in every section.
\section{Examples}
\label{examples}
The following seven examples illustrate some basic conserved densities and fluxes \eqref{conslaw-onedim}
arising in physical PDE systems in one spatial dimension.
{\bf\Ex{1}}: transport equation
\begin{gather}
u_t=(c(x,u)u)_x
\label{transport-eqn}
\end{gather}
\begin{gather}
T=u
\text{ is mass density},
\quad
X=-c(x,u)u
\text{ is mass flux \ie/, momentum}.
\end{gather}
{\bf\Ex{2}}: diffusion/heat conduction equation
\begin{gather}
u_t= (k(x,u)u_x)_x
\label{diffusionheat-eqn}
\end{gather}
\begin{gather}
T=u
\text{ is heat density (temperature)},
\quad
X=-k(x,u)u_x
\text{ is heat flux}.
\end{gather}
{\bf\Ex{3}}: telegraph equation
\begin{gather}
u_{tt}+a(t)u_t-(c(x)^2 u_x)_x = 0
\label{telegraph-eqn}
\end{gather}
\begin{gather}
\begin{aligned}
& T=\tfrac{1}{2}\exp(2{\textstyle\int} a(t)dt)(u_t{}^2+c(x)^2 u_x{}^2)
\text{ is energy density},\\
& X=-c(x)^2\exp(2{\textstyle\int} a(t)dt) u_xu_t
\text{ is energy flux} .
\end{aligned}
\end{gather}
{\bf\Ex{4}}: nonlinear dispersive wave equation
\begin{gather}
u_t+f(u)u_x+u_{xxx}=0,
\quad
f(u)\neq {\rm const.}
\label{dispersivewave-eqn}
\end{gather}
\begin{subequations}
\begin{align}
& T=u
\text{ is mass density},
\quad
X={\textstyle\int} f(u)du +u_{xx}
\text{ is mass flux \ie/, momentum};
\\
& T=u^2
\text{ is elastic energy density},
\quad
X=2{\textstyle\int} uf(u)du +2uu_{xx}-u_x{}^2
\text{ is elastic energy flux};
\\
&\begin{aligned}
T & ={\textstyle\int} g(u)du -\tfrac{1}{2}u_x{}^2
\text{ is gradient energy density},
\\
X & =\tfrac{1}{2}(g(u)+u_{xx})^2 +u_xu_t
\text{ is gradient energy flux},
\\&\quad
g(u)={\textstyle\int} f(u)du .
\end{aligned}
\end{align}
\end{subequations}
{\bf\Ex{5}}: compressible viscous fluid equations
\begin{gather}
\begin{aligned}
&\rho_t+( u\rho)_x=0 \\
&\rho(u_t+uu_x) = -p_x+\mu u_{xx}
\end{aligned}
\label{viscousfluid-eqn}
\end{gather}
\begin{subequations}
\begin{align}
& T=\rho
\text{ is mass density},
\quad
X=u\rho
\text{ is mass flux};
\\
& T=\rho u
\text{ is momentum density},
\quad
X=p-\mu u_x+Tu
\text{ is momentum flux};
\\
&\begin{aligned}
T &=\rho(tu-x)
\text{ is Galilean momentum density},
\\
X &=t(p -\mu u_x)+T u
\text{ is Galilean momentum flux}.
\end{aligned}
\end{align}
\end{subequations}
The next two examples are integrable PDE systems that possess
an infinite hierarchy of higher-order conservation laws.
{\bf\Ex{6}}: barotropic gas flow/compressible inviscid fluid equations
\begin{gather}
\begin{aligned}
& \rho_t+( u\rho)_x=0 \\
& u_t+uu_x = -p_x/\rho \\
& p=p(\rho) \text{ (barotropic equation of state) }\\
& e={\textstyle\int} p/\rho^2 d\rho \text{ (thermodynamic energy) }
\end{aligned}
\label{compressiblefluid-eqn}
\end{gather}
\begin{subequations}
\begin{align}
& T=\rho(\tfrac{1}{2}u^2+e)
\text{ is energy density},
\quad
X=(p+T)u
\text{ is energy flux};
\\
&\begin{aligned}
T & =\rho_x/(u_x^2-p'\rho_x^2/\rho^2)
\text{ is higher-derivative quantity},
\\
X & =\rho u_x/(u_x^2-p'\rho_x^2/\rho^2)
\text{ is higher-derivative flux}.
\end{aligned}
\label{higherTXcompressiblefluid}
\end{align}
\end{subequations}
{\bf\Ex{7}}: breaking wave (Camassa-Holm) equation
\begin{gather}
m_t + 2 u_xm + um_x =0,
\quad
m=u-u_{xx}
\label{breakingwave-eqn}
\end{gather}
\begin{subequations}
\begin{align}
& T = m
\text{ is momentum density},
\quad
X = \tfrac{1}{2}(u^2-u_x^2) +um-u_{tx}
\text{ is momentum flux};
\\
& T = \tfrac{1}{2}(u^2+u_x^2)
\text{ is energy density},
\quad
X = u(um-u_{tx})
\text{ is energy flux};
\\
&\begin{aligned}
T & = \tfrac{1}{2}u(u^2+u_x^2),
\text{ is energy-momentum density},
\\
X & = \tfrac{1}{2}(u_{tx}-u(m+\tfrac{1}{2}u)+\tfrac{1}{2}u_x^2)^2 -u_t(uu_x+\tfrac{1}{2}u_t)
\text{ is energy-momentum flux};
\end{aligned}
\\
& T = m^{1/2}
\text{ is Hamiltonian Casimir},
\quad
X = 2um^{1/2}
\text{ is Casimir flux};
\\
&\begin{aligned}
T & = m^{-5/2}m_x^2 + 4m^{-1/2}
\text{ is higher-derivative energy density},
\\
X & = -m^{-5/2}(2m_t+um_x)m_x -4m^{-3/2}u_xm_x -4m^{-1/2}u-8m^{1/2}
\text{ is higher-derivative flux}.
\end{aligned}
\label{higherTXbreakingwave}
\end{align}
\end{subequations}
The following three examples illustrate
some intrinsically multi-dimensional conservation laws \eqref{conslaw-multidim}
that arise in physical PDE systems in two or more dimensions.
{\bf\Ex{8}}: porous media equation
\begin{gather}
u_t= \nabla\cdot(k(u)\nabla u)
\label{porousmedia-eqn}
\end{gather}
\begin{gather}
\begin{aligned}
T &=\alpha(x)u
\text{ is a general mass-density moment},
\\
\vec X &={\textstyle\int} k(u)du \nabla\alpha(x) - \alpha(x)\nabla{\textstyle\int} k(u)du
\text{ is flux moment of mass-density},
\\&
\Delta\alpha =0
\quad\text{ (arbitrary solution of Laplace equation)}.
\end{aligned}
\end{gather}
{\bf\Ex{9}}: non-dispersive wave equation
\begin{gather}
u_{tt}-c^2\Delta u = f(u)
\label{nondispersivewave-eqn}
\end{gather}
\begin{subequations}
\begin{align}
&\begin{aligned}
T &=u_t (\mathbf{a}\cdot\vec x)\cdot\nabla u
\text{ is angular momentum density},
\\
\vec X &=(\tfrac{1}{2}c^2|\nabla u|^2 -\tfrac{1}{2}u_t^2-{\textstyle\int} f(u)du)\mathbf{a}\cdot\vec x -c^2((\mathbf{a}\cdot\vec x)\cdot\nabla u)\nabla u
\text{ is angular momentum flux},
\\&
\text{ (arbitrary constant antisymmetric tensor $\mathbf{a}$)}.
\end{aligned}
\\
&\begin{aligned}
T &=\vec b\cdot\vec x(\tfrac{1}{2} u_t^2 +\tfrac{1}{2}c^2|\nabla u|^2 -{\textstyle\int} f(u)du) +c^2tu_t \vec b\cdot\nabla u
\text{ is boost momentum density},
\\
\vec X &=c^2t(\tfrac{1}{2}c^2|\nabla u|^2 -\tfrac{1}{2} u_t^2-{\textstyle\int} f(u)du)\vec b
-c^2(\vec b\cdot\vec x u_t +c^2 t\vec b\cdot\nabla u)\nabla u
\text{ is boost momentum flux},
\\&
\text{ (arbitrary constant antisymmetric vector $\vec b$)}.
\end{aligned}
\end{align}
\end{subequations}
{\bf\Ex{10}}: inviscid (compressible/incompressible) fluid equation
\begin{gather}
\begin{aligned}
& \vec u_t + \vec u\cdot\nabla\vec u= -(1/\rho)\nabla p
\\
& e={\textstyle\int} p/\rho^2 d\rho \text{ (thermodynamic energy) }
\end{aligned}
\label{inviscidfluid-eqn}
\end{gather}
\begin{subequations}
\begin{align}
&\text{in three dimensions }
\begin{cases}
T= \vec u\cdot(\nabla\times\vec u)
\text{ is local helicity},
\\
\vec{\mathcal X}= \vec X - T\vec u = \tfrac{1}{2}(|\vec u|^2 +(p/\rho) +e)
\text{ is moving helicity flux};
\end{cases}
\\
&\text{in two dimensions }
\begin{cases}
T= \rho f(({\rm curl}\,\vec u)/\rho)
\text{ is local enstrophy},
\\
\vec{\mathcal X}= \vec X - T\vec u = 0
\text{ is moving enstrophy flux},
\\\quad
\text{(arbitrary function $f$)}.
\end{cases}
\end{align}
\end{subequations}
The last four examples illustrate spatial boundary conservation laws \eqref{conslaw-boundary}
and spatial flux conservation laws \eqref{div-conslaw}
for physical PDE systems in three dimensions.
{\bf\Ex{11}}: electric (displacement) field equation inside matter
\begin{gather}
\begin{aligned}
& \vec D_t = c\nabla\times\vec H \\
& \nabla\cdot\vec D = 4\pi \rho \\
& \vec J=0 \text{ (no currents) } \\
& \rho_t =0 \text{ (static charges) }
\end{aligned}
\end{gather}
\begin{gather}
\vec T= \vec D
\text{ is flux density of electric field lines.}
\end{gather}
{\bf\Ex{12}}: magnetohydrodynamics (infinite conductivity) equations
\begin{gather}
\begin{aligned}
& \vec u_t + \vec u\cdot\nabla\vec u= (1/\rho)(\vec J\times\vec B -\nabla p) \\
& \vec B_t = \nabla\times(\vec u\times \vec B) \\
& \nabla\times\vec B =4\pi\vec J \\
& \nabla\cdot\vec B = 0
\end{aligned}
\end{gather}
\begin{gather}
\vec T=\vec X= \vec B
\text{ is flux density of magnetic field lines.}
\end{gather}
{\bf\Ex{13}}: fluid incompressibility equation
\begin{gather}
\nabla\cdot\vec u=0 \\
\vec X= \vec u
\text{ is flux density of streamlines}.
\end{gather}
{\bf\Ex{14}}: charge source equation (in empty space)
\begin{gather}
\nabla\cdot\vec E=0 \\
\vec X= \vec E
\text{ is flux density of electric field lines}.
\end{gather}
\section{Conserved integrals, conservation laws, and symmetries}
\label{prelim}
Throughout, the following notation will be used.
Let $t$, $x=(x^1,\ldots,x^n)$ be independent variables, $n\geq1$,
and let $u=(u^1,\ldots,u^m)$ be dependent variables, $m\geq1$.
Partial derivatives of $u$ with respect to $t,x$ are denoted
$\partial u=(u_t,u_{x^1},\ldots,u_{x^n})$,
and $k$th-order partial derivatives are denoted $\partial^k u$, $k\geq 2$.
The coordinate space $J=(t,x,u,\partial u,\partial^2 u,\ldots)$ is called
the {\em jet space} associated with the variables $t,x,u$.
Partial derivatives with respect to these variables are given by
$\parderop{t}$,
$\parderop{x}=(\parderop{x^1},\ldots,\parderop{x^n})^{\rm t}$,
$\parderop{u}=(\parderop{u^1},\ldots,\parderop{u^m})^{\rm t}$,
with a superscript ``${\rm t}$'' denoting the transpose,
and similarly for partial derivatives with respect to the derivative variables in $J$.
Total derivatives with respect to $t,x$, acting by the chain rule,
are denoted $D=(D_t,D_{x^1},\ldots,D_{x^n})$.
In particular, $Du=\partial u$, $D\partial u = \partial^2 u$, and so on.
$D^k$ denotes all of the $k$th order total derivatives with respect to $t,x$.
Spatial divergences are denoted ${\rm Div\,}=D_x\cdot$,
with a dot denoting the vector dot product.
Consider an $N$th-order system of $M\geq 1$ PDEs
\begin{equation}\label{pde}
G=(G^1(t,x,u,\partial u,\ldots,\partial^N u),\ldots,G^M(t,x,u,\partial u,\ldots,\partial^N u))=0 .
\end{equation}
The space of all locally smooth solutions $u(t,x)$ of the system will be denoted ${\mathcal E}$.
This space has an embedding as a subspace in $J$,
since $u(t,x)\in{\mathcal E}$ determines $(t,x,u(t,x),\partial u(t,x),\partial^2 u(t,x),\ldots)\in J$.
(In the applied mathematics and physics literature,
${\mathcal E}$ is commonly identified with the set of equations $G=0$, $DG=0$, $D^2G=0$, $\ldots$ in $J$,
which assumes these equations are locally solvable \cite{Olv}.)
A {\em local conservation law} of a given PDE system \eqref{pde}
is a local continuity equation
\begin{equation}\label{conslaw}
(D_t T + D_x\cdot X)|_{\mathcal E} =0
\end{equation}
which holds on the whole solution space ${\mathcal E}$ of the system,
where $T(t,x,u,\partial u,\ldots,\partial^r u)$ is the {\em conserved density}
and $X=(X^1(t,x,u,\partial u,\ldots,\partial^r u),\ldots,X^n(t,x,u,\partial u,\ldots,\partial^r u))$
is the {\em spatial flux}.
The pair
\begin{equation}\label{current}
(T,X)=\Phi
\end{equation}
is called a {\em conserved current}.
Every conservation law \eqref{conslaw} can be integrated over
any given spatial domain $\Omega\subseteq\mathbb{R}^n$ to get
\begin{equation}\label{globalconslaw}
\frac d{dt} \int_{\Omega} T|_{\mathcal E} dV = -\oint_{\partial\Omega} X|_{\mathcal E}\cdot\nu dA
\end{equation}
by the divergence theorem,
where $\partial\Omega$ is the boundary of the domain
and $\nu$ denotes the outward pointing unit normal vector.
This shows that the rate of change of the quantity
\begin{equation}\label{C}
{\mathcal C}[u]= \int_{\Omega} T|_{\mathcal E} dV
\end{equation}
in the domain is balanced by the net flux escaping through the domain boundary.
The quantity \eqref{C} is called a {\em conserved integral},
and the relation \eqref{globalconslaw} is called a {\em global conservation law} or {\em global balance equation}.
Two conservation laws are {\em locally equivalent}
if they give the same global balance equation \eqref{globalconslaw} up to boundary terms.
This occurs iff their conserved densities
differ by a total spatial divergence $D_x\cdot\Theta$
on the solution space ${\mathcal E}$,
and correspondingly,
their fluxes differ by a total time derivative $-D_t\Theta$
modulo a divergence-free vector.
A conservation law is thereby called {\em locally trivial} if
\begin{equation}\label{trivconslaw}
T_{\rm triv}|_{\mathcal E} = D_x\cdot\Theta|_{\mathcal E},
\quad
X_{\rm triv}|_{\mathcal E} = - D_t\Theta|_{\mathcal E} +D_x\cdot\Gamma|_{\mathcal E}
\end{equation}
holds for some vector function $\Theta(t,x,u,\partial u,\ldots,\partial^{r-1} u)$
and some antisymmetric tensor function $\Gamma(t,x,u,\partial u,\ldots,\partial^{r-1} u)$.
The {\em differential order of a conservation law} is defined to be
the smallest differential order among all locally equivalent conserved currents.
(It is common in the mathematics literature to define a local conservation law itself
as the equivalence class of locally equivalent conserved currents.)
The global form of a locally trivial conservation law is given by
\begin{equation}\label{loctriv}
\frac d{dt} \oint_{\partial\Omega} \Theta|_{\mathcal E} \cdot\nu dA
= \oint_{\partial\Omega} D_t\Theta|_{\mathcal E}\cdot\nu dA
\end{equation}
since $\oint_{\partial\Omega} (D_x\cdot\Gamma)|_{\mathcal E}\cdot\nu dA=0$ by Stokes' theorem.
This integral equation \eqref{loctriv} is just an identity, with no physical content,
unless the spatial flux integral of $D_t\Theta|_{\mathcal E}$ vanishes.
From the divergence theorem,
this integral will vanish for all domains $\Omega$ iff
$D_x\cdot D_t\Theta|_{\mathcal E} = 0$ holds.
In such cases, the boundary integral
\begin{equation}
\int_{\Omega} T|_{\mathcal E} dV
= \oint_{\partial\Omega} \Theta|_{\mathcal E} \cdot\nu dA
\end{equation}
will be a constant of motion for solutions of the given PDE system.
This type of boundary conservation law arises for PDE systems typically
when the PDEs in the system are related by obeying a differential identity,
as will be discussed further in \secref{characteristiceqns}.
In all cases when both
$D_x\cdot\Theta|_{\mathcal E}$ and $D_x\cdot D_t\Theta|_{\mathcal E}$ do not vanish identically,
a locally trivial conservation law has no physical content.
For a given PDE system \eqref{pde},
the set of all non-trivial conservation laws (up to local equivalence)
forms a vector space on which the symmetries of the system have a natural action
\cite{Olv,2ndbook}.
An {\em infinitesimal symmetry} \cite{Olv,1stbook,2ndbook}
of a given PDE system \eqref{pde} is a generator
\begin{equation}\label{generator}
{\bf X}=\tau\parderop{t} +\xi\parderop{x} +\eta\parderop{u}
\end{equation}
whose prolongation leaves invariant the PDE system,
\begin{equation}\label{Xsymmcond}
{\rm pr}{\bf X}(G)|_{\mathcal E} =0
\end{equation}
which holds on the whole solution space ${\mathcal E}$ of the system.
Here $\tau(t,x,u,\partial u,\ldots,\partial^r u)$,
$\xi=(\xi^1(t,x,u,\partial u,\ldots,\partial^r u),\ldots,\xi^n(t,x,u,\partial u,\ldots,\partial^r u))$,
and $\eta=(\eta^1(t,x,u,\partial u,\ldots,\partial^r u),\ldots,$
$\eta^m(t,x,u,\partial u,\ldots,\partial^r u))$
are called the {\em characteristic functions} in the symmetry generator.
When acting on the solution space ${\mathcal E}$,
an infinitesimal symmetry generator can be formally exponentiated to produce
a one-parameter group of transformations $\exp(\epsilon{\rm pr}{\bf X})$,
with parameter $\epsilon$,
where the infinitesimal transformation is given by
\begin{equation}
\begin{aligned}
u(t,x) \rightarrow
u(t,x) +& \epsilon\big( \eta(t,x,u(t,x),\partial u(t,x),\ldots,\partial^r u(t,x))
\\&\qquad
-u_t(t,x)\tau(t,x,u(t,x),\partial u(t,x),\ldots,\partial^r u(t,x))
\\&\qquad
-u_x(t,x)\cdot\xi(t,x,u(t,x),\partial u(t,x),\ldots,\partial^r u(t,x)) \big) +O\big(\epsilon^2\big)
\end{aligned}
\end{equation}
for all solutions $u(t,x)$ of the PDE system.
Two infinitesimal symmetries are equivalent if they have the same action
on the solution space ${\mathcal E}$ of a given PDE system.
An infinitesimal symmetry is thereby called {\em trivial}
if it leaves all solutions $u(t,x)$ unchanged.
This occurs iff its characteristic functions satisfy the relation $\hat\eta|_{\mathcal E}=0$,
where
\begin{equation}
\hat\eta = \eta - u_t\tau -u_x\cdot\xi.
\end{equation}
The corresponding generator \eqref{generator} of a trivial symmetry
on the solution space ${\mathcal E}$
is thus given by
\begin{equation}\label{trivsymm}
{\bf X}_{\rm triv} = \tau\parderop{t} + \xi\cdot\parderop{x} +(u_t\tau +u_x\cdot\xi)\parderop{u}
\end{equation}
which has the prolongation ${\rm pr}{\bf X}_{\rm triv}=\tau D_t + \xi\cdot D_x$.
Conversely, any generator of this form \eqref{trivsymm} represents a trivial symmetry.
The {\em differential order of an infinitesimal symmetry} is defined to be
the smallest differential order among all equivalent generators.
In jet space $J$,
a group of transformations $\exp(\epsilon{\rm pr}{\bf X})$
in general will not act in a closed form
on $t,x,u$, and derivatives $\partial^k u$ up to a finite order,
except \cite{Olv,2ndbook} for point transformations acting on $(t,x,u)$,
and contact transformations acting on $(t,x,u,u_t,u_x)$.
Moreover, a contact transformation is a prolonged point transformation
when the number of dependent variables is $m=1$ \cite{Olv,2ndbook}.
A {\em point symmetry} is defined as a symmetry transformation group on $(t,x,u)$,
whose generator is given by characteristic functions of the form
\begin{equation}\label{Xpointsymm}
{\bf X}=\tau(t,x,u)\parderop{t} +\xi(t,x,u)\parderop{x} +\eta(t,x,u)\parderop{u}
\end{equation}
corresponding to the infinitesimal point transformation
\begin{equation}\label{pointgroup}
\begin{aligned}
& t\rightarrow t+\epsilon \tau(t,x,u) + O(\epsilon^2),
\\
& x\rightarrow x+\epsilon \xi(t,x,u) + O(\epsilon^2),
\\
& u\rightarrow u+\epsilon \eta(t,x,u) + O(\epsilon^2) .
\end{aligned}
\end{equation}
Likewise,
a {\em contact symmetry} is defined as a symmetry transformation group on $(t,x,u,u_t,u_x)$
whose generator corresponds to an infinitesimal transformation that preserves
the contact relations $u_t =\partial_t u$, $u_x =\partial_x u$.
The set of all admitted point symmetries and contact symmetries
for a given PDE system comprises its group of {\em Lie symmetries}.
Common examples of point symmetries
admitted by PDE systems arising in physical applications
are time translations, space translations, and scalings.
Higher-order symmetries are typically admitted only by integrable PDE systems.
However,
it is worth emphasizing that any admitted symmetry can be used
to obtain a mapping of a given solution $u=f(t,x)$ of a PDE system
into a one-parameter family of solutions
$u =\tilde f(t,x,\epsilon)
= \big(\exp(\epsilon{\rm pr}\hat{\bf X})u\big)|_{u=f(t,x)}
= \big(u + \epsilon \hat\eta + \tfrac{1}{2}\epsilon^2 {\rm pr}\hat{\bf X} \hat\eta+ \cdots\big)|_{u=f(t,x)}$
where $\hat{\bf X} = {\bf X}-{\bf X}_{\rm triv} = \hat\eta\partial/\partial u$;
and also to find symmetry-invariant solutions $u=f(t,x)$ of a PDE system
by considering the invariance condition
$(\hat{\bf X} u)|_{u=f(t,x)} = \hat\eta|_{u=f(t,x)} =0$.
Thus, for these two main purposes, symmetries of any differential order
are equally useful.
Similar remarks can be made for conservation laws.
In physical applications,
the most common examples of conserved densities admitted by PDE systems
are mass, momentum, and energy.
These densities are always of low differential order,
whereas higher-order densities are typically admitted only by integrable PDE systems.
Nevertheless, for the many purposes outlined in \secref{intro},
any admitted conservation law of a given PDE system can be useful.
\subsection{Regular PDE systems and computation of symmetry generators, conserved densities and fluxes}
To determine if a current \eqref{current} is conserved for a given PDE system,
and if a generator \eqref{generator} is an infinitesimal symmetry of a given PDE system,
it is necessary to coordinatize the solution space ${\mathcal E}$ of the system in jet space $J$.
This can be accomplished in a general way by the following steps.
First, for any PDE system \eqref{pde},
introduce an index notation for the components of $x$ and $u$:
$x^i$, $i=1,\ldots,n$; and $u^\alpha$, $\alpha=1,\ldots,m$.
Next, suppose each PDE $G^a=0$, $a=1,\ldots,M$, in the given system
can be expressed in a solved form
\begin{equation}\label{pde-solvedform}
G^a = \partial_{(\ell_a)} u^{\alpha_a} -g^a
\end{equation}
for some derivative of a single dependent variable $u^{\alpha_a}$,
after a point transformation (change of variables) if necessary,
such that all other terms in the system contain
neither this derivative nor its differential consequences,
namely
\begin{equation}\label{pde-leadingder}
\begin{gathered}
\partial_{(\ell_a)}u^{\alpha_a} \neq \partial^k \partial_{(\ell_b)}u^{\alpha_b},
\quad
a,b=1,\ldots,M,
\quad
k\geq 1 ,
\\
\parder{g^a}{(\partial^k\partial_{(\ell_b)} u^{\alpha_b})} =0,
\quad
a,b=1,\ldots,M,
\quad
k\geq 0 .
\end{gathered}
\end{equation}
Such derivatives $\{\partial_{(\ell_a)}u^{\alpha_a}\}_{a=1,\ldots, M}$
are called a set of {\em leading derivatives} for the PDE system.
Last, suppose the given PDE system is closed in the sense that
it has no integrability conditions and all of its differential consequences
produce PDEs that have a solved form in terms of differential consequences of the leading derivatives.
Note that if a PDE system is not closed then it can always be enlarged to get a closed system
by appending any integrability conditions and differential consequences that involve the introduction of more leading derivatives.
Then, coordinates for the solution space ${\mathcal E}$ of the closed PDE system in $J$
are provided by
the independent variables $t,x^i$,
the dependent variables $u^\alpha$,
and all of the non-leading derivatives of $u^\alpha$.
A closed PDE system \eqref{pde} admitting such a solved form \eqref{pde-solvedform}--\eqref{pde-leadingder}
will be called {\em regular}.
A more restrictive class of PDE systems is given by Cauchy-Kovalevskaya systems
and their generalizations.
Recall, a PDE system \eqref{pde} is of Cauchy-Kovalevskaya form \cite{Eva,Olv}
if the leading derivatives in the solved form of the system
consist of pure derivatives of $u$ with respect to a single independent variable,
namely $\partial_{(\ell_a)}u^{\alpha_a}=\partial_z^{k_a}u^{\alpha_a}$,
$a=1,\ldots,M$, $z\in \{t,x^i\}$,
and if their differential order $k_a$ is equal to the differential order $N$ of the system,
namely $k_a=N$, $a=1,\ldots,M$.
Cauchy-Kovalevskaya systems,
and their generalizations \cite{Mar-Alo} in which $k_a$ differs from $N$,
have the feature that they do not possess any differential identities
and that none of their differential consequences possess differential identities.
Such PDE systems are usually called {\em normal}.
Note that, in contrast to normal systems,
the leading derivatives in a regular PDE system can be, for instance,
a mixed derivative of all the dependent variables $u^\alpha$
or a different derivative of each of the dependent variables $u^\alpha$.
{\bf Running \Ex{(1)}}:
Generalized Korteweg-de Vries (gKdV) equation
\begin{equation}\label{gkdv-eqn}
u_t+u^pu_x+u_{xxx}=0,
\quad
p>0 .
\end{equation}
This is a regular PDE since it has the leading derivative
$u_t= -u^pu_x-u_{xxx}$.
It also has a third-order leading derivative
$u_{xxx}=-u_t-u^pu_x$.
Both of these solved forms are of generalized Cauchy-Kovalevskaya type.
{\bf Running \Ex{(2)}}:
Breaking wave equation \cite{DegHonHol}
\begin{equation}\label{b-fam-eqn}
m_t + bu_xm + um_x =0,
\quad
m=u-u_{xx},
\quad
b\neq -1 .
\end{equation}
This is a regular PDE system since it has the leading derivatives
$m_t= -bu_xm -um_x$, $u_{xx}=u-m$.
Equivalently, if $m$ is eliminated through the second PDE,
this yields a scalar equation $u_t -u_{txx} + (b+1)uu_x = bu_xu_{xx} + uu_{xxx}$
which is a regular PDE with respect to the leading derivative
\begin{equation}\label{b-fam-u-eqn}
u_{txx} =u_t + (b+1)uu_x -bu_xu_{xx} - uu_{xxx} .
\end{equation}
Neither of these solved forms are of generalized Cauchy-Kovalevskaya type.
However, the alternative solved forms
$u_{xxx}=u_x + u^{-1}(bu_x(u-u_{xx})- u_t+u_{txx})$
and $m_x= -u^{-1}(m_t +bu_x m)$, $u_{xx}=u-m$
are of generalized Cauchy-Kovalevskaya type.
{\bf Running \Ex{(3)}}:
Euler equations for constant density, inviscid fluids in two dimensions
\begin{equation}\label{fluid-eqn}
\begin{gathered}
\nabla\cdot\vec u =0,
\quad
\rho ={\rm const.} ,
\\
\vec u_t +\vec u\cdot\nabla\vec u = -(1/\rho)\nabla p ,
\\
\Delta p = -\rho (\nabla\vec u)\cdot(\nabla\vec u)^{\rm t} .
\end{gathered}
\end{equation}
In this system, the independent variables are $t$ and $(x,y)$,
and the dependent variables consist of $p$ and $\vec u=(u^1,u^2)$,
in Cartesian components.
Leading derivatives are given by writing the PDEs in the solved form
\begin{equation*}\label{fluid-eqn-solvedform}
\begin{aligned}
u^1_x & =-u^2_y ,
\\
u^1_t &= -(u^1u^1_x + u^2u^1_y +(1/\rho)p_x)\big|{}_{u^1_x=-u^2_y}
= -(u^2u^1_y -u^1u^2_y + +(1/\rho)p_x) ,
\\
u^2_t &= -(u^1u^2_x + u^2u^2_y +(1/\rho)p_y) ,
\\
p_{xx} &= -p_{yy} -\rho((u^1_x)^2 +(u^2_y)^2 +2u^1_yu^2_x)\big|{}_{u^1_x=-u^2_y}
= -p_{yy} -2\rho((u^2_y)^2 +u^1_yu^2_x) .
\end{aligned}
\end{equation*}
Thus, this system is a regular PDE system,
but it does not have a generalized Cauchy-Kovalevskaya form.
A related feature is that the PDEs in the system obey a differential identity
\begin{equation}\label{fluid-eqn-diffid}
{\rm Div\,}( \vec u_t +\vec u\cdot\nabla\vec u +(1/\rho)\nabla p )
- (D_t +\vec u\cdot\nabla)(\nabla\cdot\vec u)
=(1/\rho)\Delta p +(\nabla\vec u)\cdot(\nabla\vec u)^{\rm t} .
\end{equation}
Note that the pressure equation is often not explicitly considered in writing down the Euler equations.
However, without including the pressure equation, the system would not be closed,
since the differential identity \eqref{fluid-eqn-diffid} shows that the pressure equation
arises as an integrability condition of the other equations.
Correspondingly, the pressure equation does not have a solved form in terms of
differential consequences of the set of derivatives $\{u^1_x,u^1_t,u^2_t\}$.
{\bf Running \Ex{(4)}}:
Magnetohydrodynamics equations for a compressible, infinite conductivity plasma
in three dimensions
\begin{equation}\label{mhd-eqn}
\begin{gathered}
p=P(\rho),
\quad
\nabla\times\vec B =4\pi\vec J,
\quad
\nabla\cdot\vec B = 0 ,
\\
\rho_t +\nabla\cdot(\rho\vec u) =0 ,
\\
\vec u_t + \vec u\cdot\nabla\vec u= (1/\rho)(\vec J\times\vec B -\nabla p) ,
\\
\vec B_t = \nabla\times(\vec u\times \vec B) .
\end{gathered}
\end{equation}
The independent variables in this system are $t$ and $(x,y,z)$,
and the dependent variables consist of $\rho$,
$\vec u=(u^1,u^2,u^3)$ and $\vec B=(B^1,B^2,B^3)$,
in Cartesian components.
This is a regular PDE system,
where a set of leading derivatives is given by writing the PDEs in the solved form
\begin{equation*}\label{mhd-eqn-solvedform}
\begin{aligned}
B^1_x &=-B^2_y-B^3_z ,
\\
\rho_t &= \rho(u^1_x+u^2_y+u^3_z) +\rho_x u^1+\rho_y u^2+\rho_z u^3 ,
\\
u^1_t &= -(u^1u^1_x + u^2u^1_y + u^3u^1_z+(1/\rho)(P'(\rho)\rho_x +B^2J^3-B^3J^2)) ,
\\
u^2_t &= -(u^1u^2_x + u^2u^2_y + u^3u^2_z+(1/\rho)(P'(\rho)\rho_y +B^3J^1-B^1J^3)) ,
\\
u^3_t &= -(u^1u^3_x + u^2u^3_y + u^3u^3_z+(1/\rho)(P'(\rho)\rho_z +B^1J^2-B^2J^1)) ,
\\
B^1_t &= u^1B^2_y -u^2B^1_y +u^1B^3_z-u^3B^1_z +u^1_yB^2 -u^2_yB^1 +u^1_zB^3-u^3_zB^1 ,
\\
B^2_t & = (u^2B^3_z -u^3B^2_z +u^2B^1_x-u^1B^2_x +u^2_zB^3 -u^3_zxB^2+u^2_xB^1-u^1_xB^2)\big|{}_{B^1_x=-B^2_y-B^3_z}
\\
& = -u^2B^2_y-u^3B^2_z -u^1B^2_x +u^2_zB^3 -u^3_zB^2+u^2_xB^1-u^1_xB^2 ,
\\
B^3_t & = (u^3B^1_x -u^1B^3_x +u^3B^2_y-u^2B^3_y +u^3_xB^1 -u^1_xB^3 +u^3_yB^2-u^2_yB^3)\big|{}_{B^1_x=-B^2_y-B^3_z}
\\
& = -u^3B^3_z -u^1B^3_x -u^2B^3_y +u^3_xB^1 -u^1_xB^3 +u^3_yB^2-u^2_yB^3 ,
\end{aligned}
\end{equation*}
with
\begin{equation*}
4\pi J^1= B^3_y -B^2_z,
\quad
4\pi J^2= B^1_z-B^3_x,
\quad
4\pi J^3= B^2_x-B^1_y .
\end{equation*}
These PDEs lack a generalized Cauchy-Kovalevskaya form,
which is related to the feature that they obey a differential identity
\begin{equation}\label{mhd-eqn-diffid}
{\rm Div\,}( \vec B_t -\nabla\times(\vec u\times \vec B) )
=D_t(\nabla\cdot\vec B) .
\end{equation}
As seen from the examples here and in \secref{examples},
all typical PDE systems arising in physical applications belong to the class of regular systems.
For any given regular PDE system,
the standard approach \cite{Ovs,AndIbr,Ibr,CRC}
to look for symmetries consists of
solving the invariance condition ${\rm pr}{\bf X}(G)|_{\mathcal E} =0$ to find
the characteristic functions $\eta$, $\tau$, $\xi$ in the generator ${\bf X}$.
The computations in this approach are reasonable for finding point symmetries,
but become much more complicated for finding contact symmetries and higher-order symmetries.
{\bf Running \Ex{(1)}}
Consider the gKdV equation \eqref{gkdv-eqn}.
Since this is a scalar PDE,
its Lie symmetries are generated by point transformations and contact transformations,
with the general infinitesimal form
\begin{equation*}
{\bf X}=\tau(t,x,u,u_t,u_x)\parderop{t} +\xi(t,x,u,u_t,u_x)\parderop{x} +\eta(t,x,u,u_t,u_x)\parderop{u} .
\end{equation*}
Substitution of this generator into the determining condition
${\rm pr}{\bf X}(u_t+u^pu_x+u_{xxx})|_{\mathcal E} =0$
requires prolonging ${\bf X}$ to first-order with respect to $t$
and third-order with respect to $x$:
\begin{equation*}
{\rm pr}{\bf X}= {\bf X} + \eta^{(t)}\parderop{u_t} + \eta^{(x)}\parderop{u_x} + D_x^2\eta^{(xx)}\parderop{u_{xx}} + D_x^3\eta^{(xxx)}\parderop{u_{xxx}}
\end{equation*}
where
\begin{equation*}
\begin{aligned}
\eta^{(t)} &= D_t\eta -u_tD_t\tau -u_xD_t\xi ,
\\
\eta^{(x)} &= D_x\eta -u_tD_x\tau -u_xD_x\xi ,
\\
\eta^{(xx)} &= D_x\eta^{(x)} -u_{tx}D_x\tau -u_{xx}D_x\xi ,
\\
\eta^{(xxx)} &= D_x\eta^{(xx)} -u_{txx}D_x\tau -u_{xxx}D_x\xi .
\end{aligned}
\end{equation*}
This yields
\begin{equation*}
\begin{aligned}
& \big(
u_x\eta +D_t\eta -u_tD_t\tau -u_xD_t\xi
+uD_x\eta -(uu_t+3u_{txx})D_x\tau -(uu_x+3u_{xxx})D_x\xi
\\&\qquad
-3u_{tx}D_x^2\tau -3u_{xx}D_x^2\xi+ D_x^3\eta -u_tD_x^3\tau -u_xD_x^3\xi
\big)|_{\mathcal E} =0 .
\end{aligned}
\end{equation*}
There are two steps in solving this determining condition.
First, since the condition is formulated on the gKdV solution space ${\mathcal E}$,
a leading derivative of $u$ (and all of its differential consequences)
needs to be eliminated.
The most convenient choice is $u_{xxx}=-u_t-u^pu_x$
rather than $u_t=-u^pu_x-u_{xxx}$,
since $\tau$, $\xi$, $\eta$ depend on $u_t$.
Next, after the total derivatives of $\tau$, $\xi$, $\eta$ are expanded out,
the resulting equation needs to be split with respect to the jet variables
$u_{tt},u_{tx},u_{xx},u_{txx},u_{xxx},u_{txxx},u_{xxxx}$
which do not appear in $\tau$, $\xi$, $\eta$.
Finally, the split equations need to be simplified,
as some are differential consequences of others.
After these lengthy computations and simplifications,
a linear system of 6 determining equations is obtained for $\tau$, $\xi$, $\eta$:
\begin{gather*}
2\tau_{u_t} +u_t\tau_{u_t u_t} +u_x\xi_{u_t u_t} -\eta_{u_t u_t} =0 ,
\\
2\xi_{u_x} +u_t\tau_{u_x u_x} +u_x\xi_{u_x u_x} -\eta_{u_x u_x} =0 ,
\\
\tau_{u_x} +\xi_{u_t} +u_t\tau_{u_t u_x} +u_x\xi_{u_t u_x} -\eta_{u_t u_x} =0 ,
\\
\begin{aligned}
& 3pu_t\eta +2u(u_t\tau_t+u_x\xi_t-\eta_t)
+3p\big( u_t^2(u_t\tau_{u_t} +u_x\tau_{u_x})
\\&\qquad
+ u_tu_x(u_t\xi_{u_t} +u_x\xi_{u_x}) - u_t(u_t\eta_{u_t} +u_x\eta_{u_x}) \big)
=0 ,
\end{aligned}
\\
\begin{aligned}
& pu_x\eta +2u(u_t\tau_x+u_x\xi_x-\eta_x)
+p\big( u_tu_x(u_t\tau_{u_t} +u_x\tau_{u_x})
\\&\qquad
+ u_x^2(u_t\xi_{u_t} +u_x\xi_{u_x}) - u_x(u_t\eta_{u_t} +u_x\eta_{u_x}) \big)
=0 ,
\end{aligned}
\\
\begin{aligned}
& \eta +u(u_t\tau_u+u_x\xi_u-\eta_u)
+u_t(u_t\tau_{u_t} +u_x\tau_{u_x})
\\&\qquad
+ u_x(u_t\xi_{u_t} +u_x\xi_{u_x}) - (u_t\eta_{u_t} +u_x\eta_{u_x})
=0 .
\end{aligned}
\end{gather*}
This system can be solved, with $p$ treated as an unknown, to get
\begin{equation*}
\begin{gathered}
\tau = \tilde\tau(t,x,u,u_t,u_x),
\quad
\xi = \tilde\xi(t,x,u,u_t,u_x),
\\
\eta = u_t(\tilde\tau -c_1 -3c_3)+u_x(\tilde\xi -c_2 -c_3 x-c_4 t) -\tfrac{2}{p}c_3 u +c_4,
\quad
c_4 = 0 \text{ if $p\neq1$,}
\end{gathered}
\end{equation*}
which is a linear combination of
a time translation ($c_1$), a space translation ($c_2$),
a scaling ($c_3$), and a Galilean boost ($c_4$),
plus a trivial symmetry involving two arbitrary functions $\tilde\tau(t,x,u,u_t,u_x)$, $\tilde\xi(t,x,u,u_t,u_x)$.
Clearly, for finding higher-order symmetries,
or for dealing with PDE systems that have a high differential order
or that involve more spatial dimensions,
the previous standard approach becomes increasingly complicated,
as the general solution of the symmetry determining condition
will always contain a trivial symmetry
involving arbitrary differential functions.
In particular,
the resulting linear system of determining equations for finding $\tau$, $\xi$, $\eta$
becomes less over-determined and hence more computationally difficult to solve
when going to higher orders.
The situation for finding conservation laws is quite similar.
For any given regular PDE system,
it is possible to look for conservation laws by solving
the local continuity equation $(D_t T + D_x\cdot X)|_{\mathcal E} =0$
to find $T$ and $X$.
This approach is workable when the conserved densities $T$ and fluxes $X$
being sought have a low differential order
and when the number of spatial dimensions is low.
{\bf Running \Ex{(1)}}
Consider again the gKdV equation \eqref{gkdv-eqn}.
This is a time evolution PDE of third order in spatial derivatives,
while the conserved currents in lowest order form
for mass, energy, and $L^2$ norm
are of first order in derivatives for the densities
and of second order in derivatives for the fluxes.
Substitution of functions
\begin{equation*}
T(t,x,u,u_t,u_x),
\quad
X(t,x,u,u_t,u_x,u_{tt},u_{tx},u_{xx})
\end{equation*}
into the determining condition $(D_t T + D_x X)|_{\mathcal E} =0$ yields
\begin{equation*}
\big(
T_t + u_tT_u + u_{tx}(T_{u_x}+X_{u_t}) + u_{tt}T_{u_t}
+ X_x + u_xX_u +u_{xx}X_{u_x} +u_{txx}X_{u_{tx}} +u_{ttx}X_{u_{tt}} +u_{xxx}X_{u_{xx}}
\big)|_{\mathcal E} =0 .
\end{equation*}
The steps in solving this determining condition are similar to
those used in solving the symmetry determining equation.
First, a leading derivative of $u$ (and all of its differential consequences)
needs to be eliminated.
The most convenient choice is $u_{xxx}=-u_t-u^pu_x$
rather than $u_t=-u^pu_x-u_{xxx}$,
since $T$ and $X$ depend on $u_t$.
Next, the resulting equation needs to be split with respect to the jet variables
$u_{ttx},u_{txx}$, which do not appear in $T$, $X$.
This splitting immediately leads to a further splitting with respect to $u_{tx},u_{tt}$,
giving a linear system of 5 PDEs for $T$, $X$:
\begin{gather*}
T_{u_t} =0,
\quad
X_{u_{tt}} =0,
\quad
X_{u_{tx}} =0,
\quad
T_{u_x}+X_{u_t} =0,
\\
T_t + u_tT_u + X_x + u_xX_u +u_{xx}X_{u_x} -(u_t+u^pu_x)X_{u_{xx}} =0 .
\end{gather*}
This system can be solved, treating $p$ as an unknown, to obtain
\begin{equation*}
\begin{aligned}
T = &
c_1u^2 + c_2u + c_3(\tfrac{1}{(p+1)(p+2)}u^{p+2}-\tfrac{1}{2}u_x{}^2)
+c_4(xu-\tfrac{1}{2}tu^2)
\\&\qquad
+c_5(t(\tfrac{1}{2}u^2-3u_x{}^2)-xu^2)
+D_x\Theta(t,x,u) ,
\\
X= &
c_1(\tfrac{2}{p+2}u^{p+2}+2uu_{xx}-u_x{}^2) +c_2(\tfrac{1}{p+1}u^{p+1}+u_{xx})
\\&\qquad
+c_3(\tfrac{1}{2}(\tfrac{1}{p+1}u^{p+1}+u_{xx})^2 +u_xu_t)
+c_4(x(\tfrac{1}{2}u^2+u_{xx})-t(\tfrac{1}{3}u^3 +uu_{xx}-\tfrac{1}{2}u_x{}^2)-u_x)
\\&\qquad
+c_5(t(3(\tfrac{1}{3}u^3+u_{xx})^2+6u_tu_x)+x(u_x{}^2-2uu_{xx}-\tfrac{1}{2}u^3)+2uu_x)
-D_t\Theta(t,x,u) ,
\\
& c_4 = 0 \text{ if $p\neq1$},
\quad
c_5 = 0 \text{ if $p\neq2$}
\end{aligned}
\end{equation*}
which yields a linear combination of the densities and the fluxes
representing conserved currents for the $L^2$-norm ($c_1$),
mass ($c_2$), energy ($c_3$), Galilean momentum ($c_4$), and Galilean energy ($c_5$),
plus a term involving an arbitrary function $\Theta(t,x,u)$
which represents a locally trivial conserved current.
However, when going to higher orders or to higher spatial dimensions,
it becomes increasingly more difficult to solve the local continuity equation
$(D_t T + D_x\cdot X)|_{\mathcal E} =0$,
as the general solution will contain
a trivial density term $D_x\cdot\Theta$ in $T$
and a trivial flux term $-D_t\Theta +D_x\cdot\Gamma$ in $X$
involving a differential vector function $\Theta$
and a differential antisymmetric tensor function $\Gamma$, which are arbitrary.
In particular,
the resulting linear system of determining equations for finding $T$ and $X$
will be less over-determined and hence more computationally difficult to solve,
compared to the low order case or the one dimensional case.
These difficulties motivate introducing a characteristic form
(or canonical representation) for conserved currents
so that all locally equivalent conserved currents have the same characteristic form,
and likewise for symmetry generators.
To derive this formulation,
some tools from variational calculus will be needed.
\section{Tools in variational calculus}
\label{tools}
For working with symmetries and conservation laws of PDE systems,
the natural setting in which to apply variational calculus is
the space of {\em differential functions}
defined by locally smooth functions of finitely many variables
in jet space $J=(t,x,u,\partial u,\partial^2 u,\ldots)$.
As examples,
in the nonlinear dispersive wave equation \Ex{4},
if the constitutive nonlinearity function $f(u)$ is smooth,
then the conserved density and flux for mass and energy
are smooth functions of $u,u_x,u_{xx}$ in $J$,
but if $f(u)$ blows up when $u=0$ then these functions are singular
at points in $J$ such that $u=0$;
in the barotropic gas flow \Ex{6},
the higher-derivative density and flux are singular functions of $\rho,u,\rho_x,u_x$
at points in $J$ where $u_x^2=p(\rho)'/\rho$,
but at all other points these functions are smooth.
The basic tools that will be needed from variational calculus are
the Fr\'echet derivative and adjoint derivative,
the Euler operator,
a homotopy integral,
a total null-divergence identity,
and a scaling identity.
Throughout,
$f(t,x,u,\partial u,\ldots,\partial^k u)$ denotes a differential function of order $k\geq 0$,
and $v=(v^1(t,x,u,\partial u,\partial^2 u,\ldots),\ldots,v^m(t,x,u,\partial u,\partial^2 u,\ldots))$,
$w(t,x,u,\partial u,\partial^2 u,\ldots)$
denote differential functions of arbitrary finite order.
The {\em Fr\'echet derivative} of a differential function is
the linearization of the function as defined by
\begin{equation}\label{frechet}
\begin{aligned}
\delta_v f &
= \parder{}{\epsilon}f(t,x,u+\epsilon v,\partial (u+\epsilon v),\ldots,\partial^k(u+\epsilon v))\big|_{\epsilon =0}
\\&
= v\parder{f}{u} + Dv\cdot\parder{f}{(\partial u)} +\cdots +D^kv\cdot\parder{f}{(\partial^k u)}
\end{aligned}
\end{equation}
which can be viewed as a local directional derivative in jet space,
corresponding to the action of a generator $\hat{\bf X}=v\partial_u$ in characteristic form,
$\hat{\bf X}(f) = \delta_v f$.
It is useful also to view the Fr\'echet derivative
as a linear differential operator acting on $v$.
Then the relation
\begin{equation}\label{adjoint}
w\delta_v f - v\delta^*_w f = D\cdot\Psi(v,w;f)
\end{equation}
as obtained using integration by parts
defines the {\em Fr\'echet adjoint derivative}
\begin{equation}\label{adjfrechet}
\delta^*_w f =
w\parder{f}{u} -D\cdot\Big(w\parder{f}{(\partial u)}\Big) +\cdots
+(-D)^k\cdot\Big(w\parder{f}{(\partial^k u)}\Big)
\end{equation}
which is a linear differential operator acting on $w$.
The associated current $\Psi(v,w;f)=(\Psi^t,\Psi^x)$ is given by
\begin{equation}\label{frechetcurrent}
\begin{aligned}
\Psi(v,w;f) &
= vw \parder{f}{(\partial u)}
+ (Dv)\cdot\Big(w\parder{f}{(\partial^2 u)}\Big) -vD\cdot\Big(w\parder{f}{(\partial^2 u)}\Big) +\cdots
\\&\qquad
+\sum_{l=1}^{k} (D^{k-l}v)\cdot\Big((-D)^{l-1}\cdot\Big(w\parder{f}{(\partial^k u)}\Big)\Big) .
\end{aligned}
\end{equation}
An alternative notation for the Fr\'echet derivative and its adjoint is
$\delta_v f =f'(v)$ and $\delta^*_w f = f'{}^*(w)$,
or sometimes $\delta_v f=D_v f$ and $\delta^*_w f=D_w^* f$.
The Fr\'echet derivative of a differential function $f$
can be inverted to recover $f$
by using a line integral along any curve $C$ in $J$,
where the endpoints $\partial C$ are given by
a general point $(t,x,u,\partial u,\ldots,\partial^k u) \in J$
and any chosen point $(t,x,u_0,\partial u_0,\ldots,\partial^k u_0) \in J$
at which $f$ is non-singular.
This yields
\begin{equation}\label{line-integral}
f\big|_{\partial C} =
\int_C \parder{f}{u^{\rm t}}du^{\rm t} + \parder{f}{(\partial u^{\rm t})}\cdot d\partial u^{\rm t} +\cdots
+\parder{f}{(\partial^k u^{\rm t})}\cdot d\partial^k u^{\rm t} .
\end{equation}
If the curve $C$ is chosen so that the contact relations hold,
$d\partial u|_C = \partial du|_C,\ldots,d\partial^k u|_C = \partial^k du|_C$,
then the line integral becomes a general homotopy integral
\begin{equation}\label{invfrechet}
f = f\big|_{u=u_0} + \int_{0}^{1} (\delta_v f)\Big|_{v=\partial_\lambda u_{(\lambda)},u=u_{(\lambda)}}\;d\lambda,
\quad
u_{(1)} = u,
\quad
u_{(0)} = u_0
\end{equation}
where $u_{(\lambda)}(t,x)$ is a homotopy curve given by a parametric family of functions.
If $f$ is non-singular when $u=0$,
then the homotopy curve can be chosen simply to be a homogeneous line,
which yields a standard linear-homotopy integral \cite{Olv}
\begin{equation}
f = f\big|_{u=0} + \int_{0}^{1} (\delta_u f)\big|_{u=u_{(\lambda)}}\frac{d\lambda}{\lambda},
\quad
u_{(\lambda)} = \lambda u .
\end{equation}
The {\em Euler operator} $E_u$ is defined in terms of the Fr\'echet derivative
through the relation
\begin{equation}\label{frechet-euler}
\delta_v f = vE_u(f) +D\cdot\Upsilon_f(v)
\end{equation}
obtained from integration by parts,
which gives
\begin{equation}\label{eulerop}
E_u(f) =
\parder{f}{u} -D\cdot\Big(\parder{f}{(\partial u)}\Big) + \cdots
+ (-D)^k\cdot\Big(\parder{f}{(\partial^k u)}\Big)
\end{equation}
where
\begin{equation}\label{frechet-euler-current}
\begin{aligned}
\Upsilon_f(v) = \Psi(v,1;f) & =
v\parder{f}{(\partial u)}
+ Dv\cdot\parder{f}{(\partial^2 u)} - vD\cdot\parder{f}{(\partial^2 u)} +\cdots
\\&\qquad
+ \sum_{l=1}^{k}(D^{k-l}v)\cdot\Big((-D)^{l-1}\cdot\parder{f}{(\partial^k u)}\Big)
= \sum_{l=0}^{k-1}(D^lv)\cdot E_{\partial^{l+1}u}(f) .
\end{aligned}
\end{equation}
The Euler-Lagrange relation \eqref{frechet-euler} can be combined with the general homotopy integral \eqref{invfrechet}
to obtain the following useful formula.
\begin{lem}\label{inveulerlagr-lemma}
\begin{equation}\label{inveulerlagr}
f=\int_0^1 \partial_\lambda u_{(\lambda)} E_u(f)\big|_{u=u_{(\lambda)}}\;d\lambda
+ D\cdot F
\end{equation}
is an identity,
where
\begin{equation}\label{inveulerlagr-current}
F= \int_0^1\Upsilon_f(\partial_\lambda u_{(\lambda)})\big|_{u=u_{(\lambda)}}\;d\lambda +F_0
\end{equation}
with $F_0=(F_0^t(t,x),F_0^x(t,x))$ being any current such that $D\cdot F_0=f|_{u=u_0}$.
\end{lem}
A useful relation is
\begin{equation}\label{highereuler-current}
\Upsilon_f(v) =
vE_u^{(1)}(f) + D\cdot(vE_u^{(2)}(f)) +\cdots
+ D^{k-1}\cdot(vE_u^{(k)}(f))
\end{equation}
which arises through repeated integration by parts on the expression \eqref{frechet-euler-current},
where
\begin{equation}\label{highereulerop}
E_u^{(l)}(f) =
\parder{f}{(\partial^l u)} -\binom{l+1}{l}D\cdot\Big(\parder{f}{(\partial^{l+1} u)}\Big) + \cdots
+ \binom{k}{l}(-D)^{k-l}\cdot\Big(\parder{f}{(\partial^k u)}\Big),
\quad
l=1,\ldots,k
\end{equation}
define the {\em higher Euler operators}.
Equations \eqref{frechet-euler} and \eqref{highereuler-current}
then provide an alternative formula
for the Fr\'echet derivative
\begin{equation}\label{frechet-euler-rel}
\delta_v f = vE_u(f) + D\cdot(vE_u^{(1)}(f)) + \cdots + D^k\cdot(vE_u^{(k)}(f))
\end{equation}
which leads to a similar formula for the Fr\'echet adjoint derivative
\begin{equation}\label{adjfrechet-euler-rel}
\delta_w^* f = wE_u(f) -(Dw)\cdot E_u^{(1)}(f) + \cdots + (-D)^kw\cdot E_u^{(k)}(f)
\end{equation}
after integration by parts.
Explicit coordinate formulas for all of the Euler operators
are stated in \Ref{Olv};
coordinate formulas for the Fr\'echet derivative and its adjoint,
as well as the associated divergence, are shown in \Ref{AncKar}.
The Euler operators \eqref{eulerop} and \eqref{highereulerop}
have the following important properties.
\begin{lem}\label{eulerop-lemma}
(i) $E_u(fg) = \delta_g^* f + \delta_f^* g$ is a product rule.
(ii) $E_u(f)=0$ holds identically iff $f=D\cdot F$
for some differential current function $F=(F^t,F^x$).
(iii) $E_u^{(1)}(D\cdot F) = E_u(F^t,F^x) = (E_u(F^t),E_u(F^x))$
and
$E_u^{(l+1)}(D\cdot F) = (E_u^{(l)},E_u^{(l)})\odot (F^t,F^x)$, $l\geq 1$,
are descent rules,
where $\odot$ denotes the symmetric tensor product.
\end{lem}
The proof of (i) is an immediate consequence of the ordinary product rule
applied to each partial derivative term in $E_u(fg)$.
To prove the first part of (ii),
if $f=D\cdot F$ then
$\delta_v f = D\cdot\delta_v F$ combined with the Euler-Lagrange relation \eqref{frechet-euler}
yields $vE_u(f)= D\cdot(\delta_v F-\Upsilon_f(v))$.
Since $v$ is an arbitrary differential function, this implies $E_u(f)=0$
(and $\Upsilon_f(v) = \delta_v F$ modulo a divergence-free term).
Conversely, for the second part of (ii),
if $E_u(f)=0$ then the general homotopy integral \eqref{inveulerlagr}
shows $f=D\cdot F$ holds,
with $F$ given by the formula \eqref{inveulerlagr-current}.
The proof of (iii) starts from the property $\delta_v(D\cdot F)= D\cdot\delta_v F$.
Next, the Fr\'echet derivative relation \eqref{frechet-euler-rel} is applied separately to
$f=D\cdot F$ and $f=F$.
This yields $D\cdot(vE_u^{(1)}(D\cdot F))= D\cdot(vE_u(F))$,
$D^2\cdot(vE_u^{(2)}(D\cdot F))= D\cdot(vD\cdot E_u^{(1)}(F))$, and so on.
The expressions for $E_u^{(1)}(D\cdot F)$, $E_u^{(2)}(D\cdot F)$, and so on
are then obtained by recursively expanding out each Euler operator in components
$E_u^{(1)}= E_u^{(t,x)}= (E_u^{(t)}, E_u^{(x)})$
and $E_u^{(l+1)}= E_u^{(l,t,x)}= (E_u^{(l,t)}, E_u^{(l,x)})$, $l\geq 1$,
followed by symmetrizing over these components together with the components of $F=(F^t,F^x)$.
This completes the proof of \lemref{eulerop-lemma}.
A {\em null-divergence} is a total divergence $D\cdot\Phi=0$
vanishing identically in jet space,
where $\Phi=(\Phi^t,\Phi^x)$ is a differential current function.
Similarly to Poincar\'e's lemma,
which shows that ordinary divergence-free vectors in $\mathbb{R}^n$
can be expressed as curls,
null-divergences are total curls in jet space.
\begin{lem}\label{null-div-lemma}
If a differential current function
$\Phi=(\Phi^t(t,x,u,\partial u,\ldots,\partial^k u),\Phi^x(t,x,u,\partial u,$ $\ldots,\partial^k u))$
has a null-divergence,
\begin{equation}\label{total-null-div}
D\cdot\Phi = D_t\Phi^t +D_x\cdot\Phi^x =0
\text{ in $J$,}
\end{equation}
then it is equal to a total curl
\begin{equation}\label{total-curl}
\Phi=D\cdot\Psi = (D_x\cdot\Theta,-D_t\Theta+D_x\cdot\Gamma)
\text{ in $J$}
\end{equation}
with
\begin{equation}\label{curl-terms}
\Psi=\begin{pmatrix} 0 & \Theta\\ -\Theta & \Gamma\end{pmatrix}
\end{equation}
holding for some differential vector function
$\Theta(t,x,u,\partial u,\ldots,\partial^{k-1} u)$
and some differential antisymmetric tensor function
$\Gamma(t,x,u,\partial u,\ldots,\partial^{k-1} u)$,
both of which can be expressed in terms of $\Phi^t,\Phi^x$.
\end{lem}
The proof begins by taking the Fr\'echet derivative of the null-divergence
to get $D\cdot\delta_v\Phi = 0$.
A descent argument will be used to solve this equation.
Let the terms in $\delta_v\Phi=(\delta_v\Phi^t,\delta_v\Phi^x)$
containing highest derivatives $\partial^k v$ be denoted
$(T^{(k)}\partial^k v,X^{(k)}\partial^k v)$,
where the coefficients $T^{(k)}$ and $X^{(k)}$ of each term
are given by a differential scalar function and a differential vector function in $J$.
Then the highest derivative terms $\partial^{k+1}v$
in the equation $D\cdot\delta_v\Phi = 0$ consist of
$T^{(k)}\partial_t\partial^k v +X^{(k)}\cdot\partial_x\partial^k v$.
The coefficients of $\partial^{k+1}v$ in this expression must vanish,
which can be shown to give
$T^{(k)}\partial^k v = \theta^{(k-1)}\cdot\partial_x\partial^{k-1}v$
and
$X^{(k)}\partial^k v = -\theta^{(k-1)}\partial_t\partial^{k-1}v + \gamma^{(k-1)}\cdot\partial_x\partial^{k-1}v$,
where $\theta^{(k-1)}$ is some differential vector function,
and $\gamma^{(k-1)}$ is some differential antisymmetric tensor function.
Integration by parts on these expressions yields
\begin{gather*}
T^{(k)}\partial^k v = D_x\cdot(\theta^{(k-1)}\partial^{k-1}v)+\lot,
\\
X^{(k)}\partial^k v = -D_t(\theta^{(k-1)}\partial^{k-1}v) + D_x\cdot(\gamma^{(k-1)}\partial^{k-1}v)+\lot,
\end{gather*}
and hence
\begin{equation*}
(T^{(k)}\partial^k v,X^{(k)}\partial^k v) = D\cdot\Psi^{(k-1)}(v) +\lot
\end{equation*}
where
\begin{equation*}
\Psi^{(k-1)}(v)=
\begin{pmatrix} 0 & \Theta^{(k-1)}(v)\\ -\Theta^{(k-1)}(v) & \Gamma^{(k-1)}(v)\end{pmatrix}
\end{equation*}
with $\Theta^{(k-1)}(v) = \theta^{(k-1)}\partial^{k-1}v$
and $\Gamma^{(k-1)}(v) = \gamma^{(k-1)}\partial^{k-1}$.
This shows that the highest derivative terms in $\delta_v\Phi$
have the form of a total curl, modulo lower order terms.
Subtraction of this curl $D\cdot\Psi^{(k-1)}(v)$ from $\delta_v\Phi$
will now eliminate all terms containing $\partial^k v$, so that
\begin{equation*}
\delta_v\Phi-D\cdot\Psi^{(k)}(v) = (T^{(k-1)}\partial^{k-1}v,X^{(k)}\partial^{k-1}v) +\lot
\end{equation*}
where the coefficients $T^{(k-1)}$ and $X^{(k-1)}$ of the $\partial^{k-1}v$ terms
are again a differential scalar function and a differential vector function in $J$.
Since total curls have a vanishing total divergence,
the highest derivative terms remaining in the null-divergence equation
$0=D\cdot\delta_v\Phi$ are given by
$T^{(k-1)}\partial_t\partial^{k-1}v +X^{(k-1)}\cdot\partial_x\partial^{k-1}v$,
which has the same form as the expression obtained at highest order.
This completes the first step in the descent argument.
Continuing to lower orders,
the descent argument will terminate at the equation
$T^{(0)}\partial_tv +X^{(0)}\cdot\partial_xv = 0$,
which yields $T^{(0)}=0$ and $X^{(0)}=0$.
As a result,
the solution of the null-divergence equation $D\cdot\delta_v\Phi = 0$
is given by $\delta_v\Phi= \sum_{l=1}^{k} D\cdot\Psi^{(l-1)}(v)$.
The final step in the proof is simply to apply the general homotopy integral \eqref{invfrechet}
to the Fr\'echet derivative $\delta_v\Phi= \sum_{l=1}^{k} D\cdot\Psi^{(l-1)}(v)$,
which gives
\begin{equation*}
\Phi -\Phi\big|_{u=u_0} =
\int_{0}^{1} \big(\sum_{l=1}^{k} D\cdot\Psi^{(l-1)}(\partial_\lambda u_{(\lambda)})\big)\big|_{u=u_{(\lambda)}} \;d\lambda .
\end{equation*}
This shows $\Phi=D\cdot\Psi$ is a total curl,
where
\begin{equation*}
\Psi = \Psi_0 + \int_{0}^{1} \big(\sum_{l=1}^{k} \Psi^{(l-1)}(\partial_\lambda u_{(\lambda)})\big)\big|_{u=u_{(\lambda)}}\;d\lambda
\end{equation*}
has the form \eqref{curl-terms},
with $D\cdot\Psi_0$ being an ordinary curl determined by Poincare's lemma
applied to the vanishing divergence $D\cdot(\Phi|_{u=u_0})=0$.
This completes the proof of \lemref{null-div-lemma}.
{\em Scaling transformations} are a one-parameter Lie group whose action is given by
\begin{equation}\label{scaling}
t\rightarrow \lambda^a t,
\quad
x^i\rightarrow \lambda^{b_{(i)}} x^i,
\quad
u^\alpha\rightarrow \lambda^{c_{(\alpha)}} u^\alpha,
\quad
\lambda \neq 0
\end{equation}
prolonged to jet space,
where the constants $a,b_{(i)},c_{(\alpha)}$ are the scaling weights of $t,x^i,u^\alpha$.
Note the generator of these transformations is ${\bf X}_{\rm scal} = \tau \partial_t + \xi\partial_x + \eta \partial_u$
where
\begin{equation}
\tau = at,
\quad
\xi =(b_{(1)}x^1,\ldots,b_{(n)}x^n),
\quad
\eta = (c_{(1)} u^1,\ldots,c_{(m)} u^m) .
\end{equation}
In characteristic form,
the scaling generator is $\hat{\bf X}_{\rm scal} = P_{\rm scal} \partial_u$
with $P_{\rm scal} = \eta -u_t\tau - u_x\cdot\xi$.
Now consider a differential function $f$ that is homogeneous under the action of the scaling transformation \eqref{scaling},
such that $f\rightarrow \lambda^s f$.
Then the infinitesimal action is given by
$\hat{\bf X}_{\rm scal}(f) = \delta_{P_{\rm scal}} f = s f -\tau D_t f -\xi\cdot D_xf$.
A useful identity comes from integrating this expression by parts
and combining it with the Euler-Lagrange relation \eqref{frechet-euler},
yielding
\begin{equation}\label{inveulerlagr-scaling}
\omega f= P_{\rm scal} E_u(f) +D_t F^t +D_x\cdot F^x,
\quad
\omega= s +D_t\tau +D_x\cdot\xi = s +a + \sum_{i=1}^{n} b_{(i)}
\end{equation}
where
\begin{equation}\label{inveulerlagr-scalingcurrent}
F^t = f\tau + \Upsilon^t_f(P_{\rm scal}),
\quad
F^x = f\xi + \Upsilon^x_f(P_{\rm scal})
\end{equation}
with $\Upsilon_f=(\Upsilon^t_f,\Upsilon^x_f)$ given by expression \eqref{frechet-euler-current}.
Note here $\omega$ is equal to the scaling weight of the integral quantity
$\int_{t_0}^{t_1}\int_\Omega f\;dV\;dt$,
as defined on any given spatial domain $\Omega\subseteq\mathbb{R}^n$
and any time interval $[t_0,t_1]\subseteq\mathbb{R}$.
Finally,
for subsequent developments,
the following technical result
(which is a straightforward application of Hadamard's lemma \cite{Nes}
to the setting of jet space)
will be useful.
\begin{lem}\label{hadamard}
If a differential function $f(t,x,u,\partial u,\ldots,\partial^k u)$ vanishes
on the solution space ${\mathcal E}$ of a given regular PDE system \eqref{pde},
then
\begin{equation}\label{fRrel}
f= R_f(G)
\end{equation}
holds identically,
where
\begin{equation}\label{Rop}
R_f = R_f^{(0)} + R_f^{(1)}\cdot D + \cdots + R_f^{(k-N)}\cdot D^{k-N}
\end{equation}
is a linear differential operator, depending on $f$,
with coefficients given by differential functions
$R_f^{(0)}$, $R_f^{(1)}$, $\ldots$, $R_f^{(k-N)}$ that are non-singular
when evaluated on ${\mathcal E}$.
The operator $R_f|_{\mathcal E}$ is canonically determined by the function $f$
if the PDE system has no differential identities.
Otherwise, if the PDE system satisfies a differential identity
\begin{equation}\label{diffid}
{\mathcal D}(G)={\mathcal D}_1G^1+\cdots+ {\mathcal D}_MG^M=0
\end{equation}
with ${\mathcal D}_1,\ldots,{\mathcal D}_M$ being linear differential operators
whose coefficients are non-singular differential functions
when evaluated on ${\mathcal E}$,
then the operator $R_f|_{\mathcal E}$ is canonically determined only modulo $\chi{\mathcal D}$,
where $\chi$ is an arbitrary differential function.
\end{lem}
The proof relies heavily on the coordinatization property \eqref{pde-solvedform}
that characterizes a PDE system being regular.
For a regular PDE system $G=0$ of order $N\geq 1$,
consider its prolongation to order $k\geq 1$,
${\rm pr} G=(G,DG,\ldots,D^k G)=0$,
which has differential order $k+N$.
Let $(\zeta^1-g^1(Z),\zeta^2-g^2(Z),\ldots)$
be the solved-form derivative expressions for the PDEs in ${\rm pr} G$,
where $\zeta=(\zeta^1,\zeta^2,\ldots)$ $\in J$
denotes the leading derivatives of $u^\alpha$ chosen for the prolonged system,
and $Z=(Z^1,Z^2,\ldots)$ $\in J$
denotes the coordinates for the prolonged solution space ${\mathcal E}\subset J$ of the system.
Note that ${\rm pr} G=0$ represents ${\mathcal E}$ as a set of surfaces
$\zeta^1=g^1(Z)$, $\zeta^2=g^2(Z)$ ,$\ldots$ in $J$.
Then we have $f(\zeta,Z)|_{\mathcal E} = f(g(Z),Z)=0$.
We now use the standard line integral identity
\begin{equation*}
f(\zeta,Z)= \int_{g(Z)}^{\zeta} \partial_y f(y,Z)\cdot dy
= \int_0^1 (\zeta-g(Z))\cdot\partial_\zeta f(s\zeta+(1-s)g(Z),Z)\;ds .
\end{equation*}
This shows that $f(\zeta,Z)= F(\zeta,Z)\cdot(\zeta-g(Z))$,
with $F(\zeta,Z)= \int_0^1 \partial_\zeta f(s\zeta+(1-s)g(Z),Z)\;ds$ being a vector function.
Note $F(\zeta,Z)|_{\mathcal E} = F(g(Z),Z) =\partial_\zeta f(g(Z),Z)$
is non-singular since $f$ is a differential function.
Hence we obtain $F(\zeta,Z)\cdot(\zeta-g(Z))= R_f(G)$
where $R_f$ is a linear differential operator
whose coefficients $F^1(\zeta,Z)$, $F^2(\zeta,Z)$, $\ldots$
are non-singular when evaluated on ${\mathcal E}$.
Furthermore,
the expression for $F(\zeta,Z)$ shows that it is canonically determined by $f$,
unless the PDE system satisfies a differential identity,
whereby $0={\mathcal D}(G)=h(Z)\cdot(\zeta-g(Z))$
holds identically for some vector function $h(Z)$.
In this case, $R_f(G)$ is well-defined only modulo $\chi{\mathcal D}(G)=0$,
where $\chi$ is any differential function.
This completes the proof of \lemref{hadamard}.
\section{Characteristic forms and determining equations for\\ conservation laws and symmetries}
\label{characteristiceqns}
Consider an infinitesimal symmetry \eqref{generator} of a regular PDE system \eqref{pde}.
When acting on the solution space ${\mathcal E}$ of the PDE system in jet space $J$,
the symmetry generator is equivalent to a generator given by
\begin{equation}\label{symmchar}
\hat{\bf X}={\bf X}-{\bf X}_{\rm triv} = P\parderop{u},
\quad
P =\eta-u_t\tau -u_x\cdot\xi
\end{equation}
under which $u$ is infinitesimally transformed
while $t,x$ are invariant.
This generator \eqref{symmchar} defines the {\em characteristic form}
(or canonical representation) for the infinitesimal symmetry.
The symmetry invariance \eqref{Xsymmcond} of the PDE system
can then be expressed by
\begin{equation}\label{symmdeteq}
{\rm pr}\hat{\bf X}(G)|_{\mathcal E} = 0
\end{equation}
holding on the whole solution space ${\mathcal E}$ of the given system.
Note that the action of ${\rm pr}\hat{\bf X}$ is the same as a Fr\'echet derivative \eqref{frechet},
and hence an equivalent, modern formulation \cite{Olv,2ndbook} of this invariance \eqref{symmdeteq}
is given by the {\em symmetry determining equation}
\begin{equation}\label{modernsymmdeteq}
(\delta_P G)|_{\mathcal E} = 0 .
\end{equation}
This formulation of infinitesimal symmetries has several advantages compared to
the standard formulation shown in \secref{prelim}.
Firstly,
a symmetry is trivial iff its characteristic function $P$ vanishes on ${\mathcal E}$.
Also, the differential order of a symmetry is simply given by
the differential order of $P|_{\mathcal E}$.
Secondly,
the symmetry determining equation \eqref{modernsymmdeteq} can be set up
without doing any prolongations of the generator \eqref{symmchar},
as only total differentiation is needed.
Thirdly,
when contact symmetries or higher-order symmetries are sought,
the generator can be formulated simply as
\begin{equation}\label{symmP}
\hat{\bf X} = P(t,x,u,\partial u,\ldots,\partial^r u)\partial_u
\end{equation}
with the symmetry determining equation then being a linear PDE
for the characteristic function $P$.
This formulation \eqref{symmP} eliminates arbitrary functions
depending on all of the variables $t,x,u,\partial u,\ldots,\partial^r u$ in the solution for $P$.
Now consider a conservation law \eqref{conslaw} of a regular PDE system \eqref{pde}.
The starting point to obtain an equivalent characteristic form of the conservation law
is provided by equations \eqref{fRrel} and \eqref{Rop} in \lemref{hadamard}.
These equations show that the conservation law can be expressed
as a divergence identity
\begin{equation}\label{conslawoffE}
D_t T +D_x\cdot X = R_\Phi(G)
= R_\Phi^{(0)}G^{\rm t} + R_\Phi^{(1)}\cdot DG^{\rm t} + \cdots + R_\Phi^{(r+1-N)}\cdot D^{r+1-N}G^{\rm t}
\end{equation}
which is obtained by moving off solutions of the PDE system,
where $u(t,x)$ is an arbitrary (sufficiently smooth) function.
Here $r$ is the differential order of the conserved current $\Phi=(T,X)$,
and $N$ is the differential order of the PDE system.
The next step is to integrate by parts on the righthand side
in the divergence identity \eqref{conslawoffE},
yielding
\begin{equation}\label{chareqn}
D_t\tilde T +D_x\cdot\tilde X= GQ
\end{equation}
with
\begin{equation}\label{equivTX}
\begin{aligned}
(\tilde T, \tilde X) =
& (T,X) + R_\Phi^{(1)}G^{\rm t} + R_\Phi^{(2)}\cdot DG^{\rm t} -(D\cdot R_\Phi^{(2)})G^{\rm t}
\\&\qquad
+\cdots
+ \sum_{l=0}^{r-N}\big((-D)^{l}\cdot R_\Phi^{(r+1-N)}\big)\cdot D^{r-N-l}G^{\rm t}
\end{aligned}
\end{equation}
and
\begin{equation}\label{Q}
Q^{\rm t} = (Q_1,\ldots,Q_M)
= R_\Phi^{(0)} - D\cdot R_\Phi^{(1)} +\cdots +(-D)^{r+1-N}\cdot R_\Phi^{(r+1-N)} .
\end{equation}
On the solution space ${\mathcal E}$,
note that $(\tilde T,\tilde X)|_{\mathcal E} = (T,X)|_{\mathcal E}$ reduces to
the conserved density and the flux in the given conservation law
$(D_t T +D_x\cdot X)|_{\mathcal E}=0$,
and hence
\begin{equation}
(D_t\tilde T +D_x\cdot\tilde X)|_{\mathcal E}= 0
\end{equation}
is a locally equivalent conservation law.
The identity \eqref{chareqn} is called the {\em characteristic equation}
for the conservation law,
and the set of functions \eqref{Q} is called the {\em multiplier}.
Explicit coordinate formulas for the density $\tilde T$ and the flux $\tilde X$ in terms of $T$ and $X$
are shown in \Ref{AncKar}.
When a regular PDE system is expressed in
a solved form \eqref{pde-solvedform}--\eqref{pde-leadingder}
for a set of leading derivatives,
note that these leading derivatives (and their differential consequences)
can be eliminated from the expression for a conserved current $\Phi=(T,X)$
without loss of generality,
since this only changes the conserved current by the addition of a locally trivial current.
Then it is straightforward to derive explicit expressions for the coefficient functions in the operator $R_\Phi$ by applying the chain rule to $D_t T$ and $D_{x^i} X^i$
with the use of subleading derivatives defined by the relations
$\partial_t\partial_{(\ell_a/t)} u^{\alpha_a} = \partial_{x^i}\partial_{(\ell_a/x^i)} u^{\alpha_a}
= \partial_{(\ell_a)} u^{\alpha_a}$.
This leads to an explicit Euler-Lagrange expression
\begin{equation}\label{QfromTX}
Q^{\rm t}= \big(
E_{\partial_{(\ell_1/t)} u^{\alpha_1}}(T) +\sum_{i=1}^{n}E_{\partial_{(\ell_1/x^i)} u^{\alpha_1}}(X^i),\ldots,
E_{\partial_{(\ell_M/t)} u^{\alpha_M}}(T) +\sum_{i=1}^{n}E_{\partial_{(\ell_M/x^i)} u^{\alpha_M}}(X^i)
\big)
\end{equation}
for the components of the multiplier \eqref{Q},
where $\partial_{(\ell_a/t)} u^{\alpha_a}$ and $\partial_{(\ell_a/x^i)} u^{\alpha_a}$
denote the subleading derivatives.
As a result, the multiplier components \eqref{QfromTX}
can contain leading derivatives $\partial_{(\ell_a)} u^{\alpha_a}$
(and their differential consequences) at most polynomially.
Also note that, as asserted by \lemref{hadamard},
if a regular PDE system has no differential identities \eqref{diffid},
then the operator $R_\Phi|_{\mathcal E}$ will be canonically determined by the expression for $\Phi=(T,X)$.
This implies the relation
\begin{equation}
Q^{\rm t}|_{\mathcal E} = (Q_1,\ldots,Q_M)|_{\mathcal E}
= E_G(D_t\tilde T+D_x\cdot\tilde X)|_{\mathcal E}
= (E_{G^1}(D_tT+D_x\cdot X),\ldots,E_{G^M}(D_tT+D_x\cdot X))|_{\mathcal E}
\end{equation}
for the multiplier \eqref{Q}.
In general, for a given regular PDE system \eqref{pde},
a set of functions
\begin{equation}\label{generalQ}
Q= (Q_1(t,x,u,\partial u,\partial^2 u,\ldots\partial^r u),\ldots,Q_M(t,x,u,\partial u,\partial^2 u,\ldots\partial^r u))^{\rm t}
\end{equation}
will be a multiplier iff each function is non-singular on the PDE solution space ${\mathcal E}$
and their summed product with the expressions $G=(G^1,\ldots,G^M)$ for the PDEs
has the form of a total space-time divergence.
The characteristic equation \eqref{chareqn} establishes that, up to local equivalence,
all non-trivial conservation laws for any regular PDE system arise from multipliers.
A determining condition to find all multipliers comes from \lemref{eulerop-lemma}
applied to the characteristic equation \eqref{chareqn}, yielding
\begin{equation}\label{multrdeteq}
0=E_{u}(GQ)=\delta_Q^* G +\delta_G^* Q .
\end{equation}
This condition, which is required to hold identically in jet space,
is necessary and sufficient for $Q$ to be a multiplier.
For each solution $Q$,
a corresponding conserved current that satisfies the characteristic equation \eqref{chareqn}
can be obtained from the expression $f=GQ$ by using \lemref{inveulerlagr-lemma}.
This yields
\begin{equation}
\tilde\Phi= \int_0^1\Upsilon_{GQ}(\partial_\lambda u_{(\lambda)})\big|_{u=u_{(\lambda)}}\;d\lambda
\end{equation}
whose multiplier \eqref{Q} is $Q$.
An explicit formula for this conserved current is stated next.
\begin{lem}\label{TXfromQ}
For a regular PDE system \eqref{pde},
each multiplier \eqref{generalQ} yields a conserved current \eqref{chareqn}
which is explicitly given by a homotopy integral
\begin{align}
&
\tilde T =
\int_0^1 \Big(
\sum_{l=0}^{k-1} \partial_\lambda\partial^{l}u_{(\lambda)}\cdot\big( E_{\partial^{l}\partial_t u}(GQ) \big)\big|_{u=u_{(\lambda)}}
\Big)d\lambda + D_x\cdot\Theta,
\label{TfromQ}\\
&
\tilde X =
\int_0^1 \Big(
\sum_{l=0}^{k-1} \partial_\lambda\partial^{l}u_{(\lambda)}\cdot\big( E_{\partial^{l}\partial_x u}(GQ) \big)\big|_{u=u_{(\lambda)}}
\Big)d\lambda - D_t\cdot\Theta + D_x\cdot\Gamma
\label{XfromQ}
\end{align}
along a homotopy curve $u_{(\lambda)}(t,x)$,
with $u_{(1)}=u$ and $u_{(0)}=u_0$ such that $(GQ)|_{u=u_0}$ is non-singular.
Here $k=\max(r,N)$.
\end{lem}
Note the conserved current formula \eqref{TfromQ}--\eqref{XfromQ} can be simplified
by evaluating it on the solution space ${\mathcal E}$ of the given regular PDE system.
Modulo a locally trivial current, this yields
\begin{equation}\label{PhifromQ}
\tilde\Phi|_{\mathcal E} =
\int_0^1 \sum_{j=1}^{k}\Big(
\partial_\lambda \partial^{j-1}u_{(\lambda)}\cdot\Big(
\sum_{l=j}^{k}(-D)^{l-j}\cdot\Big(\parder{G}{(\partial^l u)}Q\Big)\Big|_{u=u_{(\lambda)}}
\Big) \Big) d\lambda
\end{equation}
where the curve $u_{(\lambda)}(t,x)$ is now in the solution space ${\mathcal E}$.
\subsection{Correspondence between conservation laws and multipliers}
As shown by the following key result,
multipliers provide a unique characteristic form (or canonical representation)
for locally equivalent conservation laws,
in analogy to the characteristic form \eqref{symmchar} for symmetries,
if a regular PDE system has no differential identities.
A generalization holding for regular PDE systems with differential identities
will be stated later.
\begin{prop}\label{correspondence-no-diffids}
For any regular PDE system \eqref{pde} that has no differential identities,
a conserved current is locally trivial \eqref{trivconslaw}
iff its corresponding multiplier \eqref{Q} vanishes
when evaluated on the solution space of the system.
\end{prop}
The proof has two parts.
For the ``only if part'',
suppose a conserved current is locally trivial \eqref{trivconslaw}.
By \lemref{hadamard},
the conserved density and the flux will have the respective forms
$T=D_x\cdot\Theta + \hat T(G)$
and
$X=-D_t\Theta +D_x\cdot\Gamma+ \hat X(G)$
for some linear differential operators $\hat T$ and $\hat X$
whose coefficients are differential functions that are non-singular
when evaluated on ${\mathcal E}$.
For this conserved current $\Phi=(T,X)$,
consider the divergence identity \eqref{conslawoffE},
where $R_\Phi(G)= D_t\hat T(G) +D_x\cdot\hat X(G)$.
As the PDE system is assumed to have no differential identities,
then the homotopy integral formula for the operator $R_\Phi$
from the proof of \lemref{hadamard}
shows that integration by parts applied to $R_\Phi(G)$
yields $\tilde T=T -\hat T(G)$, $\tilde X=X-\hat X(G)$
in the characteristic equation \eqref{chareqn}--\eqref{equivTX},
and hence $GQ=D_t\tilde T +D_x\cdot \tilde X =0$.
It is now straightforward to determine $Q$ from the equation $GQ=0$.
In the case when $G$ comprises a single PDE (\ie/, $M=1$),
then $Q=0$ is immediate.
In the case when $G$ contains more than one PDE (\ie/, $M>1$),
the equation $GQ=0$ can be solved by linear algebra as follows.
First express each PDE $G^a=0$, $a=1,\ldots, M$,
in the solved form \eqref{pde-solvedform}--\eqref{pde-leadingder}
for a leading derivative $\partial_{(\ell_a)} u^{\alpha_a}$.
Then take the Fr\'echet derivative of $GQ=0$,
which yields
\begin{equation*}
(\delta_vG)Q + G(\delta_vQ)=0 .
\end{equation*}
To solve this Fr\'echet derivative equation,
consider the terms involving $\partial^k\partial_{(\ell_a)}v^{\alpha_a}$
and let $w=(\partial_{(\ell_1)}v^{\alpha_1},\ldots,\partial_{(\ell_M)}v^{\alpha_M})$
for ease of notation.
It is easy to see the expression $\delta_vG$
contains only one term of this form,
which is simply given by $w$ itself,
as a consequence of the solved form of the PDEs $G=(G^1,\ldots,G^M)$.
The expression $\delta_vQ$ contains a sum of terms involving derivatives of $w$,
which will have the form $\sum_{k=0}^{r} Q^{(k)}\partial^kw^{\rm t}$,
where $r$ is the differential order of the highest derivatives of
the variables $\partial_{(\ell_a)} u^{\alpha_a}$ in $Q$,
and where the coefficients $Q^{(k)}$ are differential $M\times M$ matrix functions in $J$.
Hence, all of the terms involving $\partial^k\partial_{(\ell_a)}v^{\alpha_a}$
in the Fr\'echet derivative equation
consist of $wQ+\sum_{k=0}^{r} GQ^{(k)}\partial^kw^{\rm t}=0$.
Then the coefficients of each jet variable $\partial^kw$, $k=0,1,\ldots,r$,
must vanish separately.
This immediately yields $Q^{(k)}=0$ for $k=1,\ldots,r$.
The remaining terms are given by $wQ+GQ^{(0)}w^{\rm t}=0$.
This is a linear homogeneous equation in $w^{\rm t}$,
after the transpose relation $wQ=(wQ)^t=Q^tw^t$ is used,
which gives $(Q^{\rm t} +GQ^{(0)})w^t=0$.
The vanishing of the coefficient of $w^t$ yields $Q^{\rm t} =-GQ^{(0)}$,
and hence $Q|_{\mathcal E}=0$.
For the ``if part'',
suppose a multiplier satisfies $Q|_{\mathcal E}=0$.
Then, \lemref{hadamard} can be applied to get $Q= \hat Q(G)$,
where $\hat Q$ is some linear differential operator
whose coefficients are differential functions that are non-singular
when evaluated on ${\mathcal E}$.
The characteristic equation \eqref{chareqn} must now be solved to
determine the corresponding conserved density $\tilde T$ and flux $\tilde X$.
This will be done in two main steps.
For the first step,
a descent argument will be given to solve the Fr\'echet derivative equation
\begin{equation*}
D\cdot(\delta_v\tilde\Phi)=(\delta_v G)Q+G(\delta_vQ)
\end{equation*}
for $\delta_v\tilde\Phi=(\delta_v\tilde T,\delta_v\tilde X)$,
similarly to the proof of \lemref{null-div-lemma}.
Let $F(v) = (\delta_v G)Q+G(\delta_vQ)$, with $Q= \hat Q(G)$.
The terms in $F(v)$ containing highest derivatives of $v$
will be denoted $F^{(k)}\partial^k v$,
where $k$ is the larger of the differential orders of $Q$ and $G$,
and where the coefficients $F^{(k)}$ are differential functions in $J$
such that $F^{(k)}|_{\mathcal E}=0$ since $F(v)|_{\mathcal E} = ((\delta_v G)\hat Q(G))|_{\mathcal E} = (\delta_v G)|_{\mathcal E} \hat Q(0)=0$.
Note the differential order of $\delta_v\tilde\Phi$ then can be assumed to be $k-1$.
Next, let $\Upsilon(v)=(\delta_v\tilde T,\delta_v\tilde X)$,
and denote the terms containing highest derivatives of $v$
in $\Upsilon(v)$ as $\tilde T^{(k-1)}\partial^{k-1}v$ and $\tilde X^{(k-1)}\partial^{k-1}v$, respectively,
where the coefficients $\tilde T^{(k-1)}$ and $\tilde X^{(k-1)}$
are given by a set of differential scalar functions and a set of differential vector functions in $J$.
In this notation, the Fr\'echet derivative equation becomes
\begin{equation*}
D\cdot\Upsilon(v) = F(v) .
\end{equation*}
Now the highest derivative terms $\partial^k v$ in this equation
are given by
\begin{equation*}
\tilde T^{(k-1)}\partial_t\partial^{k-1}v +\tilde X^{(k-1)}\cdot\partial_x\partial^{k-1}v = F^{(k)}\partial^k v .
\end{equation*}
Expand out $F^{(k)}\partial^k v= F^{(k-1,t)}\partial_t\partial^{k-1}v + F^{(k-1,x)}\cdot\partial_x\partial^{k-1}v$,
and collect the terms $\partial_t\partial^{k-1}v$ and $\partial_x\partial^{k-1}v$ in the equation,
giving
\begin{equation*}
(\tilde T^{(k-1)}-F^{(k-1,t)})\partial_t\partial^{k-1}v +(\tilde X^{(k-1)}-F^{(k-1,x)})\cdot\partial_x\partial^{k-1}v
=0 .
\end{equation*}
The same analysis used in the proof of \lemref{null-div-lemma} then yields
\begin{equation*}
\begin{aligned}
(\tilde T^{(k-1)}\partial^{k-1}v,\tilde X^{(k-1)}\partial^{k-1}v) & =
(F^{(k-1,t)}\partial^{k-1}v,F^{(k-1,x)}\partial^{k-1}v)
\\&\qquad
+ D\cdot\Psi^{(k-2)}(v) +\lot
\end{aligned}
\end{equation*}
where
\begin{equation*}
\Psi^{(k-2)}(v)=
\begin{pmatrix} 0 & \Theta^{(k-2)}(v)\\ -\Theta^{(k-2)}(v) & \Gamma^{(k-2)}(v)\end{pmatrix}
\end{equation*}
with $\Theta^{(k-2)}(v) = \theta^{(k-2)}\partial^{k-2}v$
and $\Gamma^{(k-2)}(v) = \gamma^{(k-2)}\partial^{k-2}$
being given by some differential vector function $\theta^{(k-2)}$
and some differential antisymmetric tensor function $\gamma^{(k-2)}$.
Hence the highest derivative terms in $\Upsilon(v)$ involving $v$ have the form
\begin{equation*}
\Upsilon(v) = (F^{(k-1,t)}\partial^{k-1}v,F^{(k-1,x)}\partial^{k-1}v) + D\cdot\Psi^{(k-2)}(v) + \tilde\Upsilon(v)
\end{equation*}
where $\tilde\Upsilon(v)$ comprises all remaining terms,
which contain derivatives of $v$ up to order $\partial^{k-2}v$,
and where $D\cdot\Psi^{(k-2)}(v)$ is a total curl, which has a vanishing total divergence.
Substitution of this expression $\Upsilon(v)$ into the Fr\'echet derivative equation gives
\begin{equation*}
\begin{aligned}
& (\tilde T^{(k-2)}+D\cdot F^{(k-1,t)})\partial_t\partial^{k-2}v+(\tilde X^{(k-2)}+D\cdot F^{(k-1,x)})\cdot\partial_x\partial^{k-2}v
\\&\qquad
= F^{(k-1)}\partial^{k-1} v +\lot
\end{aligned}
\end{equation*}
where $\tilde T^{(k-2)}$ and $\tilde X^{(k-2)}$ are a set of differential scalar functions and a set of differential vector functions
given by the coefficients of the terms $\partial^{k-2}v$ in $\tilde\Upsilon(v)$.
After $F^{(k-1)}\partial^{k-1} v= F^{(k-2,t)}\partial_t\partial^{k-2}v + F^{(k-2,x)}\cdot\partial_x\partial^{k-2}v$ is expanded out,
the terms containing highest derivatives of $v$ in this equation are given by
\begin{equation*}
(\tilde T^{(k-2)}-F^{(k-2,t)}+D\cdot F^{(k-1,t)})\partial_t\partial^{k-2}v+(\tilde X^{(k-2)}-F^{(k-2,x)}+D\cdot F^{(k-1,x)})\cdot\partial_x\partial^{k-2}v = 0
\end{equation*}
which has the same form as the equation solved previously.
This completes the first step in the descent argument.
Next, continuing to all lower orders,
the descent argument yields
\begin{equation*}
\Upsilon(v) = \sum_{l=1}^{k-1}D\cdot\Psi^{(l-1)}(v) + \sum_{l=0}^{k-1}\sum_{j=l}^{k-1} ((-D)^{j-l}\cdot F^{(j,t)},(-D)^{j-l}\cdot F^{(j,x)})\partial^{l}v .
\end{equation*}
Note the terms in the first sum are a total curl,
and the terms in the second sum vanish on ${\mathcal E}$ since each $F^{(l)}|_{\mathcal E}=0$.
The final step is to apply the general line integral \eqref{line-integral}
to the Fr\'echet derivative $\delta_v\tilde\Phi=\Upsilon(v)$
evaluated on ${\mathcal E}$.
Since $\Upsilon(v)|_{\mathcal E}= \sum_{l=1}^{k-1}D\cdot\Psi^{(l-1)}(v)|_{\mathcal E}$,
this gives
\begin{equation*}
\tilde\Phi|_{\mathcal E} -\tilde\Phi|_{u=0} =
\int_{0}^{1} \sum_{l=1}^{k-1}D\cdot\Psi^{(l-1)}(v)\Big|_{u=u_{(\lambda)},v=\partial_\lambda u_{(\lambda)}}\;d\lambda
\end{equation*}
where $u_{(\lambda)}(t,x)$ is a homotopy curve in the solution space ${\mathcal E}$ of the regular PDE system,
with $u_{(1)}=u(t,x)$ being an arbitrary solution
and $u_{(0)}=u_0(t,x)$ being any particular solution.
Thus $\tilde\Phi|_{\mathcal E} -\tilde\Phi_0 =D\cdot\Psi$ is a total curl,
where
\begin{equation*}
\Psi = \int_{0}^{1} \sum_{l=1}^{k-1} \Psi^{(l-1)}(v)\Big|_{u=u_{(\lambda)},v=\partial_\lambda u_{(\lambda)}}\;d\lambda
\end{equation*}
has the form \eqref{curl-terms}.
Now, substitution of $\tilde\Phi|_{\mathcal E} =\tilde\Phi|_{u=u_0} +D\cdot\Psi$
into $D\cdot\tilde\Phi = GQ$ yields
$0=(D\cdot\tilde\Phi -GQ)|_{\mathcal E} = D\cdot(\tilde\Phi|_{u=u_0})$.
This immediately establishes that $\tilde\Phi|_{u=u_0}=D\cdot\Psi_0$ is an ordinary curl,
by Poincare's lemma.
Thus,
\begin{equation*}
\tilde\Phi|_{\mathcal E} =D\cdot(\Psi+\Psi_0)
\end{equation*}
is a locally trivial conserved current,
which completes the proof of \propref{correspondence-no-diffids}.
The correspondence stated in \propref{correspondence-no-diffids} no longer holds
when a PDE system possesses a differential identity \eqref{diffid}.
In particular, for a given differential identity,
multiplication by an arbitrary differential function $\chi$,
followed by integration by parts,
yields
\begin{equation}\label{gaugeconslaw}
0=\chi {\mathcal D}(G) = G{\mathcal D}^*(\chi) + D\cdot\Phi(\chi,G)
\end{equation}
where $\Phi(\chi,G)$ is a conserved current that vanishes on the solution space of the PDE system,
\begin{equation}\label{gaugePhi}
\Phi(\chi,G)|_{\mathcal E} =\Phi(\chi,0)=0 .
\end{equation}
Hence
\begin{equation}\label{gaugeQ}
Q={\mathcal D}^*(\chi)
\end{equation}
is a multiplier which determines a locally trivial conserved current.
This derivation can be reversed,
showing that the existence of a multiplier \eqref{gaugeQ} is necessary and sufficient
for a PDE system to possess a differential identity \eqref{diffid}.
Multipliers of the form \eqref{gaugeQ},
given by a linear differential operator acting on an arbitrary differential function $\chi$,
will be called {\em gauge multipliers} \cite{AncPoh},
in analogy with gauge symmetries.
Note that a gauge multiplier is non-vanishing on the solution space ${\mathcal E}$ of the PDE system
whenever the differential identity is non-trivial,
since ${\mathcal D}|_{\mathcal E}\neq0$ implies $Q|_{\mathcal E}\neq0$ for $\chi\neq0$.
Two multipliers that differ by a gauge multiplier will be called {\em gauge equivalent}.
{\bf Running \Ex{(3)}}:
The Euler equations for constant density, inviscid fluids in two dimensions
comprise an evolution equation for $\vec u=(u^1,u^2)$,
\begin{equation*}
\vec G = \vec u_t +\vec u\cdot\nabla\vec u +(1/\rho)\nabla p=0,
\end{equation*}
a spatial equation relating $\vec u$ to $p$,
\begin{equation*}
G^p = (1/\rho)\Delta p + (\nabla\vec u)\cdot(\nabla\vec u)^{\rm t}=0,
\end{equation*}
and a spatial constraint equation on $\vec u$,
\begin{equation*}
G^{\rm div} = \nabla\cdot\vec u =0 .
\end{equation*}
This PDE system obeys a differential identity
\begin{equation*}
{\rm Div\,}\vec G -D_t G^{\rm div} -G^p =0
\end{equation*}
which has the form \eqref{diffid} where
${\mathcal D} = {\rm diag}({\rm Div\,},-1,-D_t)$ and $G=(\vec G,G^p,G^{\rm div})$.
The corresponding gauge multiplier is given by
\begin{equation*}
Q=(\vec Q,Q^p,Q^{\rm div})^{\rm t},
\quad
\vec Q = -{\rm Grad\,}\chi,
\quad
Q^p =-\chi,
\quad
Q^{\rm div} = D_t\chi
\end{equation*}
where $\chi$ is an arbitrary differential scalar function.
The characteristic equation yields
\begin{equation*}
GQ = -({\rm Grad\,}\chi)\cdot\vec G -\chi G^p + (D_t\chi)G^{\rm div}
= D_t(\chi G^{\rm div}) +D_x\cdot(-\chi\vec G)
\end{equation*}
which is a locally trivial conservation law,
where $T=\chi G^{\rm div}$ is the conserved density
and $\vec X=-\chi\vec G$ is the spatial flux.
If $\chi$ is chosen to be a constant,
$\chi=1$,
then the conserved density becomes a total spatial divergence
$T=D_x\cdot\vec u$
which produces a boundary conservation law
\begin{equation*}
\frac{d}{dt}\int_\Omega TdV|_{\mathcal E}
=\frac{d}{dt}\oint_{\partial\Omega} \vec u\cdot\vec \nu dA|_{\mathcal E} = 0
\end{equation*}
on any closed spatial domain $\Omega\in\mathbb{R}^2$,
since the flux vanishes on the solution space of the system,
$\vec X|_{\mathcal E} = 0$.
This boundary conservation law represents conservation of streamlines
in the fluid.
{\bf Running \Ex{(4)}}:
The magnetohydrodynamics equations for a compressible, infinite conductivity plasma in three dimensions
comprise evolution equations for $\rho$, $\vec u=(u^1,u^2,u^3)$ and $\vec B=(B^1,B^2,B^3)$,
\begin{gather*}
G^\rho =
\rho_t +\nabla\cdot(\rho\vec u) =0 ,
\\
\vec G^u = \vec u_t + \vec u\cdot\nabla\vec u+(1/\rho)(P'(\rho)\nabla\rho-\vec J\times\vec B)=0,
\quad
4\pi\vec J = \nabla\times\vec B ,
\\
\vec G^B = \vec B_t -\nabla\times(\vec u\times \vec B) =0,
\end{gather*}
and a spatial constraint equation on $\vec B$,
\begin{equation*}
G^{\rm div} = \nabla\cdot\vec B =0 .
\end{equation*}
This PDE system obeys a differential identity
\begin{equation*}
{\rm Div\,}( \vec G^B ) - D_t G^{\rm div} =0
\end{equation*}
which has the form \eqref{diffid} where
${\mathcal D} = {\rm diag}(0,0,{\rm Div\,},-D_t)$ and $G=(G^\rho,\vec G^u,\vec G^B,G^{\rm div})$.
The corresponding gauge multiplier is given by
\begin{equation*}
Q=(Q^\rho,\vec Q^u,\vec Q^B,Q^{\rm div})^{\rm t},
\quad
Q^\rho =0,
\quad
\vec Q^u =0,
\quad
\vec Q^B = -{\rm Grad\,}\chi,
\quad
Q^{\rm div} = D_t\chi
\end{equation*}
where $\chi$ is an arbitrary differential scalar function.
The characteristic equation yields
\begin{equation*}
GQ = -({\rm Grad\,}\chi)\cdot\vec G^B + (D_t\chi)G^{\rm div}
= D_t(\chi G^{\rm div}) +D_x\cdot(-\chi\vec G^B)
\end{equation*}
which is a locally trivial conservation law,
where $T=\chi G^{\rm div}$ is the conserved density
and $\vec X=-\chi\vec G^B$ is the spatial flux.
If $\chi$ is chosen to be a constant, $\chi=1$,
then the conserved density becomes a total spatial divergence
$T=D_x\cdot\vec B$
which produces a boundary conservation law
\begin{equation*}
\frac{d}{dt}\int_\Omega TdV|_{\mathcal E}
= \frac{d}{dt}\oint_{\partial\Omega} \vec B\cdot\vec \nu dA|_{\mathcal E} = 0
\end{equation*}
on any closed spatial domain $\Omega\in\mathbb{R}^3$,
since the flux vanishes on the solution space of the system,
$\vec X|_{\mathcal E} = 0$.
This boundary conservation law represents conservation of magnetic flux
in the plasma.
The following natural generalization of \propref{correspondence-no-diffids} will now be established.
\begin{prop}\label{correspondence-diffids}
For any regular PDE system \eqref{pde} that possesses a differential identity \eqref{diffid},
a conserved current is locally trivial \eqref{trivconslaw}
iff its corresponding multiplier \eqref{Q} evaluated on the solution space of the system
is equal to a gauge multiplier \eqref{gaugeQ} for some differential function $\chi$.
\end{prop}
The same steps used in proof for \propref{correspondence-no-diffids} go through with only two changes.
For the ``if part'',
suppose a multiplier satisfies $Q|_{\mathcal E} = {\mathcal D}^*(\chi)$,
which implies $Q=\hat Q(G) +{\mathcal D}^*(\chi)$ by \lemref{hadamard},
where $\hat Q$ is some linear differential operator
whose coefficients are differential functions that are non-singular
when evaluated on ${\mathcal E}$.
Then the conservation law identity \eqref{gaugeconslaw}
combined with the characteristic equation \eqref{chareqn}
yields
\begin{equation*}
G\hat Q(G) = G(Q-{\mathcal D}^*(\chi)) = D_t(\tilde T+\Phi^t(\chi,G)) +D_x\cdot(\tilde X+\Phi^x(\chi,G)) .
\end{equation*}
This equation can be solved by the same steps
used in proving the ``if part'' of \propref{correspondence-no-diffids},
thus showing that $\tilde\Phi+\Phi(\chi,G)$ is a locally trivial current.
Since $\Phi(\chi,G)$ itself is a locally trivial current,
the conservation law given by $\tilde\Phi$ is therefore locally trivial \eqref{trivconslaw}.
For the ``only if'' part,
suppose a conserved current is locally trivial \eqref{trivconslaw},
so then, by \lemref{hadamard},
the conserved density and the spatial flux will have the respective forms
$T=D_x\cdot\Theta + \hat T(G)$
and
$X=-D_t\Theta +D_x\cdot\Gamma+ \hat X(G)$
for some linear differential operators $\hat T$ and $\hat X$
whose coefficients are differential functions that are non-singular
when evaluated on ${\mathcal E}$.
As the PDE system is assumed to satisfy a differential identity \eqref{diffid},
the divergence identity \eqref{conslawoffE} will be unique only up to the addition of
a multiple of this differential identity, $\chi{\mathcal D}(G)=0$.
This implies from the homotopy integral formula for the operator $R_\Phi$
that the characteristic equation \eqref{chareqn}--\eqref{equivTX} holds with
$\tilde T=T-\hat T(G)-\Phi^t(\chi,G)$, $\tilde X=X-\hat X(G)-\Phi^x(\chi,G)$,
and $GQ=G{\mathcal D}^*(\chi)$.
The equation $G(Q-{\mathcal D}^*(\chi))=0$ can be solved by the same steps
used in proving the ``only if part'' of \propref{correspondence-no-diffids},
thereby showing $(Q-{\mathcal D}^*(\chi))|_{\mathcal E}=0$,
so $Q|_{\mathcal E}$ is equal to ${\mathcal D}^*(\chi)|_{\mathcal E}$.
This completes the proof of \propref{correspondence-diffids}.
The characterization of locally trivial conservation laws
in \propref{correspondence-no-diffids} and \propref{correspondence-diffids}
establishes an important general correspondence result
which underlies the usefulness of multipliers.
For a given regular PDE system,
the set of multipliers forms a vector space
on which the symmetries of the system have a natural action
\cite{Anc16,AncKar}.
A multiplier is called {\em trivial} if yields a locally trivial conservation law,
and two multipliers are said to be {\em equivalent} if they differ by a trivial multiplier.
When the PDE system has no differential identities,
then a multiplier $Q$ is trivial iff it vanishes on the solution space, $Q|_{\mathcal E}=0$,
whereas when the PDE system possesses a differential identity \eqref{diffid},
a multiplier $Q$ is trivial iff it equals a gauge multiplier \eqref{gaugeQ}
on the solution space, $Q|_{\mathcal E} = {\mathcal D}^*(\chi)$.
A set of multipliers is linearly independent if no linear combination of the multipliers is trivial.
Likewise, a set of conservation laws is linearly independent if no linear combination of the conserved currents is locally trivial.
\begin{thm}\label{correspondence}
(i) For any regular PDE system \eqref{pde},
whether or not it possesses a differential identity,
there is a one-to-one correspondence between
its admitted equivalence classes of linearly-independent local conservation laws
and its admitted equivalence classes of linearly-independent multipliers.
(ii) An explicit formulation of this correspondence is given by
the homotopy integral formula \eqref{TfromQ}---\eqref{XfromQ}
for conserved currents in terms of multipliers.
\end{thm}
Infinitesimal symmetries have a well-known action on conserved currents \cite{Olv,2ndbook}.
This action induces a corresponding action of infinitesimal symmetries on multipliers
\cite{Anc16,AncKar},
and there are several equivalent formulas
\cite{Kha,AncBlu97,IbrKarMah,BluTemAnc,Ibr07,Ibr11,Anc16,Anc17}
for the conserved current obtained from the action of a given infinitesimal symmetry
applied to a given multiplier.
It is worth noting that this action does not preserve linear independence of equivalence classes.
For example \cite{Anc16,Anc17},
any non-trivial conserved current that does not explicitly contain at least one of the independent variables in a PDE system
is mapped into a locally trivial current under any translation symmetry.
\subsection{Low-order conservation laws}\label{loworder}
For any given regular PDE system,
the correspondence between local conservation laws and multipliers
stated in \thmref{correspondence} gives a straightforward way using the following three steps
to find all of the non-trivial local conservation laws (up to equivalence) admitted by the PDE system.
Step 1: solve the determining condition \eqref{multrdeteq} to obtain all multipliers.
Step 2: find all linearly independent equivalence classes of non-trivial multipliers.
Step 3: apply the homotopy integral formula \eqref{PhifromQ} to a representative multiplier
in each equivalence class to obtain a corresponding conserved current.
In practice, for solving the determining condition \eqref{multrdeteq},
it is very useful to know at which differential orders the non-trivial multipliers will be found.
As seen in the examples in \secref{examples},
physically important conservation laws, such as energy and momentum,
always have a low differential order for the conserved density $T$ and the spatial flux $X$,
whereas conservation laws having a high differential order are typically connected with integrability.
A general pattern emerges from these conservation law examples
when their multipliers are examined.
In \Ex{1} and \Ex{2},
mass conservation
for the transport equation \eqref{transport-eqn}
and net heat conservation
for the diffusion/heat conduction equation \eqref{diffusionheat-eqn}
both have $Q=1$
which does not involve $u$ or its derivatives.
In \Ex{3},
energy conservation
for the telegraph equation \eqref{telegraph-eqn}
has $Q=\exp(2{\textstyle\int} a(t)dt)u_t$,
while the leading derivative in this equation is $u_{tt}$ or $u_{xx}$.
In \Ex{4},
for the nonlinear dispersive wave equation \eqref{dispersivewave-eqn},
mass conservation,
$L^2$-norm conservation,
and energy conservation
respectively have
$Q=1$,
$Q=2u$,
and
$Q=g(u)+u_{xx}$.
The leading derivative in this equation is $u_t$ or $u_{xxx}$.
In \Ex{5},
for the viscous fluid equations \eqref{viscousfluid-eqn},
mass conservation has $Q^{\rm t}=(1,0)$,
momentum conservation has $Q^{\rm t}=(u,1)$,
and Galilean momentum conservation has $Q^{\rm t}=(tu,t)$,
while $\{\rho_t,u_t\}$ is a set of leading derivatives in this system.
In \Ex{6},
energy conservation for
the barotropic gas flow/compressible inviscid fluid equations \eqref{compressiblefluid-eqn},
has $Q^{\rm t}=(\tfrac{1}{2}u^2,\rho u)$,
and again $\{\rho_t,u_t\}$ is a set of leading derivatives in this system.
In \Ex{7},
momentum conservation,
energy conservation,
and momentum-energy conservation
for the breaking wave equation \eqref{breakingwave-eqn}
respectively have
$Q=1$, $Q=u$, $Q=-(u_{tx}-u(m+\tfrac{1}{2}u)+\tfrac{1}{2}u_x^2)$,
while the Hamiltonian Casimir has $Q=\tfrac{1}{2}(u-u_{xx})^{1/2}$.
The leading derivative in this equation is $u_{txx}$ or $u_{xxx}$.
In \Ex{8},
mass conservation for the porous media equation \eqref{porousmedia-eqn}
has $Q=\alpha(x)$.
In \Ex{9},
angular momentum conservation and boost momentum conservation
for the non-dispersive wave equation \eqref{nondispersivewave-eqn}
respectively have
$Q=(\mathbf{a}\cdot\vec x)\cdot\nabla u$,
$Q=\vec b\cdot x u_t +c^2 t\vec b\cdot\nabla u$,
while the leading derivative in this equation is $u_{tt}$ or $\Delta u$.
In all of these examples,
each variable $\partial^k u$ that appears in the conservation law multiplier
is related to some leading derivative of $u$ in the PDE system
by differentiation of this variable $\partial^k u$ with respect to $t,x$.
In contrast,
the conservation laws for the higher-derivative quantities
\eqref{higherTXcompressiblefluid} in \Ex{6}
and \eqref{higherTXbreakingwave} in \Ex{7}
have, respectively,
$Q^{\rm t}=\big( (2\rho_xu_x/(u_x^2-p'\rho_x^2/\rho^2)^2)_x,(u_x^2/(u_x^2-p'\rho_x^2/\rho^2)^2)_x+p'/\rho^2(\rho_x^2/(u_x^2-p'\rho_x^2/\rho^2)^2)_x \big)$
and $Q=\tfrac{5}{2}m^{-7/2}m_x^2 -2m^{-5/2}m_{xx} -2m^{-3/2}$,
which involve variables of higher differential order
than the leading derivatives.
An exceptional case is the conservation laws for local helicity
and local enstrophy in \Ex{10}.
These conservation laws for the inviscid (compressible/incompressible) fluid equation \eqref{inviscidfluid-eqn}
have, respectively,
$\vec Q = 2 \nabla\times\vec u$
which involves a variable with the same differential order
as the leading derivative $\vec u_t$,
and
$\vec Q = f''(({\rm curl}\,\vec u)/\rho)\nabla\cdot((\nabla\wedge\vec u)/\rho)$
which involves a higher-derivative variable.
Note, however, if the fluid equation is expressed as a system
for the velocity $\vec u$ and
the vorticity vector $\vec\omega =\nabla\times\vec u$ in three dimensions
or the vorticity scalar $\omega ={\rm curl}\,\vec u$ in two dimensions,
then the multipliers for helicity and enstrophy conservation
are given by, respectively,
$Q^{\rm t}=(\vec\omega,\vec u)$ and $Q^{\rm t}=(\vec 0,f'(\omega/\rho))$
in which the variables are related to the leading derivatives
$\vec u_t$ and $\vec\omega_t$ by differentiation with respect to $t$.
This pattern motivates introducing the following general class of multipliers.
A multiplier $Q$ for a regular PDE system \eqref{pde} will be called {\em low-order}
if each jet variable $\partial^k u^\alpha$ that appears in $Q|_{\mathcal E}$
is related to some leading derivative of $u^\alpha$
by differentiations with respect to $t,x^i$.
(Note that, therefore, the differential order $r$ of $Q|_{\mathcal E}$
must be strictly less than the differential order $N$ of the PDE system.)
Correspondingly, a conservation law is said to be of {\em low-order}
if its multiplier is low-order when evaluated on the solution space of the PDE system.
For a given regular PDE system,
the explicit form for low-order conservation laws can be determined from the form for low-order multipliers
by inverting the relation \eqref{Q} which defines a multiplier in terms of a conserved current.
{\bf Running \Ex{(1)}}
The gKdV equation \eqref{gkdv-eqn}
is a time evolution PDE whose leading derivative is $u_t$ or $u_{xxx}$.
Its low-order conservation laws $(D_t T+ D_x X)|_{\mathcal E}=0$
are given by multipliers that have the form
\begin{equation*}
Q(t,x,u,u_x,u_{xx})
\end{equation*}
since, in the jet space $J=(t,x,u,u_t,u_x,u_{tt},u_{tx},u_{xx},\ldots)$,
the only variables that can be differentiated with respect to $t$ or $x$
to obtain a leading derivative are $u,u_x,u_{xx}$.
To derive the corresponding form for low-order conserved currents $\Phi=(T,X)$,
the first step is to expand out $D_t T$ and $D_x X$
starting from general expressions for $T$ and $X$
in which a leading derivative $u_t$ or $u_{xxx}$ has been eliminated
along with all of its differential consequences.
If $u_t$ is chosen, then the starting expressions will be
$T(t,x,u,u_x,u_{xx},\ldots)$ and $X(t,x,u,u_x,u_{xx},\ldots)$,
which gives
\begin{equation*}
D_t T = T_t +u_tT_u + u_{tx}T_{u_x} + u_{txx}T_{u_{xx}} + \cdots,
\quad
D_x X = X_x +u_xX_u + u_{xx}X_{u_x} + u_{xxx}X_{u_{xx}} + \cdots .
\end{equation*}
The second step is to obtain the operator $R_\Phi$ from the terms in the divergence expression
$D_t T +D_x X$ containing $u_t$ (and its differential consequences).
This yields
\begin{equation*}
D_t T +D_x X= (T_u + T_{u_x}D_x + T_{u_{xx}}D_x^2 + \cdots)u_t +T_t + X_x +u_xX_u + u_{xx}X_{u_x} + u_{xxx}X_{u_{xx}} + \cdots
\end{equation*}
and hence
\begin{equation*}
R_\Phi = T_u + T_{u_x}D_x + T_{u_{xx}}D_x^2 + \cdots
\end{equation*}
since $u_t=G-u^pu_x-u_{xxx}$ is the solved form for the PDE expression.
Then the main steps are, first, to equate $Q$ with the expression $E_{G}(R_\Phi(G))$
and, next, to use the resulting equation together with the characteristic equation
$D_t\tilde T+ D_x\tilde X=GQ$
to determine the dependence of $T$ and $X$ on all jet variables that do not appear in $Q$.
This gives, first,
\begin{equation*}
R_\Phi(G) = T_uG + T_{u_x}D_xG + T_{u_{xx}}D_x^2G + \cdots
= \delta_G T = GE_u(T) + D_x\Upsilon^x(G)
\end{equation*}
by using the Euler-Lagrange relation \eqref{frechet-euler},
which yields the equation
\begin{equation*}
Q(t,x,u,u_x,u_{xx})= E_{G}(R_\Phi(G)) = E_G(\delta_G T) = E_u(T) .
\end{equation*}
Comparison of the differential order of both sides of this equation directly determines
\begin{equation*}
T= \tilde T(t,x,u,u_x) +D_x \Theta(t,x,u,u_x,\ldots) .
\end{equation*}
This implies $\Upsilon^x(G)= \tilde T_{u_x} G$.
Next, the characteristic equation then yields
\begin{equation*}
GQ = D_t T + D_x(X -\tilde T_{u_x}G) = D_t\tilde T +D_x\tilde X
\end{equation*}
which gives
\begin{equation*}
D_x(X +D_t\Theta) = D_x(\tilde X +G\tilde T_{u_x})
= -\tilde T_t +(u^pu_x+u_{xxx})\tilde T_u +(u^pu_x+u_{xxx})_x\tilde T_{u_x} .
\end{equation*}
Comparison of both sides of this equation now determines
\begin{equation*}
X= \tilde X(t,x,u,u_t,u_x,u_{xx}) -D_t \Theta(t,x,u,u_x,\ldots) - G\tilde T_{u_x}(t,x,u,u_x) .
\end{equation*}
The same result can be shown to hold if $u_{xxx}$ is chosen as the leading derivative instead of $u_t$.
Hence, all low-order conserved currents have the general form
\begin{equation*}
\Phi|_{\mathcal E}=(\tilde T(t,x,u,u_x),\tilde X(t,x,u,u_t,u_x,u_{xx}))
\end{equation*}
modulo locally trivial conserved currents.
{\bf Running \Ex{(2)}}
The breaking wave equation \eqref{b-fam-u-eqn}
is a regular PDE whose leading derivative is $u_{txx}$ or $u_{xxx}$.
All low-order conservation laws of this PDE are given by multipliers that have the second-order form
\begin{equation}\label{bfam-lowordQ}
Q(t,x,u,u_t,u_x,u_{tx},u_{xx})
\end{equation}
where $u_{tt}$ is excluded because it cannot be differentiated to obtain a leading derivative $u_{txx}$ or $u_{xxx}$.
The corresponding form for low-order conserved currents $\Phi=(T,X)$
is derived by starting from general expressions for $T$ and $X$
in which a leading derivative $u_{txx}$ or $u_{xxx}$ has been eliminated
along with all of its differential consequences.
It is simplest to use the pure derivative $u_{xxx}$,
which implies $T$ and $X$ are functions only of
$t$, $x$, $u$, $u_x$, $u_{xx}$, and their $t$-derivatives.
Then the terms in the divergence expression $D_t T +D_x X$
containing the leading derivative $u_{xxx}$ (and its differential consequences)
are given by
\begin{equation*}
\begin{aligned}
D_t T +D_x X & = (X_{u_{xx}} + X_{u_{txx}}D_t + X_{u_{ttxx}}D_t{}^2 + \cdots)u_{xxx}
\\&\qquad
+T_t +X_x +u_tT_u +u_xX_u + u_{tt}T_{u_t} + u_{tx}(T_{u_x}+X_{u_t}) + u_{xx}X_{u_x}
\\&\qquad
+ u_{ttt}T_{u_{tt}} + u_{ttx}(T_{u_{tx}} +X_{u_{tt}}) + u_{txx}(T_{u_{xx}} + X_{u_{tx}}) + \cdots .
\end{aligned}
\end{equation*}
This expression yields the operator
\begin{equation*}
R_\Phi = (X_{u_{xx}} + X_{u_{txx}}D_t + X_{u_{ttxx}}D_t{}^2 + \cdots)u^{-1}
\end{equation*}
since $u_{xxx}=-u^{-1}(G+bu_x(u_{xx}-u)-u_{txx} +u_t)+u_x$ is the solved form for the PDE expression.
Now, the main steps consist of,
first, equating $Q$ with the expression $E_{G}(R_\Phi(G))$
and, next, using the characteristic equation $D_t\tilde T+ D_x\tilde X=GQ$
to determine the dependence of $T$ and $X$ on all jet variables that do not appear in $Q$.
The first step gives
\begin{equation*}
R_\Phi(G) = (X_{u_{xx}} + X_{u_{txx}}D_t + X_{u_{ttxx}}D_t{}^2 + \cdots)(-u^{-1}G)
= -u^{-1}G E_{u_{xx}}(X) - D_t\Upsilon^t(u^{-1}G)
\end{equation*}
after using the relation \eqref{frechet-euler},
which yields the equation
\begin{equation*}
Q(t,x,u,u_t,u_x,u_{tx},u_{xx}) = E_{G}(R_\Phi(G)) = E_G(-u^{-1}G E_{u_{xx}}(X)) = -u^{-1}E_{u_{xx}}(X) .
\end{equation*}
Comparison of the differential order of both sides of this equation directly determines
\begin{equation*}
X=\tilde X(t,x,u,u_t,u_x,u_{tx},u_{xx}) -D_t\Theta(t,x,u,u_t,u_x,u_{tt},u_{tx},u_{xx},\ldots)
\end{equation*}
which implies $\Upsilon^t(u^{-1}G)=0$.
Then, for the next step,
the characteristic equation yields
\begin{equation*}
GQ = D_t T +D_x X = D_t\tilde T +D_x\tilde X
\end{equation*}
giving
\begin{equation*}
D_t(T-D_x\Theta)= D_t\tilde T
= -\tilde X_x -u_x\tilde X_u -u_{tx}\tilde X_{u_t} -u_{xx}\tilde X_{u_x} -u_{txx}\tilde X_{u_{tx}} .
\end{equation*}
Comparison of both sides of this equation now determines
\begin{equation*}
T= \tilde T(t,x,u,u_x,u_{xx}) +D_x\Theta(t,x,u,u_t,u_x,u_{tt},u_{tx},u_{xx},\ldots) .
\end{equation*}
Hence, all low-order conserved currents have the general form
\begin{equation*}
\Phi|_{\mathcal E}=(\tilde T(t,x,u,u_x,u_{xx}),\tilde X(t,x,u,u_t,u_x,u_{tx},u_{xx}))
\end{equation*}
modulo locally trivial conserved currents.
\section{Variational symmetries and Noether's theorem in modern form}
\label{Noether}
A PDE system \eqref{pde} is {\em globally variational} if it is given by
the critical points of a variational principle defined on some spatial domain $\Omega\subseteq\mathbb{R}^n$ and some time interval $[t_0,t_1]\subseteq\mathbb{R}$.
In typical applications,
this will involve specifying a function space for $u(t,x)$ with $x\in \Omega$
and also posing boundary conditions on $u(t,x)$ for $x\in\partial\Omega$.
Noether's theorem is usually formulated in this context,
where it shows that every transformation group leaving invariant the variational principle
yields a corresponding conserved integral \eqref{globalconslaw}
for solutions of the PDE system with $u(t,x)$ belonging to the specified function space.
However, for the purpose of obtaining local conservation laws \eqref{conslaw},
a global variational principle is not necessary,
and a PDE system instead needs to have just a local variational principle.
A PDE system \eqref{pde} is {\em locally variational} if it is given by
the Euler-Lagrange equations
\begin{equation}\label{ELpde}
0= G=E_u(L)^{\rm t}
\end{equation}
for some differential function $L(t,x,u,\partial u,\ldots,\partial^k u)$,
called a {\em Lagrangian}.
Note that, as shown by \lemref{eulerop-lemma},
a Lagrangian is unique up to addition of an arbitrary total divergence.
In particular, $L$ and $\tilde L= L+D_t\Psi^t +D_x\cdot\Psi^x$ have the same Euler-Lagrange equations,
for any differential scalar function $\Psi^t$ and differential vector function $\Psi^x$.
There is a well-known condition
for a given PDE system to be locally variational \cite{Olv,2ndbook}.
\begin{lem}\label{eulerlagrange}
$G=E_u(L)^{\rm t}$ holds for some Lagrangian $L(t,x,u,\partial u,\ldots,\partial^k u)$
iff
\begin{equation}\label{selfadj}
\delta_v G^{\rm t}= \delta_v^* G^{\rm t}
\end{equation}
holds for all differential functions $v(t,x)$.
\end{lem}
The ``only if'' part of the proof has two steps.
First, $\delta_v E_u(L) = E_u(\delta_v L)$ can be directly verified to hold,
due to $v$ having no dependence on $u$ and derivatives of $u$.
Next, the Euler-Lagrange relation \eqref{frechet-euler} combined with \lemref{eulerop-lemma}
yields
$E_u(\delta_v L) = E_u(vE_u(L)) = \delta_v^* E_u(L)$,
after again using the fact that $v$ has no dependence on $u$ and derivatives of $u$.
Hence, $\delta_v E_u(L) = \delta_v^* E_u(L)$,
which completes this part of the proof.
The ``if'' part of the proof proceeds by first inverting the relation $G^{\rm t}=E_u(L)$
through applying \lemref{inveulerlagr-lemma} to $f=L$.
This yields
$L= \tilde L +D\cdot F$,
with $\tilde L=\int_0^1 \partial_\lambda u_{(\lambda)} G^{\rm t}\big|_{u=u_{(\lambda)}}d\lambda$.
Then the remaining steps consist of showing that $E_u(L)=E_u(\tilde L)=G^{\rm t}$
holds for this Lagrangian when $\delta_v G^{\rm t}= \delta_v^* G^{\rm t}$.
First, the Fr\'echet derivative of $\tilde L$ gives
$\delta_v\tilde L = \int_0^1 \big( \partial_\lambda v_{(\lambda)} G^{\rm t}\big|_{u=u_{(\lambda)}} + \partial_\lambda u_{(\lambda)} \delta_{v_{(\lambda)}} G^{\rm t}\big|_{u=u_{(\lambda)}} \big)d\lambda$
where $v_{(\lambda)}= \delta_v u_{(\lambda)}$.
Next, substitute $\delta_{v_{(\lambda)}} G^{\rm t}= \delta_{v_{(\lambda)}}^* G^{\rm t}$
and use the Fr\'echet derivative relation \eqref{adjoint},
which yields
\begin{equation*}
\begin{aligned}
\delta_v\tilde L &
=\int_0^1 \big( \partial_\lambda v_{(\lambda)} G^{\rm t}\big|_{u=u_{(\lambda)}} + v_{(\lambda)}\partial_\lambda G^{\rm t}\big|_{u=u_{(\lambda)}} -D\cdot\Psi(\partial_\lambda u_{(\lambda)},v_{(\lambda)};G^{\rm t})\big|_{u=u_{(\lambda)}} \big)d\lambda
\\&
=vG^{\rm t} -v_0G^{\rm t}\big|_{u=u_0}
-D\cdot\int_0^1 \Psi(\partial_\lambda u_{(\lambda)},v_{(\lambda)};G^{\rm t})\big|_{u=u_{(\lambda)}}\;d\lambda
\end{aligned}
\end{equation*}
where $\Psi$ is given by expression \eqref{frechetcurrent}.
Finally, apply $E_v$ to $\delta_v\tilde L$ to get $E_v(\delta_v\tilde L) = G^{\rm t}$,
and use the identity $E_v(\delta_v\tilde L) = E_v(vE_u(\tilde L)) = E_u(\tilde L)$
which follows from the Euler-Lagrange relation \eqref{frechet-euler}.
This yields $E_u(\tilde L) = G^{\rm t}$,
which completes the proof.
The condition \eqref{selfadj} for a PDE system $G=0$ to be locally variational
states that the linearization of $G^{\rm t}$ must be self-adjoint.
From the relations \eqref{frechet} and \eqref{adjfrechet-euler-rel},
or equivalently \eqref{adjfrechet} and \eqref{frechet-euler-rel},
this condition splits with respect to $v,\partial v,\ldots,\partial^k v$
into a linear overdetermined system of equations on $G$:
\begin{equation}\label{helmholtz}
\parder{G}{(\partial^k u)} = (-1)^k \big(E_u^{(k)}(G)\big)^{\rm t},
\quad
k=0,1,\ldots,N
\end{equation}
where $N$ is the differential order of the PDE system $G=0$.
These equations are called the {\em Helmholtz conditions}.
Note the appearance of the transpose implies that the Helmholtz conditions cannot hold
if $u$ and $G$ have a different number of components.
Also, the expression \eqref{highereulerop} for the higher Euler operators $E_u^{(k)}$
shows that the Helmholtz condition for $k=N$ reduces to the equation
\begin{equation}
(1-(-1)^N)\Big(\parder{G}{(\partial^N u)} + \Big(\parder{G}{(\partial^N u)}\Big)^{\rm t}\Big) =0
\end{equation}
which cannot hold if $N$ is odd.
Consequently, a necessary condition for a PDE system to be locally variational is that
its differential order $N$ must be even
and the number $M$ of PDEs must be the same as the number $m$ of dependent variables.
When a PDE system satisfies the Helmholtz conditions \eqref{helmholtz},
a Lagrangian $L$ for the system can be recovered from the expressions $G=(G^1,\ldots,G^M)$
by the general homotopy integral formula
\begin{equation}\label{LfromG}
L= \int_0^1 \partial_\lambda u_{(\lambda)} G^{\rm t}\big|_{u=u_{(\lambda)}}d\lambda
\end{equation}
(as shown in the proof of \lemref{eulerlagrange}).
A total divergence can be added to this Lagrangian to obtain an equivalent Lagrangian
that has the lowest possible differential order,
which is $N/2$.
{\bf Running \Ex{(1)}}
The gKdV equation \eqref{gkdv-eqn} is an odd-order PDE.
Hence, it cannot be locally variational as it stands.
To verify there is no local variational principle,
note $G=G^{\rm t}=u_t+u^pu_x+u_{xxx}$ gives
\begin{equation*}
\delta_v G^{\rm t} = v_t + u^pv_x + pu^{p-1}u_x v + v_{xxx},
\quad
\delta_v^* G^{\rm t} = -v_t -u^pv_x -v_{xxx}
\end{equation*}
and hence $\delta_v G^{\rm t} - \delta_v^* G^{\rm t} = 2v_t + 2u^pv_x + pu^{p-1}u_x v +2 v_{xxx} \neq 0$
whereby $G^{\rm t}$ fails to have a self-adjoint linearization.
Equivalently, the Helmholtz conditions are not satisfied:
\begin{equation*}
\begin{aligned}
(k=0)\quad &
\parder{G}{u} = pu^{p-1}u_x \neq E_u(G) =0,
\\
(k=1)\quad &
\parder{G}{u_t} =1 \neq -E_u^{(t)}(G) = -1,
\quad
\parder{G}{u_x} =u^p \neq -E_u^{(x)}(G) = -u^p,
\\
(k=2)\quad &
\parder{G}{u_{xx}} = 0= E_u^{(x,x)}(G) =-D_x(1),
\\
(k=3)\quad &
\parder{G}{u_{xxx}} =1 \neq -E_u^{(x,x,x)}(G) =-1 .
\end{aligned}
\end{equation*}
However, if a potential variable $w$ is introduced by putting $u=w_x$,
then the PDE becomes $w_{tx}+w_x^pw_{xx}+w_{xxxx}=0$ which has even order.
Repetition of the previous steps, with $G=G^{\rm t}=w_{tx}+w_x^pw_{xx}+w_{xxxx}$,
now gives
\begin{equation*}
\delta_v G^{\rm t} = v_{tx} + w_x^pv_{xx} + pw_x^{p-1}w_{xx} v_x + v_{xxxx} = \delta_v^* G^{\rm t}
\end{equation*}
and
\begin{equation*}
\begin{aligned}
(k=0)\quad &
\parder{G}{w} =0 = E_w(G),
\\
(k=1)\quad &
\parder{G}{w_t} =0 = -E_w^{(t)}(G) = -D_x(1),
\\&
\parder{G}{w_x} = pw_x^{p-1}w_{xx} = -E_w^{(x)}(G) = -pw_x^{p-1}w_{xx} +2D_x(w_x^p) +D_t(1) +D_x^3(1),
\\
(k=2)\quad &
\parder{G}{w_{tx}} =1 = E_w^{(t,x)}(G),
\\
(k=3)\quad &
\parder{G}{w_{xxx}} =0 = -E_w^{(x,x,,x)}(G) = D_x(1) ,
\\
(k=4)\quad &
\parder{G}{w_{xxxx}} =1 = E_w^{(x,x,x,x)}(G) .
\end{aligned}
\end{equation*}
Hence, the potential gKdV equation is locally variational.
A Lagrangian is given by the homotopy integral
\begin{equation*}
L=\int_0^1 w (\lambda w_{tx}+ \lambda^{p+1}w_x^pw_{xx}+\lambda w_{xxxx})\;d\lambda
= \tfrac{1}{2}ww_{tx}+\tfrac{1}{p+2}ww_x^pw_{xx}+\tfrac{1}{2}ww_{xxxx}
\end{equation*}
using $w_{(\lambda)}=\lambda w$.
The addition of a total divergence $D_t\Psi^t+D_x\Psi^x$ given by
\begin{equation*}
\Psi^t = -\tfrac{1}{2}ww_x,
\quad
\Psi^x = -\tfrac{1}{2}(ww_{xxx}-w_xw_{xx}) -\tfrac{1}{(p+1)(p+2)}ww_x^{p+1}
\end{equation*}
yields an equivalent Lagrangian that has minimal differential order,
\begin{equation*}
\tilde L = -\tfrac{1}{2}w_xw_t-\tfrac{1}{(p+1)(p+2)}w_x^{p+2}+\tfrac{1}{2}w_{xx}^2 .
\end{equation*}
For a locally variational PDE system,
a global variational principle on a spatial domain $\Omega$ and a time interval $[t_0,t_1]$
can be defined in terms of a Lagrangian by
\begin{equation}\label{varprinc}
S[u]= \int_{t_0}^{t_1}\int_{\Omega} ( L(t,x,u,\partial u,\ldots,\partial^k u) +D_x\cdot\Theta(t,x,u,\partial u,\ldots) )\;dV\; dt
\end{equation}
where the spatial divergence term is chosen to let spatial boundary conditions
be posed on $u(t,x)$ for $x\in\partial\Omega$.
The critical points of the variational principle \eqref{varprinc} are given by
the vanishing of the variational derivative of $S[u]$,
\begin{equation}
\begin{aligned}
0=S'[u] & = \parder{}{\epsilon}S[u+\epsilon v]\big|_{\epsilon=0}
\\
& = \int_{t_0}^{t_1}\int_{\Omega} vE_u(L)\;dV\; dt
+ \int_{t_0}^{t_1}\oint_{\partial\Omega} ( \delta_v\Theta + \Upsilon_L(v) )\cdot\nu\;dA
\end{aligned}
\end{equation}
where $v(t,x)$ is an arbitrary differential function that satisfies the same spatial boundary conditions as $u(t,x)$.
Here $\nu$ denotes the outward unit normal vector on $\partial\Omega$,
and $\Upsilon_L$ is given by the Euler-Lagrange relation \eqref{frechet-euler}.
Provided $\Theta$ is chosen so that the boundary integral vanishes,
then $S'[u]=0$ yields the PDE system $G=E_u(L)^{\rm t}=0$ on the spatial domain $\Omega$.
\subsection{Variational symmetries}
A {\em variational symmetry} \cite{Olv,1stbook}
of a given variational principle \eqref{varprinc}
is a generator \eqref{generator} whose prolongation leaves invariant the variational principle.
This invariance condition has both a global aspect,
which involves the spatial domain and the spatial boundary conditions,
and a local aspect, which involves only the Lagrangian.
For a local variational principle \eqref{ELpde},
a {\em variational (divergence) symmetry} \cite{Olv,1stbook}
is a generator \eqref{generator} whose prolongation satisfies the invariance condition
\begin{equation}
{\rm pr}{\bf X}(L) = \tau D_t L +\xi\cdot D_x L + D_t\Psi^t +D_x\cdot\Psi^x
\end{equation}
for some for differential scalar function $\Psi^t$ and differential vector function $\Psi^x$.
This condition can be expressed alternatively as
\begin{equation}
{\rm pr}{\bf X}(L) = D_t\tilde\Psi^t +D_x\cdot\tilde\Psi^x -(D_t \tau +D_x\cdot\xi)L
\end{equation}
with $\tilde\Psi^t=\Psi^t +L\tau$ and $\tilde\Psi^x=\Psi^x +L\xi$,
where $D_t \tau +D_x\cdot\xi$ represents the infinitesimal conformal change
in the space-time volume element $dV dt$ under the symmetry generator ${\bf X}$.
A simpler formulation of a variational symmetry is given by
using the characteristic form \eqref{symmchar} for the symmetry generator.
Then an infinitesimal symmetry \eqref{symmchar} is a variational symmetry iff
its prolongation leaves invariant the Lagrangian modulo total a divergence,
\begin{equation}\label{divsymm}
{\rm pr}\hat{\bf X}(L) = D_t\Psi_P^t +D_x\cdot\Psi_P^x
\end{equation}
for some for differential scalar function $\Psi_P^t$ and differential vector function $\Psi_P^x$
depending on the characteristic function $P$ of the symmetry.
Note that, since any total divergence is annihilated by the Euler operator $E_u$,
a variational symmetry preserves the critical points of the Lagrangian $L$.
As a consequence,
every variational symmetry is an infinitesimal symmetry of the PDE system $G=E_u(L)=0$.
The converse is not true in general,
since (for example) scaling symmetries of Euler-Lagrange equations need not always preserve the Lagrangian.
There is an equivalent, modern formulation of the variational symmetry condition \eqref{divsymm}
which uses only the Euler-Lagrange equations and not the Lagrangian itself.
\begin{prop}\label{varsymm-modern}
For any locally variational PDE system \eqref{ELpde},
an infinitesimal symmetry in characteristic form
$\hat{\bf X}=P(t,x,u,\partial u,\ldots,\partial^r u)\partial_u$
is a variational symmetry iff
\begin{equation}\label{varsymmdeteq}
\delta_P G^{\rm t}= -\delta_G^* P^{\rm t}
\end{equation}
holds identically.
\end{prop}
To prove this result, first note that
$E_u({\rm pr}\hat{\bf X}(L))$ vanishes identically iff ${\rm pr}\hat{\bf X}(L)$ is a total divergence,
by \lemref{eulerop-lemma}.
Next,
$E_u({\rm pr}\hat{\bf X}(L)) = E_u(\delta_P L) = E_u(P E_u(L)) = \delta_P^* G^{\rm t} + \delta_{G^{\rm t}}^* P$
directly follows from the Euler-Lagrange relation \eqref{frechet-euler}
combined with the product rule shown in \lemref{eulerop-lemma} for the Euler operator.
Finally, $\delta_P^* G^{\rm t} = \delta_P G^{\rm t}$ holds by \lemref{eulerlagrange},
and $\delta_{G^{\rm t}}^* P = \delta_G^* P^{\rm t}$ holds as an identity.
Hence
$E_u({\rm pr}\hat{\bf X}(L)) = \delta_P G^{\rm t} + \delta_G^* P^{\rm t}$ is an identity.
This completes the proof.
An importance consequence of equation \eqref{varsymmdeteq} is that it provides
a determining condition to find {\em all} variational symmetries
for a given locally variational PDE system, without the explicit use of a Lagrangian.
In particular,
this formulation avoids the need to consider the ``gauge terms'' $D_t\Psi^t +D_x\cdot\Psi^x$
which arise in the Lagrangian formulation \eqref{divsymm}.
{\bf Running \Ex{(1)}}
The Lie symmetries of the gKdV equation \eqref{gkdv-eqn}
consist of a time translation $\hat{\bf X} = -u_t\partial_u$,
a space translation $\hat{\bf X} = -u_x\partial_u$,
a scaling $\hat{\bf X} = -(\tfrac{2}{p} u+3t u_t +xu_x)\partial_u$,
and a Galilean boost $\hat{\bf X} = (1-tu_x)\partial_u$ if $p\neq1$.
These symmetries project to corresponding Lie symmetries of the potential gKdV equation
$w_{tx}+w_x^pw_{xx}+w_{xxxx}=0$
through the relation $u=w_x$.
This yields the generator
$\hat{\bf X} = P\partial_w$ with
\begin{equation}
P=P^{\rm t}=(1-\tfrac{2}{p})c_3 w +c_4 -(c_1 +3c_3t)w_t-(c_2 +c_3 x+c_4 t) w_x .
\end{equation}
The variational Lie symmetries can be easily found by checking the condition \eqref{varsymmdeteq}.
Using $G=G^{\rm t}=w_{tx}+w_x^pw_{xx}+w_{xxxx}$,
a simple computation yields
\begin{equation}
\begin{aligned}
\delta_P G^{\rm t} &
= D_tD_x P +pw_x^{p-1}w_{xx} D_x P + w_x^p D_x^2 P +D_x^4 P
\\&
= -(c_1 +3c_3 t)D_t G -(c_2 +c_3 x +c_4 t) D_x G -(3+\tfrac{2}{p})c_3 G
\end{aligned}
\end{equation}
and also
\begin{equation}
\begin{aligned}
\delta_G^* P^{\rm t} &
= G \parder{P}{w} -D_t\Big( G \parder{P}{w_t} \Big) -D_x\Big( G\parder{P}{w_x} \Big)
\\&
= (5-\tfrac{2}{p})c_3 G +(c_1 +3c_3t)D_t G + (c_2 +c_3 x+c_4 t)D_x G .
\end{aligned}
\end{equation}
Hence,
$0=\delta_P G^{\rm t}+\delta_G^* P^{\rm t} = (2-\tfrac{4}{p}) c_3 G$
determines $(p-2)c_3=0$.
This shows that all of the Lie symmetries except the scaling symmetry are variational symmetries
for an arbitrary nonlinearity power $p\neq 0$,
and that the scaling symmetry is a variational symmetry only for the special power $p=2$.
\subsection{Noether's theorem in modern form}
Variational symmetries have a direct relationship to local conservation laws
through the variational identity
\begin{equation}\label{noetherid}
\begin{aligned}
{\rm pr}\hat{\bf X}(L) & =D_t\Psi_P^t +D_x\cdot\Psi_P^x
\\
& = \delta_P L = P E_u(L) + D\cdot\Upsilon_L(P)
\end{aligned}
\end{equation}
holding due to the Euler-Lagrange relation \eqref{frechet-euler}.
The identity \eqref{noetherid} yields
\begin{equation}\label{noetherconslaw}
P E_u(L) =D\cdot\Phi,
\quad
\Phi=(\Psi_P^t-\Upsilon_L^t(P),\Psi_P^x-\Upsilon_L^x(P))
\end{equation}
which is a conservation law in characteristic form for the PDE system given by $E_u(L)=0$.
When combined with the formula \eqref{PhifromQ} for conserved currents,
this provides a modern, local form of Noether's theorem,
which does not explicitly use the Lagrangian.
\begin{thm}\label{modernNoether}
For any locally variational PDE system $G=E_u(L)^{\rm t}=0$,
variational symmetries $\hat{\bf X}=P\partial_u$
and local conservation laws in characteristic form $D_t\tilde T+ D_x\cdot\tilde X=GQ$
have a one-to-one correspondence given by the relation
\begin{equation}
P=Q^{\rm t} .
\end{equation}
Equivalently,
this correspondence is given by the homotopy integral
\begin{equation}
\tilde\Phi = (\tilde T,\tilde X)
= \int_0^1 \sum_{j=1}^{k}\Big( \partial_\lambda \partial^{j-1}u_{(\lambda)}\cdot\Big(
\sum_{l=j}^{k}(-D)^{l-j}\cdot\Big(\parder{(PG^{\rm t})}{(\partial^l u)}\Big)\Big|_{u=u_{(\lambda)}}
\Big) \Big)d\lambda
\end{equation}
modulo a total curl,
along a homotopy curve $u_{(\lambda)}(t,x)$,
with $u_{(1)}=u$ and $u_{(0)}=u_0$ such that $(GQ)|_{u=u_0}$ is non-singular.
Here $k=\max(r,N)$.
\end{thm}
The Noether correspondence stated in \thmref{modernNoether} has a sharper formulation
using the additional correspondence between multipliers and local conservation laws
provided by \thmref{correspondence}.
This formulation depends on whether a given variational PDE system possesses differential identities or not.
In particular, when a PDE system satisfies a differential identity \eqref{diffid},
there will exist {\em gauge symmetries}
\begin{equation}\label{gaugeP}
\hat{\bf X}= ({\mathcal D}^*(\chi))^{\rm t}\partial_u
\end{equation}
corresponding to gauge multipliers \eqref{gaugeQ},
where ${\mathcal D}$ is the linear differential operator defining the given differential identity \eqref{diffid},
and $\chi$ is an arbitrary differential function.
Two symmetries that differ by a gauge symmetry will be called {\em gauge equivalent}.
Recall, for any regular PDE system,
a symmetry is trivial iff its characteristic function vanishes on the solution space of the PDE system,
and two symmetries are equivalent iff they differ by a trivial symmetry.
\begin{cor}
(i) If a locally variational, regular PDE system \eqref{ELpde} has no differential identities,
then there is a one-to-one correspondence between
its admitted equivalence classes of linearly-independent local conservation laws
and its admitted equivalence classes of linearly-independent variational symmetries.
(ii) If a locally variational, regular PDE system \eqref{ELpde} satisfies a differential identity,
then its admitted equivalence classes of linearly-independent local conservation laws
are in one-to-one correspondence with
its admitted equivalence classes of linearly-independent variational symmetries
modulo gauge symmetries.
\end{cor}
\subsection{Computation of variational symmetries and Noether conservation laws}
Whenever a locally variational PDE system \eqref{ELpde} is regular,
the determining condition \eqref{varsymmdeteq} for finding variational symmetries
$\hat{\bf X}=P(t,x,u,\partial u,\ldots,\partial^r u)\partial_u$
can be converted into a linear system of equations for $P(t,x,u,\partial u,\ldots,\partial^r u)$
by the following steps.
On the solution space ${\mathcal E}$ of the PDE system,
the Fr\'echet derivative adjoint operator $\delta_G^*|_{\mathcal E}$ vanishes.
Thus, the determining condition \eqref{varsymmdeteq} implies
$(\delta_P G^{\rm t})|_{\mathcal E} =0$
which coincides with the determining equation \eqref{modernsymmdeteq}
for an infinitesimal symmetry of the PDE system.
This shows that $P$ is the characteristic function of an infinitesimal symmetry.
From \lemref{hadamard}, it then follows that $P$ satisfies the relation
\begin{equation}\label{symmdeteqoffE}
\delta_P G^{\rm t} = R_P(G^{\rm t})
\end{equation}
for some linear differential operator
\begin{equation}\label{RPop}
R_P = R_P^{(0)} + R_P^{(1)}\cdot D + R_P^{(2)}\cdot D^2 +\cdots + R_P^{(r)}\cdot D^r
\end{equation}
whose coefficients are non-singular on ${\mathcal E}$,
as the PDE system is assumed to be regular,
where $r$ is the differential order of $P$.
Note that if the PDE system satisfies a differential identity \eqref{diffid}
then $R_P$ is determined by $P$ only up to $\chi{\mathcal D}^{\rm t}$
where $\chi$ is an arbitrary differential function
and ${\mathcal D}$ is the linear differential operator defining the identity.
Substitution of the relation \eqref{symmdeteqoffE}
into the determining condition \eqref{varsymmdeteq}
yields
\begin{equation}\label{varsymmoffE}
0=R_P(G^t)+ \delta_G^* P^{\rm t} .
\end{equation}
Note that $\delta_G^* P^{\rm t}$ can be expressed in an operator form
\begin{equation}
\delta_G^* P^{\rm t} = E_u(P)G^{\rm t} -E_u^{(1)}(P)\cdot(DG^{\rm t}) + \cdots + E_u^{(r)}(P)\cdot(-D)^rG^{\rm t}
\end{equation}
using the relation \eqref{adjfrechet-euler-rel}.
Consequently,
when the PDEs $G=(G^1,\ldots,G^M)$ are expressed in a solved form
\eqref{pde-solvedform}--\eqref{pde-leadingder} for a set of leading derivatives,
equation \eqref{varsymmoffE} can be split with respect to
these leading derivatives and their differential consequences.
This yields a linear system of equations
\begin{equation}\label{helmholtzsymmeq}
0= R_P^{(k)} +(-1)^k E_{u}^{(k)}(P),
\quad
k=0,1,\ldots,r .
\end{equation}
Note that these equations are similar in structure to the Helmholtz conditions \eqref{helmholtz}.
Hence, the following result has been established.
\begin{thm}\label{varsymmdetsys}
The determining equation \eqref{varsymmdeteq} for variational symmetries
$\hat{\bf X}=P(t,x,u,\partial u,\ldots,\partial^r u)\partial_u$
of any locally variational, regular PDE system \eqref{ELpde}
is equivalent to a linear system of equations consisting of
the determining condition \eqref{modernsymmdeteq} for $\hat{\bf X}$
to be an infinitesimal symmetry of the PDE system,
and Helmholtz-type conditions \eqref{helmholtzsymmeq} for $\hat{\bf X}$
to leave any Lagrangian of the PDE system invariant modulo a total divergence.
This linear determining system \eqref{modernsymmdeteq}, \eqref{helmholtzsymmeq} is formulated
entirely in terms of the symmetry characteristic function $P$ and the PDE expressions $G=(G^1,\ldots,G^M)$,
without explicit use of a Lagrangian.
\end{thm}
It is important to emphasize that
the determining system \eqref{modernsymmdeteq}, \eqref{helmholtzsymmeq}
can be solved computationally by the same standard procedure \cite{Olv,1stbook,2ndbook}
that is used to solve the standard determining equation \eqref{symmdeteq} for symmetries.
\section{Main results}
\label{mainresults}
For any regular PDE system \eqref{pde}, whether or not it has a variational principle,
all local conservation laws have a characteristic form given by multipliers,
as shown by the general correspondence stated in \thmref{correspondence}.
In the case of regular PDE systems that are locally variational,
the modern form of Noether's theorem given by \thmref{modernNoether}
shows that multipliers for local conservation laws
are the same as characteristic functions for variational symmetries.
These symmetries satisfy a determining equation \eqref{varsymmdeteq}
which can be split into an equivalent determining system
for the symmetry characteristic functions,
without explicit use of a Lagrangian, as shown in \thmref{varsymmdetsys}.
A similar determining system can be derived for multipliers,
by splitting the multiplier determining equation \eqref{multrdeteq} in the same way.
On the solution space ${\mathcal E}$ of a given regular PDE system \eqref{pde},
the Fr\'echet derivative adjoint operator $\delta_G^*|_{\mathcal E}$ vanishes.
Thus, the multiplier determining equation \eqref{multrdeteq} implies
\begin{equation}\label{adjsymmdeteq}
(\delta_Q^* G)|_{\mathcal E} =0
\end{equation}
which is the adjoint of the symmetry determining equation \eqref{modernsymmdeteq},
and its solutions $Q(t,x,u,\partial u,\ldots,\partial^r u)$
are called {\em adjoint-symmetries} \cite{AncBlu97,AncBlu02a,AncBlu02b}
(or sometimes {\em cosymmetries}).
Then $Q$ satisfies the identity
\begin{equation}\label{adjsymmdeteqoffE}
\delta_Q^* G = \delta_{Q^{\rm t}}^* G^{\rm t} = R_{Q^{\rm t}}(G^{\rm t})
\end{equation}
from \lemref{hadamard},
where
\begin{equation}\label{RQop}
R_{Q^{\rm t}} = R_{Q^{\rm t}}^{(0)} + R_{Q^{\rm t}}^{(1)}\cdot D + R_{Q^{\rm t}}^{(2)}\cdot D^2 +\cdots + R_{Q^{\rm t}}^{(r)}\cdot D^r
\end{equation}
is some linear differential operator whose coefficients are non-singular on ${\mathcal E}$,
and $r$ is the differential order of $Q$.
Note that if the PDE system satisfies a differential identity \eqref{diffid}
then $R_{Q^{\rm t}}$ is determined by $Q$ only up to $\chi{\mathcal D}^{\rm t}$
where $\chi$ is an arbitrary differential function
and ${\mathcal D}$ is the linear differential operator defining the identity.
The determining equation \eqref{multrdeteq} now becomes
\begin{equation}\label{multroffE}
0=R_{Q^{\rm t}}(G^{\rm t})+ \delta_G^* Q .
\end{equation}
From the relation \eqref{adjfrechet-euler-rel},
note that $\delta_G^* Q$ can be expressed in an operator form
\begin{equation}
\delta_G^* Q = E_u(Q^{\rm t})G^{\rm t} -E_u^{(1)}(Q^{\rm t})\cdot(DG^{\rm t}) + \cdots + E_u^{(r)}(Q^{\rm t})\cdot(-D)^rG^{\rm t} .
\end{equation}
Consequently,
when the PDEs $G=(G^1,\ldots,G^M)$ are expressed in a solved form
\eqref{pde-solvedform}--\eqref{pde-leadingder} in terms of a set of leading derivatives,
equation \eqref{multroffE} can be split with respect to
these leading derivatives and their differential consequences.
This yields a linear system of equations
\begin{equation}\label{helmholtzmultreq}
0= R_{Q^{\rm t}}^{(k)} +(-1)^k E_{u}^{(k)}(Q^{\rm t}),
\quad
k=0,1,\ldots,r
\end{equation}
which is similar in form to the Helmholtz conditions \eqref{helmholtz}.
Thus, the following result has been established.
\begin{thm}\label{multrdetsys}
The determining equation \eqref{multrdeteq} for conservation law multipliers
of any regular PDE system \eqref{pde}
is equivalent to the linear system of equations \eqref{adjsymmdeteq}, \eqref{helmholtzmultreq}.
In particular, multipliers are adjoint-symmetries \eqref{adjsymmdeteq}
satisfying Helmholtz-type conditions \eqref{helmholtzmultreq},
where these conditions are necessary and sufficient for an adjoint-symmetry
$Q(t,x,u,\partial u,\ldots,\partial^r u)$ to have the variational form \eqref{QfromTX}
derived from a conserved current $\Phi=(T,X^i)$.
\end{thm}
A comparison of the determining systems formulated in \thmref{multrdetsys} and \thmref{varsymmdetsys}
shows how the correspondence between the local conservation laws and the multipliers for regular PDE systems
is related to the Noether correspondence between the local conservation laws and the variational symmetries for locally variational, regular PDE systems.
\begin{cor}\label{adjnoether}
When a regular PDE system is locally variational \eqref{ELpde},
the adjoint-symmetry determining equation \eqref{adjsymmdeteq}
is the same as the symmetry determining equation \eqref{modernsymmdeteq}
and the Helmholtz-type conditions \eqref{helmholtzmultreq}
under which an adjoint-symmetry is a multiplier
are equivalent to the variational conditions \eqref{helmholtzsymmeq}
under which a symmetry is a variational symmetry.
\end{cor}
Thus, \thmrefs{correspondence}{multrdetsys} provide a direct generalization of
the modern form of Noether's theorem given by \thmrefs{modernNoether}{varsymmdetsys},
in which the role of symmetries in the derivation of local conservation laws for variational PDE systems
is replaced by adjoint-symmetries in the derivation of local conservation laws for non-variational PDE systems.
\subsection{Computation of multipliers and conserved currents}
For any given regular PDE system,
all of its non-trivial local conservation laws (up to equivalence)
can be obtained by the following three steps. \\
Step 1: Solve the determining system \eqref{adjsymmdeteq}, \eqref{helmholtzmultreq}
to obtain all multipliers. \\
Step 2: Find all linearly independent equivalence classes of non-trivial multipliers. \\
Step 3: Construct the conserved current determined by a representative multiplier
in each equivalence class.
The multiplier determining system \eqref{adjsymmdeteq}, \eqref{helmholtzmultreq}
can be solved computationally by the same standard procedure \cite{Olv,1stbook,2ndbook}
that is used to solve the determining equation \eqref{modernsymmdeteq} for symmetries.
Moreover, for multipliers of a given differential order $r$,
the multiplier determining system is, in general,
more overdetermined than is the symmetry determining equation
for infinitesimal symmetries of the same differential order $r$.
Consequently,
the computation of multipliers is typically easier than the computation of symmetries.
As an alternative to solving the whole multiplier determining system together,
only the adjoint-symmetry determining equation can be solved first,
and the Helmholtz-type conditions \eqref{helmholtzmultreq} then can be checked for
each adjoint-symmetry to obtain all multipliers.
In practice, it can be computationally hard to obtain the complete solution to
the multiplier determining system (or the adjoint-symmetry determining equation)
because this will involve going to an arbitrarily high differential order
for the dependence of the multiplier (or the adjoint-symmetry) on the derivatives of the dependent variables in the PDE system.
Moreover, for computations using computer algebra,
this differential order must be specified in advance.
The same issue arises when symmetries are being sought,
but often these obstacles are set aside by looking for just Lie symmetries,
or higher symmetries of a special form.
A similar approach can be used for multipliers,
by looking just for all low-order conservation laws
or by looking just for higher order conservation laws with a special form or with a particular differential order.
In physical applications,
there is often a specific class of conserved densities that is of interest.
The form for multipliers corresponding to a given class of conserved densities
can be derived directly by balancing derivatives on both sides of the characteristic equation,
as shown in the running examples in \secref{loworder}.
For each non-trivial multiplier,
the construction of a corresponding non-trivial conserved current
can be carried out by several different methods.
First, the homotopy integral formula \eqref{TfromQ}--\eqref{XfromQ} can be applied.
An advantage of this formula compared to the standard linear-homotopy formula in the literature \cite{Olv,AncBlu02a,AncBlu02b}
is that the homotopy curve can be adapted to the structure of the expressions
for the multiplier $Q$ and the PDE system $G$,
which allows avoiding integration singularities.
Second, the characteristic equation \eqref{chareqn} can be converted into
a linear system of determining equations for the conserved density $\tilde T$ and the flux $\tilde X$.
The determining equations are derived in a straightforward way
starting from the expression for the multiplier $Q$,
similarly to the derivation of the form for low-order conservation laws
explained in \secref{loworder}.
This method is computationally advantageous as it can be implemented in the same way as
setting up and solving the determining system for multipliers \cite{Wol02b,2ndbook}.
Third, if a given PDE system possesses a scaling symmetry
then an algebraic formula that yields a scaling multiple of the conserved current
$\tilde\Phi|_{\mathcal E}=(\tilde T,\tilde X)|_{\mathcal E}$ is available \cite{Anc03},
where the scaling multiple is simply the scaling weight of the corresponding conserved integral.
The formula can be derived by applying the scaling relation \eqref{inveulerlagr-scaling}--\eqref{inveulerlagr-scalingcurrent}
directly to the function $f=GQ$.
This gives
\begin{align}
&\begin{aligned}
T= \omega\tilde T|_{\mathcal E} & = \Big(
P\sum_{l=1}^{k}(-D)^{l-1}\cdot\Big(\parder{G}{(\partial^{l-1}\partial_t u)}Q\Big)
+(D P)\cdot \Big(\sum_{l=2}^{k}(-D)^{l-2}\cdot\Big(\parder{G}{(\partial^{l-1}\partial_t u)}Q\Big)\Big)
\\&\qquad
+\cdots
+(D^{k-1} P)\cdot\Big(\parder{G}{(\partial^{k-1}\partial_t u)}Q\Big)
\Big)\Big|_{\mathcal E} ,
\end{aligned}
\label{scalTfromQ}\\
&\begin{aligned}
X= \omega\tilde X|_{\mathcal E} & = \Big(
P\sum_{l=1}^{k}(-D)^{l-1}\cdot\Big(\parder{G}{(\partial^{l-1}\partial_x u)}Q\Big)
+(D P)\cdot\Big(\sum_{l=2}^{k}(-D)^{l-2}\cdot\Big(\parder{G}{(\partial^{l-1}\partial_x u)}Q\Big)\Big)
\\&\qquad
+\cdots
+(D^{k-1}P)\cdot\Big(\parder{G}{(\partial^{k-1}\partial_x u)}Q\Big)
\Big)\Big|_{\mathcal E} ,
\end{aligned}
\label{scalXfromQ}
\end{align}
modulo a locally trivial current $\Phi_{\rm triv} = (D_x\Theta,-D_t\Theta+D_x\cdot\Gamma)$,
where
\begin{equation}\label{scalgenerator}
P= \eta -u_t\tau - u_x\cdot\xi,
\quad
\tau = at,
\quad
\xi =(b_{(1)}x^1,\ldots,b_{(n)}x^n),
\quad
\eta = (c_{(1)} u^1,\ldots,c_{(m)} u^m)
\end{equation}
are the characteristic functions in the generator of the scaling symmetry \eqref{scaling}.
Here
\begin{equation}\label{scalweight}
\omega= s +D_t\tau +D_x\cdot\xi = s +a + \sum_{i=1}^{n} b_{(i)}
\end{equation}
is a scaling factor,
with $s$ being the scaling weight of the function $GQ$.
Note, as seen from the characteristic equation \eqref{chareqn},
$\omega$ is equal to the scaling weight of the conserved integral
$\int_\Omega \tilde T|_{\mathcal E} dV$,
as defined on any given spatial domain $\Omega\subseteq\mathbb{R}^n$.
This algebraic formula \eqref{scalTfromQ}--\eqref{scalgenerator}
has the advantage that it does not require any integrations.
However, it assumes that the scaling multiple $\omega$ is non-zero,
which means that it can be used only for constructing conserved currents
whose corresponding conserved integral has a non-zero scaling weight, $\omega\neq 0$.
A more general algebraic construction formula can be derived by utilizing dimensional analysis,
which is applicable to PDE systems without a scaling symmetry.
Any given PDE system arising in physical applications will be scaling homogeneous
under dimensional scaling transformations that act by
rescaling the fundamental physical units of all variables and all parametric constants \cite{Olv,1stbook}
(whether or not the PDE system admits a scaling symmetry).
In particular, these dimensional scaling transformations will comprise
independent rescalings of length, time, mass, charge, and so on.
For each dimensional scaling transformation,
a scaling formula will arise for $T$ and $X$,
generalizing the algebraic formula \eqref{scalTfromQ}--\eqref{scalgenerator}
in a way that involves the dependence of $Q$ and $G$
on all of the dimensionful parametric constants appearing in their expressions.
If a conserved integral represents a dimensionful physical quantity,
then the scaling multiple in the resulting formula will be non-zero.
A derivation of this general construction formula will be given elsewhere \cite{AncHer}.
Here, it will be illustrated in a running example.
{\bf Running \Ex{(1)}}
All low-order conservation laws will now be derived
for the gKdV equation \eqref{gkdv-eqn}.
As shown previously,
low-order conserved currents correspond to low-order multipliers,
which have the general form $Q(t,x,u,u_x,u_{xx})$.
Multipliers are adjoint-symmetries that satisfy Helmholtz-type conditions.
To set up the determining system for multipliers,
first note $\delta_Q^* G =-(D_tQ +D_x^3Q +u^pD_x Q)$,
where $G=u_t+u^pu_x+u_{xxx}$.
Hence the adjoint-symmetry determining equation for $Q$ is given by
\begin{equation*}
(D_tQ +D_x^3Q +u^pD_x Q)|_{\mathcal E} =0 .
\end{equation*}
Next look at the terms that contain the leading derivative $u_t$ and its $x$-derivatives
in this equation.
This yields
\begin{equation*}
-D_tQ +D_x^3Q +u^pD_x Q = -\parder{Q}{u} -\parder{Q}{u_x}D_x G -\parder{Q}{u_{xx}}D_x^2 G
= R_Q(G)
\end{equation*}
holding off of the gKdV solution space,
where the components of the operator $R_Q$ are given by
\begin{equation*}
R_Q^{(0)} =-\parder{Q}{u},
\quad
R_Q^{(x)} = -\parder{Q}{u_x},
\quad
R_Q^{(x,x)} = -\parder{Q}{u_{xx}} .
\end{equation*}
Then the Helmholtz-type equations on $Q$ consist of
\begin{equation*}
\begin{aligned}
0= & R_Q^{(0)} + E_{u}(Q) = -D_x\parder{Q}{u_x} + D_x^2\parder{Q}{u_{xx}} ,
\\
0= & R_Q^{(x)} -E_{u}^{(x)}(Q) = -2\parder{Q}{u_x} + 2D_x\parder{Q}{u_{xx}} ,
\\
0= & R_Q^{(x,x)} + E_{u}^{(x,x)}(Q^{\rm t}) = 0 ,
\end{aligned}
\end{equation*}
which reduce to a single equation
\begin{equation*}
D_x\parder{Q}{u_{xx}} -\parder{Q}{u_x} =0 .
\end{equation*}
This Helmholtz-type equation and the adjoint-symmetry equation can be split
with respect to all derivatives of $u$ which do not appear in $Q$,
with $u_t$ eliminated through the gKdV equation.
This gives, after some simplifications,
a linear overdetermined system of 8 equations:
\begin{gather*}
\parder{Q}{u_x} =0,
\quad
\parder{^2Q}{u_{xx}^2} =0,
\quad
\parder{{}^2Q}{x\partial u_{xx}}=0,
\quad
\parder{{}^2Q}{u\partial u_{xx}} =0,
\quad
\parder{{}^3Q}{x\partial u^2} =0,
\\
\\
\parder{{}^3Q}{u^3} -p(p-1)u^{p-2}\parder{Q}{u_{xx}} =0,
\quad
\parder{{}^3Q}{x^2\partial u} + u_{xx}\parder{{}^2Q}{u^2} - pu^{p-1}u_{xx}\parder{Q}{u_{xx}} =0,
\\
\parder{Q}{t} +u^p\parder{Q}{x} +\parder{{}^3Q}{x^3} +3u_{xx}\parder{{}^2Q}{x\partial u} =0 .
\end{gather*}
These equations can be solved for $Q$, with $p$ treated as an unknown, to get
\begin{equation*}
Q=c_1 +c_2 u +c_3(u_{xx} + \tfrac{1}{p+1}u^{p+1}) +c_4 (x-tu) +c_5 (t(3u_{xx} + u^3) -xu)
\end{equation*}
with $c_4=0$ if $p\neq 1$, and $c_5=0$ if $p\neq 2$.
Hence, 5 low-order multipliers are obtained,
\begin{gather*}
Q_1=1,
\quad
Q_2 = u,
\quad
Q_3 = u_{xx} + \tfrac{1}{p+1}u^{p+1},
\quad
p>0 ,
\\
Q_4 = x-tu,
\quad
p=1 ,
\\
Q_5 =t(3u_{xx} + u^3) -xu,
\quad
p=2 .
\end{gather*}
The corresponding low-order conserved currents will now be derived
using the three different construction methods.
First is the homotopy integral method.
The simplest choice for the homotopy is $u_{(\lambda)}=\lambda u$
since the gKdV equation is a homogeneous PDE, $G|_{u=0}=0$.
Hence the homotopy integral is simply given by
\begin{align*}
\tilde T & = \int_0^1 u \parder{(GQ)}{u_t}\Big|_{u=u_{(\lambda)}}\;d\lambda
\\
& = \int_0^1 u \big(
c_1 +c_4 x +(c_2 -c_4 t -c_5 x)u\lambda +(c_3+c_5 3t)u_{xx}\lambda
+ c_5 tu^3 \lambda^3 + c_3 \tfrac{1}{p+1}u^{p+1} \lambda^{p+1}
\big)\;d\lambda
\\
& = (c_1 +c_4 x)u +\tfrac{1}{2}(c_2 -c_4 t -c_5 x)u^2
+\tfrac{1}{2}(c_3 +c_5 3t) uu_{xx} + c_5 \tfrac{1}{4}tu^4
+ c_3 \tfrac{1}{(p+1)(p+2)}u^{p+2}
\end{align*}
and
\begin{align*}
\tilde X &
= \int_0^1 \Big(
u \Big( \parder{(GQ)}{u_x}\Big|_{u=u_{(\lambda)}} -D_x\parder{(GQ)}{u_{xx}}\Big|_{u=u_{(\lambda)}} +D_x^2\parder{(GQ)}{u_{xxx}}\Big|_{u=u_{(\lambda)}} \Big)
\\&\qquad
+u_x \Big( \parder{(GQ)}{u_{xx}}\Big|_{u=u_{(\lambda)}} -D_x\parder{(GQ)}{u_{xxx}}\Big|_{u=u_{(\lambda)}} \Big)
+u_{xx} \parder{(GQ)}{u_{xxx}}\Big|_{u=u_{(\lambda)}}
\Big)d\lambda
\\
& = \int_0^1 \Big(
u \big( (c_4xu-2c_5 u_x-(c_3+c_5 3t)u_{tx} -(c_4t+c_5 x-c_2)u_{xx})\lambda
\\&\qquad
-c_4tu^2\lambda^2
+c_5(3u^2u_{xx}-xu^3)\lambda^3
+c_5 tu^5\lambda^5
+c_1 u^p\lambda^p
+(c_2u^{p+1} +c_3u^pu_{xx})\lambda^{p+1}
\\&\qquad
+c_3\tfrac{1}{p+1} u^{2p+1}\lambda^{2p+1} \big)
+u_x \big( -c_4 +(c_5 u+(c_3+c_5 3t)u_t +(c_4 t+c_5 x-c_2)u_x)\lambda \big)
\\&\qquad
+u_{xx} \big( c_1 +c_4 x +(c_2 -c_4 t -c_5 x)u +(c_3+c_5 3t)u_{xx}) \lambda
+ c_5 tu^3 \lambda^3 + c_3 \tfrac{1}{p+1}u^{p+1} \lambda^{p+1} \big)
\Big)d\lambda
\end{align*}
which is easiest to evaluate when separated into the non-overlapping cases
$p=1$ with $c_5=c_1=c_2=c_3=0$, $p=2$ with $c_4=c_1=c_2=c_3=0$,
and $p>0$ with $c_4=c_5=0$.
This yields the 5 low-order conserved currents
\begin{align*}
&
\tilde T_1 = u ,
\quad
\tilde X_1 = \tfrac{1}{p+1}u^{p+1} +u_{xx}
\\
&
\tilde T_2 = \tfrac{1}{2} u^2 ,
\quad
\tilde X_2 = \tfrac{1}{p+2}u^{p+2} +uu_{xx} - \tfrac{1}{2} u_x^2
\quad
\\
&
\tilde T_3 = \tfrac{1}{2} uu_{xx} + \tfrac{1}{(p+1)(p+2)}u^{p+2} ,
\quad
\tilde X_3 = \tfrac{1}{2(p+1)^2}u^{2p+2} +\tfrac{1}{p+1}u^{p+1} u_{xx} + \tfrac{1}{2}(u_{xx}^2 +u_tu_x)-uu_{tx}
\\
&
\tilde T_4 = xu -\tfrac{1}{2} tu^2 ,
\quad
\tilde X_4 = t(\tfrac{1}{2}u_x^2-uu_{xx} -\tfrac{1}{3}u^3) +x(u_{xx}+\tfrac{1}{2}u^2)-u_x
\quad
p=1
\\
&\begin{aligned}
\tilde T_5 = \tfrac{1}{2}(3t uu_{xx} -xu^2) +\tfrac{1}{4}tu^4 ,
\quad
\tilde X_5 & = t(\tfrac{3}{2}(u_{xx}^2+u_tu_x)+u^3u_{xx} -\tfrac{3}{2}uu_{tx}+\tfrac{1}{6}u^6)
\\&\qquad
+x(\tfrac{1}{2}u_x^2-uu_{xx}-\tfrac{1}{4}u^4)-\tfrac{1}{2}uu_x,
\quad
p=2
\end{aligned}
\end{align*}
whose respective multipliers are $Q_1,\ldots,Q_5$.
Each of these conserved currents is in characteristic form,
namely $D_t\tilde T_i +D_x\tilde X_i = Q_iG$.
Second is the integration method using the characteristic equation
$D_t\tilde T +D_x\tilde X = GQ$,
where
\begin{equation*}
GQ=\big( c_1 +c_2 u +c_3(u_{xx} + \tfrac{1}{p+1}u^{p+1}) +c_4 (x-tu) +c_5 (t(3u_{xx} + u^3) -xu) \big)(u_t+u^pu_x+u_{xxx})
\end{equation*}
with $c_4=0$ if $p\neq 1$, and $c_5=0$ if $p\neq 2$.
There are three steps in this method.
First, as shown previously
from balancing derivatives on both sides of the characteristic equation,
the general form for all low-order conserved conserved currents
$\tilde\Phi=(\tilde T,\tilde X)$ is found to be given by
\begin{equation*}
\tilde\Phi|_{\mathcal E} =(\tilde T(t,x,u,u_x),\tilde X(t,x,u,u_t,u_x,u_{xx})) .
\end{equation*}
Second,
the characteristic equation can then be split with respect to $u_{tx}$ and $u_{xxx}$,
which yields a linear overdetermined system of three equations:
\begin{gather*}
\parder{\tilde T}{u_x} + \parder{\tilde X}{u_t} =0,
\\
\parder{\tilde X}{u_{xx}}
= c_3\tfrac{1}{p+1}u^{p+1} +c_1+c_4x +(c_2-c_4t-c_5x)u+(c_3+c_53t)u_{xx} +c_5tu^3,
\\
\begin{aligned}
\parder{\tilde T}{t} +\parder{\tilde X}{x} +u_t\parder{\tilde T}{u} +u_x\parder{\tilde X}{u} +u_{xx}\parder{\tilde X}{u_x}
&
= (u_t +u^pu_x) \big( c_3\tfrac{1}{p+1}u^{p+1} +c_1+c_4x
\\&\qquad
+(c_2-c_4t-c_5x)u+(c_3+c_53t)u_{xx} +c_5tu^3 \big) .
\end{aligned}
\end{gather*}
These equations can be integrated directly.
It is simplest to consider separately the non-overlapping cases
$p=1$ with $c_5=c_1=c_2=c_3=0$, $p=2$ with $c_4=c_1=c_2=c_3=0$,
and $p>0$ with $c_4=c_5=0$.
The first case is found to reproduce $\tilde T_4$ and $\tilde X_4$;
the second case yields
$\tilde T_5 -D_x\tilde\Theta_5$ and $\tilde X_5+D_t\tilde\Theta_5$
where $\tilde\Theta_5=\tfrac{3}{2}tuu_x$.
Similarly, the third case with $c_3=0$ is found to reproduce
$\tilde T_1$, $\tilde T_2$, $\tilde X_1$, $\tilde X_2$,
and with $c_3\neq 0$ it yields
$\tilde T_3 -D_x\tilde\Theta_3$ and $\tilde X_3+D_t\tilde\Theta_3$
where $\tilde\Theta_3=\tfrac{1}{2}uu_x$.
Thus, the resulting conserved currents
agree with those obtained from the homotopy integral, up to locally trivial currents.
In particular, each of these currents is in characteristic form.
Third is the scaling symmetry method.
The gKdV equation possesses a scaling symmetry
\begin{equation*}
t\rightarrow \lambda^3 t,
\quad
x\rightarrow \lambda x,
\quad
u\rightarrow \lambda^{-2/p} u,
\quad
\lambda \neq 0
\end{equation*}
with the characteristic $P=-(2/p)u - 3t u_t -x u_x$.
Note the multipliers $Q_1,\ldots,Q_5$ are each homogeneous under the scaling symmetry,
with respective scaling weights $q_1=0$, $q_2=-2/p$, $q_3=-2-2/p$, $q_4=1$, $q_5=0$.
Hence the corresponding scaling factors \eqref{scalweight} are given by
$\omega_1=1-2/p$, $\omega_2=1-4/p$, $\omega_3=-1-4/p$, $\omega_4=0$, $\omega_5=0$,
where $s_i=q_i +c-a$, $a=3$, $b=1$, $c=-2/p$.
Then the scaling symmetry formula is given by
\begin{align*}
&
\begin{aligned}
T_i= \omega_i\tilde T_i|_{\mathcal E} = \Big(P \parder{G}{u_t}Q_i\Big)\Big|_{\mathcal E},
\end{aligned}
\\
&
\begin{aligned}
X_i= \omega_i\tilde X_i|_{\mathcal E} & = \Big(
P\Big(\parder{G}{u_x}Q_i -D_x\Big(\parder{G}{u_{xx}}Q_i\Big) + D_x^2\Big(\parder{G}{u_{xxx}}Q_i\Big)\Big)
\\&\qquad
+D_x P \Big( \parder{G}{u_{xx}}Q_i -D_x\Big(\parder{G}{u_{xxx}}Q_i\Big) \Big)
+D_x^2 P \Big( \parder{G}{u_{xxx}}Q_i \Big)
\Big)\Big|_{\mathcal E} ,
\end{aligned}
\end{align*}
modulo a locally trivial current.
For $i=1,2,3$,
this yields the conserved density expressions
\begin{align*}
&
T_1= -(\tfrac{2}{p}u +3t u_t +x u_x)|_{\mathcal E}
= (1-2/p)\tilde T_1|_{\mathcal E} +D_x\Theta_1,
\quad
\Theta_1 = 3t\tilde X_1 -x\tilde T_1,
\\
&
T_2= -((\tfrac{2}{p}u +3t u_t +x u_x) u)|_{\mathcal E} = (1-4/p)\tilde T_2|_{\mathcal E} +D_x\Theta_2,
\quad
\Theta_2 = 3t\tilde X_2 -x\tilde T_2,
\\
&
\begin{aligned}
& T_3= -(\tfrac{2}{p}u +3t u_t +x u_x) (u_{xx} + \tfrac{1}{p+1}u^{p+1})|_{\mathcal E}
= (-1-4/p)\tilde T_3|_{\mathcal E} +D_x\Theta_3 ,
\\&\qquad
\Theta_3 = \tfrac{1}{2}(1+4/p)uu_x +3t(\tilde X_3 +D_t\tilde\Theta_3)-x(\tilde T_3 -D_x\tilde\Theta_3) .
\end{aligned}
\end{align*}
Note their scaling factors are non-zero when $p\neq2$, $p\neq4$, and $p\neq -4$, respectively.
When $p=2$, $T_1$ reduces to a locally trivial conserved density $D_x\Theta_1$,
and when $p=4$, $T_2$ reduces to a locally trivial conserved density $D_x\Theta_2$.
Likewise, when $p=-4$, $T_3$ reduces to a locally trivial conserved density $D_x\Theta_3$.
The expressions given by the scaling symmetry formula for $i=4,5$
yield
\begin{gather*}
T_4 = -(2u +3t u_t +x u_x)(x-tu)|_{\mathcal E} =D_x\Theta_4,
\\
\Theta_4 = (x-tu)(t(3u_{xx} +u^2)-xu) +\tfrac{3}{2}(tu_x-1)^2,
\end{gather*}
and
\begin{gather*}
T_5 = -(u +3t u_t +x u_x)(3t(u_{xx} + u^3) -xu)|_{\mathcal E} =D_x\Theta_5,
\\
\Theta_5 =\tfrac{1}{2}(t(3u_{xx}+u^3)-xu)^2.
\end{gather*}
These cases for $p>0$ in which the scaling symmetry formula
yields locally trivial currents
are called the critical powers for the corresponding conserved currents.
To obtain the conserved currents for a critical power,
it is necessary to use the more general dimensional scaling formula.
Several steps are needed to set up the dimensional scaling formula.
The first step is to introduce dimensionful constants into the gKdV equation
so that it is homogeneous under separate dimensional scalings of
$t$ $[\text{time}]$, $x$ $[\text{length}]$, and $u$ $[\text{mass}]$.
Thus, let
\begin{equation*}
\tilde G = u_t +\mu u^pu_x +\nu u_{xxx},
\quad
\mu,\nu = {\rm const.}
\end{equation*}
where $\mu$ has dimensions of
$[\text{time}]^{-1}[\text{length}][\text{mass}]^{-p}$,
and $\nu$ has dimensions of
$[\text{time}]^{-1}[\text{length}]^{3}$.
Note $\tilde G=G$ will be the gKdV equation
when these constants have the numerical values $\mu=1$ and $\nu=1$.
The next step is to insert factors of $\mu$ and $\nu$
into the expressions for the low-order multipliers so that $Q_1,\ldots,Q_5$
are each dimensionally homogeneous:
\begin{gather*}
Q_1=1,
\quad
Q_2 = u,
\quad
Q_3 = \nu u_{xx} + \tfrac{1}{p+1}\mu u^{p+1},
\quad
p>0,
\\
Q_4 = x-\mu tu,
\quad
p=1,
\\
Q_5 =t(3\nu u_{xx} + \mu u^3) -xu,
\quad
p=2 .
\end{gather*}
The main step consists of generalizing the scaling relation \eqref{inveulerlagr-scaling}--\eqref{inveulerlagr-scalingcurrent}
so that it applies to dimensional scaling transformations.
These transformations are given by
\begin{align*}
& t \rightarrow \lambda t,
\quad
\mu \rightarrow \lambda^{-1}\mu,
\quad
\nu \rightarrow \lambda^{-1}\nu;
\\
& x \rightarrow \lambda x,
\quad
\mu \rightarrow \lambda\mu,
\quad
\nu \rightarrow \lambda^3\nu;
\\
& u \rightarrow \lambda u,
\quad
\mu \rightarrow \lambda^{-p}\mu,
\quad
\nu \rightarrow \nu;
\end{align*}
as determined by the dimensions of $\mu$ and $\nu$.
Since the scaling relation \eqref{inveulerlagr-scaling}--\eqref{inveulerlagr-scalingcurrent}
only holds for variables in jet space,
the constants $\mu$ and $\nu$ now must be treated as variables
by introducing the equations
\begin{equation*}
\tilde G^{(\mu)} = (\mu_t,\mu_x)=0,
\quad
\tilde G^{(\nu)} = (\nu_t,\nu_x)=0 .
\end{equation*}
Then the augmented PDE system
\begin{equation*}
\tilde G =0,
\quad
\tilde G^{(\mu)} = 0,
\quad
\tilde G^{(\nu)} = 0
\end{equation*}
will admit each of the three scaling transformations as symmetries
formulated in the augmented jet space
$\tilde J=(t,x,u,\mu,\nu,u_t,u_x,\mu_t,\mu_x,\nu_t,\nu_x,\ldots)$.
Note that the characteristic equation for conserved currents
will have additional multiplier terms
\begin{equation*}
D_t\tilde T +D_x\tilde X =
\tilde GQ + \tilde G^{(\mu)}\tilde Q_{(\mu)} +\tilde G^{(\nu)}\tilde Q_{(\nu)}
\end{equation*}
for some expressions
$\tilde Q_{(\mu)} =(\tilde Q_{(\mu)}^t,\tilde Q_{(\mu)}^x)^{\rm t}$
and
$\tilde Q_{(\nu)} =(\tilde Q_{(\nu)}^t,\tilde Q_{(\nu)}^x)^{\rm t}$,
where $Q$ is unchanged.
These expressions can be found in a straightforward way
by setting up and solving the multiplier determining system,
with $Q=Q_i$ being the previously derived low-order multipliers
for the gKdV equation.
Since $\mu$ and $\nu$ appear linearly in each $Q_i$
as well as in the PDE expression $\tilde G$,
note $\tilde Q_{(\mu)}$ and $\tilde Q_{(\nu)}$
can have at most linear dependence on these variables
and cannot contain any derivatives of these variables.
Also, since $Q_i$ depends on $u,u_x,u_{xx},u_{xxx}$,
and $\tilde G$ depends on $u,u_t,u_x,u_{xxx}$,
note $\tilde Q_{(\mu)}$ and $\tilde Q_{(\nu)}$ can depend on only
$u,u_t,u_x,u_{tx},u_{xx}$ in addition to $t,x$ and $\mu,\nu$:
\begin{equation*}
\tilde Q_{(\mu)}(t,x,u,\mu,\nu,u_t,u_x,u_{tx},u_{xx}),
\quad
\tilde Q_{(\nu)}(t,x,u,\mu,\nu,u_t,u_x,u_{tx},u_{xx}) .
\end{equation*}
The multiplier determining system is then given by
\begin{gather*}
E_u(\tilde GQ_i + \tilde G^{(\mu)}\tilde Q_{(\mu)} +\tilde G^{(\nu)}\tilde Q_{(\nu)})=0,
\\
E_{(\mu)}(\tilde GQ_i + \tilde G^{(\mu)}\tilde Q_{(\mu)} +\tilde G^{(\nu)}\tilde Q_{(\nu)})=0,
\quad
E_{(\nu)}(\tilde GQ_i + \tilde G^{(\mu)}\tilde Q_{(\mu)} +\tilde G^{(\nu)}\tilde Q_{(\nu)})=0
\end{gather*}
for $i=1,\ldots,5$.
This system splits with respect to all derivatives of $u,\mu,\nu$ which do not appear in $\tilde Q_{(\mu)}$ and $\tilde Q_{(\nu)}$.
Integration of the resulting equations yields
\begin{gather*}
\tilde Q_{1(\mu)}^t = 0,
\quad
\tilde Q_{1(\mu)}^x = \tfrac{1}{p+1}u^{p+1},
\quad
\tilde Q_{1(\nu)}^t = 0,
\quad
\tilde Q_{1(\nu)}^x = u_{xx},
\\
\tilde Q_{2(\mu)}^t = 0,
\quad
\tilde Q_{2(\mu)}^x = \tfrac{1}{p+2}u^{p+2},
\quad
\tilde Q_{2(\nu)}^t = 0,
\quad
\tilde Q_{2(\nu)}^x = uu_{xx}-\tfrac{1}{2}u_x^2,
\\
\begin{aligned}
& \tilde Q_{3(\mu)}^t =\tfrac{1}{(p+1)(p+2)}u^{p+2},
\quad
\tilde Q_{3(\mu)}^x = \tfrac{1}{p+1}\nu u^{p+1} +\tfrac{1}{(p+1)(p+2)}\mu u^{2p+2},
\\
& \tilde Q_{3(\nu)}^t = -\tfrac{1}{2}u_x^2,
\quad
\tilde Q_{3(\nu)}^x = \nu u_{xx}^2 +u_tu_x +\mu u^{p+1}u_{xx},
\end{aligned}
\\
\begin{aligned}
& \tilde Q_{4(\mu)}^t =-\tfrac{1}{2}tu^2,
\quad
\tilde Q_{4(\mu)}^x = t\nu(\tfrac{1}{2} u_x^2 -uu_{xx}) +\tfrac{1}{2}x u^2-\tfrac{2}{3}\mu u^3,
\\
& \tilde Q_{4(\nu)}^t = 0,
\quad
\tilde Q_{4(\nu)}^x = t\mu(\tfrac{1}{2} u_x^2 -uu_{xx}) +xu_{xx} -u_x,
\end{aligned}
\\
\begin{aligned}
& \tilde Q_{5(\mu)}^t =\tfrac{1}{4}tu^2,
\quad
\tilde Q_{5(\mu)}^x = t(\nu u^3u_{xx} + \tfrac{1}{3}\mu u^6) -\tfrac{1}{4}x u^4,
\\
& \tilde Q_{5(\nu)}^t = -\tfrac{3}{2}tu_x^2,
\quad
\tilde Q_{5(\nu)}^x = t(\mu u^3u_{xx} +3\nu u_{xx}^2+3u_tu_x) +x(\tfrac{1}{2} u_x^2 -uu_{xx}) +uu_x .
\end{aligned}
\end{gather*}
The scaling relation \eqref{inveulerlagr-scaling}--\eqref{inveulerlagr-scalingcurrent}
can now be applied to the function
$f_i=\tilde G Q_i + \tilde G^{(\mu)}\tilde Q_{i(\mu)} +\tilde G^{(\nu)}\tilde Q_{i(\nu)}$
in the augmented jet space
$\tilde J=(t,x,u,\mu,\nu,u_t,u_x,\mu_t,\mu_x,\nu_t,\nu_x,\ldots)$
by using an infinitesimal scaling symmetry
given by one of the scaling transformation generators
\begin{gather*}
\hat{\bf X}_{\text{time}} = -(\mu+t\mu_t)\partial_\mu -(\nu+t\nu_t)\partial_\nu -t u_t\partial_u ,
\\
\hat{\bf X}_{\text{length}} = (\mu-x\mu_x)\partial_\mu +(3\nu-x\nu_x)\partial_\nu -x u_x\partial_u ,
\\
\hat{\bf X}_{\text{mass}} = -p\mu\partial_\mu +u\partial_u .
\end{gather*}
Let $P$, $P^{(\mu)}$, $P^{(\nu)}$ denote the characteristic functions
in the selected scaling transformation generator $\hat{\bf X}$.
Then this yields the dimensional scaling formula
\begin{align*}
& T_i=\omega_i\tilde T_i|_{\mathcal E} = \Big(
P \parder{\tilde G}{u_t}Q_i + P^{(\mu)}\tilde Q^t_{i(\mu)} + P^{(\nu)}\tilde Q^t_{i(\nu)}
\Big)\Big|_{\mathcal E} ,
\\
&\begin{aligned}
X_i =\omega_i\tilde X_i|_{\mathcal E} & = \Big(
P\Big(\parder{\tilde G}{u_x}Q_i -D_x\Big(\parder{\tilde G}{u_{xx}}Q_i\Big) + D_x^2\Big(\parder{\tilde G}{u_{xxx}}Q_i\Big)\Big)
\\&\qquad
+D_x P \Big( \parder{\tilde G}{u_{xx}}Q_i -D_x\Big(\parder{\tilde G}{u_{xxx}}Q_i\Big) \Big)
+D_x^2 P \Big( \parder{\tilde G}{u_{xxx}}Q_i \Big)
\\&\qquad
+ P^{(\mu)}\tilde Q^x_{i(\mu)} + P^{(\nu)}\tilde Q^x_{i(\nu)}
\Big)\Big|_{\mathcal E} ,
\end{aligned}
\end{align*}
modulo a locally trivial current,
where
\begin{equation}
\omega_i= q_i +s +D_t\tau+D_x\xi
\end{equation}
is a scaling factor defined in terms of
the scaling weights $q_i,s$ of $Q_i,\tilde G$
and the divergence factor $D_t\tau+D_x\xi$
arising from the selected dimensional scaling transformation.
In particular, for each $i=1,\ldots,5$,
there will be some (possibly combined) transformation
such that the scaling factor $w_i$ is non-zero,
as seen from Tables~\ref{scalproperties} and~\ref{Phiweights}.
\begin{table}[ht!]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|}
\hline
\hfill
& $P$ & $P^{(\mu)}$ & $P^{(\nu)}$ & $D_t\tau+D_x\xi$
& $s$ & $q_1$ & $q_2$ & $q_3$ & $q_4$ & $q_5$
\\
\hline
time
& $-t u_t$
& $-(\mu+t\mu_t)$
& $-(\nu+t\nu_t)$
& $1$
& $-1$
& $0$
& $0$
& $-1$
& $0$
& $0$
\\
\hline
length
& $-x u_x$
& $\mu-x\mu_x$
& $3\nu-x\nu_x$
& $1$
& $0$
& $0$
& $0$
& $1$
& $1$
& $1$
\\
\hline
mass
& $u$
& $-p\mu$
& $0$
& $0$
& $1$
& $0$
& $1$
& $1$
& $0$
& $1$
\\
\hline
\end{tabular}
\end{center}
\caption{Properties of dimensional scaling transformations for the gKdV equation and its low-order multipliers}
\label{scalproperties}
\end{table}
\begin{table}[ht!]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
\hfill & $\omega_1$ & $\omega_2$ & $\omega_3$ & $\omega_4$ & $\omega_5$
\\
\hline
time
& $0$
& $0$
& $-1$
& $0$
& $0$
\\
\hline
length
& $1$
& $1$
& $2$
& $2$
& $2$
\\
\hline
mass
& $1$
& $2$
& $2$
& $1$
& $2$
\\
\hline
\end{tabular}
\end{center}
\caption{Dimensional scaling weights for low-order conserved currents of the gKdV equation}
\label{Phiweights}
\end{table}
The dimensional scaling formula will now be used to obtain
the conserved currents $\Phi_i=(T_i,X_i)|_{\mathcal E}$ that were missed previously
by the scaling symmetry formula.
These cases are: $i=4,5$; and, $i=1,2$ when $p$ is a critical power.
From the form of the dimensional scaling generators,
the mass scaling transformation is simplest choice to use.
Then the formula becomes
\begin{align*}
T_i = \omega_i\tilde T_i & = ( u Q_i -p\tilde Q^t_{i(\mu)} )|_{\mu=\nu=1} ,
\\
X_i = \omega_i\tilde X_i & = ( u^{p+1} Q_i + uD_x^2Q_i -u_x D_xQ_i +u_{xx} Q_i -p\tilde Q^x_{i(\mu)} )|_{\mu=\nu=1},
\end{align*}
modulo a locally trivial current.
This mass scaling formula yields the conserved density and flux expressions
\begin{align*}
T_1 = \tilde T_1,
\quad
X_1 = \tilde X_1,
\\
T_2 = 2\tilde T_2,
\quad
X_2 = 2\tilde X_2,
\end{align*}
which hold for all powers $p>0$
(including the critical powers $p=2$ and $p=4$, respectively),
and also
\begin{align*}
T_4 = \tilde T_4,
\quad
X_4=\tilde X_4,
\\
T_5 = 2\tilde T_5|,
\quad
X_5 = 2\tilde X_5 .
\end{align*}
\section{Concluding remarks}
\label{remarks}
The main results presented in \secref{mainresults}
provide a broad generalization of Noether's theorem in modern form using multipliers,
yielding a general method which is applicable to all typical PDE systems
arising in physical applications.
In this generalization,
the problem of finding all conservation laws for a given PDE system
becomes an adjoint version of the problem of finding all infinitesimal symmetries of the PDE system.
For any given variational PDE system,
conservation laws arise from variational symmetries,
which are infinitesimal symmetries that satisfy variational conditions
corresponding to invariance of any variational principle for the PDE system.
Noether's theorem shows that the characteristic functions in a variational symmetry
are precisely the component functions in a multiplier.
For any given non-variational PDE system,
the role of symmetries in the derivation of conservation laws is replaced by adjoint-symmetries,
and the variational conditions under which an infinitesimal symmetry is a variational symmetry
are replaced by Helmholtz-type conditions under which an adjoint-symmetry is a multiplier.
Also,
the role of a Lagrangian in constructing a conserved integral from a variational symmetry
is replaced by several different constructions:
an explicit integral formula, an explicit algebraic scaling formula,
and a system of determining equations,
all of which use only a multiplier and the given PDE system itself.
Most importantly,
the completeness of this general method in finding all conservation laws for a given PDE system
is established by working with the system expressed in a solved-form for a set of leading derivatives without restricting it to have a generalized Cauchy-Kovalevskaya form.
This means that the method applies equally well to PDE systems that possess differential identities.
As a consequence,
there is no need to use special methods or anstazes for determining the conservation laws of any given PDE system,
just as there is no necessity to use special methods or ansatzes for finding its symmetries.
The formulation of the general method as a generalization of Noether's theorem
rests on the adjoint relationship between variational symmetries and multipliers,
which originates from the algebraic relationship between symmetries and adjoint-symmetries.
An interesting question is whether this algebraic relationship has a geometrical interpretation.
As will be shown in more detail elsewhere \cite{Anc17b},
adjoint-symmetries indeed can be given a simple geometrical meaning.
In the case of PDE systems comprised of dynamical evolution equations,
$G=\partial_t u -g(t,x,u,\partial_x u,\partial_x^2 u, \ldots, \partial_x^N u)=0$,
an adjoint-symmetry defines a 1-form (or covector field) $Q\mathbf{d}u$ that is invariant under the dynamical flow on $u(t,x)$,
similarly to how a symmetry $P\partial_u$ defines an invariant vector field.
This geometrical statement essentially relies on the number of dependent variables
being the same as the number of equations in the PDE system.
For general PDE systems $G=0$,
it seems necessary to use the well-known procedure \cite{Olv} of embedding the PDE system
into a larger, variational system defined by a Lagrangian $L=G v^{\rm t}$
where $v$ denotes additional dependent variables
which are paired with the equations $G=0$ in the given PDE system.
In this setting,
an adjoint-symmetry defines a symmetry vector field $Q\partial_v$ of the enlarged system,
$G=0$ and $G'{}^*(v)=0$,
where $G'$ is the Fr\'echet derivative of $G$, and $G'{}^*$ is its adjoint.
Then, it is straightforward to show that an adjoint-symmetry is a multiplier
precisely when $Q\partial_v$ is a variational symmetry.
\section*{Acknowledgements}
S.C. Anco is supported by an NSERC research grant.
The referees are thanked for valuable comments which have improved this work.
|
\section{Introduction}
Unless otherwise stated, we shall use small letters such as $x$ to
denote non-negative integers or set elements or functions,
capital letters such as $X$ to denote sets, and calligraphic
letters such as $\mathcal{F}$ to denote \emph{families}
(that is, sets whose members are sets themselves). The set $\{1, 2, \dots\}$ of all positive integers is denoted by $\mathbb{N}$.
For any $m,n \in \mathbb{N}$ with $m < n$, the set $\{i \in \mathbb{N} \colon m \leq i \leq n\}$ is denoted by $[m,n]$. We abbreviate $[1,n]$ to $[n]$. It is to be assumed that arbitrary sets and families are \emph{finite}. We call a set $A$ an \emph{$r$-element set}, or simply an \emph{$r$-set}, if its size $|A|$ is $r$. For a set $X$, the \emph{power set of $X$} (that is, the family of all subsets of $X$) is denoted by $2^X$, and the family of all $r$-element subsets of $X$ is denoted by ${X \choose r}$.
We say that a set $A$ \emph{$t$-intersects} a set $B$ if $A$ and $B$ contain at least $t$ common elements. A family $\mathcal{A}$ of sets is said to be \emph{$t$-intersecting} if every two sets in $\mathcal{A}$ $t$-intersect. A $1$-intersecting family is also simply called an \emph{intersecting family}.
For a family $\mathcal{F}$ and a set $T$, we denote the family $\{F \in \mathcal{F} \colon T \subseteq F\}$ by $\mathcal{F}(T)$.
We call $\mathcal{F}(T)$ a \emph{$t$-star of $\mathcal{F}$} if $|T| = t$.
A $t$-star of a family is the simplest example of a $t$-intersecting subfamily. We denote the size of a largest $t$-star of $\mathcal{F}$ by $l(\mathcal{F},t)$. We denote the set of largest $t$-stars of $\mathcal{F}$ by ${\rm L}(\mathcal{F},t)$. We say that $\mathcal{F}$ has the \emph{$t$-star property} if at least one $t$-star of $\mathcal{F}$ is a largest $t$-intersecting subfamily of $\mathcal{F}$.
One of the most popular endeavours in extremal set theory is that of determining the size or the structure of a largest $t$-intersecting subfamily of a given family $\mathcal{F}$. This originated in \cite{EKR}, which features the classical Erd\H os-Ko-Rado (EKR) Theorem. The EKR Theorem says that, for $1 \leq t \leq r$, there exists an integer $n_0(r,t)$ such that, for every $n \geq n_0(r,t)$, the size of a largest $t$-intersecting subfamily of ${[n] \choose r}$ is ${n-t \choose r-t}$, meaning that ${[n] \choose r}$ has the $t$-star property. It was also shown in \cite{EKR} that the smallest possible value of $n_0(r,1)$ is $2r$, and two of the various proofs of this fact (see \cite{Kat,K,D}) are particularly short and beautiful: Katona's \cite{K}, introducing the elegant cycle method, and Daykin's \cite{D}, using the Kruskal-Katona Theorem \cite{Kr,Ka}. If $n/2 < r < n$, then ${[n] \choose r}$ itself is intersecting.
A sequence of results \cite{EKR,F_t1,W,FF,AK1} culminated in the complete solution of the problem for $t$-intersecting subfamilies of ${[n] \choose r}$. The solution confirmed a conjecture of Frankl
\cite{F_t1} and
particularly tells us that ${[n] \choose r}$ has the $t$-star property if and only if $n \geq (t+1)(r-t+1)$ \cite{F_t1,W}. The same $t$-intersection problem for $2^{[n]}$ was solved by Katona \cite{Kat}.
These are among the most prominent results in extremal set theory.
The EKR Theorem inspired a wealth of results that establish how large a system of sets can be under certain intersection conditions; see \cite{DF,F,F2,HST,HT,Borg7}.
Two families $\mathcal{A}$ and $\mathcal{B}$ are said to be \emph{cross-$t$-intersecting} if each set in $\mathcal{A}$ $t$-intersects each set in $\mathcal{B}$. More generally, $k$ families $\mathcal{A}_1, \dots, \mathcal{A}_k$ are said to be \emph{cross-$t$-intersecting} if for every $i$ and $j$ in $[k]$ with $i \neq j$, each set in $\mathcal{A}_i$ $t$-intersects each set in $\mathcal{A}_j$. Cross-$1$-intersecting families are also simply called \emph{cross-intersecting families}.
For $t$-intersecting subfamilies of a given family $\mathcal{F}$,
the natural question to ask is how large they can be. For
cross-$t$-intersecting families, two natural parameters arise: the
sum and the product of sizes of the cross-$t$-intersecting
families (note that the product of sizes of $k$ families
$\mathcal{A}_1, \dots, \mathcal{A}_k$ is the number of $k$-tuples
$(A_1, \dots, A_k)$ such that $A_i \in \mathcal{A}_i$ for each $i
\in [k]$). It is therefore natural to consider the problem of
maximising the sum or the product of sizes of $k$ cross-$t$-intersecting subfamilies (not necessarily distinct or non-empty) of a given family $\mathcal{F}$. The paper \cite{Borg8} analyses this problem in general, particularly reducing it to the problem of maximising the size of a $t$-intersecting subfamily of $\mathcal{F}$ for $k$ sufficiently large. Solutions have been obtained for various families (see \cite{Borg8}).
Wang and Zhang \cite{WZ} solved the maximum sum problem for an important class of families that particularly includes ${[n] \choose r}$, using a striking combination of the method in \cite{Borg4,Borg3,Borg2,BL2,Borg5} and an important lemma that is found in \cite{AC,CK} and is referred to as the \emph{no-homomorphism lemma}. The solution for ${[n] \choose r}$ with $t=1$ had been obtained by Hilton \cite{H} and is the first result of this kind.
For $2^{[n]}$, the maximum sum problem was solved \cite[Theorems~3.10, 4.1]{Borg8} via the result in \cite{WZ}, and the maximum product problem was settled in \cite{MT2} for the case where $k = 2$ or $n+t$ is even (see \cite[Section~5.2]{Borg8}, which features a conjecture for the case where $k > 2$ and $n+t$ is odd).
In this paper, we address the maximum product problem for the more general setting where each $\mathcal{A}_i$ is a subfamily of a family $\mathcal{F}_i$. This has been considered for a few special families \cite{Pyber,MT,Hirschorn,Borg11,BorgBLMS}, and, as we explain below, in many cases it is enough to solve the problem for $k = 2$ (see Lemma~\ref{prodgenlemma}).
The maximum product problem for ${[n] \choose r}$ was first addressed by Pyber \cite{Pyber}, who proved that, for $r, s, n \in \mathbb{N}$ such that either $r = s \leq n/2$ or $r < s$ and $n \geq 2s + r -2$, if $\mathcal{A} \subseteq {[n] \choose r}$ and $\mathcal{B} \subseteq {[n] \choose s}$ such that $\mathcal{A}$ and $\mathcal{B}$ are cross-intersecting, then $|\mathcal{A}||\mathcal{B}| \leq {n-1 \choose r-1}{n-1 \choose s-1}$. Subsequently, Matsumoto and Tokushige \cite{MT} proved this for $r \leq s \leq n/2$ (see also \cite{Bey}).
For cross-$t$-intersecting subfamilies, we have the following.
\begin{theorem}[\cite{Borg11}] \label{nchooser} For $1 \leq t \leq r \leq s$, there exists an integer $n_0(r,s,t)$ such that, for every $n \geq n_0(r,s,t)$, if $\mathcal{A} \subseteq {[n] \choose r}$, $\mathcal{B} \subseteq {[n] \choose s}$, and $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting, then $|\mathcal{A}||\mathcal{B}| \leq {n-t \choose r-t}{n-t \choose s-t}$, and equality holds if and only if $\mathcal{A} = \{A \in {[n] \choose r} \colon T \subseteq A\}$ and $\mathcal{B} = \{B \in {[n] \choose s} \colon T \subseteq B\}$ for some $T \in {[n] \choose t}$.
\end{theorem}
Hirschorn made a Frankl-type conjecture \cite[Conjecture~4]{Hirschorn} for any $r$, $s$, $t$ and $n$. A value of $n_0(r,s,t)$ that is close to best possible is established in \cite{Borg12}. The special case $r = s$ is treated in \cite{Tok1,Tok2,FLST}, which establish values of $n_0(r,r,t)$ that are also nearly optimal.
Let $c : \mathbb{N}^3 \rightarrow \mathbb{N}$ such that, for $r,s,t \in \mathbb{N}$, $c(r,s,t) = \max \left\{ r{s \choose t}, s{r \choose t} \right\} + 1$ if $t \leq \min\{r,s\}$, and $c(r,s,t) = 1$ otherwise. Clearly, $c(r,s,t) = r{s \choose t}+1$ for $t \leq r \leq s$.
The following is our main result, proved in Section~\ref{proofmain}.
\begin{theorem}\label{main} If $r,s,t \in \mathbb{N}$, $\mathcal{F}$ is a $(\leq r)$-family with $l(\mathcal{F},t) \geq c(r,s,t)l(\mathcal{F},t+1)$, $\mathcal{G}$ is a $(\leq s)$-family with $l(\mathcal{G},t) \geq c(r,s,t)l(\mathcal{G},t+1)$, and $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting families such that $\mathcal{A} \subseteq \mathcal{F}$ and $\mathcal{B} \subseteq \mathcal{G}$, then
$$|\mathcal{A}||\mathcal{B}| \leq l(\mathcal{F},t)l(\mathcal{G},t),$$
and equality holds if and only if $\mathcal{A} = \mathcal{F}(T) \in {\rm L}(\mathcal{F},t)$ and $\mathcal{B} = \mathcal{G}(T) \in {\rm L}(\mathcal{G},t)$ for some $t$-set $T$.
\end{theorem}
As we show in Section~\ref{specialfamilies}, this solves the problem for many natural families with a sufficiently large parameter depending on $r$, $s$ and $t$.
For example, Theorem~\ref{main} yields Theorem~\ref{nchooser} by taking $n$ large enough so that ${n-t \choose r-t} \geq c(r,s,t){n-t-1 \choose r-t-1}$; see Section~\ref{lps}.
For $r, s, t \in \mathbb{N}$, let $\chi(r,s,t)$ be the smallest non-negative real number $a$ such that $|\mathcal{A}||\mathcal{B}| \leq l(\mathcal{F},t)l(\mathcal{G},t)$ for every $\mathcal{A}$, $\mathcal{B}$, $\mathcal{F}$ and $\mathcal{G}$ such that $\mathcal{F}$ is a $(\leq r)$-family with $l(\mathcal{F},t) \geq a \cdot l(\mathcal{F},t+1)$, $\mathcal{G}$ is a $(\leq s)$-family with $l(\mathcal{G},t) \geq a \cdot l(\mathcal{G},t+1)$, $\mathcal{A} \subseteq \mathcal{F}$, $\mathcal{B} \subseteq \mathcal{G}$, and $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting.
\begin{prob} What is the value of $\chi(r,s,t)$?
\end{prob}
By Theorem~\ref{main}, $\chi(r,s,t) \leq c(r,s,t)$.
In Theorem~\ref{main}, the case $\mathcal{F} = \mathcal{G}$ is of particular importance. First of all, it implies that $\mathcal{F}$ has the $t$-star property if $l(\mathcal{F},t) \geq c(r,r,t)l(\mathcal{F},t+1)$.
\begin{theorem} \label{t-intversion} If $1 \leq t \leq r$ and $\mathcal{A}$ is a $t$-intersecting subfamily of a $(\leq r)$-family $\mathcal{F}$ with $l(\mathcal{F},t) \geq c(r,r,t)l(\mathcal{F},t+1)$, then
\[|\mathcal{A}| \leq l(\mathcal{F},t),\]
and equality holds if and only if $\mathcal{A} \in {\rm L}(\mathcal{F},t)$.
\end{theorem}
\textbf{Proof.} Let $\mathcal{G} = \mathcal{F}$ and $\mathcal{B} = \mathcal{A}$. Since $\mathcal{A}$ is $t$-intersecting, $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting. By Theorem~\ref{main}, the result follows.~\hfill{$\Box$} \\
Also note that in Theorem~\ref{main} with $\mathcal{F} = \mathcal{G}$, the bound is attained by taking $\mathcal{A} = \mathcal{B} \in {\rm L}(\mathcal{F},t)$; a generalization of this fact is given by Proposition~\ref{prop1}. As we show in Example~\ref{example1},
for $\mathcal{F} \neq \mathcal{G}$, it may be that the bound is not attained, and we may also have $\mathcal{A}$ and $\mathcal{B}$ for which no $t$-set $T$ satisfies $|\mathcal{A}||\mathcal{B}| \leq |\mathcal{F}(T)||\mathcal{G}(T)|$, no matter how large $\frac{l(\mathcal{F},t)}{l(\mathcal{F},t+1)}$ and $\frac{l(\mathcal{G},t)}{l(\mathcal{G},t+1)}$ are required to be. In view of this, we will now introduce further definitions. We will also generalise Theorem~\ref{main} to a result for $k$ cross-$t$-intersecting families.
\section{The cross-$t$-star property}
If $\mathcal{A}_1, \dots, \mathcal{A}_k$ are cross-$t$-intersecting families, then we say that the tuple $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ is \emph{cross-$t$-intersecting}.
Let $\mathcal{F}_1, \dots, \mathcal{F}_k$ be families. We say that $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ is \emph{below} $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ if $\mathcal{A}_i \subseteq \mathcal{F}_i$ for each $i \in [k]$. We say that $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the
\begin {enumerate}[(a)]
\item \emph{cross-$t$-star property} if $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k l(\mathcal{F}_i,t)$ for each cross-$t$-intersecting tuple $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$.
\item \emph{strict cross-$t$-star property} if, for each cross-$t$-intersecting tuple $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$, $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k l(\mathcal{F}_i,t)$, and the inequality is strict if there exists no $t$-set $T$ such that, for each $i \in [k]$, $\mathcal{A}_i = \mathcal{F}_i(T)$.
\item \emph{strong cross-$t$-star property} if, for some $t$-set $T$, $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k |\mathcal{F}_i(T)|$ for each cross-$t$-intersecting tuple $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$.
\item \emph{extrastrong cross-$t$-star property} if there exists a $t$-set $T$ such that, for each cross-$t$-intersecting tuple $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$, $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k |\mathcal{F}_i(T)|$, and equality holds only if there exists a $t$-set $T'$ such that, for each $i \in [k]$, $\mathcal{A}_i = \mathcal{F}_i(T')$.
\end {enumerate}
Note that each of (b)--(d) implies (a), and (d) implies (a)--(c). As we demonstrate in Example~\ref{example1}, it may be that (b) holds, (c) does not hold, and hence (d) does not hold; clearly, this is the case only if $\prod_{i=1}^k |\mathcal{A}_i| < \prod_{i=1}^k l(\mathcal{F}_i,t)$ for each cross-$t$-intersecting tuple $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$.
\begin{example}\label{example1} \emph{Let $r_1, \dots, r_k, t \in \mathbb{N}$ with $k \geq 2$ and $t < r_1 \leq \dots \leq r_k$. Let $T_1, \dots, T_{k}, A_{1,1}, \dots, A_{1,q_1}, \dots, A_{k,1}, \dots, A_{k,q_k}$ be pairwise disjoint sets such that, for each $i \in [k]$, $|T_i| = t$ and $|A_{i,1}| = \dots = |A_{i,q_i}| = r_i-t$. Let $R_1, \dots, R_{k-1}$ be sets such that $|R_i| = r_i$ for each $i \in [k-1]$, $R_1 \subseteq \dots \subseteq R_{k-1}$, and $R_{k-1} \cap \bigcup_{i=1}^k\bigcup_{j=1}^{q_i} (T_i \cup A_{i,j}) = \emptyset$ (that is, no set $R_m$ intersects a set $T_i \cup A_{i,j}$).
For each $i \in [k-1]$, let $\mathcal{F}_i = \{T_i \cup A_{i,1}, \dots, T_i \cup A_{i,q_i}, R_i\}$. Let $\mathcal{F}_k = \{T \cup A_{k,j} \colon T \in {R_1 \choose t}, j \in [q_k]\}$. For each $i \in [k]$, each set in $\mathcal{F}_i$ is of size $r_i$, and clearly $l(\mathcal{F}_i,t) = q_i$. For every $i,j \in [k]$ with $i < j$, a set $A$ in $\mathcal{F}_i$ $t$-intersects a set $B$ in $\mathcal{F}_j$ if and only if $A = R_i$ and either $j < k$ and $B = R_j$ or $j = k$ and $B \in \mathcal{F}_j$.
Let $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ be a cross-$t$-intersecting tuple below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ such that $\mathcal{A}_1, \dots, \mathcal{A}_k$ are non-empty (so that $\prod_{i=1}^k |\mathcal{A}_i| \neq 0$). Then $\mathcal{A}_i = \{R_i\}$ for each $i \in [k-1]$. Thus $\prod_{i=1}^k |\mathcal{A}_i| \leq |\mathcal{F}_k|$, and equality holds if and only if $(\mathcal{A}_1, \dots, \mathcal{A}_k) = (\{R_1\}, \dots, \{R_{k-1}\}, \mathcal{F}_k)$. Therefore, if $\prod_{i=1}^{k-1}q_i > {r_1 \choose t}$, then $\prod_{i=1}^k |\mathcal{A}_i| < \prod_{i=1}^k l(\mathcal{F}_i,t)$ (since $\prod_{i=1}^k |\mathcal{A}_i| \leq |\mathcal{F}_k| = {r_1 \choose t}q_k$ and $\prod_{i=1}^k l(\mathcal{F}_i,t) = \prod_{i=1}^k q_i$), meaning that $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property. Now let $T$ be a $t$-set such that $\prod_{i=1}^k |\mathcal{F}_i(T)| \neq 0$. Then $T \subseteq R_1$ and $\mathcal{F}_i(T) = \{R_i\}$ for each $i \in [k-1]$. Thus $\prod_{i=1}^k |\mathcal{F}_i(T)| = |\mathcal{F}_k(T)| < \prod_{i=1}^k |\mathcal{A}_i|$ if $(\mathcal{A}_1, \dots, \mathcal{A}_k) = (\{R_1\}, \dots, \{R_{k-1}\}, \mathcal{F}_k)$. Therefore, $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ does not have the strong cross-$t$-star property.}
\end{example}
\begin{rem} \emph{By Example~\ref{example1}, for $1 \leq t < r \leq s$, there is no real number $a$ such that $(\mathcal{F},\mathcal{G})$ has the strong cross-$t$-star property for every $(\leq r)$-family $\mathcal{F}$ with $l(\mathcal{F},t) \geq a \cdot l(\mathcal{F},t+1)$ and every $(\leq s)$-family $\mathcal{G}$ with $l(\mathcal{G},t) \geq a \cdot l(\mathcal{G},t+1)$.}
\end{rem}
Of particular importance is the case $\mathcal{F}_1 = \dots = \mathcal{F}_k$.
\begin{prop}\label{prop1} (i) If $\mathcal{F}_1 = \dots = \mathcal{F}_k$ and $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the cross-$t$-star property, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strong cross-$t$-star property. \\
(ii) If $\mathcal{F}_1 = \dots = \mathcal{F}_k$ and $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{prop}
\textbf{Proof.} Suppose $\mathcal{F}_1 = \dots = \mathcal{F}_k$. Let $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ be a cross-$t$-intersecting tuple below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$. Let $T$ be a $t$-set such that $|\mathcal{F}_1(T)| = l(\mathcal{F}_1,t)$. Since $\mathcal{F}_1 = \dots = \mathcal{F}_k$, $|\mathcal{F}_i(T)| = l(\mathcal{F}_i,t)$ for each $i \in [k]$. If $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the cross-$t$-star property, then $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k |\mathcal{F}_i(T)|$. If $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property and $\prod_{i=1}^k |\mathcal{A}_i| = \prod_{i=1}^k |\mathcal{F}_i(T)|$, then there exists a $t$-set $T'$ such that, for each $i \in [k]$, $\mathcal{A}_i = \mathcal{F}_i(T')$.~\hfill{$\Box$}
\begin{prop}\label{prop2} The tuple $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property if it has the strict cross-$t$-star property and there exists a $t$-set $T$ such that, for each $i \in [k]$, $\mathcal{F}_i(T) \in {\rm L}(\mathcal{F}_i,t)$.
\end{prop}
\textbf{Proof.} Let $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ be a cross-$t$-intersecting tuple below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$. Under the given conditions, $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k l(\mathcal{F}_i,t) = \prod_{i=1}^k |\mathcal{F}_i(T)|$, and $\prod_{i=1}^k |\mathcal{A}_i| = \prod_{i=1}^k l(\mathcal{F}_i,t)$ only if there exists a $t$-set $T'$ such that, for each $i \in [k]$, $\mathcal{A}_i = \mathcal{F}_i(T')$.~\hfill{$\Box$}\\
By Theorem~\ref{main}, $(\mathcal{F},\mathcal{G})$ has the cross-$t$-star property if $l(\mathcal{F},t) \geq c(r,s,t)l(\mathcal{F},t+1)$ and $l(\mathcal{G},t) \geq c(r,s,t)l(\mathcal{G},t+1)$.
The following generalisation of \cite[Lemma~5.1]{Borg8} follows immediately from \cite[Lemma~5.2]{Borg8} and particularly tells us that the cross-$t$-star property is guaranteed for $k$ families if it is guaranteed for every two of them
\begin{lemma} \label{prodgenlemma} If $2 \leq p \leq k$ and $\mathcal{F}_1, \dots, \mathcal{F}_k$ are families such that $(\mathcal{F}_{i_1}, \dots, \mathcal{F}_{i_p})$ has the cross-$t$-star property for each $p$-element subset $\{i_1, \dots, i_p\}$ of $[k]$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the cross-$t$-star property.
\end{lemma}
For example,
Theorem~\ref{main} yields the following generalization.
\begin{theorem}\label{maingen} If $1 \leq t \leq r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i$ is a $(\leq r_i)$-family with $l(\mathcal{F}_i,t) \geq c(r_{k-1},r_k,t)l(\mathcal{F}_i,t+1)$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property.
\end{theorem}
We now start working towards the proofs of Theorems~\ref{main} and \ref{maingen}. Then, in Section~\ref{specialfamilies}, we apply the results above to several important families.
\section{Proof of the main result} \label{proofmain}
If a set $T$ $t$-intersects each set in a family $\mathcal{A}$, then we call $T$ a \emph{$t$-transversal of $\mathcal{A}$}.
\begin{lemma}\label{main lemma} If $T$ is a $t$-transversal of a subfamily $\mathcal{A}$ of a family $\mathcal{F}$, then
\[|\mathcal{A}| \leq {|T| \choose t} l(\mathcal{F},t).\]
\end{lemma}
\textbf{Proof.} Let $\mathcal{T} = {T \choose t}$. Since $|A \cap T| \geq t$ for all $A \in \mathcal{A}$, we have
\begin{align} |\mathcal{A}| &= \left| \bigcup_{I \in \mathcal{T}}\mathcal{A}(I) \right| \leq \sum_{I \in
\mathcal{T}}|\mathcal{A}(I)| \leq \sum_{I \in \mathcal{T}}|\mathcal{F}(I)| \leq \sum_{I \in \mathcal{T}} l(\mathcal{F},t) = |\mathcal{T}| l(\mathcal{F},t), \nonumber
\end{align}
and hence the result.~\hfill{$\Box$}
\begin{lemma} \label{mainlemma2} If $T$ is a $t$-transversal of a subfamily $\mathcal{A}$ of a family $\mathcal{F}$, $X$ is a set of size $t$, $\mathcal{A} \subseteq \mathcal{F}(X)$, and $X \nsubseteq T$, then
\[|\mathcal{A}| \leq |T \backslash X| l(\mathcal{F},t+1).\]
\end{lemma}
\textbf{Proof.} The result is trivial if $\mathcal{A} = \emptyset$. Suppose $\mathcal{A} \neq \emptyset$. For each $A \in \mathcal{A}$, we have
\[t \leq |A \cap T| = |A \cap (T \cap X)| + |A \cap (T \backslash X)| = |T \cap X| + |A \cap (T \backslash X)| \leq t-1 + |A \cap (T \backslash X)|,\]
and hence $|A \cap (T \backslash X)| \geq 1$. Together with $\mathcal{A} \subseteq \mathcal{F}(X)$, this gives us
\begin{align} \mathcal{A} &\subseteq \{F \in \mathcal{F} \colon X \subseteq F, |F \cap (T \backslash X)| \geq 1\} \nonumber \\
&= \{F \in \mathcal{F} \colon X \cup \{y\}\subseteq F \mbox{ for some } y \in T \backslash X\} = \bigcup_{y \in T \backslash X} \mathcal{F}(X \cup \{y\}). \nonumber
\end{align}
Thus $|\mathcal{A}| \leq \sum_{y \in T \backslash X} |\mathcal{F}(X \cup \{y\})| \leq \sum_{y \in T \backslash X} l(\mathcal{F},t+1) = |T \backslash X| l(\mathcal{F},t+1)$.~\hfill{$\Box$}\\
\\
We can now prove Theorem~\ref{main}. We will call a $t$-intersecting family $\mathcal{A}$ \emph{trivial} if the sets in $\mathcal{A}$ have at least $t$ common elements.\\
\\
\textbf{Proof of Theorem~\ref{main}.} Suppose $|F| < t$ for each $F \in \mathcal{F}$. Then ${\rm L}(\mathcal{F},t) = \{\emptyset\}$, and hence $l(\mathcal{F},t) = 0 = l(\mathcal{F},t+1)$. Also, $|F \cap G| < t$ for each $F \in \mathcal{F}$ and each $G \in \mathcal{G}$. Thus, since $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting, one of $\mathcal{A}$ and $\mathcal{B}$ is empty, and hence $|\mathcal{A}||\mathcal{B}| = 0 = l(\mathcal{F},t)l(\mathcal{G},t)$. Similarly, $|\mathcal{A}||\mathcal{B}| = 0 = l(\mathcal{F},t)l(\mathcal{G},t)$ if $|G| < t$ for each $G \in \mathcal{G}$.
Now suppose that each of $\mathcal{F}$ and $\mathcal{G}$ has a set of size at least $t$. Then $r \geq t$, $s \geq t$, $l(\mathcal{F},t) \geq 1$ and $l(\mathcal{G},t) \geq 1$.
If one of $\mathcal{A}$ and $\mathcal{B}$ is empty, then $|\mathcal{A}||\mathcal{B}| = 0 < l(\mathcal{F},t)l(\mathcal{G},t)$.
Suppose $\mathcal{A} \neq \emptyset$ and $\mathcal{B} \neq \emptyset$. Since $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting, each set in $\mathcal{A}$ is a $t$-transversal of $\mathcal{B}$, and each set in $\mathcal{B}$ is a $t$-transversal of $\mathcal{A}$.\medskip
\textit{Case 1: $\mathcal{A}$ is not a trivial $t$-intersecting family, and $\mathcal{B}$ is not a trivial $t$-intersecting family.} Let $D \in \mathcal{B}$. For each $X \in {D \choose t}$, let $\mathcal{A}_X = \mathcal{A}(X)$. Since $|A \cap D| \geq t$ for each $A \in \mathcal{A}$, $\mathcal{A} = \bigcup_{X \in {D \choose t}} \mathcal{A}_X$.
Consider any $X \in {D \choose t}$. Since $\mathcal{B}$ is not a trivial $t$-intersecting family,
there exists $B \in \mathcal{B}$ such that $X \nsubseteq B$. Since $B$ is a $t$-transversal of $\mathcal{A}_X$, $|\mathcal{A}_X| \leq |B \backslash X| l(\mathcal{F},t+1) \leq s \cdot l(\mathcal{F},t+1)$ by Lemma~\ref{mainlemma2}.
Therefore, we have
\begin{equation} |\mathcal{A}| = \left| \bigcup_{X \in {D \choose t}} \mathcal{A}_X \right| \leq \sum_{X \in {D \choose t}} |\mathcal{A}_X| \leq \sum_{X \in {D \choose t}} s l(\mathcal{F},t+1) = s {|D| \choose t} l(\mathcal{F},t+1), \nonumber
\end{equation}
and hence $|\mathcal{A}| \leq s {s \choose t} l(\mathcal{F},t+1)$. By a similar argument, $|\mathcal{B}| \leq r {r \choose t} l(\mathcal{G},t+1)$. It follows that
\begin{equation} |\mathcal{A}||\mathcal{B}| \leq s {s \choose t} l(\mathcal{F},t+1) r {r \choose t} l(\mathcal{G},t+1) \leq r {r \choose t} s {s \choose t} \frac{l(\mathcal{F},t)}{c(r,s,t)} \frac{l(\mathcal{G},t)}{c(r,s,t)} < l(\mathcal{F},t) l(\mathcal{G},t). \nonumber
\end{equation}
\textit{Case 2: $\mathcal{A}$ is a trivial $t$-intersecting family, and $\mathcal{B}$ is not a trivial $t$-intersecting family.} We have $\mathcal{A} \subseteq \mathcal{F}(X)$ for some set $X$ of size $t$. Since $\mathcal{B}$ is not a trivial $t$-intersecting family, there exists $B \in \mathcal{B}$ such that $X \nsubseteq B$. By Lemma~\ref{mainlemma2}, $|\mathcal{A}| \leq |B \backslash X| l(\mathcal{F},t+1) \leq s \cdot l(\mathcal{F},t+1)$. Now let $C \in \mathcal{A}$. By Lemma~\ref{main lemma}, $|\mathcal{B}| \leq {|C| \choose t} l(\mathcal{G},t) \leq {r \choose t} l(\mathcal{G},t)$. Therefore,
\begin{equation} |\mathcal{A}||\mathcal{B}| \leq s \cdot l(\mathcal{F},t+1){r \choose t} l(\mathcal{G},t) \leq s {r \choose t} \frac{l(\mathcal{F},t)}{c(r,s,t)} l(\mathcal{G},t) < l(\mathcal{F},t) l(\mathcal{G},t). \nonumber
\end{equation}
\textit{Case 3: $\mathcal{A}$ is not a trivial $t$-intersecting family, and $\mathcal{B}$ is a trivial $t$-intersecting family.} The result follows by an argument similar to that for Case~2.\medskip
\textit{Case 4: $\mathcal{A}$ and $\mathcal{B}$ are trivial $t$-intersecting families.}
Then $\mathcal{A} \subseteq \mathcal{F}(X)$ for some $t$-set $X$, and $\mathcal{B} \subseteq \mathcal{G}(Y)$ for some $t$-set $Y$. Thus $|\mathcal{A}| \leq l(\mathcal{F},t)$, $|\mathcal{B}| \leq l(\mathcal{G},t)$, and hence $|\mathcal{A}||\mathcal{B}| \leq l(\mathcal{F},t)l(\mathcal{G},t)$. Suppose $|\mathcal{A}||\mathcal{B}| = l(\mathcal{F},t)l(\mathcal{G},t)$. Then $|\mathcal{A}| = l(\mathcal{F},t)$ and $|\mathcal{B}| = l(\mathcal{G},t)$. Therefore, $\mathcal{A} = \mathcal{F}(X) \in {\rm L}(\mathcal{F},t)$ and $\mathcal{B} = \mathcal{G}(Y) \in {\rm L}(\mathcal{G},t)$.
Suppose $X \neq Y$. Then $X \backslash Y \neq \emptyset$ since $|X| = |Y| = t$. Let $x \in X \backslash Y$. Suppose $x \in B$ for all $B \in \mathcal{B}$. Then $\mathcal{B} \subseteq \mathcal{G}(Y \cup \{x\})$, and hence $|\mathcal{B}| \leq l(\mathcal{G},t+1) \leq \frac{l(\mathcal{G},t)}{c(r,s,t)} < l(\mathcal{G},t)$, a contradiction. Thus $x \notin D$ for some $D \in \mathcal{B}$. Thus $X \nsubseteq D$. By Lemma~\ref{mainlemma2}, $|\mathcal{A}| \leq |D \backslash X| l(\mathcal{F},t+1) \leq s \frac{l(\mathcal{F},t)}{c(r,s,t)} < l(\mathcal{F},t)$, a contradiction.
Therefore, $X = Y$.~\hfill{$\Box$}\\
\\
\textbf{Proof of Theorem~\ref{maingen}.} For $k = 2$, the result is given by Theorem~\ref{main}. Consider $k \geq 3$. Let $(\mathcal{A}_1, \dots, \mathcal{A}_k)$ be a cross-$t$-intersecting tuple below $(\mathcal{F}_1, \dots, \mathcal{F}_k)$. Then, for every $i, j \in [k]$ with $i \neq j$, $\mathcal{A}_i$ and $\mathcal{A}_j$ are cross-$t$-intersecting, and, since $r_1 \leq \dots \leq r_k$, we have $c(r_i,r_j,t) \leq c(r_{k-1},r_k,t)$. By Theorem~\ref{main} and Lemma~\ref{prodgenlemma}, $\prod_{i=1}^k |\mathcal{A}_i| \leq \prod_{i=1}^k l(\mathcal{F}_i,t)$. Suppose equality holds.
Suppose $|\mathcal{A}_h| < l(\mathcal{F}_h,t)$ for some $h \in [k]$. By Theorem~\ref{main} and Lemma~\ref{prodgenlemma}, $\prod_{i \in [k] \backslash \{h\}} |\mathcal{A}_i| \leq \prod_{i \in [k] \backslash \{h\}} l(\mathcal{F}_i,t)$. Thus $\prod_{i=1}^k |\mathcal{A}_i| < \prod_{i=1}^k l(\mathcal{F}_i,t)$, a contradiction.
Therefore, $|\mathcal{A}_i| \geq l(\mathcal{F}_i,t)$ for each $i \in [k]$. Since $\prod_{i=1}^k |\mathcal{A}_i| = \prod_{i=1}^k l(\mathcal{F}_i,t)$, $|\mathcal{A}_i| = l(\mathcal{F}_i,t)$ for each $i \in [k]$. For each $i \in [2,k]$, we have $|\mathcal{A}_1||\mathcal{A}_i| = l(\mathcal{F}_1,t)l(\mathcal{F}_i,t)$, and hence, by Theorem~\ref{main}, there exists a $t$-set $T_{1,i}$ such that $\mathcal{A}_1 = \mathcal{F}_1(T_{1,i}) \in {\rm L}(\mathcal{F}_1,t)$ and $\mathcal{A}_i = \mathcal{F}_i(T_{1,i}) \in {\rm L}(\mathcal{F}_i,t)$. By the argument in Case~4 of the proof of Theorem~\ref{main}, $T_{1,i} = T_{1,2}$ for each $i \in [2,k]$.
Thus $\mathcal{A}_i = \mathcal{F}_i(T_{1,2})$ for each $i \in [k]$.~\hfill{$\Box$}
\section{Classes of families} \label{specialfamilies}
In this section, we apply Theorem~\ref{maingen} to important classes of families. Thus, for each family $\mathcal{F}$, we need to obtain an upper bound for $\frac{l(\mathcal{F},t)}{l(\mathcal{F},t+1)}$ and compare it with $c(r,s,t)$.
Much of the work done on the $t$-intersection problem for the families treated here is outlined in \cite{Borg7}. Much less is known about the product cross-$t$-intersection problem because it takes the $t$-intersection problem to a deeper level; most of the main results are outlined in \cite{Borg8}.
We will show that Theorem~\ref{maingen} provides a solution for many of the most natural and mostly studied classes of families. For each class, Theorem~\ref{t-intversion} provides a solution for the $t$-intersection problem.
\subsection{Levels of power sets} \label{lps}
For a family $\mathcal{F}$ and a non-negative integer $r$, the family of all $r$-element sets in $\mathcal{F}$ is called the \emph{$r$-th level of $\mathcal{F}$}. For a set $X$, ${X \choose r}$ is the $r$-th level of $2^X$.
Consider $\mathcal{F} = {[n] \choose p}$ with $1 \leq p \leq n$. Suppose $l(\mathcal{F},t+1) > 0$ for some $t \geq 1$. Then $p \geq t+1$. We have
\begin{equation} \frac{l(\mathcal{F},t)}{l(\mathcal{F},t+1)} =
\frac{ {n-t \choose p-t} }{ {n-t-1 \choose p-t-1} } = \frac{n-t}{p-t}. \label{lps1}
\end{equation}
Therefore, $l(\mathcal{F},t) \geq c(r,s,t)l(\mathcal{F},t+1)$ if $n \geq (p-t)c(r,s,t) + t$.
The following is a generalization of Theorem~\ref{nchooser}.
\begin{theorem} \label{nchoosergen} If $1 \leq t \leq r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i = {[n_i] \choose r_i}$ with $n_i \geq (r_i-t)c(r_{k-1},r_k,t) + t$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{theorem}
\textbf{Proof.} By (\ref{lps1}), for each $i \in [k]$, $l(\mathcal{F}_i,t) \geq c(r_{k-1},r_k,t)l(\mathcal{F}_i,t+1)$ as $n_i \geq (r_i-t)c(r_{k-1},r_k,t) + t$. By Theorem~\ref{maingen}, $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property. Since $\mathcal{F}_i([t]) \in {\rm L}(\mathcal{F}_i,t)$ for each $i \in [k]$, the result follows by Proposition~\ref{prop2}.~\hfill{$\Box$}
\subsection{Families of integer sequences}
For an $r$-element set $X = \{x_1, \dots, x_r\}$
and an integer $m \geq 1$, we define
\[\mathcal{S}_{X,m} = \{\{(x_1,y_1), \dots, (x_r,y_r)\}
\colon y_1, \dots, y_r \in [m]\}.\]
Note that $\mathcal{S}_{X,m}$ is isomorphic to the set $[m]^r$, that is, the set of all sequences $(y_1, \dots, y_r)$ such that $y_i \in [m]$ for each $i \in [r]$. We take $\mathcal{S}_{\emptyset,m}$ to be $\emptyset$. With a slight abuse of notation, for a family $\mathcal{F}$, we define
\[\mathcal{S}_{\mathcal{F},m} = \bigcup_{F \in
\mathcal{F}}\mathcal{S}_{F,m}.\]
The $t$-intersection problem for $\mathcal{S}_{[n],m}$ was solved by Ahlswede and Khachatrian \cite{AK2} and by Frankl and Tokushige \cite{FT2}, and that for $\mathcal{S}_{\mathcal{F},m}$ is solved in \cite{Borg6} for $m$ sufficiently large. The product cross-$t$-intersection problem for $\mathcal{S}_{[n],m}$ was solved by Moon \cite{Moon} for $m \geq t+2$, and by Frankl et al.~\cite{FLST} and Pach and Tardos \cite{PT} for $m \geq t+1$. We solve the problem for $\mathcal{S}_{\mathcal{F},m}$ with $m$ sufficiently large depending only on $t$ and the size of a largest set in $\mathcal{F}$.
\begin{theorem} \label{intseqresult} If $1 \leq t \leq r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i$ is a $(\leq r_i)$-family and $m_i \geq c(r_{k-1},r_k,t)$, then $(\mathcal{S}_{\mathcal{F}_1,m_1}, \dots, \mathcal{S}_{\mathcal{F}_k,m_k})$ has the strict cross-$t$-star property.
\end{theorem}
\begin{lemma} \label{intseqlemma} If $1 \leq t \leq r$ and $\mathcal{F}$ is a $(\leq r)$-family, then $l(\mathcal{S}_{\mathcal{F},m},t) \geq m \cdot l(\mathcal{S}_{\mathcal{F},m},t+1)$.
\end{lemma}
\textbf{Proof.} Suppose $l(\mathcal{S}_{\mathcal{F},m},t+1) > 0$. Then $r \geq t+1$. Let $\mathcal{A}$ be a $(t+1)$-star of $\mathcal{S}_{\mathcal{F},m}$ of size $l(\mathcal{S}_{\mathcal{F},m},t+1)$. Then $\mathcal{A} = \mathcal{S}_{\mathcal{F},m}(Z)$ for some $(t+1)$-element set $Z$. Let $\mathcal{G} = \{F \in \mathcal{F} \colon \mathcal{S}_{F,m}(Z) \neq \emptyset\}$. Let $T \in {Z \choose t}$.
We have
\begin{align} l(\mathcal{S}_{\mathcal{F},m},t+1) &= |\mathcal{S}_{\mathcal{F},m}(Z)| = \sum_{F \in \mathcal{F}} |\mathcal{S}_{F,m}(Z)| = \sum_{F \in \mathcal{G}} |\mathcal{S}_{F,m}(Z)| = \sum_{F \in \mathcal{G}} m^{|F|-t-1} \nonumber \\
&= \frac{1}{m} \sum_{F \in \mathcal{G}} m^{|F|-t} = \frac{1}{m} \sum_{F \in \mathcal{G}} |\mathcal{S}_{F,m}(T)| \leq \frac{1}{m} \sum_{F \in \mathcal{F}} |\mathcal{S}_{F,m}(T)| \nonumber \\
&=\frac{1}{m} |\mathcal{S}_{\mathcal{F},m}(T)| \leq \frac{1}{m}l(\mathcal{S}_{\mathcal{F},m},t), \nonumber
\end{align}
and hence the result.~\hfill{$\Box$}\\
\\
\textbf{Proof of Theorem~\ref{intseqresult}.} For any $i \in [k]$, $l(\mathcal{S}_{\mathcal{F}_i,m_i},t) \geq c(r_{k-1},r_k,t) l(\mathcal{S}_{\mathcal{F}_i,m_i},t+1)$ by Lemma~\ref{intseqlemma} and the given condition $m_i \geq c(r_{k-1},r_k,t)$. The result follows by Theorem~\ref{maingen}.~\hfill{$\Box$}
\begin{theorem} \label{intseqresultcor} If $1 \leq t \leq r$, $\mathcal{F}$ is a $(\leq r)$-family, $m \geq c(r,r,t)$, and $\mathcal{F}_1 = \dots = \mathcal{F}_k = \mathcal{S}_{\mathcal{F},m}$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{theorem}
\textbf{Proof.} The result follows by Theorem~\ref{intseqresult} and Proposition~\ref{prop1}.~\hfill{$\Box$} \\
We make the following conjecture, which is analogous to \cite[Conjecture~2.1]{Borg6}.
\begin{conj} \label{fisconj} For any $t \geq 1$, there
exists a positive integer $m_0(t)$ such that $(\mathcal{S}_{\mathcal{F},m}, \mathcal{S}_{\mathcal{F},m})$ has the strong cross-$t$-star property for any family $\mathcal{F}$ and any $m \geq m_0(t)$.
\end{conj}
We also conjecture that the smallest possible $m_0(t)$ is $t+1$, and that $(\mathcal{S}_{\mathcal{F},m}, \mathcal{S}_{\mathcal{F},m})$ has the extrastrong cross-$t$-star property if $m > t+1$. By Lemma~\ref{prodgenlemma} and Proposition~\ref{prop1}, this would imply a strengthening of Theorem~\ref{intseqresultcor}. The conjecture does not hold for $m < t+1$. Indeed, it can be checked that, if $m \leq t$, $n \geq t+2$, $\mathcal{F} = \{[n]\}$, and $\mathcal{A} = \mathcal{B} = \{A \in \mathcal{S}_{[n],m} \colon |A \cap \{(1,1), \dots, (t+2,1)\}| \geq t+1\}$, then $\mathcal{A}$ and $\mathcal{B}$ are cross-$t$-intersecting subfamilies of $\mathcal{S}_{\mathcal{F},m}$ and $|\mathcal{A}||\mathcal{B}| > (m^{n-t})^2 = (l(\mathcal{S}_{\mathcal{F},m}))^2$.
\subsection{Families of permutations}
For an $r$-set $X = \{x_1, \dots, x_r\}$ and an integer $m \geq 1$, we define $\mathcal{S}_{X,m}^*$ to be the special subfamily of
$\mathcal{S}_{X,m}$ given by
$$\mathcal{S}_{X,m}^* = \left\{\{(x_1, y_1), \dots, (x_r, y_r)\}
\colon y_1, \dots, y_r \mbox{ are distinct elements of } [m] \right\}.$$
Note that $\mathcal{S}_{X,m}^* \neq \emptyset$ if and only if $r \leq m$. The family $\mathcal{S}_{X,m}^*$ can be interpreted as the set of permutations of sets in ${[m] \choose r}$; indeed, a member $\{(x_1, y_1), \dots, (x_r, y_r)\}$ of $\mathcal{S}_{X,m}^*$ corresponds uniquely to the permutation $(y_1, \dots, y_r)$ of the $r$-element subset $\{y_1, \dots, y_r\}$ of $[m]$. We take $\mathcal{S}_{\emptyset,m}$ to be $\emptyset$. With a slight abuse of notation, for a family $\mathcal{F}$, we define
$\mathcal{S}_{\mathcal{F},m}^*$ to be the special subfamily of
$\mathcal{S}_{\mathcal{F},m}$ given by
\[\mathcal{S}_{\mathcal{F},k}^* = \bigcup_{F \in
\mathcal{F}}\mathcal{S}_{F,k}^*.\]
In \cite{DF1}, Deza and Frankl established the $1$-star property of $\mathcal{S}_{[m],m}^*$ and conjectured that $\mathcal{S}_{[m],m}^*$ has the $t$-star property for $m$ sufficiently large depending on $t$.
Ellis, Friedgut and Pilpel \cite{EFP} proved the conjecture together with the product cross-$t$-intersection version. The $t$-intersection problem for $\mathcal{S}_{\mathcal{F},m}^*$ is solved in \cite{Borg6} for $m$ sufficiently large depending only on $t$ and the size of a largest set in $\mathcal{F}$. For this case, we have the following analogous result for the product cross-$t$-intersection problem.
\begin{theorem} \label{permresult} If $1 \leq t \leq r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i$ is a $(\leq r_i)$-family and $m_i \geq c(r_{k-1},r_k,t)+t$, then $(\mathcal{S}_{\mathcal{F}_1,m_1}^*, \dots, \mathcal{S}_{\mathcal{F}_k,m_k}^*)$ has the strict cross-$t$-star property.
\end{theorem}
Similarly to Theorem~\ref{intseqresult}, this follows from the fact that if
$\mathcal{S}_{F,m}^*(Z) \neq \emptyset$ for some $(t+1)$-element $Z$, then $|\mathcal{S}_{F,m}^*(T)| = \frac{(m-t)!}{(m-|F|)!} = (m-t)|\mathcal{S}_{F,m}^*(Z)|$ for any $T \in {Z \choose t}$.
\begin{theorem} \label{permresultcor} If $1 \leq t \leq r$, $\mathcal{F}$ is a $(\leq r)$-family, $m \geq c(r,r,t) + t$, and $\mathcal{F}_1 = \dots = \mathcal{F}_k = \mathcal{S}_{\mathcal{F},m}^*$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{theorem}
\textbf{Proof.} The result follows by Theorem~\ref{permresult} and Proposition~\ref{prop1}.~\hfill{$\Box$} \\
We make the following conjecture, which is analogous to \cite[Conjecture~2.4]{Borg6}.
\begin{conj} \label{fisconj} For any $t \geq 1$, there
exists a positive integer $m_0^*(t)$ such that $(\mathcal{S}_{\mathcal{F},m}^*, \mathcal{S}_{\mathcal{F},m}^*)$ has the extrastrong cross-$t$-star property for any family $\mathcal{F}$ and any $m \geq m_0^*(t)$.
\end{conj}
By Lemma~\ref{prodgenlemma} and Proposition~\ref{prop1}, this would imply a strengthening of Theorem~\ref{permresultcor}.
\subsection{Families of multisets}
A \emph{multiset} is a collection $A$ of objects such that each object possibly appears more than once in $A$. Thus the difference between a multiset and a set is that a multiset may have repetitions of its members. The \emph{multiplicity} of a member $a$ of a multiset $A$ is the number of instances of $a$ in $A$, and is denoted by $m_A(a)$. If $a_1, \dots, a_r$ are the distinct members of a multiset $A$, then we can represent $A$ uniquely by the set $\{(a_i,j) \colon i \in [r], j \in [m_A(a)]\}$, which we denote by $S_A$. Let $M_{n,r}$ denote the set of all multisets $A$ such that the members of $A$ are in $[n]$ and amount to $r$ with repetitions included. An elementary counting result is that \[|M_{n,r}| = {n+r-1 \choose r}.\] Let $\mathcal{M}_{n,r}$ denote the family $\{S_A \colon A \in M_{n,r}\}$. Note that two multisets $A$ and $B$ have exactly $q$ common members (with repetitions included) if and only if $|S_A \cap S_B| = q$.
The $t$-intersection problem for $\mathcal{M}_{n,r}$ was solved by Meagher and Purdy \cite{MP} for $t = 1$, and by F\"{u}redi, Gerbner and Vizer \cite{FGV} for $n \geq 2r-t$. Here we solve the product cross-$t$-intersection problem for $n$ sufficiently large depending on $r$ and $t$.
Consider $\mathcal{F} = \mathcal{M}_{n,p}$. Suppose $l(\mathcal{F},t+1) > 0$ for some $t \geq 1$. Then $p \geq t+1$. We have
\begin{equation} \frac{l(\mathcal{F},t)}{l(\mathcal{F},t+1)} =
\frac{ {n+p-t-1 \choose p-t} }{ {n+p-t-2 \choose p-t-1} } = \frac{n+p-t-1}{p-t}. \label{fm1}
\end{equation}
Therefore, $l(\mathcal{F},t) \geq c(r,s,t)l(\mathcal{F},t+1)$ if $n \geq (p-t)c(r,s,t) - p + t + 1$.
\begin{theorem} \label{fmresult} If $1 \leq t \leq r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i = \mathcal{M}_{n_i,r_i}$ with $n_i \geq (r_i-t)c(r_{k-1},r_k,t) - r_i + t + 1$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{theorem}
\textbf{Proof.} For each $i \in [k]$, $l(\mathcal{F}_i,t) \geq c(r_{k-1},r_k,t)l(\mathcal{F}_i,t+1)$ by (\ref{fm1}) and the given condition $n_i \geq (r_i-t)c(r_{k-1},r_k,t) - r_i + t + 1$. By Theorem~\ref{maingen}, $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property. Let $T = \{(1,i) \colon i \in [t]\}$. Since $\mathcal{F}_i(T) \in {\rm L}(\mathcal{F}_i,t)$ for each $i \in [k]$, the result follows by Proposition~\ref{prop2}.~\hfill{$\Box$}
\subsection{Families of compositions}
If $a_1, a_2, \dots, a_r$ and $n$ are positive integers such that $n = a_1 + a_2 + \dots + a_r$, then the tuple $(a_1, a_2, \dots, a_r)$ is said to be a \emph{composition of $n$} of \emph{length $r$}. Let $C_{n,r}$ denote the set of all compositions of $n$ of length $r$.
An elementary counting result is that \[|C_{n,r}| = {n-1 \choose n-r} = {n-1 \choose r-1}.\] We can represent a composition ${\bf a} = (a_1, \dots, a_r)$ uniquely by the set $\{(1,a_1), \dots, (r,a_r)\}$, which we denote by $S_{\bf a}$. Let $\mathcal{C}_{n,r}$ denote the family $\{S_{\bf a} \colon {\bf a} \in C_{n,r}\}$.
We say that a composition ${\bf a} = (a_1, \dots, a_r)$ \emph{strongly $t$-intersects} a composition ${\bf b} = (b_1, \dots, b_s)$ if there exists a $t$-element subset $T$ of $[\min\{r,s\}]$ such that $a_i = b_i$ for each $i \in T$. Note that ${\bf a}$ \emph{strongly $t$-intersects} ${\bf b}$ if and only if $|S_{\bf a} \cap S_{\bf b}| \geq t$.
Ku and Wong \cite{KW1} solved the $t$-intersection problem for $\mathcal{C}_{n,r}$ with $n$ sufficiently large. In \cite{KW2}, they also proved Theorem~\ref{fcresult} below for sufficiently large values of $n_1, \dots, n_r$.
Consider $\mathcal{F} = \mathcal{C}_{n,p}$ with $t+1 < p \leq n$. It is straightforward that $\mathcal{F}(\{(i,1) \colon i \in [t]\})$ is a largest $t$-star of $\mathcal{C}_{n,p}$. We have
\begin{equation} \frac{l(\mathcal{F},t)}{l(\mathcal{F},t+1)} =
\frac{ {n-t-1 \choose p-t-1} }{ {n-t-2 \choose p-t-2} } = \frac{n-t-1}{p-t-1}. \label{fc1}
\end{equation}
Therefore, $l(\mathcal{F},t) \geq c(r,s,t)l(\mathcal{F},t+1)$ if $n \geq (p-t-1)c(r,s,t) + t + 1$.
\begin{theorem} \label{fcresult} If $2 \leq t+1 < r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i = \mathcal{C}_{n_i,r_i}$ with $n_i \geq (r_i-t-1)c(r_{k-1},r_k,t) + t + 1$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{theorem}
\textbf{Proof.} For each $i \in [k]$, $l(\mathcal{F}_i,t) \geq c(r_{k-1},r_k,t)l(\mathcal{F}_i,t+1)$ by (\ref{fc1}) and the given condition $n_i \geq (r_i-t-1)c(r_{k-1},r_k,t) + t + 1$. By Theorem~\ref{maingen}, $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property. Let $T = \{(i,1) \colon i \in [t]\}$. Since $\mathcal{F}_i(T) \in {\rm L}(\mathcal{F}_i,t)$ for each $i \in [k]$, the result follows by Proposition~\ref{prop2}.~\hfill{$\Box$}
\subsection{Families of set partitions}
If $X_1, X_2, \dots, X_r$ are pairwise disjoint non-empty sets and $X = \bigcup_{i=1}^r X_i$, then the set $\{X_1, X_2, \dots, X_r\}$ is called a \emph{partition of $X$} of \emph{length $r$}, and $X_1, X_2, \dots, X_r$ are called the \emph{parts} of the partition. Let $\mathsf{P}_{n,r}$ denote the family of all partitions of $[n]$ of length $r$, and let $s_{n,r} = |\mathsf{P}_{n,r}|$. Trivially, $s_{n,1} = 1 = s_{n,n}$. An elementary result is that
\begin{equation} s_{n,r} = s_{n-1,r-1} + rs_{n-1,r} \quad \mbox{if } 2 \leq r \leq n-1. \nonumber
\end{equation}
It follows that
\begin{equation} s_{m,r} \leq s_{n,r} \quad \mbox{if } 1 \leq m \leq n. \label{fsp1}
\end{equation}
\begin{lemma} \label{fsplemma1} If $1 < r < n$, then $s_{n,r} \geq \frac{n-1}{r-1}s_{n-1,r-1}$.
\end{lemma}
\textbf{Proof.} Consider any $X \in \mathsf{P}_{n-1,r-1}$. For any $i \in [n-1]$, let $A_i$ be the part of $X$ that contains $i$, and let $X_i$ be the member of $\mathsf{P}_{n,r}$ obtained by replacing $i$ by $n$ in $A_i$, and adding $\{i\}$ as a part; that is, $X_i = (X \backslash \{A_i\}) \cup \{(A_i \backslash \{i\}) \cup \{n\}\} \cup \{\{i\}\}$. For any $Y \in \mathsf{P}_{n,r}$, let $f(X_i,Y) = 1$ if $X_i = Y$, and let $f(X_i,Y) = 0$ if $X_i \neq Y$. If $Y$ has no parts of size $1$, then $f(X_i,Y) = 0$. Suppose that $\{y_1\}, \dots, \{y_p\}$ are the distinct parts of $Y$ of size $1$. Since $n > r$, $p \leq r-1$. Let $B$ be the part of $Y$ that contains $n$. Then $f(X_i,Y) = 1$ if and only if $i \in \{y_1, \dots, y_p\}$ and $X = (Y \backslash \{B, \{i\}\}) \cup \{(B \backslash \{n\}) \cup \{i\}\}$.
Therefore, we have
\begin{align} (n-1) s_{n-1,r-1} &= \sum_{X \in \mathsf{P}_{n-1,r-1}} \sum_{i = 1}^{n-1} 1 = \sum_{X \in \mathsf{P}_{n-1,r-1}} \sum_{i = 1}^{n-1} \sum_{Y \in \mathsf{P}_{n,r}} f(X_i,Y) \nonumber \\
&= \sum_{Y \in \mathsf{P}_{n,r}} \sum_{X \in \mathsf{P}_{n-1,r-1}} \sum_{i = 1}^{n-1} f(X_i,Y) \leq \sum_{Y \in \mathsf{P}_{n,r}} (r-1) = (r-1) s_{n,r}, \nonumber
\end{align}
and hence the result.~\hfill{$\Box$}\\
Erd\H{o}s and Sz\'{e}kely \cite{ES} solved the $t$-intersection problem for $\mathsf{P}_{n,r}$ with $n$ sufficiently large (see \cite{KR} for a related result). Using the results above, we prove the following cross-$t$-intersection result.
\begin{theorem} \label{fspresult} If $2 \leq t+1 < r_1 \leq \dots \leq r_k$ and, for each $i \in [k]$, $\mathcal{F}_i = \mathsf{P}_{n_i,r_i}$ with $n_i \geq (r_i-t-1)c(r_{k-1},r_k,t) + t + 1$, then $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the extrastrong cross-$t$-star property.
\end{theorem}
\begin{lemma} \label{fsplemma2} If $1 \leq t < r \leq n$, then $l(\mathsf{P}_{n,r},t) = s_{n-t,r-t}$.
\end{lemma}
\textbf{Proof.} Let $T = \{\{i\} \colon i \in [t]\}$. We have $l(\mathsf{P}_{n,r},t) \geq |\mathsf{P}_{n,r}(T)| = s_{n-t,r-t}$. Let $\mathcal{A}$ be a largest $t$-star of $\mathsf{P}_{n,r}$. There exist $t$ pairwise disjoint non-empty subsets $X_1, \dots, X_t$ of $[n]$ such that $\mathcal{A} = \mathsf{P}_{n,r}(\{X_1, \dots, X_t\})$. Thus $l(\mathsf{P}_{n,r},t) = |\mathcal{A}| = s_{n',r-t}$, where $n' = n - \sum_{i=1}^t |X_i| \leq n-t$. By (\ref{fsp1}), $l(\mathsf{P}_{n,r},t) \leq s_{n-t,r-t}$. Since $l(\mathsf{P}_{n,r},t) \geq s_{n-t,r-t}$, the result follows.~\hfill{$\Box$}
\begin{lemma} \label{fsplemma3} If $2 \leq t+1 < r < n$, then $l(\mathsf{P}_{n,r},t) \geq \frac{n-t-1}{r-t-1}l(\mathsf{P}_{n,r},t+1)$.
\end{lemma}
\textbf{Proof.}
By Lemma~\ref{fsplemma2}, $l(\mathsf{P}_{n,r},t) = s_{n-t,r-t}$ and $l(\mathsf{P}_{n,r},t+1) = s_{n-t-1,r-t-1}$. Thus, by Lemma~\ref{fsplemma1}, $l(\mathsf{P}_{n,r},t) \geq \frac{n-t-1}{r-t-1} l(\mathsf{P}_{n,r},t+1)$.~\hfill{$\Box$}\\
\\
\textbf{Proof of Theorem~\ref{fspresult}.} For each $i \in [k]$, $l(\mathcal{F}_i,t) \geq c(r_{k-1},r_k,t)l(\mathcal{F}_i,t+1)$ by Lemma~\ref{fsplemma3} and the given condition $n_i \geq (r_i-t-1)c(r_{k-1},r_k,t) + t + 1$. By Theorem~\ref{maingen}, $(\mathcal{F}_1, \dots, \mathcal{F}_k)$ has the strict cross-$t$-star property. Let $T = \{\{i\} \colon i \in [t]\}$. For each $i \in [k]$, we have $|\mathcal{F}_i(T)| = s_{n_i-t,r_i-t}$, and hence $\mathcal{F}_i(T) \in {\rm L}(\mathcal{F}_i,t)$ by Lemma~\ref{fsplemma2}. The result follows by Proposition~\ref{prop2}.~\hfill{$\Box$}
\footnotesize
|
\section{INTRODUCTION}
Over the past decades, robotic technologies and robot systems have been steadily becoming more sophisticated and complex. Programming of an advanced robot system such as the PR2 robot remains a challenging task. There are three long standing issues that still inhibit rapid advancement in robot software development. First, programming for robots usually requires specialist knowledge of the robot platform involved. Robot programmers need to have in-depth understanding of both robot hardware and software. Programmers must have detailed knowledge of the existing software modules written specifically for the platform. In some cases, programmers may even have to learn the specialised programming language used for the robot platform\cite{Baillie2005}. These requirements introduce a steep learning curve for novice programmers and create barriers for expert robotic programmers to transfer their expertises and software to different robot platforms. Second, highly customised software development environments are required before the start of robotic software development. Setting up such a development environment is both technically challenging and time consuming for unexperienced users. The combination of these issues and the difficulty of gaining access to physical robots deters the entry of novice robotic programmers and constraints the growth of the robot software development community. Finally, integration of disparate software modules to produce practically useful robot functions is also nontrivial and time consuming.
The recent development and wide adoption of Robot Operating System (ROS)\cite{Quigley2009} has partly elevated first and third issue by ``standardising" the software module packaging and providing a common data communication framework between different modules. Software implementation of common robotic functions and algorithms can now be shared among different robot platforms. ROS uses C++ and Python as the main programming languages and provides a set of APIs for both languages so that programmers with background of these languages can adopt ROS quickly. The problem, however, is that programming in ROS is still a non-trivial exercise and setting up the ROS development environment (on a system other than Ubuntu Linux) can be both challenging and time consuming. On the other hand, there are development environments \cite{Aldebaran,Jackson2007,Resnick2009} that are easy to setup and use visually based programming techniques to ease the difficulties of programming robot routines. However, these systems are often limited to certain platforms and far less flexible than traditional programming approaches.
In this paper, we introduce a lightweight middleware, Python based Robot \textit{Interactive} Development Environment (PyRIDE), that can considerably ease the burdens of programming advanced robots, specifically the PR2 robot. PyRIDE is designed to address the previously discussed issues in small and practically useful steps: a) provide an abstraction layer to integrate the existing robot software functionalities and expose them through a unified Python robotic programming interface. It is no longer necessary for novice PR2 programmers to learn specific details of ROS and how to communicate with low level PR2 software modules before they can start developing complicate robot skills and behaviours on PR2. Expert PR2/ROS developers can also easily integrate existing ROS modules with PyRIDE through this abstraction layer.
b) provide an interactive shell facility that allows programmers to experiment and test code on robots remotely and interactively in realtime. This also makes client side development software installation redundant. The overhead of accessing and administering a PR2 robot is minimised because users no longer require local user accounts on the robot's computers before they can install and execute their code on the robot\footnote{We assume that the robot resides within an isolated and secure network environment for its software development.}, and c) provide a client-sever mechanism so that programmers can develop remote client programs quickly without dealing with tasks such as data communication between the client and robot.
In the following sections, we will first present the system architecture of PyRIDE and discuss the key design choices we made for the system. We will then use four practical use cases to demonstrate the effectiveness of the system. Finally, we conclude with a brief discussion on the limitations and possible future improvements for PyRIDE. We would like to emphasise that PyRIDE is not limited to PR2/ROS. PyRIDE also operates on NAO robots. It can be easily ported to other robot platforms.
\section{SYSTEM ARCHITECTURE}
The system architecture of PyRIDE is shown in Fig. \ref{sys_arch}. PyRIDE can be viewed as a lightweight middleware that sits on the top of an existing robot software environment. It provides an abstraction layer for the existing software modules and unify their functionalities into a common Python programming interface. The PyRIDE software architecture is agnostic to the underlying robot platform. PyRIDE can be packaged in accordance to the underlying platform requirements. Fig. \ref{sys_arch} shows that PyRIDE for PR2 works under the ROS environment as a standard ROS node. On the NAO robot, PyRIDE is packaged as a loadable runtime NAOQi module that automatically connects to other runtime modules when it is loaded. Since the core of PyRIDE is written in standard portable C++ and C, it can be ported to other robot platforms with relative ease.
\begin{figure}
\centering
\includegraphics[width=1\linewidth]{pyride_arch}
\caption{A System architecture diagram for PyRIDE on ROS/PR2. Note: dashed arrow lines are network communication.}
\label{sys_arch}
\end{figure}
\subsection{System Components}
At the centre of PyRIDE is an embedded Python scripting engine that provides a self contained programming environment for its users. Functional access to the robot system is provided through a single \textbf{PyPR2} Python module interface. PR2 programmers can interact with the robot system by calling appropriate methods in \textbf{PyPR2} and setting up necessary callback functions to receive information from the rest of system and other third-party ROS nodes. There are five key system components attached to the embedded Python engine:
\begin{itemize}
\item \textbf{Interactive shell interface} that provides remote programming access to the embedded Python engine. The interface supports basic telnet network communication protocol so that a programmer can interact with the Python engine remotely using a simple telnet client. This facility provides the same interactivity as a standard interactive Python interpreter (Fig. \ref{telnet_console}).
\\
\item \textbf{Robot function wrapper and Python extension module}. This is the only system component in PyRIDE that is tailored to the specifics of a robot platform. The robot function wrapper such as PR2 function wrapper manages the lower level communication with the PR2 specific subsystems, e.g. joint\_state\_controller. It provides the high level wrapper functions to the underlying functionalities. These wrapper functions are further exposed to the embedded Python engine through \textbf{PyPR2} extension module (see Fig. \ref{telnet_console}). For example, one can call \code{PyPR2.moveArmWithJointPos} method in the remote interactive shell to move a PR2 arm joints to specific joint angles. PyRIDE defines an abstract programming interface so that this platform specific functional wrapper module can be tailored to a different robot platforms.
\\
\item \textbf{Client communication layer}. PyRIDE uses the standard client server model to provide remote user access to the robot platform. This communication subsystem enables data exchanges between the robot and remote client programs, for example, real-time video and audio streaming to a remote client.
All remote client communications are managed through the scripts running on the embedded Python engine.
\\
\item \textbf{Scripts} loaded and executed in the embedded Python engine. These custom Python scripts are developed by programmers for advanced robot behaviours using the services provided by PyRIDE. Each script (or a set of functionally related scripts) can be considered as a small application that runs on the top of PyRIDE middleware. These applications can be loaded and executed dynamically within PyRIDE. Note that, PyRIDE automatically bootstraps a Python startup script \code{py\_main.py}.
\\
\item \textbf{Clients} are the remote user programs for PyRIDE. Remote users can use the data exchange functionalities provided by the client to communicate with the PyRIDE service running on a robot platform. The PyRIDE client component supports multiple operating systems and it offers a set of programming interface so that programmers can customise and extend the clients accordingly. Fig. \ref{ipad_client} shows an example of PyRIDE remote client running on an iPad device. Camera images are streaming to the client in realtime.
\end{itemize}
\begin{figure}
\centering
\includegraphics[width=0.95\linewidth]{pyride_console}
\caption{A telnet session of the PyRIDE interactive interface. It shows the array of PR2 specific functions provided through \textbf{PyPR2} extension module.}
\label{telnet_console}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.95\linewidth]{ipad_client}
\caption{A live screenshot of a PyRIDE remote iPad client. The screenshot shows a testing session for a human detection subsystem running on our PR2 robot.}
\label{ipad_client}
\end{figure}
\subsection{Design Choices}
The design and development of PyRIDE represents a small important step to further improve the programming environments for the PR2 (and other) robots. We attempted to balance simplicity and easy to use with the need of programming flexibility and support for extensive testing. A number of deliberate design decisions were made to achieve our goals.
First, similar to ROS and other systems\cite{Blank2004}, we adopt Python as the main programming language due to its clear syntax and is easy to learn and use. There is already a large population of Python developers. Unlike the existing systems such as ROS that incorporate Python through the standard module extension mechanism, PyRIDE embeds a Python interpreter directly at its core and it becomes a self-contained development environment. Combined with a remote interactive shell service that uses the standard telnet communication protocol\cite{Postel1983}, PyRIDE provides the same interactive programming facility as the standard interactive Python interpreter (Fig. \ref{telnet_console}). This means that a programmer has direct remote access to the functionalities provided by PyRIDE without installing additional software on his/her local development machine, apart from telnet. No user specific account needs to be created and maintained on the PR2 robot's computer in order for a new user to start programming on the robot. It also means that one can develop programs on PyRIDE from (almost any) personal computer with a network connection to the robot\footnote{The interactive shell service can be switched off so to prevent unauthorised access after software development moves to the deployment phase.}. PyRIDE helps a PR2 programmer to focus on his/her work by minimising communication and administrative overhead and removing distractions such as working with the underlying Linux system.
Second, the integration framework with robot function wrapper brings the existing robot functions together under one unified programming interface so that programmers can immediately utilise them without any additional efforts to set up correct access. For example, to obtain the current joint positions of a PR2 robot arm, one can simply call \code{PyPR2.getArmJointPositions} method in PyRIDE Python environment. In contrast, the standard approach in ROS requires many lines of boilerplate code to set up necessary communication proxy channel to the appropriate ROS node and run an event loop in order to retrieve the relevant data. The standard ROS programming approach becomes cumbersome and error prone when a programmer starts to develop complex robot behaviours that involve the coordination and execution of multiple underlying ROS nodes. With the PyRIDE function integration framework, advanced PR2 developers can integrate their existing software modules together under PyRIDE and expose their functionalities to a significantly broader programmer audience.
Third, PyRIDE is designed to automatically search and load a default Python script on startup. Through this bootstrap mechanism, any scripts developed in the embedded Python environment can be immediately deployed. PyRIDE enables seamless transition between software development and software deployment. A typical software development to deployment in PyRIDE starts with the development and testing of code using the interactive mode. Afterwards, the functioning code is packaged into modules and linked to the startup script. PyRIDE will automatically load and execute the code on start up. There is no complicate deployment procedure.
Finally, we developed a client server subsystem in PyRIDE to enable human robot interactions with multiple remote devices and user interfaces. Remote client component, e.g. \code{pyride\_remote}, can be used to develop custom remote client applications to control and monitor robot operation in realtime. Such remote access is integrated and managed directly through scripts running in PyRIDE. This subsystem enhances the accessibility of the robot considerably from both user and programmer perspective.
In summary, we have taken the aforementioned steps to make programming a PR2 robot significantly easier. Each individual features and ideas that we have adopted in PyRIDE is not completely new by itself. Our innovations focused on the seamless integration of these features and ideas to provide practical improvements in terms of robot accessibility, ease of use and easy integration of disparate software components. In the next section, we will use four real life use cases to show that the developer focused improvements in PyRIDE increase the efficiency of robot experimentation, programming and testing. We will show PyRIDE is robust and flexible enough to have highly complex robot software built upon it.
\section{USE CASES}
In this section, we present four real life scenarios to demonstrate the capabilities of PyRIDE described in the previous sections. We use the PR2 robot in our research lab in these examples. Each scenario is shown in the video clips accompanied with this paper. The video is available at \url{http://youtu.be/0DTB62lm8z4}.
\subsection{Experimental Setup and Data Collection}
One of the common tasks in robotic research is setting up an experimental environment for a robot to collect various sensor data, e.g. video stream, for offline analysis and modeling. This experiment preparation and data collection process is usually composed of disparate arrays of operating procedures that are often put together in an `ad-hoc' fashion. For instance, suppose we would like to manually move the PR2 robot to a location, then make the robot arm perform some pre-defined actions, while the robot sensor data is being recorded. To manipulate the robot using the standard approach involves manual execution of various command line tools and custom made scripts under multiple terminal consoles in a specific pre-planned sequence. Minor adjustments in the experiment setup will quickly disrupt the pre-planned experimental sequence, and making small changes to the plan, e.g. modifying action sequence, is both time consuming and error prone.
This experimental setup procedure can be performed in a straight forward fashion with PyRIDE through its interactive Python console:
\begin{enumerate}
\item \code{telnet pr2 27005}.\\Connect to the interactive shell service remotely.
\item \code{PyPR2.tuckBothArms()}.\\Fold the arms and ready for moving the robot.
\item \code{PyPR2.startJoystickControl()}.\\Enable PS3 joystick control of PR2 and move the robot to the experimental location.
\item \code{PyPR2.stopJoystickControl()}.\\Disable the joystick control of PR2.
\item \code{PyPR2.moveTorsoBy(0.02)}.\\Adjust the height of the robot.
\item \code{PyPR2.moveHeadTo(`base\_link',5.0,0.0, 1.2)}.\\Adjust the head position.
\item \code{PyPR2.setToMannequinMode(True)}.\\Set to mannequin mode so that robot arms can be move manually to a pre-defined position.
\item \code{PyPR2.setToMannequinMode(False)}.\\Turn off mannequin mode and restore normal arm control.
\item \code{PyPR2.setTiltLaserPeriodic(0.5,2.0)}.\\Activate the tilt laser scanner.
\item \code{PyPR2.startDataRecording(REC\_CAM| REC\_SCAN|REC\_TF)}.\\Start recording sensory data.
\item Perform some predefined actions.
\item \code{PyPR2.stopDataRecording()}.\\Stop the data recording.
\end{enumerate}
All steps above are demonstrated live in the video clips 1 and 2. It is clear that the interactive programmable control of various robot functions has removed many nuances in operating the PR2 robot. This significantly improves the management of the robot and streamlines experimental setup and data collection processes.
\subsection{An Inverse Kinematics System}
The previous example shows some of the basic functionalities offered by PyRIDE. Highly sophisticated software can be built entirely within PyRIDE. In fact, it has been used to build a new optimised inverse kinematic control system, \textbf{S-PR2}, for the PR2 robot\cite{Taghiabadi2015}. This system is written entirely in Python\footnote{S-PR2 uses scientific computing Python extension \code{numpy} that is written largely in C.}. It uses the joint redundancies to generate optimised smooth joint trajectories in 2D and 3D world spaces. PyRIDE provides the realtime joint information to the S-PR2 system and takes computed low-level joint control commands from S-PR2 to drive the robot arms. Combined with a \textit{PyRIDE remote iPad client}, we were able to develop an integrated system that allows the PR2 to copy human handwriting from the iPad client inputs to a whiteboard. The third video clip shows a demonstration of this handwriting system.
\subsection{Realtime Code Debugging}
Diagnosing and testing robot functions is one of the most challenging tasks in programming robots. Due to the realtime nature of robotic systems, it is difficult to pause and diagnose code related problems while the robot programs are running continuously. A crash or a restart of a subsystem often means that the entire (or a large part of) robot system needs to be reinitialised. PyRIDE offers an alternative approach to the traditional code and test software development cycle. First, PyRIDE is fault tolerate since crashes of individual scripts will not bring down the entire system. All debugging and crash logs are saved in a log file through a realtime logging subsystem in the PyRIDE. The interactive nature of PyRIDE encourages agile code development so that individual functions are coded and tested separately in rapid iterations. With a proper modular design, it is possible to execute, (manually) pause, fix code bugs, (manually) reload the portion of code that contains the fixes and finally continue the execution without completely restarting the application and losing vital system state information. This iterative code and test approach was adopted in the development of S-PR2. Consequently, the development productivity was dramatically improved.
The PyRIDE remote clients also provide additional facilities to support realtime diagnosis of code problems. In particular, a remote client on a mobile device can greatly increase the mobility of programmers when they conduct field tests. Custom robot operational data can be streamed to the mobile device, the programmer is no longer bonded to a computer console that is often difficult to carry during testing. For example, when we test a human detection algorithm on the PR2 robot, the detection results are streamed to an iPad remote client (Fig. \ref{ipad_client}), while we are in front of the robot as the testing subjects (see video clip 4). It would be awkward to use a standard laptop computer in these situations.
\subsection{Third-party ROS Module Integration}
As an easy to use middleware, PyRIDE is not ideal platform for advanced programming tasks that require intensive computational resource or very low level realtime motor control. These tasks should be implemented in native C++ code for computational efficiency. However, one key use of PyRIDE is system integration. That is, advanced PR2/ROS developers can use PyRIDE as an integration tool for the existing ROS modules. Integration of a third-party ROS module with PyRIDE on PR2 involves modifications of PyRIDE PR2 function wrapper (\code{PR2ProxyManager.cpp}\footnote{See PyRIDE on PR2/ROS source code for details}) and the associated Python extension module (\code{PyPR2Module.cpp}). One can follow the standard ROS node initialisation procedure to set up the proxy to the third-party ROS node, e.g. PointHeadAction client, or subscribe to the topics provided by the node under \code{PR2ProxyManager::initWithNodeHandle} method. Public visible methods should be added to \code{PR2ProxyManager} to access the services provided by the ROS node. For example, \code{PR2ProxyManager::pointHeadTo} method uses the PointHeadAction client to control PR2 head direction. These publicly accessible methods can then be wrapped under a Python C extension function in \code{PyPR2Module.cpp} to be accessed from the embedded Python engine in PyRIDE. See \code{PyModule\_PR2PointHeadTo} for example.
For ROS node data subscription, all data messages from a subscribed ROS node publisher are processed through callback functions and passed on to Python via \code{PyPR2Module::invokeCallback} function after data transformation. See \code{PR2ProxyManager::tiltScanDataCB} for example. A simpler alternative for sending data to PyRIDE is publishing the data to the topic of \code{pyride\_pr2/node\_status} as formatted strings. PyRIDE continuously listens on this topic and pass received messages to a Python callback function linked to \code{PyPR2.onNodeStatusUpdate}. Messages can be then decoded in the callback function for further processing. In short, integrating a third-party ROS node with PyRIDE is very much a straight forward process.
\section{DISCUSSION}
We have presented a software development middleware for programming high-level robot skills and behaviours on PR2 robot. The design philosophy of PyRIDE is to take small and practical steps to improve the software development process for the robot. Key features such as the remote interactive Python shell service allow a PR2 programmer to operate, program and perform experiments on PR2 with significantly greater efficiency compared with the standard ROS programming approach. PyRIDE lowers the entry level for robot programming by minimising the steep robot programming learning curve and avoiding cumbersome development environment setup. A Python programmer can directly apply his/her existing skills to program a PR2 without dealing with the specific details of ROS and the underlying operating system. In comparison, a combination of rosh shell\cite{rosh2010} and IPython\cite{Perez2007} offer an interactive programming experience similar to PyRIDE. However, one still needs to install the necessary software on his/her local development machine or have access to a well maintained user account on the robot. One still requires in-depth knowledge of ROS and the PR2 specific ROS modules, e.g. what plugin should be used, what ROS service/message type should be used, before he/she can begin his/her work on PR2. PyRIDE helps PR2 programmers to concentrate on the programming tasks at hands by removing all these unnecessary ``distractions".
PyRIDE works as a smart integration tool for combining the functionalities provided by various ROS modules. A third-party ROS node integrated with PyRIDE can be dynamically managed and utilised from the embedded Python engine with minimum setup. Sophisticate robot behaviours can be developed using relatively simple scripts that aggregate various services provided by the external ROS nodes. Third-party integration with PyRIDE is no more complicated than the standard ROS programming. PyRIDE could be a very useful tool even for the most advanced PR2 developers.
The modular design of PyRIDE means it is agnostic to the underlying robot platform. In addition to PR2/ROS, PyRIDE supports NAO humanoid robot. In fact, PyRIDE was used to develop our NAO robot soccer system that competed in RoboCup 2011\cite{Stanton2011}. PyRIDE can be ported to other robot systems similar to PR2 and NAO with minimal efforts. Furthermore, with the popularity of mobile devices and associated robot toys like Romo\cite{Romotive}, we have ported PyRIDE onto the iOS/Romo platform\footnote{We plan to release PyRIDE on Romo as an app in iOS App Store in near future.} so that even robot hobbyists and school children can develop programs for their devices. PyRIDE on Romo/IOS provides the same coding environment as PyRIDE on PR2/ROS and PyRIDE on NAO. Script modules that are not specifically tied to the underlying platform can be directly shared among the systems running PyRIDE. The common programming approach under PyRIDE means even code that uses platform specific functions can be adapted to different platforms without lengthy rework. In another words, PyRIDE establishes accessible pathway for robot programs and robot software developers to migrate between different platforms. It is our hope that PyRIDE can help to attract more developers to robot software development because it significantly reduces the usual high-bar of expertise needed and make robotics more accessible to a much larger audience of software engineers and programmers.
It is important to highlight that features and ideas embodied within PyRIDE maybe be found in other systems. The innovation of PyRIDE is the seamless and intelligent combination of these features to improve programming efficiency and provide a much better programming experiences for both novice and advanced PR2 developers. Like any software systems, there are several areas where PyRIDE can be further improved. First, PyRIDE only provides a minimal user interface. This could be considered as a drawback for the developers who are comfortable or prefer with Graphical User Interfaces. We will integrate PyRIDE with popular IPython/Jupyter\cite{Ragan-Kelley2014} web interface and investigate possible integration with development environments such as Eclipse. Second, even though PyRIDE has provided a software integration framework for robot developers, the task of integrating disparate software module requires some work. We plan to develop a software integration plugin system and working examples for developers who wish to integrate their systems with PyRIDE. Similarly, the PyRIDE remote client programming interface needs to be further consolidated for third party developers to write their own remote client tools upon PyRIDE remote client component. To this end, we are in the process of documenting and releasing parts of PyRIDE to the public domain so that people from the robot development community can help us improving the current software. In the current initial release phase, we have made PyRIDE on PR2/ROS source code available at \url{http://github.com/uts-magic-lab/pyride_pr2}, PyRIDE on NAO at \url{http://github.com/uts-magic-lab/pyride_nao}. We invite feedbacks and suggestions from the community.
\addtolength{\textheight}{-12cm}
\bibliographystyle{IEEEtran}
|
\section{Introduction}
\subsection{The Boltzmann equation}
We consider a 3-dimensional spatially homogeneous Boltzmann equation, which depicts the density $f_t(v)$ of particles in a gas, moving with velocity $v\in\mathbb{R}^3$ at time $t\geq 0$. The density $f_t(v)$ solves \begin{equation}}\def\eeq{\end{equation}\label{Bol}
\partial_tf_t(v)=\int_{\mathbb{R}^3}dv_{\ast}\int_{\mathbb{S}^2}d\sigma B(|v-v_{\ast}|,\theta)[f_{t}(v^{\prime})f_{t}(v_{\ast}^{\prime})-f_t(v)f_t(v_{\ast})],
\eeq
where
\begin{equation}}\def\eeq{\end{equation}\label{Bol1}
v^{\prime}=\frac{v+v_{\ast}}{2}+\frac{|v-v_{*}|}{2}\sigma,~v_{\ast}^{\prime}
=\frac{v+v_{\ast}}{2}-\frac{|v-v_{\ast}|}{2}\sigma,
\eeq
and $\theta$ is the \emph{deviation angle} defined by $\cos\theta=\frac{v-v_{\ast}}{|v-v_{\ast}|}\cdot\sigma$. The \emph{collision Kernel} $B(|v-v_{\ast}|,\theta)\geq 0$ depends on the type of interaction between particles. It only depends on $|v-v_{\ast}|$ and on the cosine of the deviation angle $\theta$.
Conservations of mass, momentum and kinetic energy hold for reasonable solutions and we may assume without loss of generality that $\int_{\mathbb{R}^3}f_{t}(v)dv=1$ for all $t\ge0$.
\subsection{Assumptions}
We will assume that there is a measurable function $\beta:(0,\pi]\rightarrow\mathbb{R}_+$ such that
\begin{equation}}\def\eeq{\end{equation}\label{con}
\left\{\begin{array}{l} B(|v-v_{\ast}|,\theta)\sin\theta={|v-v_{\ast}|}^{\gamma}\beta(\theta),\\
\exists~0<c_0<c_1,~\forall~\theta\in(0,\pi/2), ~c_0\theta^{-1-\nu}\leq\beta(\theta)\leq c_1\theta^{-1-\nu},\\
\forall~\theta\in[\pi/2,\pi],~\beta(\theta)=0,
\end{array}
\right.\eeq
for some $\nu\in(0,1)$, and $\gamma\in(-1,0)$ satisfying $\gamma+\nu>0$.
\vskip1mm
The last assumption $\beta=0$ on $[\pi/2,\pi]$ is not a restriction and can be obtained by symmetry as noted in the introduction of \cite{MR1765272}.
This assumption corresponds to a classical physical example, inverse power laws interactions:
when particles collide by pairs due to a repulsive force proportional to $1/r^s$ for some $s>2$, assumption \eqref{con} holds with $\gamma=(s-5)/(s-1)$ and $\nu=2/(s-1)$.
Here we will focus on the case of moderately soft potentials, i.e. $s\in (3,5)$.
\subsection{Some notations}
Let us denote by $\mathcal{P}(\mathbb{R}^3)$ the set of probability measures on $\mathbb{R}^3$ and by $Lip(\mathbb{R}^3)$ the set of bounded globally Lipschitz functions $\phi: \mathbb{R}^3 \mapsto \mathbb{R}$. When $f\in\mathcal{P}(\mathbb{R}^3)$ has a density, we also denote this density by $f$ .
For $q>0$, we set
\[\mathcal{P}_q(\mathbb{R}^3)=\{f\in\mathcal{P}(\mathbb{R}^3): m_q(f)<\infty\} \quad\text{with}~~m_q(f):=\int_{\mathbb{R}^3}|v|^q f(dv).\]
We now introduce, for $\theta\in(0,\pi/2)$ and $z\in[0,\infty)$,
\begin{equation}}\def\eeq{\end{equation}\label{nota}
H(\theta)=\int_\theta^{\pi/2}\beta(x)dx\quad\text{and}\quad G(z)=H^{-1}(z).
\eeq
Under \eqref{con}, it is clear that $H$ is a continuous decreasing function valued in $[0,\infty)$, so it has an inverse function $G:[0,\infty)\mapsto(0,\pi/2)$ defined by $G(H(\theta))=\theta$ and $H(G(z))=z$. Furthermore, it is easy to verify that there exist some constants $0<c_2<c_3$ such that for all $z>0$,
\begin{equation}}\def\eeq{\end{equation}\label{con1}
c_2(1+z)^{-1/\nu}\le G(z)\le c_3(1+z)^{-1/\nu},
\eeq
and we know from \cite{MR2398952} that there exists a constant $c_4>0$ such that for all $x,y\in\mathbb{R}_+$,
\begin{equation}}\def\eeq{\end{equation}\label{ineq}
\int_0^{\infty}(G(z/x)-G(z/y))^2dz\le c_4\frac{(x-y)^2}{x+y}.
\eeq
Let us now introduce the Wasserstein distance with quadratic cost on $\mathcal{P}_2(\mathbb{R}^3)$.
For $g,\tilde{g}\in\mathcal{P}_2(\mathbb{R}^3)$, let $\mathcal{H}(g,\tilde{g})$ be the set of probability measures on $\mathbb{R}^3\times\mathbb{R}^3$ with first marginal $g$ and second marginal $\tilde{g}$. We then set
\[ \mathcal{W}_2(g,\tilde{g}) = \inf\left\{\Big(\int_{\mathbb{R}^3\times\mathbb{R}^3}|v-\tilde{v}|^2R(dv,d\tilde{v})\Big)^{1/2},~~R\in\mathcal{H}(g,\tilde{g})\right\}. \]
For more details on this distance, one can see \cite[Chapter 2]{MR1964483}.
\subsection{Weak solutions}
We now introduce a suitable spherical parameterization of \eqref{Bol1} as in \cite{MR1885616}. For each $x \in\mathbb{R}^{3}\setminus{\{0\}}$,
we consider a vector $I(x)\in\mathbb{R}^3$ such that $|I(x)|=|x|$ and $I(x)\perp x$. We also set $J(x)=\frac{x}{|x|}\wedge I(x)$, where $\wedge$ is the vector product. Then the triplet $(\frac{x}{|x|},\frac{I(x)}{|x|},\frac{J(x)}{|x|})$ is an orthonormal basis of $\mathbb{R}^3$. Then for $x,v,v_\ast\in\mathbb{R}^3$,
$\theta\in(0,\pi],~\varphi\in[0,2\pi)$, we set
\begin{equation}}\def\eeq{\end{equation}\label{para}
\left\{\begin{array}{l} \Gamma(x,\varphi):=(\cos\varphi) I(x)+(\sin\varphi) J(x),\\
v^\prime(v,v_\ast,\theta,\varphi):=v-\frac{1-\cos\theta}{2}(v-v_\ast)
+\frac{\sin\theta}{2}\Gamma(v-v_\ast,\varphi),\\
a(v,v_\ast,\theta,\varphi):=v^\prime(v,v_\ast,\theta,\varphi)-v,
\end{array}
\right.\eeq
then we write $\sigma\in\mathbb{S}^2$ as $\sigma=\frac{v-v_*}{|v-v_*|}\cos\theta + \frac{I(v-v_*)}{|v-v_*|}\sin\theta\cos\varphi + \frac{J(v-v_*)}{|v-v_*|}\sin\theta\sin\varphi$, and observe at once that $\Gamma(x,\varphi)$ is orthogonal to $x$ and has the same norm as $x$,
from which it is easy to check that
\begin{equation}}\def\eeq{\end{equation}\label{num}
|a(v,v_*,\theta,\varphi)|=\sqrt{\frac{1-\cos\theta}{2}}|v-v_*|.
\eeq
Let us now give the definition of weak solutions to \eqref{Bol}.
\begin{defi}\sl{}\def\edefi{\end{defi}}\label{dfw}
Assume \eqref{con} is true for some $\nu\in(0,1), \gamma\in(-1,0)$ with $\gamma+\nu>0$.
A measurable family of probability measures $(f_t)_{t\geq 0}$
is called a \textit{weak solution} to \eqref{Bol} if it satisfies the following two conditions:
\begin{itemize}
\item For all $t\geq 0$,
\begin{equation}}\def\eeq{\end{equation}\label{conservation}
\int_{\mathbb{R}^3}vf_t(dv)=\int_{\mathbb{R}^3}vf_0(dv)~~\text{and}~~
\int_{\mathbb{R}^3}|v|^2f_t(dv)=\int_{\mathbb{R}^3}|v|^2f_0(dv)<\infty.
\eeq
\item For any bounded globally Lipschitz function $\phi\in Lip(\mathbb{R}^3)$, any $t\in [0,T]$,
\begin{equation}}\def\eeq{\end{equation}\label{weak}
\int_{\mathbb{R}^3}\phi(v)f_t(dv)=\int_{\mathbb{R}^3}\phi(v)f_0(dv)+
\int_0^t\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\mathcal{A}\phi(v,v_\ast)f_s(dv_\ast)f_s(dv)ds
\eeq
where
\[\mathcal{A}\phi(v,v_\ast)=|v-v_*|^\gamma\int_{0}^{\pi/2}\beta(\theta) d\theta\int_0^{2\pi}
[\phi(v + a(v,v_*,\theta,\varphi))-\phi(v)]d\varphi.\]
\end{itemize}
\edefi
We observe that $|\mathcal{A}\phi(v,v_*)|\le C_\phi|v-v_*|^{1+\gamma}\le C_\phi(1+|v-v_*|^2)$ from $|a(v,v_*,\theta,\varphi)|\le C\theta|v-v_*|$ and $\int_0^{\pi/2}\theta\beta(\theta)d\theta<\infty$, \eqref{weak} is thus well-defined.
\vskip1mm
Let us now recall the well-posedness result of
\eqref{Bol} in \cite[Corollary 2.4]{MR2511651} (more general existence results can be found in
\cite{MR1650006}).
\begin{thm}\def\ethm{\end{thm}}\label{well-posedness} Assume \eqref{con} for some $\gamma\in(-1,0)$,
$\nu\in(0,1)$ with $\gamma+\nu>0$. Let $q\ge2$ such that $q>\gamma^2/(\gamma+\nu)$. Let
$f_0\in\mathcal{P}_q(\mathbb{R}^3)$
with $\int_{\mathbb{R}^3}f_0(v)|\log{f_0(v)}| dv< \infty$ and let
$p\in(3/(3+\gamma),p_0(\gamma,\nu,q))$, where
\begin{equation}\label{pz}
p_0(\gamma,\nu,q)=\frac{q-\gamma}{q(3-\nu)/3-\gamma} \in (3/(3+\gamma),3/(3-\nu)).
\end{equation}
Then
\eqref{Bol} has a unique weak solution
$f\in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)\cap L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$.
\ethm
The explicit value of $p_0(\gamma,\nu,q)$ are not properly stated in \cite[Corollary 2.4]{MR2511651}.
However, following its proof (see the end of Step 3), we see that
$f\in L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$ as soon as $1<p<3/(3-\nu)$ and
$-\gamma(p-1)/(1-p(3-\nu)/3)<q$. This precisely rewrites as
$p\in (1,p_0(\gamma,\nu,q))$.
\subsection{The particle system}
Let us now recall the Nanbu particle system introduced by \cite{nanbu1983interrelations}. It is
the $(\mathbb{R}^3)^N$-valued Markov process with infinitesimal generator $\mathcal{L}_N$ defined as
follows: for any bounded Lipschitz function $\phi: (\mathbb{R}^3)^N\mapsto \mathbb{R}$ and
${\bf v}=(v_1,...,v_N)\in(\mathbb{R}^3)^N$,
$$
\mathcal{L}_{N}\phi({\bf v})=\frac{1}{N}\sum_{i\ne j}\int_{\mathbb{S}^2}[\phi({\bf v}+(v^\prime(v_i,v_j,\sigma)-v_i){\bf e}_i)-\phi({\bf v})]B(|v_i-v_j|, \theta)d\sigma,
$$
where $v{\bf e}_i=(0,...,v,...,0)\in(\mathbb{R}^3)^N$ with $v$ at the $i$-th place for $v\in\mathbb{R}^3$.
\vskip1mm
In other words, the system contains $N$ particles with velocities ${\bf v}=(v_1,...,v_N)$.
Each pair of particles (with velocities $(v_i,v_j)$), interact, for each $\sigma\in\mathbb{S}^2$,
at rate $B(|v_i-v_j|, \theta)/N$. Then one changes the velocity $v_i$ to
$v^\prime(v_i,v_j,\sigma)$ given by \eqref{Bol1} but $v_j$ remains unchanged.
That is, only one particle is changed at each collision.
\vskip1mm
The fact that $\int_0^{\pi}\beta(\theta) d\theta=\infty$ (i.e. $\beta$ is non cutoff) means that
there are infinitely many jumps with a very small deviation angle. It is thus impossible to
simulate it directly. For this reason, we will study a truncated version of Nanbu's particle system
applying a cutoff procedure as \cite{MR3456347}, who were studying the Nanbu system for
\emph{hard potentials} and \emph{Maxwell molecules}, and \cite{Cortez:2015aa}, who were dealing with
the Kac system for \emph{Maxwell molecules}. Our particle system with cutoff
corresponds to the generator $\mathcal{L}_{N,K}$ defined,
for any bounded Lipschitz function $\phi: (\mathbb{R}^3)^N\mapsto \mathbb{R}$
and ${\bf v}=(v_1,...,v_N)\in(\mathbb{R}^3)^N$, by
\begin{align}\label{generator-cut}
\mathcal{L}_{N,K}\phi({\bf v})
&=\frac{1}{N}\sum_{i\ne j}\int_{\mathbb{S}^2}[\phi({\bf v}+(v^\prime(v_i,v_j,\sigma)-v_i){\bf e}_i)-\phi({\bf v})]
B(|v_i-v_j|, \theta)\mathbf{1}_{\{\theta\ge G(K/|v_i-v_j|^\gamma)\}}d\sigma,
\end{align}
with $G$ defined by \eqref{nota}.
\vskip1mm
The generator $\mathcal{L}_{N,K}$ uniquely defines a strong Markov process with values
in $({\mathbb{R}^3})^N$. This comes from the fact that the corresponding jump rate
is finite and constant: for any configuration ${\bf v}=(v_1,...,v_N)\in(\mathbb{R}^3)^N$, it holds
that $N^{-1}\sum_{i\ne j}\int_{\mathbb{S}^2}B(|v_i-v_j|, \theta)\mathbf{1}_{\{\theta\ge G(K/|v_i-v_j|^\gamma)\}}d\sigma
=2\pi (N-1)K$.
Indeed, for any $z\in [0,\infty)$, we have
$\int_{\mathbb{S}^2}B(x, \theta)\mathbf{1}_{\{\theta\ge G(K/x^\gamma)\}}d\sigma=2\pi K$, which is easily checked
recalling that $B(x,\theta)=x^\gamma \beta(\theta)$ and the definition of $G$.
\subsection{Main results}
Now, we give our uniqueness result for the Boltzmann equation.
\begin{thm}\def\ethm{\end{thm}}\label{thm-uniqueness}
Assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ satisfying $\gamma+\nu>0$. Let $q\ge 2$
such that $q>\gamma^2/(\gamma+\nu)$. Assume that $f_0\in\mathcal{P}_q({\mathbb{R}^3})$
with a finite entropy,
i.e. $\int_{\mathbb{R}^3} f_0(v) |\log f_0(v)| dv<\infty$. Let $p\in (3/(3+\gamma),p_0(\gamma,\nu,q))$,
recall \eqref{pz},
and $(f_t)_{t\geq 0}\in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)\cap L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$ be the unique weak solution to \eqref{Bol} given by Theorem \emph{\ref{well-posedness}}. Then for any other weak solution $(\tilde{f}_t)_{t\geq 0}\in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)$ to \eqref{Bol}, we have, for any $t\geq 0$,
\[
\mathcal{W}_2^2(f_t,\tilde{f}_t)\le \mathcal{W}_2^2(f_0,\tilde{f}_0)\exp{\Big(C_{\gamma,p}\int_0^t(1+\|f_s\|_{L^p})ds}\Big).\]
In particular, we have uniqueness for \eqref{Bol} when starting from $f_0$ in the space of all weak solutions in the sense
of Definition \emph{\ref{dfw}}.
\ethm
The novelty of Theorem \ref{thm-uniqueness} is that no regularity at all is assumed
concerning $\tilde f$. In particular, we have uniqueness among all weak solutions,
while in \cite{MR2511651}, uniqueness is proved only in the class of weak solutions
lying in $L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)\cap L^1_{loc}\big([0,\infty),
L^{p}({\mathbb{R}^3})\big)$ for some $p>3/(3+\gamma)$.
\vskip1mm
Next, we write the following conclusion concerning the particle system.
\begin{thm}\def\ethm{\end{thm}}\label{main-result}
Assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ with $\gamma+\nu>0$.
Let $q>6$ such that $q>\gamma^2/(\gamma+\nu)$ and let $f_0\in\mathcal{P}_q({\mathbb{R}^3})$
with a finite entropy.
Let $(f_t)_{t\ge0}$ be the unique weak solution to \eqref{Bol} given by Theorem \emph{\ref{well-posedness}}.
For each $N\ge 1$, $K\in[1,\infty)$, let $(V_t^{i})_{i=1,...,N}$ be the Markov process with generator
$\mathcal{L}_{N,K}$ (see \eqref{generator-cut}) starting from an i.i.d. family $(V_0^{i})_{i=1,...,N}$ of
$f_0$-distributed random variables. We denote
the associated empirical measure by $\mu_{t}^{N,K}=N^{-1}\sum_{i=1}^N \delta_{V_t^{i}}$. Then for all $T>0$,
\[
\sup_{[0, T]}\mathbb{E}[\mathcal{W}_2^2(\mu_{t}^{N,K}, f_t)]\le C_{T,q}\Big(N^{-(1-6/q)(2+2\gamma)/3}+K^{1-2/\nu}+N^{-1/2}\Big).
\]
\ethm
We thus obtain a quantitive rate of chaos for the Nanbu's system with a singular interaction.
To our knowledge, this is the first result in this direction.
However, there is no doubt this rate is not the hoped optimal rate $N^{-1/2}$ like in the
hard potential case \cite{MR3456347}.
\subsection{Known results, strategies and main difficulties}
Let us give a non-exhaustive overview of the known results on the well-posedness of \eqref{Bol} for different potentials. First, the global existence of weak solution for the Boltzmann equation concerning all potentials was concluded by Villani in \cite{MR1650006}, with rather few assumptions on the initial data (finite energy and entropy),
using some compactness methods.
However, the uniqueness results are less well-understood. For \emph{hard potentials} ($\gamma\in(0,1)$) with \emph{angular cutoff} ($\int_0^\pi \beta(\theta) d\theta<\infty$), there are some optimal results obtained by Mischler-Wennberg \cite{MR1697562}, where they gave the existence of a unique weak $L^1$ solution to \eqref{Bol} with the minimal assumption that $\int_{\mathbb{R}^3}(1+|v|^2)f_0(v)dv<\infty$.
This was extended to weak measure solutions by Lu-Mouhot \cite{MR2871802}.
For the difficult case \emph{without} angular cutoff, the first uniqueness result was obtained by Tanaka \cite{MR512334} concerning \emph{Maxwell molecules} ($\gamma=0$).
See also Toscani-Villani \cite{MR1675367}, who proved uniqueness
for Maxwell molecules imposing that $\int_0^\pi \theta \beta(\theta) d\theta<\infty$
and that $\int_{\mathbb{R}^3}(1+|v|^2)f_0(dv)<\infty$.
Subsequently, Desvillettes-Mouhot \cite{MR2525118} (relying on a weighted $W_1^1$
space) and Fournier-Mouhot \cite{MR2511651} (using the Wasserstein distance ${\cal W}_1$) successively gave the uniqueness and stability for both \emph{hard potentials} ($\gamma\in(0,1]$) and \emph{moderately soft potentials} ($\gamma \in (-1,0)$ and $\nu \in (0,1)$) under different assumptions on initial data. For \emph{moderately soft potentials}, the result in \cite{MR2511651} is much better since they use less assumptions on the initial condition than \cite{MR2525118}. Finally, let us mention another work \cite{MR2398952}, where Fournier-Gu\'{e}rin proved a local (in time) uniqueness result with $f_0\in L^p(\mathbb{R}^3)$ for some $p>3/(3+\gamma)$ for the \emph{very soft potentials} ($\gamma \in (-3,0)$ and $\nu \in (0,2)$).
\vskip1mm
In this paper (Theorem \ref{thm-uniqueness}), we obtain a better uniqueness result in the case of a collision kernel without angular cutoff when $\gamma\in(-1,0)$ and $\nu\in(0,1-\gamma)$, that is, the uniqueness holds in the class of all measure solutions in $L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)$. This is very important when studying particle
systems. For example, a convergence result without rate would be almost immediate from our uniqueness: the tightness of the empirical measure of the particle system is not very difficult, as well as the fact that any limit point is a weak solution to \eqref{Bol}.
Since such a weak solution is unique by Theorem \ref{thm-uniqueness}, the convergence follows. Such a conclusion would be very difficult to obtain when using the uniqueness
proved in \cite{MR2511651}, because one would need to check that any limit point
of the empirical measure belongs to $L^1_{loc}([0,\infty,L^p(\mathbb{R}^3))$
for some $p>3/(3+\gamma)$,
which seems very difficult.
\vskip1mm
In order to extend the uniqueness result for all measure solutions, extra difficulty
is inevitable and the methods of \cite{MR2398952,MR2511651} will \emph{not} work.
However, Fournier-Hauray \cite{Fournier:2015aa} provide some ideas to overcome this,
in the simpler case of the Laudau equation for moderately soft potentials.
Here we follow these ideas, which rely on coupling methods.
Consider two weak solutions $f$ and $\tilde f$ in
$L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)$ to \eqref{Bol}, with possibly two
different initial conditions and assume that $f$ is {\it strong}, so that it belongs to
$L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$.
First, we associate to the weak solution $\tilde f$
a {\it weak} solution $(X_t)_{t\geq 0}$ to some Poisson-driven SDE.
This uses a smoothing procedure as in
\cite{MR2375067,Fournier:2015aa}, but the situation is consequently
more complicated because we deal with jump processes.
Next, we try to associate to the strong solution $\tilde f$
a strong solution $(W_t)_{t\geq 0}$ to another SDE (driven by the same Poisson measure),
as \cite{Fournier:2015aa} did. But we did not manage to do this properly and we had
to use a truncation procedure which though complicates our computation.
Then, roughly, we estimate ${\cal W}_2^2(f_t,\tilde f_t)$ by computing
$\mathbb{E}[|X_t-W_t|^2]$ as precisely as possible.
\vskip1mm
The terminology \emph{propagation of chaos}, which is equivalent to the convergence of the empirical measure of a particle system to the solution to a nonlinear equation, was first formulated by Kac \cite{MR0084985}. He was studying the convergence of a {\it toy} particle system as a step to the rigorous derivation of the Boltzmann equation.
Afterwards, McKean \cite{MR0224348} and Gr\"{u}nbaum \cite{MR0334788} extended Kac's ideas to study the chaos property for different models with bounded collision kernels. Sznitman \cite{MR753814} then showed the chaos property (without rate) for the \emph{hard spheres} ($\gamma=1$ and $\nu=0$). Following Tanaka's probabilistic interpretation for the Boltzmann equation with \emph{Maxwell molecules}, Graham-M\'{e}l\'{e}ard \cite{MR1428502} were the first to give a rate of chaos for \eqref{Bol}, concerning both Kac and Nanbu models, for Maxwell molecules with cutoff ($\gamma=0$ and $\int_0^\pi \beta(\theta)d\theta<\infty$), using the total variation distance. Recently, some important progresses have
been made. First, Mischler-Mouhot \cite{MR3069113} obtained a uniform (in time) rate of convergence of Kac's particle system of order $N^{-\epsilon}$ (for Maxwell molecules without cutoff) and $(\log N)^{-\epsilon}$ (for hard spheres, i.e. $\gamma=1$ and $\nu=0$), with some
small $\epsilon>0$.
This result, entirely relying on analytic methods, is noticeable, although the rates
are clearly not sharp.
Then, Fournier-Mischler \cite{MR3456347} proved the propagation of chaos at rate
$N^{-1/4}$ for the Nanbu system and for hard potentials without cutoff
($\gamma \in [0,1]$ and $\nu \in (0,1)$).
Finally, as mentioned in Section 1.5, Cortez-Fontbona \cite{Cortez:2015aa} used
two coupling techniques for Kac's binary interaction system and obtained
a uniform in time estimate for the Boltzmann equation with \emph{Maxwell molecules}
($\gamma=0$) under some suitable moments assumptions on the initial datum.
Let us mention that the time-uniformity uses the recent nice results of
Rousset \cite{Rousset:2014aa}.
\vskip1mm
In this paper (Theorem \ref{main-result}), we obtain, to our knowledge, the first chaos result (with rate)
for soft potentials (which are, of course, more difficult), but it is a bit
unsatisfying: (1) we cannot study Kac's system (which is physically more reasonnable
than Nanbu's system) because it is not readily to exhibit a suitable coupling;
(2) our consideration is merely for $\gamma\in(-1,0)$, since some basic estimates
in Section 2 do not hold any more if $\gamma \le -1$;
(3) our rate is not sharp.
However, since the interaction is singular, it seems hopeless to get a perfect result.
\vskip1mm
In terms of the propagation of chaos with a singular interaction, there are only very few results. Hauray-Jabin \cite{MR3377068} considered a deterministic system of particles interacting through a force of the type $1/|x|^\alpha$ with $\alpha< 1$,
in dimension $d \ge 3$, and proved the mean field limit and the propagation of chaos
to the Vlasov equation. Also, Fournier-Hauray-Mischler \cite{MR3254330} proved the convergence of the vortex model to the 2D Navier-Stokes equation with a singular Biot-Savart kernel using some entropy dissipation technique. Following the method of \cite{MR3254330}, Godinho-Qui\~{n}inao \cite{MR3365970} proved the propagation of chaos of some particle system to the 2D subcritical Keller-Segel equation.
Recently, Fournier-Hauray \cite{Fournier:2015aa} proved propagation of chaos for the Landau equation with a singular interaction ($\gamma\in(-2,0)$). Actually, they gave a quantitative rate of chaos when $\gamma\in(-1,0)$, while the convergence without rate was checked
when $\gamma\in(-2,0)$ by the entropy dissipation technique.
\vskip1mm
Roughly speaking, to prove our propagation of chaos result,
we consider an approximate version of our stability principle,
with a discrete $L^p$ norm as in \cite{Fournier:2015aa}. Here, we list the main
difficulties: The trajectory of a typical particle related to the Boltzmann
equation is a jump process so that all the continuity arguments used in
\cite{Fournier:2015aa} have to be changed. In particular, a detailed
study of small and large jumps is required.
Also, the solution to the Landau equation lies in $L^1_{loc}\big([0,\infty),L^2(\mathbb{R}^3)\big)$,
while the one of the Boltzmann equation lies in $L^1_{loc}\big([0,\infty),L^p(\mathbb{R}^3)\big)$
for some $p$ smaller than $2$. This causes a few difficulties in Section \ref{ttt},
because working in $L^p$ is slightly more complicated.
\subsection{Arrangement of the paper and final notations}
In Section 2, we give some basic estimates. In Section 3, we establish the strong/weak stability principle for \eqref{Bol}. In Section 4, we construct the suitable coupling.
In Section 5, we bound the $L^p$ norm of an empirical measure in terms of $L^p$ norm of the weak solution.
Finally, in Section 6, we prove the convergence of the particle system.
\vskip 1mm
In the sequel, $C$ stands for a positive constant whose value may change from line to line.
When necessary, we will indicate in subscript the parameters it depends on.
\vskip1mm
In the whole paper, we consider two probability spaces by Tanaka's idea for the probabilistic interpretation of the Boltzmann equation in Maxwell molecules case: the first space is the abstract space $(\Omega, \mathcal{F}, \mathbb{P})$ and the second is $([0,1],\mathcal{B}([0,1]),d\alpha)$. A stochastic process defined on the latter space is called an $\alpha$-processes and we denote the expectation on $[0,1]$ by $\mathbb{E}_\alpha$ and the laws by $\mathcal{L}_\alpha$.
\section{Preliminaries}
Above all, let us recall that for $\gamma\in(-1,0)$, $p>3/(3+\gamma)$ and
$f\in\mathcal{P}(\mathbb{R}^3)\cap L^p(\mathbb{R}^3)$, it holds that
\begin{align}\label{norm-inequality}
\sup_{v\in\mathbb{R}^3}\int_{\mathbb{R}^3}|v-v_*|^\gamma f(dv_*)
&\le \sup_{v\in\mathbb{R}^3}\int_{|v-v_*|\le1}|v-v_*|^\gamma f(dv_*) + \sup_{v\in\mathbb{R}^3}\int_{|v-v_*|\ge1}|v-v_*|^\gamma f(dv_*)\nonumber\\
&\le 1+ C_{\gamma,p}\|f\|_{L^{p}(\mathbb{R}^3)},
\end{align}
where $C_{\gamma,p}=\sup_{v\in\mathbb{R}^3}[\int_{|v-v_*|\le1}|v-v_*|^{p\gamma/(p-1)}dv_*]^{(p-1)/p}=[\int_{|v_*|\le1}|v_*|^{p\gamma/(p-1)}dv_*]^{(p-1)/p}<\infty$, since $p>3/(3+\gamma)$ by assumption.
\vskip1mm
Let us now classically rewrite the collision operator by making disappear the velocity-dependence $|v-v_*|^{\gamma}$ in the \emph{rate} using a substitution.
\begin{lem}\def\elem{\end{lem}}\label{llxx}
We assume \eqref{con} and recall \eqref{nota} and \eqref{para}. For $z\in [0,\infty)$, $\varphi\in[0,2\pi)$, $v,v_*\in\mathbb{R}^3$ and $K\in[1,\infty)$, we define
\begin{equation}}\def\eeq{\end{equation}\label{rewrite}
c(v,v_*,z,\varphi):=a[v,v_*,G(z/|v-v_*|^{\gamma}),\varphi]~~~\text{and} ~~~c_K(v,v_*,z,\varphi):=c(v,v_*,z,\varphi)\mathbf{1}_{\{z\le K\}}.
\eeq
For any $\phi\in Lip(\mathbb{R}^3)$, any $v,v_*\in\mathbb{R}$,
\begin{equation}}\def\eeq{\end{equation}\label{opera}
\mathcal{A}\phi(v,v_*)=\int_0^\infty dz\int_0^{2\pi}d\varphi [\phi(v+c(v,v_*,z,\varphi))-\phi(v)].
\eeq
For any $N\ge 1$, $K\in[1,\infty)$, ${\bf v}=(v_1,...,v_N)\in(\mathbb{R}^3)^N$, any bounded measurable $\phi:
(\mathbb{R}^3)^N\mapsto\mathbb{R}$,
\begin{equation}}\def\eeq{\end{equation}\label{rewpartopera}
\mathcal{L}_{N,K}\phi({\bf v})=\frac{1}{N}\sum_{i\ne j}\int_0^\infty
dz\int_0^{2\pi}d\varphi[\phi({\bf v}+c_K(v_i,v_j,z,\varphi){\bf e}_i)-\phi({\bf v})].
\eeq
\elem
This lemma is stated in \cite[Lemma 2.2]{MR3456347} when $\gamma\in[0,1]$, but the proof does not use this fact:
it actually holds true for any $\gamma\in\mathbb{R}$.
Next, let us recall Lemma 2.3 in \cite{MR3456347} which is an accurate version of Tanaka's trick in \cite{MR512334}. Here, we adopt the notation \eqref{para}.
\begin{lem}\def\elem{\end{lem}}\label{Tanaka}
There exists some
measurable function $\varphi_0: \mathbb{R}^3\times\mathbb{R}^3\mapsto[0,2\pi)$
such that for all $X, Y\in\mathbb{R}^3$, all $\varphi\in[0,2\pi)$,
\begin{align*}
|\Gamma(X,\varphi) - \Gamma(Y,\varphi+\varphi_0(X,Y)|\le|X-Y|.
\end{align*}
\elem
The rest of the section is an adaption of \cite[Section 3]{MR3456347},
which assumes that $\gamma\in[0,1]$, to the case where $\gamma\in(-1,0)$.
When compared with \cite{MR2398952}, what is new is that in the inequalities \eqref{ee2} and \eqref{ee3} below,
only $|v-v_*|^\gamma$ appears (while in \cite{MR2398952}, there is $|v-v_*|^\gamma+|\tilde v-\tilde v_*|^\gamma$).
This is very useful to get a strong/weak stability estimate:
we will be able to use the regularity of only one of the two solutions to be compared.
Let us mention that it seems impossible to extend our ideas to the more singular case where $\gamma \leq -1$.
\begin{lem}\def\elem{\end{lem}}\label{estimategeneral}
There is a constant $C$ such that for any $v, v_*,\tilde{v}, \tilde{v}_*\in\mathbb{R}^3$, any $K\geq 1$,
\begin{align}
\int_0^{\infty}\int_0^{2\pi}|c(v,v_*,z,\varphi)&-c(\tilde{v}, \tilde{v}_*,z,\varphi+\varphi_0(v-v_*, \tilde{v}-\tilde{v}_*))|^2d\varphi dz\label{ee2}\\
\leq& C(|v-\tilde{v}| ^2 + |v_*-\tilde{v}_*|^2) |v-v_*|^\gamma,\notag\\
\int_0^{\infty}\int_0^{2\pi}\Big(|v+c(v,v_*,z,\varphi)-&\tilde{v}-c_K(\tilde{v}, \tilde{v}_*,z,\varphi+\varphi_0(v-v_*, \tilde{v}-\tilde{v}_*))|^2-|v-\tilde{v}|^2 \Big)d\varphi dz\label{ee3}\\
\leq& C (|v-\tilde{v}|^2+|v_*-\tilde{v}_*|^2 )|v-v_*|^\gamma + C |v-v_*|^{2+2\gamma/\nu}K^{1-2/\nu} .\notag\\
\int_0^\infty \int_0^{2\pi}|c_K(v,v_*,z,\varphi)|^2 d\varphi dz \leq &C|v-v_*|^{\gamma+2}, \quad
\int_0^\infty \left | \int_0^{2\pi} c_K(v,v_*,z,\varphi) d\varphi\right | dz \le C|v-v_*|^{\gamma+1}\label{ee4}\\
\int_0^\infty \int_0^{2\pi}|c(v,v_*,z,\varphi)|^2 d\varphi dz \leq &C|v-v_*|^{\gamma+2}, \quad
\int_0^\infty \left | \int_0^{2\pi} c(v,v_*,z,\varphi) d\varphi\right | dz \le C|v-v_*|^{\gamma+1}\label{ee5}
\end{align}
\elem
\begin{proof}}\def\epf{\end{proof}
For $x>0$, we set $\Phi_K(x)= \pi\int_0^{K}(1-\cos G(z/x^\gamma))dz$ and
$\Psi_K(x)= \pi\int_{K}^\infty(1-\cos G(z/x^\gamma))dz$.
We introduce the shortened notation
$x=|v-v_*|$, $\tilde x=|\tilde v-\tilde v_*|$, $\varphi_0=\varphi_0(v-v_*, \tilde{v}-\tilde{v}_*)$,
$c=c(v,v_*,z,\varphi)$, $c_K=c_K(v,v_*,z,\varphi)=c\mathbf{1}_{\{z\leq K\}}$,
$\tilde c=c(\tilde{v}, \tilde{v}_*,z,\varphi+\varphi_0)$ and
$\tilde c_K=c_K(\tilde{v}, \tilde{v}_*,z,\varphi+\varphi_0)=\tilde c\mathbf{1}_{\{z\le K\}}$.
\vskip1mm
{\it Step 1.} We first verify that $\Phi_K(x) \leq C x^\gamma$ and that $|\Phi_K(x)-\Phi_K(\tilde x)|\leq
C |x^\gamma-\tilde x^\gamma|$. First, we immediately see that $\Phi_K(x)\leq \pi \int_0^\infty G^2(z/x^\gamma)dz
=x^\gamma \pi \int_0^\infty G^2(z)dz$ which implies the first point (recall \eqref{con1}).
To check the second point, it suffices to verify that $F_K(x)=\int_0^K(1-\cos G(z/x))dz$
has a bounded derivative (uniformly in $K\geq 1$). But
we have $F_K(x)=x \int_0^{K/x}(1-\cos G(z))dz$ so that
\begin{align*}
|F_K^\prime(x)|
\le \int_0^\infty (1-\cos G(z)) dz+ x (K/x^2)(1-\cos G(K/x))
\le C+ (K/x) G^2(K/x),
\end{align*}
which is uniformly bounded by \eqref{con1}.
\vskip1mm
{\it Step 2.} Proceeding as in the proof of \cite[Lemma 3.1]{MR3456347}, we see that
$\int_0^\infty \int_0^{2\pi}|c_K|^2 d\varphi dz=x^2\Phi_K(x)$, which is bounded by $Cx^{\gamma+2}$
by Step 1. Also, recalling \eqref{para} and \eqref{rewrite}, using that $\int_0^{2\pi} \Gamma(X,\varphi)d\varphi=0$,
we see that $\int_0^{2\pi} c_K d\varphi= -\pi (v-v_*) (1-\cos G(z/x^\gamma))$,
whence $\int_0^\infty |\int_0^{2\pi} c_K d\varphi |dz=x \Phi_K(x) \leq C x^{\gamma+1}$ by Step 1.
All this proves \eqref{ee4}, from which \eqref{ee5} follows by letting $K$ increase to infinity.
\vskip1mm
{\it Step 3.}
Let us denote by $I_K=\int_0^{K}\int_0^{2\pi} |c - \tilde c|^2 d\varphi dz $, by
$J_K=\int_0^K\int_0^{2\pi} (|v+c-\tilde v - \tilde c|^2-|v-\tilde v|^2) d\varphi dz$
and by $L_K=\int_K^\infty\int_0^{2\pi} (|v+c-\tilde v|^2-|v-\tilde v|^2) d\varphi dz$.
Proceeding exactly as in the proof of \cite[Lemma 3.1]{MR3456347}, we see that
$J_K \leq A_1^K+A_2^K$ and $L_K \leq A^K_3$, where
\begin{align*}
&A_1^K=2 x \tilde x \int_0^{K}\Big(G(z/x^\gamma) - G(z/\tilde{x}^\gamma)\Big)^2 dz,\\
&A_2^K=\big[|v-\tilde{v}|+|v_*- \tilde{v}_*|\big]|(v-v_*)\Phi_K(x)-(\tilde{v} - \tilde{v}_*)\Phi_K(
\tilde{x})|,\\
&A_3^K=(x^2+2|v-\tilde{v}| x)\Psi_K(x).
\end{align*}
Also, $I_K=J_K -2 (v-\tilde v) \cdot \int_0^{K}\int_0^{2\pi} (c - \tilde c) d\varphi dz $ and, as seen in
the proof of \cite[Lemma 3.1]{MR3456347}, $\int_0^K\int_0^{2\pi} c d\varphi dz = -(v-v_*)\Phi_K(x)$, so that
$I_K \leq J_K+A_4^K$ with
$$
A_4^K = 2 |v-\tilde v| |(v-v_*)\Phi_K(x)-(\tilde v-\tilde v_*)\Phi_K(\tilde x) |.
$$
First, we immediately deduce from \eqref{ineq} that
\begin{align*}
A_1^K&\le 2c_4 x \tilde{x} \frac{(x^\gamma - \tilde{x}^\gamma)^2}{x^\gamma + \tilde{x}^\gamma}
\le 2 c_4 (x-\tilde x)^2 \min{(x^\gamma, \tilde{x}^\gamma)}
\leq C (|v-\tilde{v}| ^2 + |v_*-\tilde{v}_*|^2) |v-v_*|^\gamma.
\end{align*}
For the second inequality, we used that $|x^{\gamma} - \tilde x^{\gamma}|\le|x^{-1}-\tilde x^{-1}|(x\land \tilde x)^{1+\gamma}$
(because $\gamma \in (-1,0)$) so that
\begin{align*}
x\tilde x\frac{|x^\gamma - \tilde x^\gamma|^2}{x^\gamma+\tilde x^\gamma}
\le (x\tilde x)^{1+|\gamma|}\frac{|x^{-1}-\tilde x^{-1}|^2 (x\land \tilde x)^{2\gamma+2}}{x^{|\gamma|}+\tilde x^{|\gamma|}}
\le (x\tilde x)^{|\gamma|-1}\frac{|x-\tilde x|^2 (x\tilde x)^{1+\gamma}}{x^{|\gamma|}+\tilde x^{|\gamma|}} =
\frac{|x-\tilde x|^2} {x^{|\gamma|}+\tilde x^{|\gamma|}},
\end{align*}
which is indeed bounded by $(x-\tilde x)^2 \min{(x^\gamma, \tilde{x}^\gamma)}$.
\vskip1mm
We now verify that
$A_2^K \leq C\big(|v-\tilde{v}|^2 + |v_*- \tilde{v}_*|^2\big)|v-v_*|^\gamma$.
By Step 1, for any $X,Y\in\mathbb{R}^3$,
\[\left|X\Phi_K(|X|)- Y\Phi_K(|Y|)\right|\le |Y||\Phi_K(|X|)-\Phi_K(|Y|)|+|X-Y|\Phi_K(|X|)\le
C |Y|\Big ||X|^\gamma-|Y|^\gamma\Big|+C|X-Y| |X|^\gamma.\]
Since again $|x^{\gamma} - \tilde x^{\gamma}|\le|x^{-1}-\tilde x^{-1}|(x\land \tilde x)^{1+\gamma}$,
we conclude that
$\left|X\Phi_K(|X|)- Y\Phi_K(|Y|)\right| \leq C |X-Y| |X|^\gamma$, whence
\begin{align*}
A_2^K &\le C\, \big[|v-\tilde{v}|+|v_*- \tilde{v}_*|\big]|(v-v_*)-(\tilde{v} - \tilde{v}_*)|\min\{x^\gamma, \tilde x^\gamma \}
\end{align*}
as desired.
\vskip1mm
We next observe that $A_4^K \leq 2A_2^K$.
\vskip1mm
Finally, we see that $\Psi_K(x)\le C\int_K^\infty G^2(z/x^\gamma)dz\le
C\int_K^\infty(z/x^\gamma)^{-2/\nu}dz=Cx^{2\gamma/\nu}K^{1-2/\nu}$ and that $\Psi_K(x)\le C\int_0^\infty G^2(z/x^\gamma)dz\le C\int_0^\infty (1+z/x^\gamma)^{-2/\nu}dz =C x^\gamma$ according to \eqref{con1} , which imply $\Psi_K(x)\le C\min\{x^\gamma, x^{2\gamma/\nu}K^{1-2/\nu}\}$. Hence,
\[
A_3^K=(x^2+2|v-\tilde v| x ) \Psi_K(x) \leq C |v-\tilde{v}|^2 |v-v_*|^\gamma
+ C |v-v_*|^{2+2\gamma/\nu}K^{1-2/\nu},
\]
because $2|v-\tilde v| x \le |v-\tilde v|^2 + x^2$ and $x^2\Psi_K(x)\le C x^{2+2\gamma/\nu}K^{1-2/\nu}$.
\vskip1mm
The left hand side of \eqref{ee3} is nothing but
$J_K+L_K$, which is bounded by $A_1^K+A_2^K+A_3^K$: \eqref{ee3} is proved.
Finally, the left hand side of \eqref{ee2} equals $\lim_{K\to \infty} I_K$ and we know that
$I_K \leq A_1^K+A_2^K+A_4^K$, which is (uniformly in $K$) bounded by
$(|v-\tilde{v}|^2 + |v_*- \tilde{v}_*|^2) |v-v_*|^\gamma$ as desired.
\epf
\section{Stability}
In this section, our goal is to prove Theorem \ref{thm-uniqueness}.
For this, we first need two important propositions.
The first one is the most delicate part in the whole proof, which consists in showing that we can associate a {\it weak}
solution to some SDE to {\it any} weak solution to \eqref{Bol}.
It extends Proposition B.1 in \cite{Fournier:2015aa} to our situation, see also
\cite[Theorem 2.6]{MR2375067} (both concerning diffusion-type PDEs and Brownian SDEs). We will prove it later.
\begin{prop}\def\eprop{\end{prop}}\label{prop-PDE}
Assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ with $\gamma+\nu>0$.
Consider any weak solution $(\tilde f_t)_{t\ge0}\in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)$ to
\eqref{Bol}. Then there exists, on some probability space, a random variable $X_0$ with law $\tilde f_0$, independent of
a Poisson measure
$M(ds, d\alpha,dz, d\varphi)$ on $[0, \infty) \times [0,1]\times [0,\infty)\times[0,2\pi)$ with intensity
$ds d\alpha dz d\varphi$, a measurable family $(X^*_t)_{t\geq 0}$ of $\alpha$-random variables and a c\`adl\`ag
adapted process $(X_t)_{t\ge0}$ solving
\begin{equation}}\def\eeq{\end{equation}\label{SDE}
X_t=X_0+\int_0^t\int_0^1\int_0^\infty\int_0^{2\pi} c \big(X_{s-}, X_s^*(\alpha), z,\varphi\big)
M(ds,d\alpha, dz,d\varphi)
\eeq
and such that for all $t\geq 0$, ${\mathcal L}(X_t)={\mathcal L}_\alpha(X^*_t)=\tilde f_t$.
\eprop
The second one is the following proposition.
\begin{prop}\def\eprop{\end{prop}}\label{prop-PDE2}
Assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ with $\gamma+\nu>0$,
that $f_0\in \mathcal{P}_q(\mathbb{R}^3)$ for some $q\ge2$ such that $q>\gamma^2/(\gamma+\nu)$ and that
$f_0$ has a finite entropy. Fix $p\in(3/(3+\gamma),p_0(\gamma,\nu,q))$.
Let $(f_t)_{t\ge0} \in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)
\cap L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$ be the corresponding unique weak solution to \eqref{Bol} given by Theorem \emph{\ref{well-posedness}}.
Consider also the Poisson measure $M$, the process $(X_t)_{t\geq 0}$ and the family $(X^*_t)_{t\geq 0}$ built in
Proposition \emph{\ref{prop-PDE}} (associated to another weak solution $(\tilde f_t)_{t\ge0}\in L^\infty\big([0,\infty),
\mathcal{P}_2({\mathbb{R}^3})\big)$. Let $W_0\sim f_0$ (independent of $M$) be such that
$\mathbb{E}[|W_0-X_0|^2]={\mathcal W}_2^2(f_0,\tilde f_0)$ and, for each $t\geq 0$, an $\alpha$-random variable $W^*_t$
such that ${\mathcal L}_\alpha(W^*_t)=f_t$ and $\mathbb{E}_\alpha[|W_t^*-X_t^*|^2]={\mathcal W}_2^2(f_t,\tilde f_t)$.
Then for $K \ge 1$, the equation
\begin{equation}}\def\eeq{\end{equation}\label{Bprocess}
W_t^K=W_0+\int_0^t\int_0^1\int_0^\infty\int_0^{2\pi}c_K(W_{s-}^K, W_s^*(\alpha),z,\varphi+\varphi_{s,\alpha,K})M(ds,d\alpha,dz,d\varphi),
\eeq
with $\varphi_{s,\alpha,K}=\varphi_0(X_{s-}-X_s^*(\alpha),W_{s-}^K-W_s^*(\alpha))$,
has a unique solution. Moreover, setting $f_t^K={\mathcal L}(W^K_t)$ for each $t\ge 0$, it holds that for all $T>0$,
\begin{equation}}\def\eeq{\end{equation}\label{prop-law-convergence}
\lim_{K\rightarrow\infty}\sup_{[0,T]}\mathcal{W}_2^2(f_t^K, f_t)=0.
\eeq
\eprop
\begin{proof}}\def\epf{\end{proof}
For any $K\geq 1$,
the Poisson measure involved in \eqref{Bprocess} is actually finite
(because $c_K=c\mathbf{1}_{\{z\le K\}}$), so the existence and uniqueness for this equation is obvious.
It only remains to prove \eqref{prop-law-convergence}, which has already been done
in \cite[Lemma 4.2]{MR2398952}, where the formulation of the equation
is slightly different. But one easily checks that $(W^K_t)_{t\geq 0}$ is a (time-inhomogeneous)
Markov process with the same generator as the one defined by \cite[Eq. (4.1)]{MR2398952},
because for all bounded measurable function
$\phi:\mathbb{R}^3\mapsto \mathbb{R}$ and all $t\geq 0$, a.s.,
\begin{align*}
&\int_0^1\int_0^\infty\int_0^{2\pi}\Big[\phi(w+c_K(w, W_t^*(\alpha),z,\varphi+
\varphi_0(X_{t-}-X^*_t(\alpha),w-W^*_t(\alpha)) )-\phi(w)\Big] d\varphi dz d\alpha\\
=& \int_0^1\int_0^\infty\int_0^{2\pi}\Big[\phi(w+c_K(w,v,z,\varphi))-\phi(w)\Big]
d\varphi dz f_t(dv)
\end{align*}
by the $2\pi$-periodicity of $c_K$ (in $\varphi$) and since ${\mathcal L}_\alpha(W^*_t)=f_t$.
\epf
Now, we use these coupled processes to conclude the
\begin{proof}}\def\epf{\end{proof}[Proof of Theorem \ref{thm-uniqueness}]
We consider a weak solution $(\tilde f_t)_{t\geq 0}$ to \eqref{Bol}, with which we associate the objects
$M$, $(X_t)_{t\geq 0}$, $(X_t^*)_{t\geq 0}$ as in Proposition \ref{prop-PDE}. We then consider $f_0$
satisfying the assumptions of Theorem \ref{well-posedness} and the corresponding unique weak solution
$(f_t)_{t\geq 0}$ belonging to $L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)
\cap L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$ (with $p\in(3/(3+\gamma),p_0(\gamma,\nu,q))$)
and we consider $(W_t^K)_{t\geq 0}$, $(W_t^*)_{t\geq 0}$
built in Proposition \ref{prop-PDE2} for any $K\ge 1$. We know that
${\mathcal W}_2^2(f_0,\tilde f_0)=\mathbb{E}[|W_0-X_0|^2]$ and that
${\mathcal W}_2^2(f_t,\tilde f_t)=\mathbb{E}_\alpha[|W_t^*-X_t^*|^2]$
for all $t\geq 0$. Using that $W_t^K\sim f_t^K$ and $X_t \sim \tilde f_t$ for each $t\ge0$,
we deduce from \eqref{prop-law-convergence} that for all $t\ge0$,
\begin{equation}\label{dfjt}
\mathcal{W}_2^2(f_t,\tilde{f}_t)\le \limsup_{K\to\infty}\mathbb{E}[|W_t^K-X_t|^2]=:J_t.
\end{equation}
Next, we focus on the time interval $[0, T]$ for any fixed $T>0$, and split the proof into several steps.
\vskip1mm
{\it Step 1.}
By the It\^{o} formula, we know that
\begin{align*}
\mathbb{E}[|W_t^K-X_t|^2]=\mathbb{E}[|W_0-X_0|^2]+\mathbb{E}\left[ \int_0^t \int_0^1 \Delta_s^K(\alpha) d\alpha ds \right],
\end{align*}
where
\begin{align*}
\Delta_s^K(\alpha):=&\int_0^\infty\int_0^{2\pi}\Big(|W_{s}^K-X_{s}+c_{K,W}(s) - c_X(s)|^2-|W_{s}^K-X_{s}|^2\Big) d\varphi dz
\end{align*}
with the shortened notation $c_{K,W}(s):=c_K \big(W_{s}^K,W_s^*(\alpha),z,\varphi+\varphi_{s,\alpha,K}\big)$
and $c_X(s):=c \big(X_{s},X_s^*(\alpha),z,\varphi\big).$
We then show that
\begin{align}
\Delta_s^K(\alpha) \le& C (|W_s^K-X_s| ^2 + |W_s^*(\alpha)-X_s^*(\alpha)|^2) |W_s^K - W_s^*(\alpha)|^\gamma
+ C |W_s^K - W_s^*(\alpha)|^{2+2\gamma/\nu} K^{1-2/\nu}, \label{rrr1}
\end{align}
and
\begin{align}
\Delta_s^K(\alpha) \leq & C|W_s^K-W_s^*(\alpha)|^{\gamma+2}+C|X_s-X_s^*(\alpha)|^{\gamma+2} \notag \\
&+C |W_{s}^K-X_{s}|\big(|W_s^K-W_s^*(\alpha)|^{\gamma+1}+|X_s-X_s^*(\alpha)|^{\gamma+1}\big).\label{rrr2}
\end{align}
First, Lemma \ref{estimategeneral} (inequality \eqref{ee3}) precisely tells us that \eqref{rrr1} holds true.
Next, we observe that
\begin{align*}
\Delta_s^K(\alpha)
\le 2\int_0^\infty\int_0^{2\pi} (|c_{K,W}(s)|^2+|c_X(s)|^2) d\varphi dz+ 2|W_{s}^K-X_{s}|\Big|\int_0^\infty\int_0^{2\pi}(c_{K,W}(s) - c_X(s)) d\varphi dz \Big|.
\end{align*}
Hence, using \eqref{ee4} and \eqref{ee5}, the proof of \eqref{rrr2} is concluded.
\vskip1mm
{\it Step 2.} Set $\kappa(\gamma)=\min((\gamma+1)/|\gamma|,|\gamma|/2)>0$. We verify that there exists a constant $C(T, f_0, \tilde{f}_0, f)>0$ (depending on $T$, $m_2(f_0)$, $m_2(\tilde{f}_0)$, $\int_0^t \|f_s\|_{L^{p}}ds$), such that for all $\ell \ge 1$ (and all $K\geq 1$),
$$
I_t^{i,\ell} \le C(T, f_0, \tilde{f}_0, f)\ell^{-\kappa(\gamma)}, \quad i=1,2,3,4,
$$
where
\begin{align*}
&I_t^{1,\ell}:=\mathbb{E}\Big[\int_0^t \int_0^1|W_s^K-W_s^*(\alpha)|^{\gamma+2}\,\mathbf{1}_{\{|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell\}} d\alpha ds\Big],\\
&I_t^{2, \ell}:=\mathbb{E}\Big[\int_0^t \int_0^1|X_s-X_s^*(\alpha)|^{\gamma+2}\, \mathbf{1}_{\{|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell\}} d\alpha ds\Big],\\
&I_t^{3,\ell}:=\mathbb{E}\Big[\int_0^t \int_0^1|W_{s}^K-X_{s}||W_s^K-W_s^*(\alpha)|^{\gamma+1}\,\mathbf{1}_{\{|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell\}} d\alpha ds\Big],\\
&I_t^{4, \ell}:=\mathbb{E}\Big[\int_0^t \int_0^1|W_{s}^K-X_{s}||X_s-X_s^*(\alpha)|^{\gamma+1}\, \mathbf{1}_{\{|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell\}} d\alpha ds\Big].
\end{align*}
Since $\gamma\in(-1,0)$ and $\kappa(\gamma)\le (\gamma+2)/|\gamma|$, we have
\[
I_t^{1,\ell}\leq \ell^{-(\gamma+2)/|\gamma|}T\le \ell^{-\kappa(\gamma)}T.
\]
Similarly,
\begin{align*}
I_t^{3,\ell} &\le \ell^{-(\gamma+1)/|\gamma|} \int_0^t \mathbb{E}\Big[ |W_{s}^K-X_{s}|\Big] ds.
\end{align*}
Using \eqref{conservation} for $(f_t)_{t\ge0}$ and $(\tilde{f}_t)_{t\ge0}$, \eqref{prop-law-convergence}, and that $m_2(f_s^K)\le 2m_2(f_s)+2\mathcal{W}_2^2(f_s,f_s^K)$, we know that $\mathbb{E}\Big[ |W_{s}^K-X_{s}|\Big]\le
C(1+m_2(f_s^K)+m_2(\tilde{f}_s))\le C(T, f_0, \tilde{f}_0)$. Hence,
\[I_t^{3,\ell}\le C(T, f_0, \tilde{f}_0)\ell^{-\kappa(\gamma)}. \]
Since $\gamma+2\in(1,2)$, it follows from the H\"{o}lder inequality that
\begin{align*}
I_t^{2, \ell}
&\le \mathbb{E}\left[\Big(\int_0^t \int_0^1|X_s-X_s^*(\alpha)|^{2}d\alpha ds\Big)^{\frac{\gamma+2}{2}}\Big( \int_0^t \int_0^1\mathbf{1}_{\{|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell\}} d\alpha ds\Big)^{\frac{|\gamma|}{2}}\right]\\
&\le C\mathbb{E}\left[\Big(\int_0^t (|X_s|^2+m_2(\tilde{f}_s)) ds\Big)^{\frac{\gamma+2}{2}}\Big( \int_0^t \int_0^1\frac{|W_s^K - W_s^*(\alpha)|^\gamma}{\ell} d\alpha ds\Big)^{\frac{|\gamma|}{2}}\right]
\end{align*}
Since ${\mathcal L}_\alpha(W_s^*)=f_s$, we have
$\int_0^1|W_s^K - W_s^*(\alpha)|^\gamma d\alpha= \int_{\mathbb{R}^3} |W_s^K - v|^\gamma f_s(dv)\leq
1+ C_{\gamma,p}\|f_s\|_{L^{p}}$ by \eqref{norm-inequality},
so that
\begin{align*}
I_t^{2, \ell}
\le& \ell^{\gamma/2}\Big(1+ \int_0^t \big(\mathbb{E}[|X_s|^2]+m_2(\tilde{f}_s)\big)ds \Big)\Big(\int_0^t \big(1+C_{\gamma,p}\|f_s\|_{L^{p}}\big)ds\Big)^{\frac{|\gamma|}{2}}\\
\le& \ell^{\gamma/2}\Big(1+ 2m_2(\tilde{f}_0)T \Big)\Big(1+ \int_0^t \big(1+C_{\gamma,p}\|f_s\|_{L^{p}}\big)ds\Big) \le C(T,\tilde{f}_0, f)\ell^{-\kappa(\gamma)}.
\end{align*}
For $I_t^{4,\ell}$, we use the triple H\"{o}lder inequality to write
\begin{align*}
I_t^{4,\ell}
\le& \mathbb{E}\Big[\int_0^t |W_{s}^K-X_{s}|^2 ds\Big]^{\frac{1}{2}}
\mathbb{E}\Big[\int_0^t \int_0^1|X_s-X_s^*(\alpha)|^2d\alpha ds\Big]^{\frac{1+\gamma}{2}}
\mathbb{E} \Big[\int_0^t\int_0^1\mathbf{1}_{\{|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell\}}d\alpha ds\Big]^{\frac{|\gamma|}{2}}.
\end{align*}
Thus $I^{4,\ell}_t\leq C(T, f_0, \tilde{f}_0, f)\ell^{-\kappa(\gamma)}$:
use that $\mathbb{E}[|X_s|^2]=\mathbb{E}_\alpha[|X_s^*|^2]=m_2(\tilde f_0)$, that
$m_2(f_s^K)\le 2m_2(f_s)+2\mathcal{W}_2^2(f_s,f_s^K)$
as before
and treat the last term of the product the same as we study $I_t^{2, \ell}$.
\vskip1mm
{\it Step 3.}
According to Step 1, we now bound $\Delta_s^K(\alpha)$ by \eqref{rrr1} when $|W_s^K - W_s^*(\alpha)|^\gamma\le \ell$
and by \eqref{rrr2} when $|W_s^K - W_s^*(\alpha)|^\gamma\ge \ell$:
\begin{align*}
\mathbb{E}[|W_t^K-X_t|^2] &\le \mathbb{E}[|W_0-X_0|^2] + C\sum_{i=1}^4 I_t^{i,\ell} + C K^{1-2/\nu} \mathbb{E}\Big[\int_0^t \int_0^1|W_s^K - W_s^*(\alpha)|^{2+2\gamma/\nu} d\alpha ds\Big]\\
&\quad~ + C \mathbb{E}\Big[\int_0^t \int_0^1 (|W_s^K-X_s| ^2 + |W_s^*(\alpha)-X_s^*(\alpha)|^2)\min{\big(|W_s^K - W_s^*(\alpha)|^\gamma, \ell \big)} d\alpha ds\Big].
\end{align*}
It then follows from Step 2 that for all $\ell\ge1$, all $K\ge1$,
\begin{align}
\mathbb{E}[|W_t^K-X_t|^2]
&\le {\mathcal W}_2^2 (f_0,\tilde f_0) + C(T, f_0, \tilde{f}_0, f) \ell^{-\kappa(\gamma)} \label{rara}\\
&+ C K^{1-2/\nu} \mathbb{E}\Big[\int_0^t \int_0^1|W_s^K - W_s^*(\alpha)|^{2+2\gamma/\nu} d\alpha ds\Big]\notag\\
& + C \mathbb{E}\Big[\int_0^t \int_0^1 |W_s^K-X_s| ^2 |W_s^K - W_s^*(\alpha)|^\gamma d\alpha ds\Big]\notag\\
&+ C\mathbb{E}\Big[ \int_0^t \int_0^1 |W_s^*(\alpha)-X_s^*(\alpha)|^2 \min{\big(|W_s^K - W_s^*(\alpha)|^\gamma, \ell \big)} d\alpha ds\Big].\notag
\end{align}
Since $\gamma+\nu>0$, it holds that $2+2\gamma/\nu>0$. As a consequence, like in Step 2,
\[\mathbb{E}\Big[\int_0^t \int_0^1 |W_s^K-W_s^*(\alpha)|^{2+2\gamma/\nu} d\alpha ds\Big]\le C_T[1+\mathbb{E}[|W_s^K|^2]+m_2(f_0)]\le C(T,f_0, \tilde{f}_0),\]
which gives \[\lim_{K\to\infty}K^{1-2/\nu} \mathbb{E}\Big[\int_0^t \int_0^1|W_s^K - W_s^*(\alpha)|^{2+2\gamma/\nu} d\alpha ds\Big]=0.\]
Moreover, we recall that a.s.
$\int_0^1|W_s^K - W_s^*(\alpha)|^\gamma d\alpha \leq 1+ C_{\gamma,p}\|f_s\|_{L^{p}}$ as in Step 2, whence
\[
\mathbb{E}\Big[\int_0^t \int_0^1 |W_s^K-X_s| ^2 |W_s^K - W_s^*(\alpha)|^\gamma d\alpha ds\Big]
\le \int_0^t \mathbb{E}[|W_s^K-X_s| ^2](1+C_{\gamma,p}\|f_s\|_{L^p})ds.
\]
Letting $K\to\infty$, by dominated convergence, we find (recall \eqref{dfjt})
\[
\limsup_{K}\mathbb{E}\Big[\int_0^t \int_0^1 |W_s^K-X_s| ^2 |W_s^K - W_s^*(\alpha)|^\gamma d\alpha ds\Big]
\le \int_0^t J_s(1+C_{\gamma,p}\|f_s\|_{L^p})ds.
\]
Next, it is obvious that for each $\ell\geq 1$ fixed, for all $s\in[0, T]$, all $\alpha\in[0,1]$,
the function $v\mapsto \min(|v-W_s^*(\alpha)|^\gamma, \ell)$ is bounded and continuous. By
\eqref{prop-law-convergence}, we conclude that
$\lim_{K\to \infty} \mathbb{E}\big[\min\big(|W_s^K - W_s^*(\alpha)|^\gamma, \ell \big) \big]
=\mathbb{E}\big[\min\big(|W_s- W_s^*(\alpha)|^\gamma, \ell \big)\big]$ and, by dominated convergence, that,
still for $\ell\geq 1$ fixed,
\begin{align*}
&\lim_{K\to\infty}\mathbb{E}\Big[ \int_0^t \int_0^1 |W_s^*(\alpha)-X_s^*(\alpha)|^2 \min\big(|W_s^K - W_s^*(\alpha)|^\gamma, \ell \big) d\alpha ds\Big]\\
&= \int_0^t \int_0^1 |W_s^*(\alpha)-X_s^*(\alpha)|^2
\mathbb{E}\big[\min{\big(|W_s- W_s^*(\alpha)|^\gamma, \ell \big)} \big] d\alpha ds.
\end{align*}
But since $W_s \sim f_s$, we have, for each $\alpha$ fixed,
$\mathbb{E}[\min{(|W_s- W_s^*(\alpha)|^\gamma, \ell )} ]\leq \int_{\mathbb{R}^3} |W_s^*(\alpha) - v|^\gamma f_s(dv)\leq
1+ C_{\gamma,p}\|f_s\|_{L^{p}}$ by \eqref{norm-inequality}. Furthermore, we have
$\int_0^1|W_s^*(\alpha)-X_s^*(\alpha)|^2 d\alpha=\mathbb{E}_\alpha[|W_s^*-X_s^*|^2]={\mathcal W}_2^2(f_s,\tilde f_s)\le J_s$.
All in all, we have checked that
\begin{align*}
&\lim_{K\to\infty}\mathbb{E}\Big[ \int_0^t \int_0^1 |W_s^*(\alpha)-X_s^*(\alpha)|^2 \min\big(|W_s^K - W_s^*(\alpha)|^\gamma, \ell \big) d\alpha ds\Big] \leq C \int_0^t J_s(1+\|f_s\|_{L^p})ds.
\end{align*}
Gathering all the previous estimates to let $K\to \infty$ in \eqref{rara}: for each $\ell\geq 1$
fixed,
$$
J_t \leq {\mathcal W}_2^2(f_0,\tilde f_0) + C(T, f_0, \tilde{f}_0, f) \ell^{-\kappa(\gamma)} +C\int_0^t J_s(1+\|f_s\|_{L^p})ds.
$$
Letting now $\ell\to \infty$ and using the Gr\"{o}nwall lemma, we find
\[J_t \le\mathcal{W}_2^2(f_0, \tilde{f}_0)
\exp{\left(C_{\gamma,p}\int_0^t\big(1+ \|f_s\|_{L^{p}}\big)ds\right)}.\]
Since $\mathcal{W}_2^2(f_t, \tilde{f}_t)\le J_t$, this completes the proof.
\epf
It remains to prove Proposition \ref{prop-PDE}. We start with a technical result.
\begin{lem}\def\elem{\end{lem}}\label{lem-local-condition}
Assume \eqref{con} for some $\gamma\in(-1,0)$, some $\nu\in(0,1)$ with $\gamma+\nu>0$ and recall that the deviation function $c$ was defined by \eqref{rewrite}. Consider $f\in \mathcal{P}_2({\mathbb{R}^3})$
and $\phi_\epsilon(x)=(2\pi\epsilon)^{-3/2} e^{-|x|^2/(2\epsilon)}$. Set $f^\epsilon(w)=(f * \phi_\epsilon)(w)$.
\begin{enumerate}} \def\eenu{\end{enumerate}[label=\emph{(\roman*)}]
\item There exists a constant $C>0$ such that for all $x\in\mathbb{R}^3$, all $\epsilon\in(0,1)$, \begin{align*}
\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty \int_0^{2\pi} |c(v,v_*,z, \varphi)| \frac{\phi_\epsilon(v-x)} {f^\epsilon(x)} d\varphi dz f(dv) f(dv_*)\le C \left(1+\sqrt{m_2(f)}+|x| \right),
\end{align*}
\item For all $\epsilon\in(0,1)$, all $R>0$, there is a constant $C_{R, \epsilon}>0$ (depending only on
$m_2(f)$) such that for all $x,y\in B(0, R)$,
\begin{align*}
\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty \int_0^{2\pi} |c(v,v_*,z, \varphi)| \left|\frac{\phi_\epsilon(v-x)}{f^\epsilon(x)} - \frac{ \phi_\epsilon(v-y)}{f^\epsilon(y)}\right| d\varphi dz f(dv) f(dv_*)\le C_{R, \epsilon}|x-y|.
\end{align*}
\eenu
\elem
\begin{proof}}\def\epf{\end{proof}
We start with (i) and set $I_\epsilon(x)=\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty \int_0^{2\pi}
|c(v,v_*,z, \varphi)| \frac{\phi_\epsilon(v-x)} {f^\epsilon(x)} d\varphi dz f(dv) f(dv_*)$.
Using \eqref{num} and \eqref{con1}, we see that $|c(v,v_*,z, \varphi)|\leq G(z/|v-v_*|^\gamma)|v-v_*|
\leq C(1+z/|v-v_*|^\gamma)^{-1/\nu}|v-v_*|$. Hence
\begin{align*}
I_\epsilon(x)\leq& C \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty (1+z/|v-v_*|^\gamma)^{-1/\nu}|v-v_*|\frac{\phi_\epsilon(v-x)}{f^\epsilon(x)} dz f(dv) f(dv_*)\\
=&C \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} |v-v_*|^{1+\gamma}\frac{\phi_\epsilon(v-x)}{f^\epsilon(x)} f(dv) f(dv_*).
\end{align*}
Using now that $|v-v_*|^{1+\gamma} \le 1+|v|+|v_*|$, we find
\begin{align*}
I_\epsilon(x)\leq & C \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} (1+|v|+|v_*|)\frac{\phi_\epsilon(v-x)}{f^\epsilon(x)} f(dv) f(dv_*)
\leq
C \Big(1+\sqrt{m_2(f)}+ \frac{\int_{\mathbb{R}^3}|v|\phi_\epsilon(v-x) f(dv)}{f^\epsilon(x)} \Big).
\end{align*}
To conclude the proof of (i), it remains to study $J_\epsilon(x)= (f^\epsilon(x))^{-1}\int_{\mathbb{R}^3}|v|\phi_\epsilon(v-x)f(dv)$.
We introduce $L:=\sqrt{2m_2(f)}$, for which $f(B(0,L))\ge 1/2$ (because $f(B(0,L)^c)\leq m_2(f)/L^2$).
Using that $\{v \in \mathbb{R}^3:|v|\leq 2|x|+L\}\cup\{v \in \mathbb{R}^3:|v-x|\geq|x|+L\}=\mathbb{R}^3$, we write
\[
J_\epsilon(x)= \frac{\int_{\mathbb{R}^3}|v|\phi_\epsilon(v-x) f(dv)}{\int_{\mathbb{R}^3}\phi_\epsilon(v-x) f(dv)} \le 2|x|+L+\frac{\int_{|v-x|\ge |x|+L}|v|\phi_\epsilon(v-x) f(dv)}{\int_{|v-x|\le |x|+L} \phi_\epsilon(v-x) f(dv)}.
\]
Since $\phi_\epsilon$ is radial and decreasing,
\[\int_{|v-x|\ge |x|+L}|v|\phi_\epsilon(v-x) f(dv) \le \phi_\epsilon(|x|+L)\sqrt{m_2(f)}\]
and
\[\int_{|v-x|\le |x|+L} \phi_\epsilon(v-x) f(dv) \ge \phi_\epsilon(|x|+L) f(B(x, |x|+L)) \ge \phi_\epsilon(|x|+L)/2\]
owing to the fact that $B(0,L)\subset B(x, |x|+L)$. Hence, $J_\epsilon(x)\leq 2|x|+L+2\sqrt{m_2(f)}
\leq 2|x|+4\sqrt{m_2(f)}$
and this completes the proof of (i).
\vskip1mm
For point (ii), we set $\Delta_\epsilon(x,y)=\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty \int_0^{2\pi} |c(v,v_*,z, \varphi)|
|F_\epsilon(x,v)-F_\epsilon(y,v)| d\varphi dz f(dv) f(dv_*)$, where
$F_{\epsilon}(v,x):= (f^\epsilon(x))^{-1} \phi_\epsilon(v-x)$.
Exactly as in point (i), we start with
\begin{align*}
\Delta_\epsilon(x,y)
&\le C\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} |v-v_*|^{1+\gamma} |F_{\epsilon}(v,x) - F_{\epsilon}(v,y)| f(dv) f(dv_*)\\
&\le C\int_{\mathbb{R}^3}(1+\sqrt{m_2(f)} + |v|) |F_{\epsilon}(v,x) - F_{\epsilon}(v,y)| f(dv)\\
&\le C|x-y| \int_{\mathbb{R}^3}(1+\sqrt{m_2(f)} + |v|)\Big(\sup_{a\in B(0, R)} | \triangledown_x F_{\epsilon}(v,a)| \Big) f(dv)
\end{align*}
for all $x,y \in B(0,R)$.
But we have
\begin{align}\label{gg}
\triangledown_x F_{\epsilon}(v,a) =\frac{1}{\epsilon}\frac{\phi_\epsilon(v-a)\int_{\mathbb{R}^3} (v-u)\phi_\epsilon(u-a) f(du)}{(f^\epsilon(a))^2}.
\end{align}
Indeed, recalling that $\phi_\epsilon(x)=(2\pi\epsilon)^{-3/2} e^{-|x|^2/(2\epsilon)}$, we observe that
\[ \triangledown_x \phi_\epsilon(v-x)= \frac{1}{\epsilon} (v-x) \phi_\epsilon(v-x) \ \text {and}\ \triangledown_x f^\epsilon(x)=\frac{1}{\epsilon} \int_{\mathbb{R}^3}\phi_\epsilon(u-x)(u-x) f(du). \]
Since $F_{\epsilon}(v,a):= (f^\epsilon(a))^{-1}\phi_\epsilon(v-a)$, we have
\begin{align*}
\triangledown_x F_{\epsilon}(v,a)
&=\frac{ \triangledown_x \phi_\epsilon(v-a) f^\epsilon(a) - \phi_\epsilon(v-a) \triangledown_x f^\epsilon(a)}{(f^\epsilon(a))^2}\\
&=\frac{\phi_\epsilon(v-a)}{\epsilon} \frac{(v-a) f^\epsilon(a) - \int_{\mathbb{R}^3}\phi_\epsilon(u-a)(u-a) f(du)}{(f^\epsilon(a))^2}\\
&=\frac{\phi_\epsilon(v-a)}{\epsilon} \frac{ \int_{\mathbb{R}^3}\phi_\epsilon(u-a)(v-a) f(du) - \int_{\mathbb{R}^3}\phi_\epsilon(u-a)(u-a) f(du)}{(f^\epsilon(a))^2},
\end{align*}
whence \eqref{gg}. Using now that $J_\epsilon(a)= (f^\epsilon(a))^{-1}\int_{\mathbb{R}^3} |u| \phi_\epsilon(u-a) f(du)
\le 2|a|+ 4\sqrt{m_2(f)}$ as proved in (i),
\begin{align*}
| \triangledown_x F_{\epsilon}(v,a)|
&\le \frac{1}{\epsilon}\frac{\phi_\epsilon(v-a)}{f^\epsilon(a)}\frac{\int_{\mathbb{R}^3} (|v|+|u|) \phi_\epsilon(u-a) f(du)}{f^\epsilon(a)}
\le \frac{1}{\epsilon}\frac{\phi_\epsilon(v-a)}{f^\epsilon(a)} \left(|v| +2|a|+ 4\sqrt{m_2(f)}\right).
\end{align*}
But we know that $\phi_\epsilon(x) \le (2\pi\epsilon)^{-3/2}$ and that
\begin{align*}
f^\epsilon(a) \ge \int_{|v-a|\le |a|+L} \phi_\epsilon(v-a) f(dv) \ge \phi_\epsilon(|a|+L) f(B(a, |a|+L)) \ge \phi_\epsilon(|a|+L)/2
\end{align*}
since $B(0,L)\subset B(a, |a|+L)$. Hence,
\[ \sup_{a\in B(0, R)} | \triangledown_x F_{ \epsilon}(v,a)| \le \frac{2}{\epsilon} \ e^{(R+L)^2/(2\epsilon)} \left(|v| +2R+ 4\sqrt{m_2(f)}\right).\]
Consequently, for all $x,y \in B(0,R)$,
\begin{align*}
\Delta_\epsilon(x,y)
&\le \frac{2C}{\epsilon} \ e^{(R+L)^2/(2\epsilon)} |x-y| \int_{\mathbb{R}^3}\left(1+\sqrt{m_2(f)} + |v|\right) \left(|v| +2R+ 4\sqrt{m_2(f)}\right) f(dv)\leq C_{R,\epsilon} |x-y|,
\end{align*}
where $C_{R,\epsilon}$ depends only on $R,\epsilon$ and $m_2(f)$ (recall that $L:=\sqrt{2m_2(f)}$).
\epf
Finally, we end the section with the
\begin{proof}}\def\epf{\end{proof}[Proof of Proposition \ref{prop-PDE}]
We consider any given weak solution $(\tilde f_t)_{t\geq 0}\in L^\infty([0,\infty),\mathcal{P}_2(\mathbb{R}^3))$ to \eqref{Bol}
and we write the proof in several steps.
\vskip1mm
{\it Step 1.} We introduce $\phi_\epsilon(x)=(2\pi\epsilon)^{-3/2} e^{-|x|^2/(2\epsilon)}$
and $\tilde f_t^\epsilon(w)=(\tilde f_t * \phi_\epsilon)(w)$.
For each $t\ge 0$, we see that $\tilde f_t^\epsilon$ is a positive smooth function. We claim that
for any $\psi \in Lip(\mathbb{R}^3)$,
$$
\frac{\partial}{\partial t} \int_{\mathbb{R}^3} \psi(w) \tilde f_t^\epsilon(dw)= \int_{\mathbb{R}^3} \tilde f_t^\epsilon (dw)
\mathcal{\tilde{A}}_{t,\epsilon} \psi(w),
$$
where
\begin{equation}}\def\eeq{\end{equation}\label{pro-generator}
\mathcal{\tilde{A}}_{t,\epsilon} \psi(w)= \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty\int_0^{2\pi}
[\psi(w+c(v, v_*, z, \varphi)) - \psi(w)] \frac{ \phi_\epsilon(v-w)}{\tilde f_t^{\epsilon}(w)}
d\varphi dz \tilde f_t(dv_*)\tilde f_t(dv).
\eeq
Indeed, $\tilde f_t^\epsilon(w)=\int_{\mathbb{R}^3} \phi_\epsilon(v-w) \tilde f_t(dv)$ since $ \phi_\epsilon(x)$ is even.
According to \eqref{weak} and \eqref{opera}, we have
\begin{align*}
\frac{\partial}{\partial t} \tilde f_t^\epsilon(w)
&=\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty \int_0^{2\pi} [\phi_\epsilon(v-w+c(v,v_*,z,\varphi))-\phi_\epsilon(v-w)] d\varphi dz \tilde f_t(dv_*) \tilde f_t(dv)\\
&= \int_{\mathbb{R}^3} \int_0^K \int_0^{2\pi} \left[\int_{\mathbb{R}^3}\phi_\epsilon(v-w+c(v,v_*,z,\varphi)) \tilde f_t(dv) - \tilde f_t^\epsilon(w) \right] d\varphi dz\tilde f_t(dv_*)\\
&\quad~~ + \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_K^\infty \int_0^{2\pi} [\phi_\epsilon(v-w+c(v,v_*,z,\varphi))-\phi_\epsilon(v-w)] d\varphi dz \tilde f_t(dv_*) \tilde f_t(dv)
\end{align*}
for any $K\geq 1$. We thus have, for any $\psi \in Lip(\mathbb{R}^3)$,
\begin{align*}
\frac{\partial}{\partial t} \int_{\mathbb{R}^3}\psi(w) &\tilde f_t^\epsilon(dw)
=\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^K \int_0^{2\pi} \int_{\mathbb{R}^3}\phi_\epsilon(v-w+c(v,v_*,z,\varphi))\psi(w)
\tilde f_t(dv) d\varphi dz \tilde f_t(dv_*) dw\\
&- \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^K \int_0^{2\pi} \psi(w)\tilde f^\epsilon_t(w) d\varphi dz \tilde f_t(dv_*)dw\\
&+ \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_K^\infty \int_0^{2\pi} [\phi_\epsilon(v-w+c(v,v_*,z,\varphi))-\phi_\epsilon(v-w)] \psi(w) d\varphi dz \tilde f_t(dv_*) \tilde f_t(dv) dw.
\end{align*}
Using the change of variables $w-c(v,v_*,z,\varphi) \mapsto w$, we see that the first integral of the RHS equals
$$
\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^K \int_0^{2\pi} \int_{\mathbb{R}^3}\phi_\epsilon(v-w) \psi(w
+ c(v,v_*,z,\varphi))\tilde f_t(dv)d\varphi dz \tilde f_t(dv_*) dw.
$$
Consequently,
\begin{align*}
&\frac{\partial}{\partial t} \int_{\mathbb{R}^3}\psi(w) \tilde f_t^\epsilon(dw)\\
&= \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^K \int_0^{2\pi} \left[\int_{\mathbb{R}^3} \psi(w+c(v,v_*,z,\varphi))\frac{\phi_\epsilon(v-w) }{\tilde f_t^\epsilon(w)} \tilde f_t(dv) -\psi(w) \right] \tilde f_t^\epsilon(w) d\varphi dz \tilde f_t(dv_*) dw \\
&\quad + \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_K^\infty \int_0^{2\pi} [\phi_\epsilon(v-w+c(v,v_*,z,\varphi))-\phi_\epsilon(v-w)] \psi(w) d\varphi dz \tilde f_t(dv_*) \tilde f_t(dv) dw\\
&=\int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^K \int_0^{2\pi} \int_{\mathbb{R}^3} \left[\psi(w+c(v,v_*,z,\varphi)) -\psi(w) \right] \frac{\phi_\epsilon(v-w) }{\tilde f_t^\epsilon(w)} \tilde f_t(dv) d\varphi dz \tilde f_t(dv_*) \tilde f_t^\epsilon(dw)\\
&\quad + \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_K^\infty \int_0^{2\pi} [\phi_\epsilon(v-w+c(v,v_*,z,\varphi))-\phi_\epsilon(v-w)] \psi(w) d\varphi dz \tilde f_t(dv_*) \tilde f_t(dv) dw.
\end{align*}
Letting $K$ increase to infinity, one easily ends the step.
\vskip1mm {\it Step 2.} We set $F_{t,\epsilon}(v,x)= (\tilde f_t^\epsilon(x))^{-1} \phi_\epsilon(v-x)$.
For a given $X_0^\epsilon$ with law $\tilde f_0^\epsilon$, and a given independent Poisson measure $N(ds, dv, dv_*, dz, d\varphi, du)$ on $[0,\infty) \times \mathbb{R}^3 \times \mathbb{R}^3 \times [0, \infty) \times [0, 2\pi) \times [0, \infty)$ with intensity $ds \tilde f_s(dv) \tilde f_s(dv_*) dz d\varphi du$, there exists a pathwise unique solution to
\begin{equation}}\def\eeq{\end{equation}\label{pro-convolution-process}
X_t^\epsilon=X_0^\epsilon + \int_0^t \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \int_0^\infty \int_0^{2\pi} \int_0^\infty c(v, v_*, z, \varphi)\mathbf{1}_{\left \{u\le F_{s,\epsilon}(v, X_{s-}^\epsilon)\right \}} N(ds, dv, dv_*, dz, d\varphi, du).
\eeq
This classically follows from Lemma \ref{lem-local-condition}, which precisely tells us that
the coefficients of this equation satisfy some
{\it at most linear growth condition} (point (i)) and some {\it local Lipschitz condition} (point (ii)).
\vskip1mm
{\it Step 3.} We now prove that $\mathcal{L}(X_t^\epsilon)=\tilde f_t^\epsilon$ for each $t\ge0$.
We thus introduce $g^\epsilon_t=\mathcal{L}(X_t^\epsilon)$. By the It\^o formula, we see that for all
$\psi \in Lip({\mathbb{R}^3})$,
\begin{align*}
&\frac{\partial}{\partial t} \int_{\mathbb{R}^3} \psi(w) g_t^\epsilon(dw)\\
=& \int_{\mathbb{R}^3} g_t^\epsilon (dw)
\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\int_0^\infty\int_0^{2\pi} \Big(\psi(w+c(v,v_*,z,\varphi))-\psi(w) \Big)
F_{t,\epsilon}(v,w) d\varphi dz \tilde f_t(dv_*)\tilde f_t(dv)\\
=& \int_{\mathbb{R}^3} g_t^\epsilon (dw) \mathcal{\tilde{A}}_{t,\epsilon} \psi(w).
\end{align*}
Thus $(\tilde f^\epsilon_t)_{t\geq 0}$ and $(g^\epsilon_t)_{t\geq 0}$ satisfy the same equation and we of course have
$g^\epsilon_0=\tilde f^\epsilon_0$ by construction. The following uniqueness result allows us to conclude the step:
for any $\mu_0 \in \mathcal{P}_2({\mathbb{R}^3})$, there exists at most one family
$(\mu_t)\in L_{loc}^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)$ such that for any
$\psi \in Lip(\mathbb{R}^3)$, any $t\ge0$,
\begin{equation}}\def\eeq{\end{equation}\label{linear-SIE}
\int_{\mathbb{R}^3} \psi(w) \mu_t(dw)= \int_{\mathbb{R}^3} \psi(w) \mu_0 (dw)
+ \int_0^t ds\int_{\mathbb{R}^3} \mu_s(dw) \tilde{\mathcal{A}}_{s,\epsilon}\psi(w).
\eeq
This must be classical (as well as Step 2 is), but we find no precise reference
and thus make use of martingale problems.
A c\`{a}dl\`{a}g adapted $\mathbb{R}^3$-valued process $(Z_t)_{t\ge0}$ on some filtered probability space
$(\Omega, \mathcal{F}, \mathcal{F}_t, \mathbb{P})$ is said to solve the martingale problem $MP(\tilde{\mathcal{A}}_{t,\epsilon}, \mu_0,
Lip(\mathbb{R}^3))$ if $\mathbb{P}\circ Z_0=\mu_0$ and if for all $\psi \in Lip(\mathbb{R}^3)$,
$(M_t^{\psi,\epsilon})_{t\ge0}$ is a $(\Omega, \mathcal{F}, \mathcal{F}_t, \mathbb{P})$-martingale, where
\[M_t^{\psi, \epsilon}=\psi(Z_t) - \int_0^t \tilde{\mathcal{A}}_{s,\epsilon}\psi(Z_s)ds.\]
According to \cite[Theorem 5.2]{MR1245309} (see also \cite[Remark 3.1, Theorem 5.1]{MR1245309}
and \cite[Theorem B.1]{MR1042343}), it suffices to check the following points to conclude the uniqueness
for \eqref{linear-SIE}.
\begin{enumerate}} \def\eenu{\end{enumerate}[label=(\roman*)]
\item there exists a countable family $(\psi_k)_{k\ge1}\subset Lip(\mathbb{R}^3)$ such that for all $t\ge0$,
the closure (for the bounded pointwise convergence) of $\{(\psi_k, \tilde{\mathcal{A}}_{t,\epsilon}\psi_k), k\ge 1\}$
contains $\{(\psi, \tilde{\mathcal{A}}_{t,\epsilon}\psi), \psi\in Lip(\mathbb{R}^3)\}$,
\item for each $w_0\in\mathbb{R}^3$, there exists a solution to $MP( \tilde{\mathcal{A}}_{t,\epsilon}, \delta_{w_0}, Lip(\mathbb{R}^3))$,
\item for each $w_0\in\mathbb{R}^3$, uniqueness (in law) holds for $MP( \tilde{\mathcal{A}}_{t,\epsilon}, \delta_{w_0}, Lip(\mathbb{R}^3))$.
\eenu
The fact that \eqref{pro-convolution-process} has a pathwise unique solution proved in Step 2
(there we can of course replace $X_0^\epsilon$ by any deterministic point $w_0\in\mathbb{R}^3$) immediately
implies (ii) and (iii). Point (i) is very easy (recall that $\epsilon>0$ is fixed here).
\vskip1mm {\it Step 4.} In this step, we check that the family $((X_t^\epsilon)_{t\ge0})_{\epsilon>0}$ is tight in $\mathbb{D}([0,\infty), \mathbb{R}^3)$. To do this, we use the Aldous criterion \cite{MR0474446}, see also \cite[p 321]{MR1943877}, i.e. it suffices to prove that for all $T>0$,
\begin{equation}}\def\eeq{\end{equation}\label{aldous}
\sup_{\epsilon \in(0,1)}\mathbb{E}\big [\sup_{[0,T]}|X_t^\epsilon| \big ]< \infty, \quad
\lim_{\delta\to 0}\sup_{\epsilon \in(0,1)}\sup_{S, S^\prime\in \mathcal{S}_T(\delta)}\mathbb{E}\big [|X_{S^\prime}^\epsilon-X_S^\epsilon| \big]=0,
\eeq
where $\mathcal{S}_T(\delta)$ is the set containing all pairs of stopping times $(S, S^\prime)$ satisfying $0\le S\le S^\prime \le S+\delta\le T$.
\vskip1mm
First, since $X^\epsilon_t\sim \tilde f^\epsilon_t= \tilde f_t \star \phi_\epsilon$,
we have $\mathbb{E}[|X^\epsilon_t|^2]\leq 2 (m_2(\tilde f_t) + 3\epsilon) \leq 2m_2(\tilde f_0) + 6$.
Thus for any $T>0$, using Lemma \ref{lem-local-condition}-(i),
\begin{align*}
\mathbb{E} \Big[\sup_{[0,T]}|X_t^\epsilon| \Big]
&\le \mathbb{E} \big[|X_0^\epsilon| \big]
+ \mathbb{E} \Big[\int_0^T\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\int_0^\infty\int_0^{2\pi}|c(v,v_*,z, \varphi)|
\frac{\phi_\epsilon(v-X_s^{\epsilon})}{\tilde f_s^\epsilon(X_s^{\epsilon})} d\varphi dz \tilde f_s(dv) \tilde f_s(dv_*) ds \Big]\\
&\le \mathbb{E} \big[|X_0^\epsilon| \big] + C \mathbb{E}\left[\int_0^T (1+|X_s^\epsilon|) \, ds \right] \leq C_T.
\end{align*}
Furthermore, for any $T>0$, $\delta>0$ and $(S, S^\prime)\in\mathcal{S}_T(\delta)$, using again
Lemma \ref{lem-local-condition}-(i),
\begin{align*}
\mathbb{E} \big [|X_{S^\prime}^\epsilon-X_S^\epsilon| \big]
&\le \mathbb{E}\left[\int_S^{S+\delta}\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\int_0^\infty\int_0^{2\pi}|c(v,v_*,z, \varphi)| \frac{\phi_\epsilon(v-X_s^{\epsilon})}{\tilde f_s^\epsilon(X_s^{\epsilon})} d\varphi dz \tilde f_s(dv) \tilde f_s(dv_*) ds \right]\\
&\le C\mathbb{E}\left[\int_S^{S+\delta}(1+|X_s^\epsilon|) ds \right] \le C\mathbb{E}\left[\delta \sup_{[0,T]}(1+|X_s^\epsilon |) \right] \le C_T \delta.
\end{align*}
Hence \eqref{aldous} holds true and this completes the step.
\vskip1mm
{\it Step 5.} We thus can find some $(X_t)_{t\ge0}$ which is the limit in law (for the Skorokhod topology)
of a sequence $(X_t^{\epsilon_n})_{t\ge0}$ with $\epsilon_n\searrow 0$. Since $\mathcal{L}(X_t^{\epsilon_n})=
\tilde f_t^{\epsilon_n}$ by Step 3 and since $\tilde f_t^{\epsilon_n} \to \tilde f_t$ by definition,
we have $\mathcal{L}(X_t)=\tilde f_t$ for each $t\ge0$. It only remains to show that $(X_t)_{t\ge0}$ is a (weak) solution
to \eqref{SDE}. Using the theory of martingale problems, it classically suffices to prove that
for any $\psi \in C^1_b(\mathbb{R}^3)$, the process $\psi(X_t)-\psi(X_0) - \int_0^t \mathcal{B}_s\psi(X_s)ds$ is a martingale,
where
$$
\mathcal{B}_t\psi(x)=\int_0^1\int_0^\infty\int_0^{2\pi}\Big(\psi(x+c(x,X_t^*(\alpha),z,\varphi))-\psi(x)\Big)
d\varphi dz d\alpha.
$$
But since ${\mathcal L}_\alpha(X_t^*)=\tilde f_t$, this rewrites (recall \eqref{opera})
$$
\mathcal{B}_t\psi(x)=\int_{\mathbb{R}^3}\int_0^\infty\int_0^{2\pi}\Big(\psi(x+c(x,v_*,z,\varphi)-\psi(x)\Big)
d\varphi dz \tilde f_t(dv_*)= \int_{\mathbb{R}^3} \mathcal{A} \psi (x,v_*) \tilde f_t(dv_*).
$$
We thus have to prove that for any $0\le s_1\le ... \le s_k \le s\le t\le T$,
any $\psi_1,..., \psi_k\in C_b^1(\mathbb{R}^3)$, and any $\psi\in C_b^1(\mathbb{R}^3)$,
$$
\mathbb{E}[\mathcal{F}(X)]=0,
$$
where $\mathcal{F}: \mathbb{D}([0,\infty), \mathbb{R}^3)\mapsto \mathbb{R}$ is defined by
\[\mathcal{F}(\lambda)=\Big(\prod_{i=1}^{k}\psi_i(\lambda_{s_i})\Big) \Big(\psi(\lambda_t)-\psi(\lambda_s)-
\int_s^t \mathcal{B}_r\psi(\lambda_r) dr\Big).\]
We of course start from $\mathbb{E}[\mathcal{F}_{\epsilon_n}(X^{\epsilon_n})]=0$, where, recalling
\eqref{pro-generator},
\[\mathcal{F}_\epsilon(\lambda)=\Big(\prod_{i=1}^{k}\psi_i(\lambda_{s_i})\Big) \Big(\psi(\lambda_t)-\psi(\lambda_s)-\int_s^t \tilde{\mathcal{A}}_{r,\epsilon} \psi(\lambda_r) dr\Big).\]
We then write
\begin{align*}
\Big| \mathbb{E}[\mathcal{F}(X)] \Big| \le \Big |\mathbb{E}[\mathcal{F}(X)]-\mathbb{E}[\mathcal{F}(X^{\epsilon_n})] \Big |
+\Big |\mathbb{E}[\mathcal{F}(X^{\epsilon_n})]- \mathbb{E}[\mathcal{F}_{\epsilon_n}(X^{\epsilon_n})]\Big |.
\end{align*}
On the one hand,
we know from \cite[Lemma 3.3]{MR3313757} that $(x,v_*)\mapsto \mathcal{A} \psi(x,v_*)$ is continuous on
$\mathbb{R}^3\times\mathbb{R}^3$ and bounded by $C\, |x-v_*|^{\gamma+1}$.
We thus easily deduce that $\mathcal{F}$ is continuous at each $\lambda \in\mathbb{D}([0,\infty), \mathbb{R}^3)$
which does not jump at $s_1,..., s_k, s, t$ (this is a.s. the case of $X\in\mathbb{D}([0,\infty), \mathbb{R}^3)$
because it has no deterministic time jump by the Aldous criterion). We also
deduce that $|\mathcal{F} (\lambda)|\leq C(1+\int_0^t \int_{\mathbb{R}^3} |\lambda_r - v_*|^{\gamma+1} \tilde f_r(dv_*)dr)$.
Using that $0<\gamma+1<1$, that $\sup_{\epsilon\in(0,1)}\mathbb{E}[\sup_{[0,T]} |X^{\epsilon}_t|]<\infty$ by Step 4
and recalling that $X^{\epsilon_n}$ goes in law to $X$, we easily conclude that
$|\mathbb{E}[\mathcal{F}(X)]-\mathbb{E}[\mathcal{F}(X^{\epsilon_n})] |$ tends to $0$ as $n\to \infty$.
\vskip1mm
On the other hand, since $|\mathcal{F}(\lambda)-\mathcal{F}_\epsilon(\lambda)|\leq C
|\int_s^t (\mathcal{B}_r\psi(\lambda_r)-\tilde{\mathcal{A}}_{r,\epsilon}\psi(\lambda_r))dr|$
and $X^\epsilon_r\sim \tilde f^\epsilon_r$,
\begin{align*}
&\Big |\mathbb{E}[\mathcal{F}(X^{\epsilon_n})]- \mathbb{E}[\mathcal{F}_{\epsilon_n}(X^{\epsilon_n})]\Big |\\
\le& C \int_s^t \mathbb{E}\Big[ \Big |
\int_{\mathbb{R}^3}\int_0^\infty \int_0^{2\pi} \int_{\mathbb{R} ^3}\psi(X_r^{\epsilon_n}+c(v, v_*, z, \varphi))
\Big[\frac{ \phi_{\epsilon_n}(v-X_r^{\epsilon_n})}{\tilde f_r^{\epsilon_n}(X_r^{\epsilon_n})}\tilde f_r(dv) -
\boldsymbol{\delta}_{X_r^{\epsilon_n}}(dv)\Big]
d\varphi dz \tilde f_r(dv_*) \Big|\Big] dr \\
=&C \int_s^t \Big| \int_{\mathbb{R}^3}\int_0^\infty \int_0^{2\pi} \int_{\mathbb{R}^3} \int_{\mathbb{R} ^3}\psi(w+c(v, v_*, z, \varphi))
\Big[ \phi_{\epsilon_n}(v-w)\tilde f_r(dv)- \tilde f_r^{\epsilon_n}(w) \boldsymbol{\delta}_{w}(dv)\Big]
dw d\varphi dz \tilde f_r(dv_*) \Big| dr.
\end{align*}
But we can write $\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}\psi(w+c(v,v_*,z,\varphi))\tilde f_r^{\epsilon_n}(w) \boldsymbol{\delta}_{w}(dv)dw
=\int_{\mathbb{R}^3}\psi(w+c(w,v_*,z,\varphi))\tilde f_r^{\epsilon_n}(w)dw=\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}
\psi(w+c(w,v_*,z,\varphi))\phi_{\epsilon_n}(v-w)\tilde f_r(dv)dw$, so that
\begin{align*}
&\Big |\mathbb{E}[\mathcal{F}(X^{\epsilon_n})]- \mathbb{E}[\mathcal{F}_{\epsilon_n}(X^{\epsilon_n})]\Big |\\
\leq & C\int_s^t \Big|\int_{\mathbb{R}^3}\int_0^\infty \int_0^{2\pi} \int_{\mathbb{R}^3}\int_{\mathbb{R} ^3} \Big[ \psi(w+c(v, v_*, z, \varphi)) - \psi(w+c(w, v_*, z, \varphi))\Big]\\
&\hskip7cm \phi_{\epsilon_n}(v-w)\tilde f_r(dv) dw d\varphi dz \tilde f_r(dv_*) \Big| dr\\
=& C\int_s^t \Big|\int_{\mathbb{R}^3}\int_0^\infty \int_0^{2\pi} \int_{\mathbb{R}^3}\int_{\mathbb{R} ^3} \Big[ \psi(w+c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*))) \\
&\hskip 4cm - \psi(w+c(w, v_*, z, \varphi))\Big] \phi_{\epsilon_n}(v-w)\tilde f_r(dv) dw d\varphi dz \tilde f_r(dv_*)\Big | dr.
\end{align*}
The last equality uses the $2\pi$-periodicity of $c$.
We now put
$$
R_n(v,v_*,z,\varphi):=\int_{\mathbb{R} ^3} \Big[ \psi(w+c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*))) - \psi(w+c(w, v_*, z, \varphi))\Big] \phi_{\epsilon_n}(v-w) dw,
$$
and show the following two things:
\begin{enumerate}} \def\eenu{\end{enumerate}[label=(\alph*)]
\item for all $v, v_*\in\mathbb{R}^3$, all $z\in[0,\infty)$ and $\varphi\in[0,2\pi)$,
$\lim_{n\to\infty}R_n(v,v_*,z,\varphi)=0$;
\item there is a constant $C>0$ such that for all $n\geq 1$, all
$v, v_*\in\mathbb{R}^3$, all $z\in[0,\infty)$ and $\varphi\in[0,2\pi)$,
$$
|R_n(v,v_*,z,\varphi)|\le C\big(1+|v-v_*|\big)(1+z)^{-1/\nu},
$$
which belongs to
$L^1([0,T]\times\mathbb{R}^3\times\mathbb{R}^3\times[0,\infty)\times[0,2\pi),dr \tilde f_r(dv_*)\tilde f_r(dv)
dz d\varphi)$ because $(\tilde f_t)_{t\geq 0} \in L^\infty([0,T],\mathcal{P}_2(\mathbb{R}^3))$ by assumption.
\eenu
By dominated convergence, we will deduce that
$\lim_{n\rightarrow \infty}\Big |\mathbb{E}[\mathcal{F}(X^{\epsilon_n})]- \mathbb{E}[\mathcal{F}_{\epsilon_n}(X^{\epsilon_n})]\Big |=0$
and this will conclude the proof.
\vskip1mm
We first study (a). Since $\psi\in C^1_b(\mathbb{R}^3)$, we immediately observe that
\begin{align}\label{startagain}
& \Big| \psi(w+c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*))) - \psi(w+c(w, v_*, z, \varphi))\Big | \\
&\le C_\psi \Big |c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*)) - c(w, v_*, z, \varphi)\Big|.\notag
\end{align}
Recalling that
\[c(v,v_*,z, \varphi)=-\frac{1-\cos G(z/|v-v_*|^\gamma)}{2}(v-v_*)+\frac{ \sin G(z/|v-v_*|^\gamma)) }{2} \Gamma(v-v_*, \varphi) ,\]
we have
\begin{align*}
& \Big |c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*)) - c(w, v_*, z, \varphi)\Big|\\
\le& \frac{|\cos G(z/|v-v_*|^\gamma) - \cos G(z/|w-v_*|^\gamma)|} {2} |v-v_*|+ \frac{ |1 - \cos G(z/|w-v_*|^\gamma)|}{2} |v-w|\\
& + \frac{|\sin G(z/|v-v_*|^\gamma) - \sin G(z/|w-v_*|^\gamma)|}{2} |\Gamma(v-v_*, \varphi+\varphi_0)| \\
& + \frac{|\sin G(z/|w-v_*|^\gamma)|}{2}|\Gamma(v-v_*, \varphi+\varphi_0) - \Gamma(w-v_*, \varphi)|.
\end{align*}
Using that $|\Gamma(v-v_*, \varphi+\varphi_0)|=|v-v_*|$ and Lemma \ref{Tanaka}, we obtain
\begin{align*}
& \Big |c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*)) - c(w, v_*, z, \varphi)\Big|\\
\le& C |G(z/|v-v_*|^\gamma) - G(z/|w-v_*|^\gamma)||v-v_*|+ C |v-w|.
\end{align*}
We deduce from \eqref{nota} that $|G^\prime(z)| =1/\beta(G(z)) \leq C$ by \eqref{con}, whence
\begin{align*}
& \Big |c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*)) - c(w, v_*, z, \varphi)\Big|\\
&\le C z \big | |v-v_*|^{|\gamma|} - |w-v_*|^{|\gamma|}\big| |v-v_*|+ C |v-w|.
\end{align*}
Using again the inequality $|x^{\alpha} - y^{\alpha}|\le|x-y|(x\vee y)^{\alpha-1}$ for $\alpha\in(0,1)$, and $x,y\ge0$, we have
\[\big | |v-v_*|^{|\gamma|} - |w-v_*|^{|\gamma|}\big| \le |v-w| |v-v_*|^{|\gamma|-1}.\]
We thus get
\[\Big |c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*)) - c(w, v_*, z, \varphi)\Big|
\le C (z |v-v_*|^{|\gamma|}+1) |v-w|.
\]
Consequently,
\begin{align*}
R_n(v,v_*,z,\varphi)
\le &C_\psi (z |v-v_*|^{|\gamma|}+1) \int_{\mathbb{R} ^3} |v-w| \phi_{\epsilon_n}(v-w) dw,
\end{align*}
which clearly tends to $0$ as $n\to \infty$.
This ends the proof of (a).
\vskip1mm
For (b), start again from \eqref{startagain} to write
\begin{align*}
& \Big| \psi(w+c(v, v_*, z, \varphi+\varphi_0(v-v_*, w-v_*))) - \psi(w+c(w, v_*, z, \varphi))\Big | \\
&\le\Big| \psi(w+c(v, v_*, z, \varphi)) -\psi(w) \Big |+ \Big |\psi(w)- \psi(w+c(w, v_*, z, \varphi))\Big | \\
&\le C_\psi ( |c(v, v_*, z, \varphi)| + |c(w, v_*, z, \varphi)|).
\end{align*}
Moreover, since $|c(v, v_*, z, \varphi)|\le G(z/|v-v_*|^\gamma)|v-v_*| \leq C |v-v_*| (1+ |v-v_*|^{|\gamma|}z)^{-1/\nu}$
by \eqref{num} and \eqref{con1}, we observe that
\begin{align*}
R_n(v,v_*,z,\varphi)\leq C |v-v_*| (1+ |v-v_*|^{|\gamma|}z)^{-1/\nu} + C \int_{\mathbb{R} ^3} |w-v_*|
(1+ |w-v_*|^{|\gamma|}z)^{-1/\nu}\phi_{\epsilon_n}(v-w) dw.
\end{align*}
Since $1+ |v-v_*|^{|\gamma|}z\ge \big(1\wedge |v-v_*|^{|\gamma|}\big)(1+z)$ for $z\in[0,\infty)$,
\begin{align*}
|v-v_*| (1+ |v-v_*|^{|\gamma|}z)^{-1/\nu} &\le |v-v_*|(1+z)^{-1/\nu}\big(1\wedge |v-v_*|^{|\gamma|} \big)^{-1/\nu}.
\end{align*}
Using that $|\gamma|/\nu<1$, we deduce that
$$
|v-v_*| (1+ |v-v_*|^{|\gamma|}z)^{-1/\nu} \le \big(1+|v-v_*| \big) (1+z)^{-1/\nu}.
$$
As a conclusion,
$$
R_n(v,v_*,z,\varphi)\leq C \Big(1+|v-v_*|+\int_{\mathbb{R} ^3} |w-v_*| \phi_{\epsilon_n}(v-w) dw \Big) (1+z)^{-1/\nu},
$$
which is easily bounded (recall that $\epsilon_n\in (0,1)$)
by $C(1+|v|+|v_*|)(1+z)^{-1/\nu}$ as desired.
\epf
\section{The coupling}
To get the convergence of the particle system, we construct a suitable coupling between the particle
system with generator $\mathcal{L}_{N,K}$ defined by \eqref{rewpartopera}
and the realization of the weak solution to \eqref{Bol}, following the ideas of \cite{MR3456347}.
\begin{lem}\def\elem{\end{lem}}\label{coupling-systems}
Assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ with $\gamma+\nu\in(0,1)$.
Let $N\ge1$ be fixed. Let $q\ge2$ such that $q>\gamma^2/(\gamma+\nu)$. Let $f_0\in\mathcal{P}_q(\mathbb{R}^3)$
with a finite entropy and let $(f_t)_{t\ge0} \in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)
\cap L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$ (with $p\in(3/(3+\gamma),p_0(\gamma,\nu,q))$)
be the unique weak solution to \eqref{Bol} given by
Theorem \emph{ \ref{well-posedness}}. Then there exists, on some probability space, a family of i.i.d. random
variables $(V_0^i)_{i=1,...,N}$ with common law $f_0$, independent of a family of i.i.d. Poisson measures
$(M_i(ds, d\alpha, dz, d\varphi))_{i=1,...,N}$ on $[0,\infty)\times[0,1]\times[0,\infty)\times[0,2\pi)$, with
intensity $ds d\alpha dz d\varphi$, a measurable family $(W^*_t)_{t\geq 0}$ of $\alpha$-random variables with
$\alpha$-law $(f_t)_{t\ge0}$ and $N$ i.i.d. c\`adl\`ag
adapted processes $(W_t^i)_{t\ge0}$ solving, for each $i=1, \cdots, N$,
\begin{equation}}\def\eeq{\end{equation}\label{B-particle}
W_t^i=V_0^i+\int_0^t\int_0^1\int_0^\infty\int_0^{2\pi}c(W_{s-}^{i},W_s^*(\alpha),z,\varphi)M_i(ds,d\alpha,dz,d\varphi).
\eeq
Moreover, $W_t^i \sim f_t$ for all $t\ge0$, all $i=1,\dots,N$.
Also, for all $T>0$,
\begin{equation}}\def\eeq{\end{equation}\label{lem-momen}
\mathbb{E}\big[\sup_{[0,T]} |W_t^{1}|^q\big]\le C_{T,q}.
\eeq
\elem
\begin{proof}}\def\epf{\end{proof}
Except for the moment estimate \eqref{lem-momen}, it suffices to apply Proposition \ref{prop-PDE}.
A simpler proof could be handled here because we deal with the {\it strong} solution
$f \in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)
\cap L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$.
We now prove \eqref{lem-momen}, which is more or less classical. We thus fix $q\geq 2$.
It is clear that
\begin{align*}
\big||v+c(v,v_*,z,\varphi)|^q-|v|^q\big| \le C_q \big( |v|^{q-1}+|c(v,v_*,z,\varphi)|^{q-1} \big) |c(v,v_*,z,\varphi)| .
\end{align*}
Due to \eqref{num} and \eqref{con1},
$|c(v,v_*,z,\varphi)|\le |v-v_*|$, $|c(v,v_*,z,\varphi)|\le (1+z/|v-v_*|^\gamma)^{-1/\nu}|v-v_*|$, whence
\begin{align}\label{q-increment}
&\int_0^\infty \int_0^{2\pi} \big||v+c(v,v_*,z,\varphi)|^q-|v|^q\big| d\varphi dz \notag \\
&\le C_q \int_0^\infty \int_0^{2\pi} \big( 1+|v|^{q-1}+|v_*|^{q-1}\big) (1+z/|v-v_*|^\gamma)^{-1/\nu}
|v-v_*| d\varphi dz \notag \\
&= C_q \big( 1+|v|^{q-1}+|v_*|^{q-1}\big) |v-v_*|^{1+\gamma} \notag\\
&\le C_q \big( 1+|v|^{q}+|v_*|^{q}\big),
\end{align}
because $0<1+\gamma < 1$. It then easily
follows from the It\^{o} formula and $\mathcal{L}_\alpha(W_t^*)=f_t={\mathcal L}(W_t^1)$ that
\begin{align*}
\mathbb{E}\big[\sup_{[0,t]}|W_s^{1}|^q\big]
&\le \mathbb{E}[|V_0^{1}|^q] + C_q \int_0^t\int_0^1\mathbb{E}\Big[ 1+|W_s^{1}|^q+|W_s^*(\alpha)|^q \Big] d\alpha ds\\
&\le \mathbb{E}[|V_0^{1}|^q] + C_q \int_0^t \Big(1+\mathbb{E}[\sup_{[0,s]}|W_u^{1}|^q]\Big) ds.
\end{align*}
We thus conclude \eqref{lem-momen} by the Gr\"{o}nwall lemma.
\epf
Next, let us recall \cite[Lemma 4.3]{MR3456347} below in order to construct our coupling.
\begin{lem}\def\elem{\end{lem}}\label{coupling}
Consider $(f_t)_{t\geq 0}$ and $(W^*_t)_{t\geq 0}$ introduced in Lemma \emph{\ref{coupling-systems}}
and fix $N\ge1$.
For ${\bf v}=(v_1, v_2, ...,v_N)\in(\mathbb{R}^3)^N$,
we introduce the empirical measure $\mu_{{\bf v}}^N:=N^{-1}\sum_{i=1}^{N}\delta_{v_i}$.
Then for all $t\ge0$, all ${\bf v}\in (\mathbb{R}^3)^N$ and all
${\bf w}\in (\mathbb{R}^3)^N_{\bullet}$, with $(\mathbb{R}^3)^N_{\bullet}:=\{{\bf w}\in(\mathbb{R}^3)^N: w_i\ne w_j \; \forall \,
i\ne j \}$,
there are $\alpha$-random variables
$Z_t^*({\bf w},\alpha)$ and $V_t^*({\bf v},{\bf w},\alpha)$ such that the $\alpha$-law of $(Z_t^*({\bf w},\cdot), V_t^*({\bf v},{\bf w},\cdot))$ is $N^{-1}\sum_{i=1}^{N}\delta_{(w_i, v_i)}$ and $\int_0^1|W_t^*(\alpha)-Z_t^*({\bf w},\alpha)|^2 d\alpha=\mathcal{W}_2^2(f_t, \mu_{{\bf w}}^N)$.
\elem
Observe that $\mathcal{L}_\alpha(Z_t^*({\bf w},\cdot))=\mu_{{\bf w}}^N$ and
$\mathcal{L}_\alpha(V_t^*({\bf v},{\bf w},\cdot))=\mu_{{\bf v}}^N$.
Owing to technical reasons, we need to introduce some more notations.
\begin{nota}\label{nnn}
We consider an $\alpha$-random variable $Y$ with uniform distribution on $B(0,1)$ (independent of everything else)
and, for $\epsilon\in (0,1)$, $t\ge0$,
$\alpha\in[0,1]$, ${\bf v}\in (\mathbb{R}^3)^N$ and ${\bf w}\in (\mathbb{R}^3)^N_{\bullet}$, we set
$W_t^{*,\epsilon}(\alpha)=W_t^*(\alpha)+\epsilon Y(\alpha)$ and $V_t^{*,\epsilon}({\bf v},{\bf w},\alpha)=
V_t^*({\bf v},{\bf w},\alpha)+\epsilon Y(\alpha)$.
It holds that $\mathcal{L}_\alpha(W_t^{*,\epsilon})=f_t*\psi_\epsilon$ and $\mathcal{L}_\alpha(V_t^{*,\epsilon}({\bf v},{\bf w},\cdot))
=\mu_{{\bf v}}^N*\psi_\epsilon$, where $\psi_\epsilon(x)=(3/(4\pi\epsilon^3))\mathbf{1}_{\{|x|\le \epsilon\}}$.
\end{nota}
At last, we built a suitable realisation for the particle system.
\begin{lem}\def\elem{\end{lem}}\label{aaaa}
Consider all the objects introduced in Lemmas \emph{\ref{coupling-systems}}-\emph{\ref{coupling}} and Notation \emph{\ref{nnn}}.
Set ${\bf {W}}_{s}=(W_{s}^{1},...,W_{s}^{N})$, which a.s. belongs to $(\mathbb{R}^3)^N_\bullet$
(because $f_s$ has a density for all $s\geq 0$). Fix $K\geq 1$ and $\epsilon\in (0,1)$.
There is a unique strong solution
$({\bf {V}}_{t})_{t\geq 0}=(V_{t}^{1}, ... , V_{t}^{N})_{t\geq 0}$ to
\begin{equation}}\def\eeq{\end{equation}\label{particle-system}
V_t^{i}=V_0^i+\int_0^t\int_0^1\int_0^\infty\int_0^{2\pi}c_K(V_{s-}^{i}, V_s^*({\bf {V}}_{s-}, {\bf {W}}_{s-}, \alpha), z, \varphi+\varphi_{i,\alpha,s})M_i(ds,d\alpha,dz,d\varphi), ~ i=1,..., N,
\eeq
where $\varphi_{i,\alpha,s}:=\varphi_{i,\alpha,s}^1+\varphi_{i,\alpha,s}^2+\varphi_{i,\alpha,s}^3$ with
\begin{align*}
\varphi_{i,\alpha,s}^1=&\varphi_{0}\big(W_{s-}^{i} - W_s^*(\alpha), W_{s-}^{i} - W_s^{*,\epsilon}(\alpha)\big),\\
\varphi_{i,\alpha,s}^2=&\varphi_{0}\big(W_{s-}^{i} - W_s^{*,\epsilon}(\alpha),
V_{s-}^{i} - V_s^{*,\epsilon}({\bf {V}}_{s-}, {\bf {W}}_{s-}, \alpha)\big),\\
\varphi_{i,\alpha,s}^3=&\varphi_{0}\big( V_{s-}^{i} - V_s^{*,\epsilon}({\bf {V}}_{s-}, {\bf {W}}_{s-}, \alpha),
V_{s-}^{i} - V_s^*({\bf {V}}_{s-}, {\bf {W}}_{s-}, \alpha)\big).
\end{align*}
Moreover, $({\bf {V}}_{t})_{t\geq 0}$ is a Markov process with generator
$\mathcal{L}_{N,K}$.
If $f_0\in\mathcal{P}_q(\mathbb{R}^3)$
for some $q\ge2$, then $\mathbb{E}\big[\sup_{[0,T]}|V_t^{1}|^q\big]\le C_{T,q}$
(this last constant not depending on $N,K$ nor $\epsilon\in(0,1)$).
\elem
\begin{proof}}\def\epf{\end{proof}
Since $c_K=c\mathbf{1}_{\{z\le K\}}$, the Poisson measures involved in \eqref{particle-system} are finite.
Hence the existence and uniqueness results hold for \eqref{particle-system}.
Using that $\mathcal{L}_\alpha(V_t^*({\bf v},{\bf w},\cdot))=\mu_{{\bf v}}^N$
and the $2\pi$-periodicity of $c_K$ in $\varphi$, one easily checks that $({\bf {V}}_{t})_{t\geq 0}$ is a
Markov process with generator $\mathcal{L}_{N,K}$: for all bounded measurable function
$\phi:(\mathbb{R}^3)^N\mapsto \mathbb{R}$, all $t\geq 0$, a.s.,
\begin{align*}
&\sum_{i=1}^N \int_0^1\int_0^\infty\int_0^{2\pi}\Big[\phi({\bf v}+c_K(v_i, V_t^*({\bf v},{\bf w},\alpha),z,\varphi+
\varphi_{i,\alpha,t} ){\bf e}_i)-\phi({\bf v})\Big] d\varphi dz d\alpha\\
=&\sum_{i=1}^N N^{-1}\sum_{j=1}^N \int_0^\infty\int_0^{2\pi}\Big[\phi({\bf v}+c_K(v_i, v_j, z,\varphi){\bf e}_i)-\phi({\bf v})\Big] d\varphi dz \\
=&N^{-1}\sum_{i \ne j} \int_0^\infty\int_0^{2\pi}\Big[\phi({\bf v}+c_K(v_i, v_j, z,\varphi){\bf e}_i)-\phi({\bf v})\Big] d\varphi dz,
\end{align*}
because $c_K(v_i,v_i,z,\varphi)=0$. This is nothing but $\mathcal{L}_{N,K}\phi({\bf v})$, recall Lemma \ref{llxx}.
\vskip1mm
Finally, we verify that $\sup_{[0,T]}\mathbb{E}\big[|V_t^{1}|^q\big]\le C_{T,q}$ if $f_0\in\mathcal{P}_q(\mathbb{R}^3)$
for some $q\ge2$: it immediately follows from the It\^o formula, \eqref{q-increment} and exchangeability
that
\begin{align*}
\mathbb{E}\big[|V_t^{1}|^q\big]
&\le \mathbb{E}[|V_0^{1}|^q] + C_q \int_0^t\int_0^1\mathbb{E}\Big[1+|V_s^{1}|^q+|V_s^*({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^q\Big] \ d\alpha ds\\
&\le \mathbb{E}[|V_0^{1}|^q] + C_q N^{-1}\sum_{i=1}^{N} \int_0^t\mathbb{E}\Big[1+|V_s^{1}|^q+|V_s^{i}|^q\Big] \ ds\\
&\le \mathbb{E}[|V_0^{1}|^q] + C_q \int_0^t\mathbb{E}\Big[1+|V_s^{1}|^q\Big] \ ds,
\end{align*}
The Gr\"{o}nwall lemma allows us to complete the proof.
\epf
\begin{rem}\def\erem{\end{rem}}
The exchangeability holds for the family $\{(W_t^i, V_t^{i})_{t\ge 0}, i=1,...,N\}$.
Indeed, the family $\{(W_t^i)_{t\ge 0}, i=1,...,N\}$ is i.i.d. by construction, so that the
exchangeability follows from the symmetry and pathwise uniqueness for \eqref{particle-system}.
\erem
\section{Bound of the Lebesgue norm of an empirical measure}\label{ttt}
An empirical measure cannot be in some $L^p$ space with $p>1$, so we will consider
a blob approximation, inspired by \cite[Proposition 5.5]{Fournier:2015aa} and \cite{MR3377068}.
But we deal with a jump process, so we need to overcome a few additional difficulties.
\vskip1mm
First, the following fact can be checked as
Lemma 5.3 in \cite{Fournier:2015aa} (the norm and the step of the subdivision are different,
but this obviously changes nothing).
\begin{lem}\def\elem{\end{lem}}\label{dicrete-norm}
Let $p\in(1,2)$ and $(f_t)_{t\geq 0} \in L^\infty \big([0,\infty),{\cal P}_2({\mathbb{R}^3})\big)\cap
L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$
such that $m_2(f_t)=m_2(f_0)$ for all $t\geq 0$.
\begin{enumerate}} \def\eenu{\end{enumerate}[label=\emph{(\roman*)}]
\item There is a constant $M_p>0$, such that for all $t\ge0$, $\|f_t\|_{L^p}\ge M_p$.
\item For any $T>0$, we can find a subdivision $(t_{\ell}^N)_{\ell=0}^{K_N+1}$ satisfying $0=t_0^N<t_1^N<\dots<t_{K_N}^N \le T \le t_{K_N+1}^N$, such that $\sup_{\ell=0,...,K_N}(t_{\ell+1}^N-t_\ell^N)\le N^{-2}$ with $K_N \le 2TN^{2}$
and
\[ \int_0^Th_N(t)dt\le 2\int_0^T\|f_t\|_{L^p} dt, \]
with $h_N(t)=\sum_{\ell=1}^{K_N+1}\|f_{t_\ell^N}\|_{L^p}\mathbf{1}_{\{t\in(t_{\ell-1}^N, t_\ell^N]\}}$.
\eenu
\elem
The goal of the section is to prove the following crucial fact.
\begin{prop}\def\eprop{\end{prop}}\label{norm-bound}
Assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ with $\gamma+\nu>0$.
Let $q\geq 2$ such that $q>\gamma^2/(\gamma+\nu)$ and let $p\in(3/(3+\gamma),p_0(\gamma,\nu,q))\subset
(1,3/2)$. Consider $f_0 \in \mathcal{P}_q(\mathbb{R}^3)$ with a finite entropy and
$(f_t)_{t\ge0}\in L^\infty \big([0,\infty),{\cal P}_2({\mathbb{R}^3})\big)\cap
L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$
the corresponding unique solution to \eqref{Bol} given by Theorem \emph{\ref{well-posedness}}.
Consider $(W_t^i)_{i=1,...,N, t\ge 0}$ the solution to \eqref{B-particle} and set
$\mu_{{\bf {W}}_t}^N=N^{-1}\sum_1^N \delta_{W^i_t}$.
Fix $\delta\in(0,1)$, set $\epsilon_N=N^{-(1-\delta)/3}$ and define
$\bar\mu_{{\bf {W}}_t}^N=\mu_{{\bf {W}}_t}^N*\psi_{\epsilon_N}$, where $\psi_\epsilon$ was defined in Notation
\emph{\ref{nnn}}. Finally, fix $T>0$ and consider $h_N$ built in Lemma \emph{\ref{dicrete-norm}}.
We have
$$
\mathbb{P}\Big(\forall t\in[0,T], \|\bar\mu_{{\bf {W}}_t}^N\|_{L^p}\le 13500 \big(1+h_N(t)\big)\Big)\ge 1- C_{T,q,\delta}N^{1-\delta q/3}.
$$
\eprop
Throughout the section, we fix $N\ge 1$, $\delta\in(0,1)$, and $\epsilon_N=N^{-(1-\delta)/3}$ and
adopt the assumptions and notations of Proposition \ref{norm-bound}. We also put $r=p/(p-1)$.
\vskip1mm
In order to extend \cite[Proposition 5.5]{Fournier:2015aa}, it is necessary to study some
properties of the paths of the processes defined by \eqref{B-particle}.
Following Lemma 3.11 in \cite{Xu:2015aa}, we introduce, for each $i=1,\dots,N$,
\begin{align}\label{newprocess}
\widetilde{W}_t^i &=V_0^i+\int_0^t\int_0^1\int_0^\infty\int_0^{2\pi}c(W_{s-}^i,W_s^*(\alpha),z,\varphi)\mathbf{1}_{\left\{|c(W_{s-}^i,W_s^*(\alpha),z,\varphi)|\le N^{-1/3}\right\}}M_i(ds,d\alpha,dz,d\varphi).
\end{align}
\begin{lem}\def\elem{\end{lem}}\label{lem-small-jump}
For all $T>0$,
$$
\mathbb{P}\Big[\sup_{[0,T]}|W_t^1|\le N^{\delta/3}, \sup_{s,t\in[0,T], |s-t|\le N^{-2}}|\widetilde{W}_t^1 - \widetilde{W}_s^1 |\ge \epsilon_N\Big]\le C_{T}N^2 e^{-N^{\delta/3}}.
$$
\elem
\begin{proof}}\def\epf{\end{proof}
Let us denote by $\tilde p$ the probability we want to bound.
\vskip1mm
{\it Step 1.}
We introduce
\begin{align*}
Z_t^1 &=\int_0^t\int_0^1\int_0^\infty\int_0^{2\pi} G\big(z/|W_{s-}^1-W_s^*(\alpha)|^\gamma\big)|W_{s-}^1-W_s^*(\alpha)| \\
& \hskip 5cm \times \mathbf{1}_{\left\{G\big(z/|W_{s-}^1 - W_s^*(\alpha)|^\gamma\big)|W_{s-}^1 -W_s^*(\alpha)|/4\le N^{-1/3}\right\}} M_1(ds,d\alpha,dz,d\varphi).
\end{align*}
It is clear that $Z_t^1$ is almost surely increasing in $t$, and that a.s., for all $s, t\in[0, T]$,
\begin{equation}}\def\eeq{\end{equation}\label{pro-inequality}|\widetilde{W}_t^1 -\widetilde{W}_s^1 |\le |Z_t^1-Z_s^1|,
\eeq
since for any $v, v_*\in\mathbb{R}^3$ (recall \eqref{num})
$$
G\big(z/|v-v_*|^\gamma\big)|v-v_*|/4\le|c(v, v_* ,z,\varphi)|\le G\big(z/|v-v_*|^\gamma\big)|v-v_*|.$$
We now consider the stopping time $\tau_N=\inf{\{t\ge0: |W_t^1|>N^{\delta/3}\}}$ and deduce from \eqref{pro-inequality} and the Markov inequality that
\begin{align*}
\tilde p
&\le\mathbb{P}\Big[\sup_{[0,T]}|W_t^1|\le N^{\delta/3}, \sup_{s,t\in[0,T], |s-t|\le N^{-2}}|Z_t^1-Z_s^1|\ge \epsilon_N\Big]\\
&\le \mathbb{P}\Big[\sup_{s,t\in[0,T], |s-t|\le N^{-2}}|Z_{t\wedge\tau_N}^1-Z_{s\wedge\tau_N}^1|\ge \epsilon_N\Big].
\end{align*}
Since $[0, T] \subset \bigcup_{k=0}^{\lfloor N^{2}T\rfloor} [k/N^{2}, (k+1)/N^{2})$ and $Z_t^N$ is almost surely increasing in $t$, we deduce that on $\{\sup_{s,t\in[0,T], |s-t|\le N^{-2}}|Z_{t\wedge\tau_N}^1-Z_{s\wedge\tau_N}^1|\ge \epsilon_N\}$, there exists $k\in\{0,1,...,\lfloor N^{2}T\rfloor \}$ for which there holds $\big(Z_{((k+1)N^{-2})\wedge\tau_N}^1-Z_{(kN^{-2})\wedge\tau_N}^1\big)\ge \epsilon_N/3$. Hence,
\begin{align*}
\tilde p \le & \sum_{k=0}^{\lfloor N^{2}T\rfloor}\mathbb{P}\left[\Big(Z_{((k+1)N^{-2})\wedge\tau_N}^1-Z_{(kN^{-2})\wedge\tau_N}^1\Big)\ge
\frac{\epsilon_N}{3}\right]\\
&\le \sum_{k=0}^{\lfloor N^{2}T\rfloor}e^{-N^{\delta/3}}\mathbb{E}\left[\exp{\left\{3N^{1/3}\Big(Z_{((k+1)N^{-2})\wedge\tau_N}^1-Z_{(kN^{-2})\wedge\tau_N}^1\Big)\right\}}\right]\\
&=:\sum_{k=0}^{\lfloor N^{2}T\rfloor}e^{-N^{\delta/3}}I_k.
\end{align*}
{\it Step 2.} We now prove that $I_k$ is (uniformly) bounded, which will complete the proof.
We put
\[J_k(t)=:\mathbb{E}\left[\exp{\left\{3N^{1/3}\Big(Z_{(t+kN^{-2})\wedge\tau_N}^1-Z_{(kN^{-2})\wedge\tau_N}^1\Big)\right\}}\right].\]
It is obvious that $I_k=J_k(N^{-2})$. Applying the It\^{o} formula, we find
\begin{align*}
J_k(t)
&=1+2\pi \mathbb{E}\bigg[\int_{(kN^{-2})\wedge\tau_N}^{(t+kN^{-2})\wedge\tau_N} \int_0^1 \int_0^\infty \exp{\left\{3N^{1/3}\Big(Z_s^1-Z_{(kN^{-2})\wedge\tau_N}^1\Big)\right\}} \\
&\quad \times \Big(e^{3N^{1/3}G\big(z/|W_{s}^1-W_s^*(\alpha)|^\gamma\big)|W_{s}^1-W_s^*(\alpha)|}-1\Big) \mathbf{1}_{\left\{G\big(z/|W_{s}^1 - W_s^*(\alpha)|^\gamma\big)|W_{s}^1 -W_s^*(\alpha)|/4\le N^{-1/3}\right\}} dz d\alpha ds\bigg].
\end{align*}
Since $3N^{1/3}G\big(z/|W_{s}^1-W_s^*(\alpha)|^\gamma\big)|W_{s}^1-W_s^*(\alpha)|\le 12$
(thanks to the indicator function), we have
$$
e^{3N^{1/3}G\big(z/|W_{s}^1-W_s^*(\alpha)|^\gamma\big)|W_{s}^1-W_s^*(\alpha)|}-1 \le CN^{1/3}G\big(z/|W_{s}^1-W_s^*(\alpha)|^\gamma\big)|W_{s}^1-W_s^*(\alpha)|
$$ for a positive constant $C$. Then using \eqref{con1}, we see that
$$
\mathbf{1}_{\left\{G\big(z/|W_{s}^1 - W_s^*(\alpha)|^\gamma\big)|W_{s}^1 -W_s^*(\alpha)|/4\le N^{-1/3}\right\}} \le \mathbf{1}_{\left\{z \ge CN^{\nu/3}|W_{s}^1-W_s^*(\alpha)|^{\gamma+\nu}-|W_{s}^1-W_s^*(\alpha)|^{\gamma}\right\}}.
$$
Hence,
\begin{align*}
J_k(t) &\le 1+ CN^{1/3}\mathbb{E}\bigg[\int_{(kN^{-2})\wedge\tau_N}^{(t+kN^{-2})\wedge\tau_N} \int_0^1 \int_0^\infty \exp{\left\{3N^{1/3}\Big(Z_s^1-Z_{(kN^{-2})\wedge\tau_N}^1\Big)\right\}} \\
& \quad \times \Big(1+z/|W_{s}^1 - W_s^*(\alpha)|^\gamma\Big)^{-1/\nu} |W_{s}^1 -W_s^*(\alpha)| \mathbf{1}_{\left\{z \ge CN^{\nu/3}|W_{s}^1-W_s^*(\alpha)|^{\gamma+\nu}-|W_{s}^1-W_s^*(\alpha)|^{\gamma}\right\}} dz d\alpha ds\bigg].
\end{align*}
But, we have
\begin{align*}
&|W_{s}^1-W_s^*(\alpha)| \int_0^\infty \Big(1+z/|W_{s}^1 - W_s^*(\alpha)|^\gamma\Big)^{-1/\nu} \mathbf{1}_{\left\{z \ge CN^{\nu/3}|W_{s}^1-W_s^*(\alpha)|^{\gamma+\nu}-|W_{s}^1-W_s^*(\alpha)|^{\gamma}\right\}} dz\\
=& C N^{-(1-\nu)/3}|W_{s}^1-W_s^*(\alpha)|^{\nu+\gamma} \\
\le& CN^{-(1-\nu)/3}(1+|W_{s}^1|^2+|W_s^*(\alpha)|^2 )
\end{align*}
since $\gamma+\nu\in(0,1)$.
Using now that $\int_0^1|W_s^*(\alpha)|^2 d\alpha=m_2(f_0)$ and that $|W_s^1|\le N^{\delta/3}$ for all $s\le \tau_N$,
we conclude that
\begin{align*}
J_k(t)&\le 1+CN^{\nu/3}(1+m_2(f_0)+N^{2\delta/3})\int_0^t J_k(s)ds\\
&\le 1+CN^{(\nu+2\delta)/3}\int_0^t J_k(s)ds.
\end{align*}
It follows from the Gr\"{o}nwall lemma that $J_k(t)\le \exp{(CN^{(\nu+2\delta)/3}t)}$, and thus
that $I_k=J_k(N^{-2})$ is uniformly bounded, because $(\nu+2\delta)/3< 2$
(recall that $\nu\in(0,1)$ and $\delta\in(0,1)$).
\epf
Next, we study the {\it large} jumps of $(W_t^1)_{t\geq 0}$.
\begin{lem}\def\elem{\end{lem}}\label{lem-big-jump}
There exists $C>0$ such that for any $\ell\in\{1,..., K_N+1\}$,
$$
\mathbb{P}\Big[\exists~ t\in(t_{\ell-1}^N, t_{\ell}^N]: |\Delta W^1_t|>N^{-1/3}\Big] \le CN^{-(2-\nu/3)}.
$$
\elem
\begin{proof}}\def\epf{\end{proof}
Let us fix $\ell$ and set $A=\{\exists~ t\in(t_{\ell-1}^N, t_{\ell}^N]: |\Delta W^1_t|>N^{-1/3}\}$.
After noting that
\begin{align*}
A&=\left\{\int_{t_{\ell-1}^N}^{t_\ell^N}\int_0^1\int_0^\infty\int_0^{2\pi}\mathbf{1}_{\left\{|c(W_{s-}^i,W_s^*(\alpha),z,\varphi)|> N^{-1/3}\right\}}M_1(ds, d\alpha, dz, d\varphi)\ge 1\right\},
\end{align*}
we have
\begin{align*}
\mathbb{P}(A)
&\le \mathbb{E}\left[\int_{t_{\ell-1}^N}^{t_\ell^N}\int_0^1\int_0^\infty\int_0^{2\pi}\mathbf{1}_{\left\{|c(W_{s-}^1,W_s^*(\alpha),z,\varphi)|> N^{-1/3}\right\}}M_1(ds, d\alpha, dz, d\varphi) \right]
\end{align*}
by the Markov inequality.
But, \eqref{num} and \eqref{con1} tell us that $|c(v,v_*,z, \varphi)|
\leq C(1+z/|v-v_*|^\gamma)^{-1/\nu}|v-v_*|$. Hence,
\begin{align*}
\mathbb{P}(A)
& \le 2\pi \mathbb{E}\Big[\int_{t_{\ell-1}^N}^{t_\ell^N}\int_0^1\int_0^\infty \mathbf{1}_{\left\{C ( 1+z/|W_s^1 -W_s^*(\alpha)|^\gamma)^{-1/\nu} |W_s^1-W_s^*(\alpha)|> N^{-1/3}\right\}} dz d\alpha ds\Big]\\
& \le 2\pi \mathbb{E}\Big[\int_{t_{\ell-1}^N}^{t_\ell^N}\int_0^1\int_0^\infty \mathbf{1}_{\left\{ z< CN^{\nu/3} |W_s^1-W_s^*(\alpha)|^{\gamma+\nu} \right\}} dz d\alpha ds\Big]\\
&= C N^{\nu/3} \mathbb{E}\Big[\int_{t_{\ell-1}^N}^{t_\ell^N}\int_0^1|W_s^1-W_s^*(\alpha)|^{\gamma+\nu} d\alpha ds\Big].
\end{align*}
Finally, using that $|W_s^1-W_s^*(\alpha)|^{\gamma+\nu}\le 1+|W_s^1|^2+|W_s^*(\alpha)|^2$
and that $\int_0^1|W_s^*(\alpha)|^2 d\alpha=\mathbb{E}[|W_s^1|^2]<\infty$,
we conclude that $\mathbb{P}(A)\leq CN^{\nu/3}(t_{\ell+1}^N-t_\ell^N)\le C N^{\nu/3-2}$ as desired.
\epf
\begin{lem}\def\elem{\end{lem}}\label{event}
For $\ell=1, ..., K_N+1$, we introduce
\begin{equation}}\def\eeq{\end{equation}\label{Number}
I_{\ell}=\{i\in\{1,...,N\}: \exists~ t\in (t_{\ell-1}^N, t_{\ell}^N] ~\text{such that}~ |\Delta W_t^i|>N^{-1/3}\},
\eeq
and the event
\begin{align*}
\Omega_{T,N}^1= & \left\{\forall i\in\{1,..,N\},\; \sup_{[0,T]}|W_t^i|\le N^{\delta/3}
\text{ and }
\sup_{s,t\in[0,T], |s-t|\le N^{-2}}|\widetilde{W}_t^i - \widetilde{W}_s^i|\le \epsilon_N\right\}\\
&\ \bigcap \, \left\{\forall \ell=1,...,K_N+1, \;\#(I_\ell) \le N \epsilon_N^{3/r}\right\}.
\end{align*}
Then we have
\[\mathbb{P}[\Omega_{T,N}^1] \ge 1-C_{T,q,\delta}N^{1-q\delta/3}.\]
\elem
\begin{proof}}\def\epf{\end{proof}
We write $\Omega_{T,N}^1=\Omega_{T,N}^{1,1} \cap \Omega_{T,N}^{1,2}$, where
\begin{align*}
\Omega_{T,N}^{1,1} &:= \left\{\forall i\in\{1,..,N\},\; \sup_{[0,T]}|W_t^i|\le N^{\delta/3}\text{ and } \sup_{s,t\in[0,T], |s-t|\le N^{-2}}|\widetilde{W}_t^i - \widetilde{W}_s^i|\le \epsilon_N\right\},\\
\Omega_{T,N}^{1,2} &:= \left\{\forall \ell=1,...,K_N+1,\; \#(I_\ell) \le N \epsilon_N^{3/r}\right\},
\end{align*}
{\it Step 1.} Here we estimate $\mathbb{P}[(\Omega_{T,N}^{1,1})^c]$.
Using the Markov inequality, \eqref{lem-momen} and Lemma \ref{lem-small-jump}, we get
\begin{align*}
\mathbb{P}[(\Omega_{T,N}^{1,1})^c]
&\le N \, \mathbb{P}\Big[\Big\{\sup_{[0,T]}|W_t^{1}|\le N^{\delta/3} \text{ and } \sup_{|s-t|\le N^{-2}} |\widetilde{W}_t^{1}-\widetilde{W}_s^1|\le \epsilon_N\Big\}^c\Big]\\
&= N\, \mathbb{P}\Big[\sup_{[0,T]}|W_t^{1}|\ge N^{\delta/3}\Big] + N \,\mathbb{P}\Big[\sup_{[0,T]}|W_t^{1}|\le N^{\delta/3} \text{ and } \sup_{|s-t|\le N^{-2}} |\widetilde{W}_t^{1}-\widetilde{W}_s^1|\ge \epsilon_N\Big]\\
&\le N\, \mathbb{E}\Big[\sup_{[0,T]} |W_t^1|^q \Big] \, N^{-q\delta/3} + C_{T} \, N^3 e^{-N^{\delta/3}}\le C_{T,q} \,N^{1-q\delta/3}.
\end{align*}
\vskip1mm
{\it Step 2.} We now prove that $\mathbb{P}[(\Omega_{T,N}^{1,2})^c]\le C_T\exp{(-N^{\delta})}$. For any fixed $\ell\in\{1,...,K_N+1\}$, we introduce $A_N^\ell=\{\exists~ t\in(t_{\ell-1}^N, t_{\ell}^N]: |\Delta W^1_t|>N^{-1/3}\}$. Then we observe that $\#(I_\ell)$ follows a Binomial distribution with parameters $N$ and $\mathbb{P}(A_N^\ell)$. Using again the Markov inequality, we observe that
\begin{align*}
\mathbb{P}[(\Omega_{T,N}^{1,2})^c]\le \sum_{\ell=1}^{K_N+1}\mathbb{P}\Big[ \#(I_\ell) \ge N\epsilon_N^{3/r}\Big]\le \sum_{\ell=1}^{K_N+1}\mathbb{E}[\exp \big( \#(I_\ell) \big)] \exp{(-N\epsilon_N^{3/r})}.
\end{align*}
But, \[\mathbb{E}[\exp{( \#(I_\ell))}]=\exp{\big(N \log(1 + (e-1) \mathbb{P}(A_N^{\ell}))\big)}\le \exp{\big(N(e-1) \mathbb{P}(A_N^{\ell})\big)}.
\]
Hence,
\[\mathbb{P}[(\Omega_{T,N}^{1,2})^c]\le \sum_{\ell=1}^{K_N+1}\exp{\big(N(e-1) \mathbb{P}(A_N^{\ell})\big)}\exp{(-N\epsilon_N^{3/r})}.\]
We know from Lemma \ref{lem-big-jump} that
$\mathbb{P}(A_N^\ell)\le C N^{-(2-\nu/3)}$, hence $N\mathbb{P}(A_N^\ell) \le CN^{-1+\nu/3}\le C$. We thus deduce that
$$
\mathbb{P}[(\Omega_{T,N}^{1,2})^c]\le C (K_N+1)\exp(-N\epsilon_N^{3/r}) \le C(2TN^2+1)\exp(-N\epsilon_N^{3/r})\le C_T \exp(-N^{\delta}),
$$
since $N\epsilon_N^{3/r}=N^{1/p+\delta/r}$ and since $1/p+\delta/r>\delta$. This ends the proof.
\epf
We now give the
\begin{proof}}\def\epf{\end{proof}[Proof of Proposition \ref{norm-bound}]
Consider the partition $\mathscr{P}_N$ of $\mathbb{R}^3$ in cubes with side length $\epsilon_N$ and its subset $\mathscr{P}_N^\delta$ consisting of cubes that have non-empty intersection with $B(0, N^{\delta/3})$. Then we observe that $\# (\mathscr{P}_N^\delta)\le (2(N^{\delta/3}+\epsilon_N)\epsilon_N^{-1})^3\le 64 N^\delta\epsilon_N^{-3}=64N$. We split the proof into several steps.
\vskip1mm
{\it Step 1.} For $(x_1,..., x_N)\in(B(0,N^{\delta/3}))^N$ and $(y_1,..., y_N)\in(B(0,N^{\delta/3}))^N$, we set
\[ I= \{i\in\{1,.., N\}: |x_i-y_i|> \epsilon_N\},\]
and denote the empirical measure of ${\bf y}=(y_1,..., y_N)\in(\mathbb{R}^3)^N$ by $\mu_{{\bf y}}^N=N^{-1}\sum_{i=1}^N\delta_{y_i}$.
The goal of this step is to show that
\[\| \mu_{{\bf y}}^N*\psi_{\epsilon_N} \|_{L^p}\le \big(\frac{3}{4\pi}\big)^{1/r} \frac{\#(I)}{N\epsilon_N^{3/r}} + 3375 \Big(N^{-p}\epsilon_N^{-3(p-1)}\sum_{D\in\mathscr{P}_N^\delta}(\#\{i\in\{1,...,N\} : x_i\in D\})^p\Big)^{1/p}.\]
Indeed, recalling that $\psi_\epsilon(x)=(3/(4\pi\epsilon^3))\mathbf{1}_{\{|x|\le \epsilon\}}$, we observe that
\begin{align*}
\mu_{{\bf y}}^N*\psi_{\epsilon_N}(v)
& = \frac{1}{N} \sum_{i=1}^{N} \psi_{\epsilon_N}(v-y_i)\mathbf{1}_{\{|x_i-y_i|>\epsilon_N\}}+N^{-1} \sum_{i=1}^{N} \psi_{\epsilon_N}(v-y_i)\mathbf{1}_{\{|x_i-y_i| \le \epsilon_N\}}\\
& = \frac{1}{N} \sum_{i\in I}\psi_{\epsilon_N}(v-y_i)
+ \frac{3}{4\pi N\epsilon_N^3} \# \big\{ i\in\{1,...,N\}: y_i\in B(v,\epsilon_N), |y_i-x_i|\le \epsilon_N \big\} \\
& \le \frac{1}{N} \sum_{i\in I} \psi_{\epsilon_N}(v-y_i) + \frac{3}{4\pi N\epsilon_N^3}\# \big \{ i\in\{1,...,N\}: x_i\in B(v , 2\epsilon_N)\big \}.
\end{align*}
Hence,
$$
\mu_{{\bf y}}^N*\psi_{\epsilon_N}(v) \le \frac{1}{N} \sum_{i\in I} \psi_{\epsilon_N}(v-y_i) +\frac{3}{4\pi N\epsilon_N^3} \sum_{D\in\mathscr{P}_N^\delta}\# \big \{ i\in\{1,...,N\} : x_i\in D \big \} \mathbf{1}_{\{D\cap B(v, 2\epsilon_N) \neq \emptyset\}}.
$$
We then deduce that
\begin{align*}
&\|\mu_{{\bf y}}^N*\psi_{\epsilon_N}\|_{L^p}\\
&\le \frac{1}{N} \| \sum_{i\in I} \psi_{\epsilon_N}(\cdot -y_i)\|_{L^p}
+ \frac{3}{4\pi N\epsilon_N^3} \| \sum_{D\in\mathscr{P}_N^\delta}\# \{ i\in\{1,...,N\}: x_i\in D\} \mathbf{1}_{\{D\cap B(\cdot, 2\epsilon_N)\neq \emptyset\}} \|_{L^p}.
\end{align*}
Since $\| \psi_{\epsilon_N}(\cdot -y_i) \|_{L^p}=(\frac{3}{4\pi})^{1/r} \epsilon_N^{-3/r}$, we have
\begin{align*}
\frac{1}{N} \| \sum_{i\in I} \psi_{\epsilon_N}(\cdot -y_i) \|_{L^p}
\le \frac{1}{N}\sum_{i\in I} \| \psi_{\epsilon_N}(\cdot -y_i) \|_{L^p} &\le \big(\frac{3}{4\pi} \big)^{1/r} \frac{\#(I)}{N\epsilon_N^{3/r}}.
\end{align*}
On the other hand, let $A := \| \sum_{D\in\mathscr{P}_N^\delta}\# \{i\in\{1,...,N\} : x_i\in D\} \mathbf{1}_{\{D\cap B(\cdot, 2\epsilon_N)\neq \emptyset\}} \|_{L^p}$, then
\begin{align*}
A^p
&=\int_{\mathbb{R}^3}\Big(\sum_{D\in\mathscr{P}_N^\delta} \# \{i: x_i\in D\} \mathbf{1}_{\{D\cap B(v, 2\epsilon_N)\neq \emptyset\}}\Big)^p dv\\
&=\int_{\mathbb{R}^3}\Big(\sum_{D,D^\prime\in\mathscr{P}_N^\delta} \# \{i: x_i\in D\}\# \{i: x_i\in D^\prime\} \mathbf{1}_{\{D\cap B(v, 2\epsilon_N)\neq \emptyset, D^\prime\cap B(v, 2\epsilon_N)\neq \emptyset\}}\Big)^{p/2} dv\\
&\le \int_{\mathbb{R}^3} \sum_{D,D^\prime\in\mathscr{P}_N^\delta} \big(\# \{i: x_i\in D\} \big)^{p/2} \big(\# \{i: x_i\in D^\prime\} \big)^{p/2} \mathbf{1}_{\{D\cap B(v, 2\epsilon_N)\neq \emptyset, D^\prime\cap B(v, 2\epsilon_N)\neq \emptyset\}} dv
\end{align*}
because $p\in(1,2)$.
From $x^2+y^2\ge 2xy$ and a symmetry argument, we see that
\begin{align*}
A^p \le \sum_{D\in\mathscr{P}_N^\delta} (\# \{i: x_i\in D\})^p \int_{\mathbb{R}^3}
\mathbf{1}_{\{D\cap B(v, 2\epsilon_N)\neq \emptyset\}} \sum_{D^\prime\in\mathscr{P}_N^\delta} \mathbf{1}_{\{D^\prime \cap B(v, 2\epsilon_N)\neq \emptyset \}} dv.
\end{align*}
But, for each $v\in\mathbb{R}^3$,
$\sum_{D^\prime\in\mathscr{P}_N^\delta} \mathbf{1}_{\{D^\prime \cap B(v, 2\epsilon_N)\neq \emptyset \}}=\#\{D^\prime\in\mathscr{P}_N^\delta: D^\prime \cap B(v, 2\epsilon_N)\neq \emptyset\} \le 5^3.$
And for each $D\in\mathscr{P}_N^\delta$, $\{v\in\mathbb{R}^3: D\cap B(v, 2\epsilon_N)\neq \emptyset\}$ is included by a ball of radius $3\epsilon_N$. Therefore,
$\int_{\mathbb{R}^3}
\mathbf{1}_{\{D\cap B(v, 2\epsilon_N)\neq \emptyset\}} dv \le 4\pi(3\epsilon_N)^3/3$. Hence,
$$
A^p \le \frac{5^3 4\pi (3\epsilon_N)^3}{3}\sum_{D\in\mathscr{P}_N^\delta} \big(\# \{i: x_i\in D\} \big)^p.
$$
Consequently,
\begin{align*}
\|\mu_{{\bf y}}^N*\psi_{\epsilon_N}(v)\|_{L^p}
&\le \big(\frac{3}{4\pi} \big)^{1/r} \frac{\#(I)}{N\epsilon_N^{3/r}}
+ \frac{3}{4\pi N\epsilon_N^3} A \\
&\le \big(\frac{3}{4\pi} \big)^{1/r} \frac{\#(I)}{N\epsilon_N^{3/r}} + \big(\frac{3}{4\pi} \big)^{1/r} (15)^{3/p}
\Big(N^{-p}\epsilon_N^{-3(p-1)}\sum_{D\in\mathscr{P}_N^\delta} \big(\#\{ i: x_i\in D\} \big)^p\Big)^{1/p}.
\end{align*}
Since $(15)^{3/p}\leq 15^3=3375$, this ends the step.
\vskip1mm
{\it Step 2.} In this step, we extend the proof of \cite[Step 3-Proposition 5.5]{Fournier:2015aa}
to show that there are some constants $C>0$ and $c>0$ (depending on $\delta$ and $M_p$ of
Lemma \ref{dicrete-norm}) such that for all fixed $t\in[0, T+1]$,
\[\mathbb{P}[(\Omega_{t,N}^2)^c]\le C \exp{(-cN^{\delta/r})},\]
where \[\Omega_{t,N}^2=\left\{N^{-p}\epsilon_N^{-3(p-1)}\sum_{D\in\mathscr{P}_N^\delta}\Big(\#\{i\in\{1,...,N\} : W_t^i\in D\}\Big)^p \le 2^{p+1} \| f_t \|_{L^p}^p \right\}.\]
To this end, we introduce, for $D\in\mathscr{P}_N^\delta$, $A_D=\#\{i: W_t^i\in D\}$. Then $A_D\sim B(N, f_t(D))$
and we have
\begin{equation}}\def\eeq{\end{equation}\label{Prop-proba}
\mathbb{P}(A_D\ge x)\le \exp(-x/8) \quad \text{for\, all}\quad x\ge 2Nf_t(D).
\eeq
Indeed, $\mathbb{P}(A_D\ge x)\le e^{-x} \mathbb{E}[\exp(A_D)]=e^{-x}\exp[N\log(1+f_t(D)(e-1))] \le e^{-x}\exp[N(e-1)f_t(D)]$. If $x\ge 2Nf_t(D)$, we thus have
$$
\mathbb{P}(A_D\ge x) \le \exp[-x+x(e-1)/2] \le \exp(-x/8).
$$
Next, it follows from the H\"{o}lder inequality that
$$
\|f_t\|_{L^p}^p \ge \sum_{D\in\mathscr{P}_N^\delta} \int_D |f_t(v)|^p dv \ge \epsilon_N^{-3p/r} \sum_{D\in\mathscr{P}_N^\delta}(f_t(D))^p.
$$
On the other hand, we observe from $\# (\mathscr{P}_N^\delta)\le 64 N^\delta\epsilon_N^{-3}$ that
$$
\|f_t\|_{L^p}^p \ge 64^{-1}N^{-\delta}\epsilon_N^3 \sum_{D\in\mathscr{P}_N^\delta} \|f_t\|_{L^p}^p.
$$
Using the two previous inequalities, we find that
$$
2^{p+1}\|f_t\|_{L^p}^p \ge \sum_{D\in\mathscr{P}_N^\delta} \big(2^p\epsilon_N^{-3p/r}(f_t(D))^p+
2^p 64^{-1}N^{-\delta}\epsilon_N^3\|f_t\|_{L^p}^p \big).
$$
Consequently, on $(\Omega_{t,N}^2)^c$, we have
$$
\sum_{D\in\mathscr{P}_N^\delta} A_D^p >N^p\epsilon_N^{3(p-1)}2^{p+1}\|f_t\|_{L^p}^p \geq N^p\epsilon_N^{3(p-1)}
\sum_{D\in\mathscr{P}_N^\delta} \big(2^p\epsilon_N^{-3p/r}(f_t(D))^p+2^p 64^{-1}N^{-\delta}\epsilon_N^3\|f_t\|_{L^p}^p \big),
$$
so that there is
at least one $D\in\mathscr{P}_N^\delta$ with $A_D^p \ge N^p \epsilon_N^{3(p-1)}
\big[2^p\epsilon_N^{-3p/r}(f_t(D))^p+2^p 64^{-1}N^{-\delta}\epsilon_N^3\|f_t\|_{L^p}^p \big] $. Hence,
$$
\mathbb{P}[(\Omega_{t,N}^2)^c] \le \sum_{D\in\mathscr{P}_N^\delta} \mathbb{P}\Big(A_D \ge N \epsilon_N^{3/r}
\big[2^p\epsilon_N^{-3p/r}(f_t(D))^p+2^p 64^{-1}N^{-\delta}\epsilon_N^3\|f_t\|_{L^p}^p\big]^{1/p}\Big).
$$
But we can apply \eqref{Prop-proba}, because $x_N:=N \epsilon_N^{3/r} \big[2^p\epsilon_N^{-3p/r}(f_t(D))^p+2^p64^{-1}N^{-\delta}\epsilon_N^3\|f_t\|_{L^p}^p\big]^{1/p}$ enjoys
the property that $x_N\ge N \epsilon_N^{3/r} [2^p\epsilon_N^{-3p/r}(f_t(D))^p]^{1/p}=2Nf_t(D)$:
$$ \mathbb{P}[(\Omega_{t,N}^2)^c] \le \sum_{D\in\mathscr{P}_N^\delta} \exp(-x_N/8).$$
Using that $x_N \ge N \epsilon_N^{3/r}(2^p 64^{-1}N^{-\delta}\epsilon_N^3\|f_t\|_{L^p}^p)^{1/p}= cN^{\delta/r}\|f_t\|_{L^p}$,
that $\#(\mathscr{P}_N^\delta) \le 64N$ and that $\|f_t\|_{L^p}\ge M_p$, we deduce that
$$
\mathbb{P}[(\Omega_{t,N}^2)^c] \le \sum_{D\in\mathscr{P}_N^\delta} \exp(- cN^{\delta/r}\|f_t\|_{L^p}/8)
\le 64N \exp(-c M_pN^{\delta/r}/8) \le C\exp(-cM_p N^{\delta/r}/10).
$$
This ends the step.
\vskip1mm
{\it Step 3.} We finally consider the event
\[\Omega_{T,N}=\Omega_{T,N}^1\cap(\cap_{\ell=1}^{K_N+1}\Omega_{t_\ell^N, N}^2),
\]
where $\Omega_{T,N}^1$ is defined in Lemma \ref{event}, and the sequence $(t_{\ell}^N)_{\ell=0}^{K_N+1}$ satisfying $0=t_0^N < t_1^N<...<t_{K_N}^N \le T\le T_{K_N+1}^N$, with $K_N\le 2TN^2$ and $\sup_{i=0,...,K_N}(t_{\ell+1}^N-t_\ell^N)\le N^{-2}$ is built in Lemma \ref{dicrete-norm}. We also recall that $h_N(t)=\sum_{\ell=1}^{K_N+1}\|f_{t_\ell^N}\|_{L^p}\mathbf{1}_{\{t\in(t_{\ell-1}^N, t_\ell^N]\}}$.
\vskip1mm
According to Lemma \ref{event} and Step 2, we see that
\[\mathbb{P}[\Omega_{T,N}^c] \leq \mathbb{P}[(\Omega_{T,N}^1)^c]+\sum_{\ell=1}^{K_N+1}\mathbb{P}[(\Omega_{t_\ell^N, N}^2)^c]\le C_{T,q,\delta}N^{1-q\delta/3}+C(K_N+1)\exp{(-cN^{\delta/r})}\le C_{T,q,\delta}N^{1-q\delta/3}.\]
\vskip1mm
Finally, we show that on $\Omega_{T,N}$, for all $t\in[0,T]$, $\|\bar\mu_{{\bf {W}}_t}^N\|_{L^p}\le 13500 (1+h_N(t))$.
Recall that $\widetilde{W}_t^i$ is defined by \eqref{newprocess} and that $I_\ell$ is given by \eqref{Number}, we have
\begin{enumerate}} \def\eenu{\end{enumerate}[label=(\roman*)]
\item for all $i=1,...,N$, and for all $t\in [0, T+1]$, $W_t^i\in B(0, N^{\delta/3})$
(according to $\Omega_{T,N}^1$);
\item for all $\ell=1,..., K_N+1$, all $t\in(t_{\ell-1}^N, t_{\ell}^N]$, all $i\in\{1,...,N\}\setminus I_\ell$,
$|W_t^i-W_{t_\ell^N}^i|=|\widetilde{W}_t^i-\widetilde{W}_{t_\ell^N}^i|\le \epsilon_N$, and
$ \#(I_\ell) \le N \epsilon_N^{3/r}$ (by definition of $\widetilde W^i$ and $I_\ell$ and thanks to $\Omega_{T,N}^1$);
\item For all $\ell=1,..., K_N+1$, $N^{-p}\epsilon_N^{-3(p-1)}\sum_{D\in\mathscr{P}_N^\delta}\Big(\#\big\{i\in\{1,...,N\} : \, W_{t_\ell^N}^i\in D\big \}\Big)^p\le 2^{p+1} \| f_{t_\ell^N} \|_{L^p}^p$ (according to $\cap_{\ell=1}^{K_N+1}\Omega_{t_\ell^N, N}^2$).
\eenu
Using Step 1 with $\bar\mu_{{\bf {W}}_t}^N=\mu_{{\bf {W}}_t}^N*\psi_{\epsilon_N}$,
we deduce that on $\Omega_{T,N}$, for all $t\in[0,T]$, choosing $\ell$ such that $t\in(t_{\ell-1}^N, t_\ell^N]$, we
have
\begin{align*}
\|\bar\mu_{{\bf {W}}_t}^N\|_{L^p}\le& \big(\frac{3}{4\pi}\big)^{1/r} \frac{\#(I_\ell)}{N\epsilon_N^{3/r}}
+ 3375 \Big(N^{-p}\epsilon_N^{-3(p-1)}\sum_{D\in\mathscr{P}_N^\delta}(\#\{i\in\{1,...,N\} : W^i_{t^N_{\ell}}\in D\})^p\Big)^{1/p}\\
\leq & 1+ 3375. 2^{(p+1)/p}\| f_{t_\ell^N} \|_{L^p}\\
=& 1+ 3375. 2^{(p+1)/p}h_N(t).
\end{align*}
This completes the proof, since $3375. 2^{(p+1)/p}\leq 3375.4=13500$.
\epf
\section{Estimate of the Wasserstein distance}
This last section is devoted to the proof of Theorem \ref{main-result}.
In the whole section, we assume \eqref{con} for some $\gamma\in(-1,0)$, $\nu\in(0,1)$ with $\gamma+\nu>0$.
We consider $q>6$ such that $q>\gamma^2/(\gamma+\nu)$, $f_0\in\mathcal{P}_q({\mathbb{R}^3})$
with a finite entropy, and $(f_t)_{t\ge0}$ the unique weak solution to \eqref{Bol}
given by Theorem \ref{well-posedness}. We fix $p\in (3/(3+\gamma),p_0(\gamma,\nu,q))$
and know that $(f_t)_{t\geq 0}\in L^\infty\big([0,\infty), \mathcal{P}_2({\mathbb{R}^3})\big)\cap
L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$.
\vskip1mm
We fix $N\geq 1$, $K\geq 1$ and put $\epsilon_N=N^{-(1-\delta)/3}$ with $\delta=6/q$.
Consider $(V_t^{i})_{t\ge0}$ for $i=1,\dots,N$, defined by \eqref{particle-system}
with the choice $\epsilon=\epsilon_N$.
We know by Lemma \ref{aaaa} that $(V_t^{i})_{i=1,\dots,N,t\ge0}$ is a Markov process with generator
${\cal L}_{N,K}$, see \eqref{generator-cut}, starting from $(V_0^{i})_{i=1,\dots,N}$, which is an i.i.d. family
of $f_0$-distributed random variables. We set $\mu^N_{{\bf{V}}_t}=N^{-1}\sum_{1}^N \delta_{V^i_t}$.
So the goal of the section is to prove that
\begin{align}\label{ggooal}
\sup_{[0, T]}\mathbb{E}[\mathcal{W}_2^2(\mu^N_{{\bf{V}}_t}, f_t)]\le C_{T,q}\Big(N^{-(1-6/q)(2+2\gamma)/3}+K^{1-2/\nu}+N^{-1/2}\Big).
\end{align}
We consider $(W_t^i)_{t\ge0}$, for $i=1,\dots,N$ defined by \eqref{B-particle} and recall that
for all $t\geq0$, the family $(W_t^i)_{i=1,\dots, N}$ is i.i.d. and $f_t$-distributed.
\vskip1mm
First, we introduce the following shortened notations:
\begin{align*}
&c_W(s):=c(W_{s}^{1},W_s^*(\alpha), z,\varphi),\\
&c_W^{N}(s):=c(W_{s}^{1},W_s^{*,\epsilon_N}(\alpha), z,\varphi+\varphi_{1,\alpha, s}^1),\\
&c_{V}^{N}(s):=c(V_{s}^{1},V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha), z,\varphi+\varphi_{1,\alpha, s}^1+\varphi_{1,\alpha, s}^2),\\
&c_{K,V}^{N}(s):=c_K(V_{s}^{1},V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha), z,\varphi+\varphi_{1,\alpha, s}^1+\varphi_{1,\alpha, s}^2),\\
&c_{K,V}(s):=c_K(V_{s}^{1}, V_s^*({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha), z, \varphi+\varphi_{1,\alpha,s}),
\end{align*}
with the notations of
Section 4.
Let us now prove an intermediate result.
\begin{lem}\def\elem{\end{lem}}\label{intermediate-lemma}
There is $C>0$ such that a.s.,
\begin{align*}
&I_0^N(s)+I_1^N(s)+I_2^N(s)+I_3^N(s) \\
\le& C \epsilon_N^{2+2\gamma} + C|W_s^1-V_s^{1}|^2+ C
K^{1-2/\nu}\int_0^1 |W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{2+2\gamma/\nu} d\alpha\\
&+C\int_0^1\Big(|W_s^1-V_s^{1}|^2+|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2\Big)|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma d\alpha.
\end{align*}
where
\begin{align*}
& I_0^N(s):=\int_0^1\int_0^\infty\int_0^{2\pi}\Big(2(W_{s}^1-V_{s}^{1})\cdot(c_W^{N}(s)-c_{K,V}^{N}(s))+|c_W^N(s)-c_{K,V}^N(s)|^2\Big)d\varphi dzd\alpha,\\
& I_1^N(s):=\int_0^1\int_0^\infty\int_0^{2\pi}2(W_{s}^1-V_{s}^{1})\cdot\big(c_W(s)-c_W^{N}(s)+c_{K,V}^{N}(s)-c_{K,V}(s)\big) d\varphi dzd\alpha,\\
& I_2^N(s):=\int_0^1\int_0^\infty\int_0^{2\pi}|c_W(s)-c_W^{N}(s)+c_{K,V}^{N}(s)-c_{K,V}(s)|^2 d\varphi dzd\alpha,\\
& I_3^N(s):=\int_0^1\int_0^\infty\int_0^{2\pi} 2\big(c_W^N(s)-c_{K,V}^N(s)\big)\cdot\big(c_W(s)-c_W^{N}(s)+c_{K,V}^{N}(s)-c_{K,V}(s)\big) d\varphi dzd\alpha.
\end{align*}
\elem
\begin{proof}}\def\epf{\end{proof} First recall that $|W_s^{*,\epsilon_N}(\alpha)-V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2
=|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2$, see Notation \ref{nnn}.
It thus follows from \eqref{ee3} (with $v=W^1_s$, $v_*=W_s^{*,\epsilon_N}(\alpha)$, $\tilde v=V^1_s$
and $\tilde v_*=V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)$) that
\begin{align*}
I_0^N(s) &\le C\int_0^1\Big(|W_s^1-V_s^{1}|^2+|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2\Big)|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma d\alpha \\
&\hskip 5cm +C
K^{1-2/\nu}\int_0^1 |W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{2+2\gamma/\nu} d\alpha.
\end{align*}
Next, we study $I_1 ^N(s)$. As seen in the proof of Lemma \ref{estimategeneral},
$$
\int_0^\infty\!\!\int_0^{2\pi}\!\!c(v,v_*,z,\varphi) d\varphi dz=-(v-v_*)\Phi(|v-v_*|)
\;\hbox{ and }\;\int_0^\infty\!\!\int_0^{2\pi}\!\!c_K(v,v_*,z,\varphi) d\varphi dz=-(v-v_*)\Phi_K(|v-v_*|),
$$
where $\Phi(x)= \pi\int_0^{\infty}(1-\cos G(z/x^\gamma))dz$ and $\Phi_K(x)= \pi\int_0^{K}(1-\cos G(z/x^\gamma))dz$.
Then,
\begin{align*}
I_1^N(s)
&=2(W_s^1-V_s^{1}) \cdot \int_0^1\Big[-\big(W_{s}^{1}-W_s^*(\alpha)\big)\Phi\big(|W_{s}^{1}-W_s^*(\alpha)|\big)\\
&\hskip3cm~ +\big(W_s^1-W_s^{*,\epsilon_N}(\alpha)\big)\Phi\big(|W_s^1-W_s^{*,\epsilon_N}(\alpha)|\big)\\
&\hskip3cm~ -\big(V_s^{1}-V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)\big)\Phi_{K}\big(|V_s^{1}-V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|\big)\\
&\hskip3cm~ +\big(V_s^{1}-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)\big)\Phi_{K}\big(|V_s^{1}-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|\big)\Big]d\alpha.
\end{align*}
But we have checked that
$\left|X\Phi_K(|X|)- Y\Phi_K(|Y|)\right| \leq C |X-Y||X|^\gamma$ for any $X,Y\in\mathbb{R}^3$
in the proof of Lemma \ref{estimategeneral}, and it of course also holds true that
$\left|X\Phi(|X|)- Y\Phi(|Y|)\right| \leq C |X-Y||X|^\gamma$. Thus
\begin{align*}
I_1^N(s)\leq & C |W_s^1-V_s^{1}| \int_0^1
\Big[ |W_s^*(\alpha)-W_s^{*,\epsilon_N}(\alpha)||W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma\\
&\hskip3cm+|V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)||V_s^{1}-V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^\gamma
\Big]d\alpha\\
=& C |W_s^1-V_s^{1}| \int_0^1 |\epsilon_NY(\alpha)| \Big[|W_s^1-W_s^{*}(\alpha)-\epsilon_NY(\alpha)|^\gamma
+|V_s^{1}-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)-\epsilon_NY(\alpha)|^\gamma
\Big]d\alpha\\
\leq & C |W_s^1-V_s^{1}|^2\\
&+C \epsilon_N^2 \int_0^1 |Y(\alpha)|^2 \Big[|W_s^1-W_s^{*}(\alpha)-\epsilon_NY(\alpha)|^{2\gamma}
+|V_s^{1}-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)-\epsilon_NY(\alpha)|^{2\gamma}
\Big]d\alpha.
\end{align*}
But $Y$ is independent of $(W_s^{*},V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s},\cdot))$
and it holds that $\sup_{x \in \mathbb{R}^3} \int_0^1|x-\epsilon_N Y(\alpha)|^{2\gamma}|Y(\alpha)|^2d\alpha\leq
\int_0^1|\epsilon_N Y(\alpha)|^{2\gamma}|Y(\alpha)|^2d\alpha=C \epsilon_N^{2\gamma}$
(recall that $\gamma \in (-1,0)$ and that $Y$ is uniformly distributed on $B(0,1)$), so that finally,
$$
I_1^N(s)\leq C |W_s^1-V_s^{1}|^2 + C \epsilon_N^{2+2\gamma}.
$$
For $I_2^N (s)$, we first write $I_2^N (s)\leq A+B$, where
$$
A=2\int_0^1\int_0^\infty\int_0^{2\pi} |c_W(s)-c_W^{N}(s)|^2d\varphi dzd\alpha
\;\hbox{ and}\;B=2\int_0^1\int_0^\infty\int_0^{2\pi}|c_{K,V}^{N}(s)-c_{K,V}(s)|^2d\varphi dzd\alpha.
$$
We first apply \eqref{ee2} with with $v=W^1_s$, $v_*=W_s^{*,\epsilon_N}(\alpha)$, $\tilde v=W^1_s$
and $\tilde v_*=W_s^{*}(\alpha)$:
$$
A\leq C \int_0^1 |W_s^*(\alpha)-W_s^{*,\epsilon_N}(\alpha)|^2 |W_s^{1}-W_s^{*,\epsilon_N}(\alpha)|^\gamma d\alpha
=C \epsilon_N^2 \int_0^1 |Y(\alpha)|^2|W_s^{1}-W_s^{*}(\alpha)-\epsilon_NY(\alpha)|^\gamma d\alpha.
$$
Using that
$\sup_{x \in \mathbb{R}^3} \int_0^1|x-\epsilon_N Y(\alpha)|^{\gamma}|Y(\alpha)|^2d\alpha\leq
\int_0^1|\epsilon_N Y(\alpha)|^{\gamma}|Y(\alpha)|^2d\alpha=C \epsilon_N^{\gamma}$ and
arguing as in the study of $I^N_1(s)$, we conclude that
$A \leq C \epsilon_N^{2+\gamma} \leq C \epsilon_N^{2+2\gamma}$.
The other term $B$ is treated in the same way
(observe that \eqref{ee2} obviously also holds when replacing $c$ by $c_K=c{\bf 1}_{\{z\leq K\}}$).
\vskip1mm
We finally treat $I_3^N (s)$. It is obvious that
\begin{align*}
I_3^N(s)&\le \int_0^1\int_0^\infty\int_0^{2\pi} |c_W^N (s)-c_{K,V}^N (s)|^2 d\varphi dzd\alpha+I_2^N(s).
\end{align*}
But
\begin{align*}
\int_0^\infty\int_0^{2\pi} |c_W^N (s)-c_{K,V}^N (s)|^2 d\varphi dz
=& \int_0^K\int_0^{2\pi} |c_W^N(s)-c_{V}^N(s)|^2 d\varphi dz+ \int_K^\infty\int_0^{2\pi} |c_W^N(s)|^2 d\varphi dz.
\end{align*}
Applying first \eqref{ee2} with $v=W^1_s$, $v_*=W_s^{*,\epsilon_N}(\alpha)$, $\tilde v=V^1_s$
and $\tilde v_*=V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)$, we find that
\begin{align*}
&\int_0^K\int_0^{2\pi} |c_W^N(s)-c_{V}^N (s)|^2 d\varphi dz\\
\le& C\big(|W_s^1-V_s^{1}|^2+|W_s^{*,\epsilon_N}(\alpha)-V_s^{*,\epsilon_N}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2\big) |W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma\\
=& C\big(|W_s^1-V_s^{1}|^2+|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2\big) |W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma.
\end{align*}
Moreover, as seen in the proof of Lemma \ref{estimategeneral},
$\int_K^\infty\int_0^{2\pi} |c_W^N (s)|^2 d\varphi dz=|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^2
\Psi_K(|W_s^1-W_s^{*,\epsilon_N}(\alpha)|)$, where
$\Psi_K(x)=\Phi(x)-\Phi_K(x)\le C\int_K^\infty G^2(z/x^\gamma)dz\le Cx^{2\gamma/\nu}K^{1-2/\nu}$. Hence,
$$
\int_K^\infty\int_0^{2\pi} |c_W^N(s)|^2 d\varphi dz \leq C |W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{2+2\gamma/\nu}K^{1-2/\nu}.
$$
All this shows that
\begin{align*}
I_3^N(s)\leq& I_2^N(s) + C\int_0^1\big(|W_s^1-V_s^{1}|^2+|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2\big) |W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma d\alpha\\
& + CK^{1-2/\nu}\int_0^1
|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{2+2\gamma/\nu} d\alpha
\end{align*}
and this ends the proof.
\epf
To prove our main result, we need the following estimate which can be found in \cite[Theorem 1]{MR3383341}.
\begin{lem}\def\elem{\end{lem}}\label{lem-empirical}
Fix $A>0$ and $q>4$. There is a constant $C_{A,q}$ such that for all
$f\in\mathcal{P}_q(\mathbb{R}^3)$ verifying $\int_{\mathbb{R}^3}|v|^q f(dv)\le A$, all i.i.d. family
$(X_i)_{i=1,...,N}$ of $f$-distributed random variables,
\begin{align*}
\mathbb{E}\left[\mathcal{W}_2^2\Big(f, N^{-1}\sum_{i=1}^N\delta_{X_i}\Big)\right]\le C_{A,q} N^{-1/2}.
\end{align*}
\elem
\begin{prop}\def\eprop{\end{prop}}\label{pro-distance}
Fix $T>0$ and recall that $h_N$ was defined in Lemma \emph{\ref{dicrete-norm}}. Consider the stopping time
\[\sigma_N=\inf\{t\ge0: \|\bar\mu_{{\bf {W}}_t}^N\|_{L^p}\ge 13500(1+h_N(t))\},\]
where $\bar\mu_{{\bf {W}}_t}^N=\mu_{{\bf {W}}_t}^N *\psi_{\epsilon_N}$ with $\psi_{\epsilon_N}(x)=(3/(4\pi\epsilon_N^3))\mathbf{1}_{\{|x|\le \epsilon_N\}}$ and $\mu_{{\bf {W}}_t}^N=N^{-1}\sum_{1}^N \delta_{W^i_t}$. We have
for all $T>0$,
\[\sup_{[0,T]}\mathbb{E}[|W_{t\wedge\sigma_N} ^1-V_{t\wedge\sigma_N} ^{1}|^2] \le C_{T}(\epsilon_N^{2+2\gamma}+K^{1-2/\nu}+N^{-1/2}).\]
\eprop
\begin{proof}}\def\epf{\end{proof}
We fix $T>0$ and set $u_t^{N}=\mathbb{E}[|W_{t\wedge\sigma_N} ^1-V_{t\wedge\sigma_N} ^{1}|^2]$ for all $t\in [0,T]$.
By the It\^{o} formula, we have
\begin{align*}
u_t^{N}
&=\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1\int_0^\infty \int_0^{2\pi}\Big(|W_{s}^1-V_{s}^{1}+c_W(s)-c_{K,V}(s)|^2-|W_{s}^1-V_{s}^1|^2\Big)d\varphi dzd\alpha\Big]\\
&=\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1\int_0^\infty \int_0^{2\pi}\Big(2(W_{s}^1-V_{s}^{1})\cdot(c_W(s)-c_{K,V}(s))+|c_W(s)-c_{K,V}(s)|^2\Big)d\varphi dzd\alpha\Big]\\
&=\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \big(I_0^{N}(s)+I_1^{N}(s)+I_2^{N}(s)+I_3^{N}(s)\big) ds\Big],
\end{align*}
where $I_i^{N}(s)$ was introduced in Lemma \ref{intermediate-lemma} for $i=0,1,2,3$.
We know from Lemma \ref{intermediate-lemma} that
\begin{align*}
u_t ^{N} \le& Ct\epsilon_N^{2+2\gamma} + C\int_0^t u_s^N ds + C(J^N_1(t)+J^N_2(t)+J^N_3(t)),
\end{align*}
where
\begin{align*}
J^N_1(t)=&
\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1 |W_s^1-V_s^{1}|^2|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma d\alpha ds\Big],\\
J^N_2(t)=&\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^\gamma d\alpha ds\Big],\\
J^N_3(t)=&K^{1-2/\nu}\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{2+2\gamma/\nu} d\alpha ds\Big].
\end{align*}
First, we have
$$
J^N_3(t)\leq C K^{1-2/\nu} t.
$$
Indeed, it suffices to use that $|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{2+2\gamma/\nu}
\leq C(1+|W^1_s|^2+|W_s^{*,\epsilon_N}(\alpha)|^2)$ (because $2+2\gamma/\nu \in (0,2)$),
that $|W_s^{*,\epsilon_N}(\alpha)|^2\leq 2+2|W_s^{*}(\alpha)|^2$ (because $\epsilon_N\in (0,1)$
and $Y$ takes its values in $B(0,1)$) and finally that $\mathbb{E}[|W^1_s|^2]=\int_0^1 |W_s^{*}(\alpha)|^2d\alpha
=m_2(f_0)$.
\vskip1mm
Next, $\mathcal{L}_\alpha(W_s^{*,\epsilon_N})=f_s*\psi_{\epsilon_N}$, so that
$\int_0^1|W_s^1-W_s^{*,\epsilon_N}(\alpha)|^{\gamma} d\alpha \le 1+C_{\gamma,p}\|f_s*\psi_{\epsilon_N}\|_{L^p}$
by (\ref{norm-inequality}) (recall that $p>3/(3+\gamma)$ is fixed since the begining of the section).
Of course, $\|f_s*\psi_{\epsilon_N}\|_{L^p}\le \|f_s\|_{L^p}$, and we conclude that
\begin{align*}
J^N_1(t)\leq
C_{\gamma,p} \int_0^{t} (1+\|f_s\|_{L^p}) \, u_s^{N}\ ds.
\end{align*}
On the other hand, using the exchangeability and that
$W_s^{*,\epsilon_N}(\alpha)=W_s^{*}(\alpha)+\epsilon_N Y(\alpha)$, with $Y(\alpha)$
independent of $W_s^{*}(\alpha)$ and $V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)$ introduced
in Notation \ref{nnn},
we have
\begin{align*}
J^N_2(t)
&=\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2
N^{-1}\sum_{i=1}^N \Big|W_s^i-\epsilon_N Y(\alpha)-W_s^{*}(\alpha) \Big|^\gamma d\alpha ds\Big]\\
&=\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1
|W_s^{*}(\alpha)-V_s^{*} ({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2
\Big(\int_{\mathbb{R}^3}\int_{\mathbb{R}^3}|w-x-W_s^{*}(\alpha)|^\gamma \psi_{\epsilon_N}(x)\mu_{{\bf {W}}_s}^N (dw) dx \Big) d\alpha ds\Big]\\
&=\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1
|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2
\Big(\int_{\mathbb{R}^3}|w-W_s^{*}(\alpha)|^\gamma \bar\mu_{{\bf {W}}_s}^N(dw) \Big) d\alpha ds\Big].
\end{align*}
But $\int_{\mathbb{R}^3}|W_s^{*}(\alpha)-w|^\gamma \bar\mu_{{\bf {W}}_s}^N(dw) \le C_{\gamma,p} (1+\|\bar\mu_{{\bf {W}}_s}^N\|_{L^p})$
by \eqref{norm-inequality}, so that
\begin{align*}
J^N_2(t)
&\le C_{\gamma,p}\mathbb{E}\Big[\int_0^{t\wedge\sigma_N} \int_0^1(1+\|\bar\mu_{{\bf {W}}_s}^N\|_{L^p})|W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2 d\alpha ds \Big].
\end{align*}
We now deduce from Lemma \ref{coupling} that
\begin{align*}
\int_0^1 |W_s^{*}(\alpha)-V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2 d\alpha
& \le 2
\int_0^1\Big(|W_s^{*}(\alpha)-Z_s^{*}({\bf {W}}_{s}, \alpha)|^2
+ |Z_s^{*}( {\bf {W}}_{s}, \alpha) - V_s^{*}({\bf {V}}_{s}, {\bf {W}}_{s}, \alpha)|^2\Big) d\alpha\\
& = 2 \mathcal{W}_2^2( f_s, \mu_{{\bf {W}}_s}^N ) + 2 \frac{1}{N}\sum_{i=1}^{N}|W_s^i - V_s^i|^2.
\end{align*}
Using the exchangeability and that $\|\bar\mu_{{\bf {W}}_s}^N\|_{L^p}\le 13500(1+h_N(s))$ for all $s\le \tau_N$,
it holds that
\begin{align*}
J^N_2(t)
&\le C \int_0^t(1+h_N(s))\mathbb{E}[\mathcal{W}_2^2( f_s, \mu_{{\bf {W}}_s}^N )]ds+ C \int_0^t(1+h_N(s))\, u_s^N ds.
\end{align*}
We thus have checked that
$$
u_t^{N}
\le C t (\epsilon_N^{2+2\gamma} + K^{1-2/\nu} ) + C \int_0^t \big(1+h_N(s) \big)\mathbb{E}\big[\mathcal{W}_2^2( f_s, \mu_{{\bf {W}}_s}^N )\big]ds + C \int_0^t \big( 1+\|f_{s}\|_{L^p}+h_N(s) \big) u_s^{N}ds.
$$
But for each $t\geq 0$, the family
$(W^i_t)_{i=1,\dots,N}$ is i.i.d. and $f_t$-distributed. Furthermore, $\sup_{[0,T]} \mathbb{E}[|W^1_t|^q]<\infty$
($q>6$) by \eqref{lem-momen}. Hence Lemma \ref{lem-empirical} tells us that
\begin{equation}\label{ddd}
\sup_{[0,T]}\mathbb{E}\big[\mathcal{W}_2^2( f_s, \mu_{{\bf {W}}_s}^N )\big] \leq C_{T}N^{-1/2}.
\end{equation}
Using the Gr\"{o}nwall lemma, we deduce that
\begin{align*}
\sup_{[0,T]} u_t^{N}
& \le C_T\left(\epsilon_N^{2+2\gamma}+K^{1-2/\nu}+N^{-1/2} \int_0^T(1+h_N(s))ds \right)\exp{\Big(C\int_0^T (1+\|f_s\|_{L^p}+h_N(s))ds\Big)}.
\end{align*}
But $\int_0^T h_N(s)ds \leq 2\int_0^T \|f_s\|_{L^p} ds$ by Lemma \ref{dicrete-norm}-(ii).
And we know that $f\in L^1_{loc}\big([0,\infty), L^{p}({\mathbb{R}^3})\big)$. We thus conclude that
\begin{align*}
\sup_{[0,T]} u_t^{N} & \le C_T\left(\epsilon_N^{2+2\gamma}+K^{1-2/\nu}+N^{-1/2}\right)
\end{align*}
as desired.
\epf
Now, we give the
\begin{proof}}\def\epf{\end{proof}[Proof of Theorem \ref{main-result}]
As explained at the beginning of the section, we only have to prove \eqref{ggooal}.
Recall that $\sigma_N=\inf\{t\ge0: \|\bar\mu_{{\bf {W}}_t}^N\|_{L^p}\ge 13500(1+h_N(t))\}$,
that $q>6$ and that $\delta=6/q$. It is clear that
$\mathbb{P}[\sigma_N\le T]\le C_{T,q,\delta}N^{1-q\delta/3}=C_{T,q}N^{-1}$ from Proposition \ref{norm-bound}.
Then for $t\in [0,T]$, we write
\begin{align*}
\sup_{[0, T]}\mathbb{E}[\mathcal{W}_2^2(\mu_{{\bf {V}}_t}^N, f_t)]
\le 2\sup_{[0, T]}\mathbb{E}[\mathcal{W}_2^2(\mu_{{\bf {V}}_t}^N, \mu_{{\bf {W}}_t}^N)+\mathcal{W}_2^2(\mu_{{\bf {W}}_t}^N, f_t)]
\leq 2\sup_{[0, T]}\mathbb{E}[\mathcal{W}_2^2(\mu_{{\bf {V}}_t}^N, \mu_{{\bf {W}}_t}^N)]+C_T N^{-1/2}
\end{align*}
by \eqref{ddd}.
But, by exchangeability, we have
$$
\mathbb{E}[\mathcal{W}_2^2(\mu_{{\bf {V}}_t}^N, \mu_{{\bf {W}}_t}^N)]
\le \mathbb{E}\Big[N^{-1}\sum_{i=1}^N|W_{t} ^i-V_{t} ^{i}|^2\Big]=\mathbb{E}[|W_{t} ^1-V_{t} ^{1}|^2].
$$
Moreover,
\begin{align*}
\mathbb{E}[|W_{t} ^1-V_{t} ^{1}|^2]
&\le \mathbb{E}[|W_{t\wedge\sigma_N} ^1-V_{t\wedge\sigma_N} ^{1}|^2] +\mathbb{E}[|W_{t} ^1-V_{t} ^{1}|^2\mathbf{1}_{\{\sigma_N \le T\}}]\\
&\le C_T(\epsilon_N^{2+2\gamma}+K^{1-2/\nu}+N^{-1/2})+ C\mathbb{E}[|W_{t} ^1|^4+|V_{t} ^{1}|^4]^{1/2}(\mathbb{P}(\sigma_N \le T))^{1/2},
\end{align*}
by Proposition \ref{pro-distance} , and the Cauchy-Schwarz inequality.
Noting that $\mathbb{E}[|W_{t} ^1|^4]\le C_T$ by \eqref{lem-momen}, and that $\mathbb{E}[|V_{t} ^{1}|^4]\le C_T \mathbb{E}[|V_0^1|^4]$ by Lemma \ref{aaaa}, we deduce that
$$
\mathbb{E}[|W_{t} ^1-V_{t} ^{1}|^2]
\le C_{T,q}(\epsilon_N^{2+2\gamma}+K^{1-2/\nu}+N^{-1/2}).
$$
All in all, we have proved that
\begin{align*}
\sup_{[0, T]}\mathbb{E}[\mathcal{W}_2^2(\mu_{{\bf {V}}_t}^N, f_t)]\leq C_{T,q}(\epsilon_N^{2+2\gamma}+K^{1-2/\nu}+N^{-1/2}).
\end{align*}
This is precisely \eqref{ggooal}, since $\epsilon_N^{2+2\gamma}=N^{-(1-6/q)(2+2\gamma)/3}$, with
$\epsilon_N=N^{-(1-\delta)/3}$ and $\delta=6/q$.
\epf
\section*{Acknowledgements}
\quad~~
I would like to thank greatly Nicolas Fournier for continuous and generous supports in this research and especially Maxime Hauray for inspiring discussion about the proof of Proposition 3.1-Step 1.
\bibliographystyle{abbrv}
|
\section{Introduction}
In this paper, we consider the problem of \textit{best arm identification} with a \textit{fixed budget $T$}, in the $K$-armed stochastic bandit setting. Given $K$ distributions (or arms) that take value in $[0,1]$, and given a fixed number of samples $T>0$ (or budget) that can be collected sequentially and adaptively from the distributions, the problem of the learner in this setting is to identify the set of distributions with the highest mean, denoted $\mathcal{A}^*$. This setting was introduced in~\cite{bubeck2009pure,audibert2010best}, and is a variant of the best arm identification problem with \textit{fixed confidence} introduced in~\cite{even2002pac,mannor2004sample}.
The best arm identification problem is an important problem in practice as well as in theory, as it is the simplest setting for stochastic non-convex and discrete optimization. It was therefore extensively studied, see~\cite{even2002pac,mannor2004sample,bubeck2009pure,audibert2010best, gabillon2012best,kalyanakrishnan2012pac, jamieson2014best, jamieson2013lil,karnin2013almost,chen2015optimal} and also the full literature review in Section~\ref{sec:lit} for more references and a presentation of the existing results.
Although this problem has been extensively studied, and the results in the \textit{fixed confidence setting} (see see Section~\ref{sec:lit} for a definition and for a presentation of existing results in this setting) have been refined to a point where the optimality gap between best strategies and known lower bounds is really small, see~\cite{chen2015optimal}, there is to the best of our knowledge a major gap between upper and lower bounds in the fixed budget setting. In order to recall this gap, let us write $\mu_k$ for the means of each of the $K$ distributions, $\mu_{(k)}$ for the mean of the arm that has $k$-th highest mean and $\mu^*$ for the highest of these means. Let us define the quantities
$H = \sum_{k \not \in \mathcal{A}^*}(\mu^* - \mu_k)^{-2}$ and $H_2 = \sup_{k > |\mathcal{A}^*|}k(\mu^* - \mu_{(k)})^{-2}$.
The tightest known lower bound for the probability of not identifying an arm with highest mean after using the budget $T$ is of order
$$\exp\Big(-\frac{T}{H}\Big),$$
while the tightest known upper bounds corresponding to existing strategies for $K \geq 3$ are either
$$ \exp\Big(-\frac{T}{18a}\Big)~~~~~\mathrm{or}~~~~~\exp\Big(-\frac{T}{2\log(K)H_2}\Big),$$
depending on whether the learner has access to an upper bound $a$ on $H$ (first bound) or not (second bound).
Since $H_2 \leq H \leq 2\log(K) H_2$, this highlights a gap in the scenario where the learner does not have access to a tight upper bound $a$ on $H$. See~\cite{audibert2010best} for the seminal paper where these state of the art results are proven, and~\cite{gabillon2012best, jamieson2013lil,karnin2013almost,chen2014combinatorial} for papers that propose among other results (generally in the fixed confidence setting) alternative strategies for this fixed budget problem, and~\cite{kaufmann2014complexity} for the lower bound
In this paper, we close this gap, improving the lower bound and proving that the strategies developed in~\cite{audibert2010best} are optimal, in both cases (i.e.~when the learner has access to an upper bound $a$ on $H$ or not). Namely, we prove that there exists no strategy that misidentifies the optimal arm with probability smaller than
$$\exp\Big(-\frac{T}{a}\Big),$$
uniformly over the problems that have complexity $a$, and that there exists no strategy that misidentifies the optimal arm with probability smaller than
$$\exp\Big(-\frac{T}{\log(K)H}\Big),~~~~~~~~~\Big[\text{and note that}\exp\Big(-\frac{T}{\log(K)H}\Big) \geq \exp\Big(-\frac{T}{\log(K)H_2}\Big)\Big]$$
uniformly over all problems. The first lower bound of order $\exp(-\frac{T}{a})$ is not surprising when one considers the lower bounds results in the \textit{fixed confidence setting} by~\cite{even2002pac,mannor2004sample,gabillon2012best, kalyanakrishnan2012pac, jamieson2014best,jamieson2013lil,karnin2013almost,chen2015optimal}, and was already implied by the results of~\cite{kaufmann2014complexity}, but the second lower bound of order $\exp(-\frac{T}{\log(K)H})$ is on the other hand quite unexpected in light of the results in the fixed confidence setting. In fact it is often informally stated in the fixed confidence literature that since the sample complexity in the fixed confidence setting is $H$, the same should hold for the fixed budget setting, and that therefore the right complexity should be $H$ and not $H\log(K)$, i.e.~it is often conjectured that the right bound should be $\exp(-\frac{T}{H})$ and not $\exp(-\frac{T}{\log(K)H})$. In this paper, we disprove formally this conjecture and prove that in the fixed budget setting, \textit{unlike in the fixed confidence setting}, there is an additional $\log(K$) price to pay for adaptation to $H$ in the absence of knowledge over this quantity. Moreover, our lower bound proofs are very simple, short, and based on ideas that differ from previous results, in the sense that we consider a class of problems with different complexities.
In Section~\ref{sec:set}, we present formally the setting, and in Section~\ref{sec:lit}, we present the existing results in a more detailed fashion. Section~\ref{sec:main} contains our main results and Section~\ref{sec:proofg} their proofs.
\section{Setting}\label{sec:set}
\paragraph{Learning setting} We consider a classical $K$ armed stochastic bandit setting with fixed horizon $T$. Let $K>1$ be the number of arms that the learner can choose from. Each of these arms is characterized by a distribution $\nu_k$ that we assume to be defined on $[0,1]$. Let us write $\mu_k$ for its mean.
Let $T>0$. We consider the following dynamic game setting with horizon $T$, which is common in the bandit literature. For any time $t \geq 1$ and $t \leq T$, the learner chooses an arm $I_t$ from $\mathbb{A} = \{1,...,K\}$. It receives a noisy reward drawn from the distribution $\nu_{I_t}$ associated to the chosen arm. An adaptive learner bases its decision at time $t$ on the samples observed in the past. At the end of the game $T$, the learner returns an arm
$$\hat k_T \in \{1, \ldots, K\}.$$
\paragraph{Objective} In this paper, we consider the problem of \textit{best arm identification}, i.e.~we consider the learning problem of finding dynamically, in $T$ iterations of the game mentioned earlier, one of the arms with the highest mean. Let us define the set of optimal arms as
$$\mathcal{A}^* = \arg\max_k \mu_k,$$
and $\mu^* = \mu_{k^*}$ with $k^* \in \mathcal{A}^*$ as the highest mean of the problem. Then we define the \textit{expected loss} of the learner as the probability of not identifying an optimal arm, i.e.~as
$$\mathbb P\Big(\hat k_T \not\in \mathcal{A}^*\Big),$$
where $\mathbb P$ is the probability according to the samples collected during the bandit game. The aim of the learner is to follow a strategy that minimizes this expected loss.
This is known as the \textit{best arm identification} problem in the \textit{fixed budget setting}, see~\cite{audibert2010best}. As was explained in~\cite{audibert2010best}, it is linked to the notion of \textit{simple regret}, where the simple regret is the expected sub-optimality of the chosen arm with respect to the highest mean, i.e.~it is $\mathbb E(\mu^* - \mu_{\hat k_T})$, where $\mathbb E$ is the expectation according to the samples collected during the bandit game.
\paragraph{Problem dependent complexity} We now define two important problem dependent quantities, following e.g.~\cite{even2002pac,mannor2004sample,audibert2010best, gabillon2012best,kalyanakrishnan2012pac, jamieson2014best, jamieson2013lil,karnin2013almost,chen2015optimal}. We will characterize the \textit{complexity} of bandit problems by the quantities
\begin{equation}\label{eq:comp}
H = \sum_{k \not \in \mathcal{A}^*}\frac{1}{(\mu^* - \mu_k)^2} ~~~~\mathrm{and} ~~~ H_2 = \sup_{k > |\mathcal{A}^*|}\frac{k}{(\mu^* - \mu_{(k)})^2},
\end{equation}
where for any $k \leq K$, $\mu_{(k)}$ is the $k$-th largest mean of the arms. As noted in~\cite{audibert2010best}, the following inequalities hold $H_2 \leq H \leq \log(2K) H_2 \leq 2\log(K) H_2$.
\section{Literature review}\label{sec:lit}
The problem of \textit{best arm identification} in the $K$ armed stochastic bandit problem has gained wide interest in the recent years. It can be cast in two settings, \textit{fixed confidence}, see~\cite{even2002pac,mannor2004sample}, and \textit{fixed budget}, see~\cite{bubeck2009pure,audibert2010best}, which is the setting we consider in this paper. In the \textit{fixed confidence setting}, the learner is given a precision $\delta$ and aims at returning an optimal arm, while collecting as few samples as possible. In the \textit{fixed budget setting}, the objective of the learner is to minimize the probability of not recommending an optimal arm, given a fixed budget of $T$ pulls of the arms. The links between these two settings are discussed in details in~\cite{gabillon2012best,karnin2013almost}: the fixed confidence setting is a stopping time problem and the fixed budget setting is a problem of optimal resource allocation. It is argued in~\cite{gabillon2012best} that these problems are equivalent. But as noted in~\cite{karnin2013almost,kaufmann2014complexity}, this equivalence holds only if some additional information e.g. $H$ is available in the fixed budget setting, otherwise it appears that the fixed budget setting problem is significantly harder. This fact is highlighted in the literature review below.
\paragraph{Fixed confidence setting} The fixed confidence setting has been more particularly investigated, with papers proposing strategies that are more and more refined and clever. The papers~\cite{even2002pac,mannor2004sample} introduced the problem and proved the first upper and lower bounds for this problem (where $\gtrsim$ and $\lesssim$ are $\geq$ and $\leq$ up to a constant)).
\begin{itemize}
\item Upper bound : There exists an algorithm that returns, after $\hat T$ number of pulls, an arm $\hat k_{\hat T}$ that is optimal with probability larger than $1-\delta$, and is such that the number of pulls $\hat T$ satisfies
$$\mathbb E \hat T \lesssim H \Big(\log(\delta^{-1}) + \log(K) + \log\big((\max_{k\not\in A^*}(\mu^* - \mu_k)^{-1})\big)\Big).$$
\item Lower bound : For any algorithm that returns an arm $\hat k_{\hat T}$ that is optimal with probability larger than $1-\delta$, the number of pulls $\hat T$ satisfies
$$\mathbb E \hat T \gtrsim H \Big(\log(\delta^{-1})\Big).$$
\end{itemize}
These first results already showed that the quantity $H$ plays an important role for the best arm identification problem. These results are tight in the multiplicative terms $H$ but are not tight in the second order logarithmic terms - and there were several interesting works on how to improve both upper and lower bounds to make these terms match, see~\cite{gabillon2012best, kalyanakrishnan2012pac, jamieson2014best, jamieson2013lil,karnin2013almost,kaufmann2014complexity,chen2015optimal}. To the best of our knowledge, the most precise upper bound is in~\cite{chen2015optimal}, and the most precise lower bound in the case of the two armed problem is in~\cite{kaufmann2014complexity}. These bounds, although not exactly matching in general, are matching up to a multiplicative constant for $\delta$ small enough with respect to $H,K$, i.e.~for $\delta$ small enough with respect to $H,K$, it holds that both upper and lower bounds on $\mathbb E \hat T$ are of order
$$H\log(\delta^{-1}).$$
Note that this can already be seen from the two bounds reported in this paper, i.e.~for $\delta$ smaller than $\min\Big(K^{-1}, \max_{k\not\in \mathcal{A}^*}(\mu^* - \mu_k)\Big)$.
\paragraph{Fixed budget setting} The fixed budget has also been studied intensively, but to the best of our knowledge, an important gap still remains between upper and lower bound results. The best known (up to constants) upper bounds are in the paper~\cite{audibert2010best}, while the best lower bound can be found in~\cite{kaufmann2014complexity}, and they are as follows.
\begin{itemize}
\item Upper bound : Assume that an upper bound $a$ on the complexity $H$ of the problem is known to the learner. There exists an algorithm that, at the end of the budget $T$, fails selecting an optimal arm with probability upper bounded as
$$\mathbb P\Big(\hat k_T \not\in \mathcal A^*\Big) \leq 2TK\exp\Big(-\frac{T-K}{18a}\Big).$$
Even if no upper bound on the complexity $H$ is known to the learner, there exists an algorithm that, at the end of the budget $T$, fails selecting an optimal arm with probability upper bounded as
$$\mathbb P\Big(\hat k_T \not\in \mathcal A^*\Big)\leq \frac{K(K-1)}{2}\exp\Big(-\frac{T-K}{\log(2K)H_2}\Big).$$
\item Lower bound : Even if an upper bound on $H, H_2$ is known to the learner, any algorithm, at the end of the budget $T$, fails selecting an optimal arm with probability lower bounded as
$$\mathbb P\Big(\hat k_T \not\in \mathcal A^*\Big)\geq \exp\Big(-\frac{4T}{H}\Big).$$
\end{itemize}
Several papers exhibit other strategies for the fixed budget problem (in general in combination with a fixed confidence strategy), see e.g.~\cite{gabillon2012best, jamieson2013lil,karnin2013almost}, but their theoretical results do not outperform the ones recalled here and coming from~\cite{audibert2010best}. Note that these results highlight a gap between upper and lower bounds. In the case where an upper bound $a$ on the complexity $H$ is known to the learner, the gap is related to the distance between $a$ and $H$. Beyond the fact that $H_2$ is always smaller than $H$, we would like to emphasize here that if the upper bound $a$ on $H$ is not tight enough, the algorithm's performance will be sub-optimal compared to the hypothetical performance of an oracle algorithm that has access to $H$ - as the non-oracle algorithm will over explore.
Now in the case where one does not want to assume the knowledge of $H$, the gap between known upper and lower bounds becomes even larger and is related to the distance between $H$ and $\log(2K) H_2$. Unlike in the fixed confidence setting, this gap remains also for $T$ large (which corresponds to $\delta$ small in the fixed confidence setting).
We would like to emphasize that although this gap is often belittled in the literature, as it is ``only" a a gap up to a $\log(K)$ factor, this $\log(K)$ factor has an effect in the exponential, and in some sense it is much larger than the gap that was remaining in the fixed confidence setting after the seminal papers~\cite{even2002pac,mannor2004sample}, and over which many valuable works have further improved. Indeed, in order to compare the bounds in the fixed confidence setting with the bounds in the fixed budget setting, one can set
$\delta := \mathbb P\Big(\hat k_T \not\in \mathcal A^*\Big)$, and compute the fixed budget $T$ for which a precision of at least $\delta$ is achieved for both upper and lower bounds. Inverting the upper bounds in the fixed budget setting, one would get the upper bounds on $T$
$$T \lesssim a\log(KT/\delta),~~~~\mathrm{or} ~~~ T \lesssim H_2\log(K)\log(K/\delta)),$$
when respectively an upper bound $a$ on $H$ is known by the learner or when no knowledge of $H$ is available. Conversely, the lower bound in the fixed budget setting yields that the fixed budget $T$ must be of order higher than
$$T \gtrsim H\log(1/\delta).$$
As mentioned, this gap also remains for $\delta$ small. This highlights the fact that the gap in the fixed budget setting is much more acute than the gap in the fixed confidence setting, and that this $\log(K)$ factor is not negligible if one looks at the fixed budget setting problem from the fixed confidence setting perspective. This knowledge gap between the fixed confidence and fixed budget setting was underlined in the papers~\cite{karnin2013almost,kaufmann2014complexity} where the authors explain that closing the gap in the fixed budget setting is a difficult problem that goes beyond known techniques for the fixed confidence setting.
We close this review of literature by mentioning related works on the more involved TopK bandit problem, where the aim is to find $k$ arms that have the highest means, see~\cite{bubeckmultiple,gabillon2012best, kaufmann2014complexity, zhou2014optimal,cao2015top}, and also the more general pure exploration bandit setting introduced in~\cite{chen2014combinatorial}. These results apply to the best arm identification problem considered in this paper, which is a special case of their settings, but they do not improve on the mentioned results for the best arm identification problem.
\section{Main results} \label{sec:main}
We state our results in two parts. First, we provide a weaker version of our results in Subsection~\ref{ss:wr}, which has the advantage of not requiring the introduction of too many additional technical notations We then propose in Subsection~\ref{ss:tf} a technical and stronger formulation of our results.
\subsection{First formulation of our results}\label{ss:wr}
We state the following lower bound for the bandit problem introduced in Section~\ref{sec:set}.
\begin{theorem}\label{thm:main3}
Let $K>1$, $a>0$. Let $\mathbb B_a$ be the set of all bandit problems with distributions in $[0,1]$ and complexity $H$ bounded by $a$. For $\mathcal G \in \mathbb B_a$, we write $\mathcal A^*(\mathcal G)$ for the set of arms with highest mean of problem $\mathcal G$, and $H(\mathcal G)$ for the complexity defined in Equation~\eqref{eq:comp} as $H$ (first quantity) and associated to problem $\mathcal G$.
If $T\geq a^2 \big(4\log(6TK)\big)/ (60)^2$, for any bandit strategy that returns arm $\hat k_T$ at time $T$, it holds that
$$\sup_{\mathcal G \in \mathbb B(a)} \mathbb P_{\mathcal G^{\otimes T}} (\hat k_T \not \in \mathcal A^*(\mathcal G)) \geq \frac{1}{6}\exp\Big(-120\frac{T}{a} \Big).$$
If in addition $a \geq 11K^2$ and if $K \geq 2$, then for any bandit strategy that returns arm $\hat k_T$ at time $T$, it holds that
$$\sup_{\mathcal G \in \mathbb B(a)} \Bigg[\mathbb P_{\mathcal G^{\otimes T}} (\hat k_T \not \in \mathcal A^*(\mathcal G)) \times \exp\Big(400\frac{T}{\log(K)H(\mathcal G)} \Big)\Bigg]\geq \frac{1}{6}.$$
\end{theorem}
This theorem implies what we described in the introduction:
\begin{itemize}
\item Even when an upper bound $a$ on the complexity $H$ of the target bandit problem is known, any learner will misidentify the arm with highest mean with probability larger than
$$ \frac{1}{6}\exp\Big(-120\frac{T}{a} \Big),$$
on at least one of the bandit problems with complexity $H$ bounded by $a$.
\item For $T, a,K$ large enough - $T$ of larger order than $a^2\log(K)$, $a$ of larger order than $K^2$ and $K$ larger than $2$ - any learner will misidentify the arm with highest mean with probability larger than
$$ \frac{1}{6}\exp\Big(-400\frac{T}{\log(K)H(\mathcal G)} \Big),$$
on at least one of the bandit problems $\mathcal G\in \mathbb B_a$ which is associated to some complexity $H(\mathcal G)$ bounded by $a$.
\end{itemize}
The first result is expected when one looks at the lower bounds in the fixed confidence setting, see~\cite{even2002pac,mannor2004sample,gabillon2012best, kalyanakrishnan2012pac, jamieson2014best,jamieson2013lil,karnin2013almost,kaufmann2014complexity,chen2015optimal}. On the other hand, the second result cannot be conjectured from lower bounds in the fixed confidence setting. We remind that in order to obtain a precision $\delta>0$ in the fixed confidence setting, even if the learner does not know $H$, it only requires $$O(H\log(\delta^{-1})),$$ samples for $\delta$ small enough. The natural conjecture following from this is that the probability of error in the fixed budget setting is $$\exp(-T/H),$$ for $T$ large enough. We proved that this does not hold and that the probability of error in the fixed budget setting is lower bounded for any strategy in at least one problem by $$\exp(-T/(\log(K)H)),$$ for $T$ large enough - which corresponds to a higher sample complexity $$H \log(K)\log(1/\delta),$$ in the fixed confidence setting. This lower bound highlights a fundamental difference between the fixed confidence setting - where one does not need to know $H$ in order to adapt to it - and the fixed budget setting - where in the absence of the knowledge of $H$, one pays a price of $\log(K)$ for the adaptation.
Moreover, this lower bound proves that the Successive Reject strategy introduced in~\cite{audibert2010best} is optimal, as its probability of error is upper bounded by a quantity of order
$$\exp(-T/(\log(K)H_2)),$$
which is always smaller in order than our lower bound of orde
$$\exp(-T/(\log(K)H)).$$
This might seem contradictory as the lower bound might seem higher than the upper bound. It is of course not and this only highlights that the problems on which all strategies won't perform well are problems such that $H_2$ is of same order as $H$ - problems having many sub-optimal arms close to the optimal ones. These problems are the most difficult problems in the sense of adapting to the complexity $H$, and for them, a $\log(K)$ adaptation price is unavoidable. This kind of phenomenon, i.e.~the necessity of paying a price for not knowing the model (here the complexity $H$), is not very much studied in the bandit literature, but arises in many fields of high dimensional statistics and non-parametric statistics, see e.g.~\cite{lepski1997optimal, bunea2007sparsity}.
\subsection{Technical and stronger formulation of the results}\label{ss:tf}
We will now present the technical version of our results. This is a lower bound that will hold in the much easier (for the learner) problem where the learner knows that the bandit setting it is facing is one of only $K$ given bandit settings (and where it has all information about these settings). This lower bound ensures that even in this much simpler case, the learner, however good it is, will nevertheless make a mistake.
Before stating the main technical theorem, let us introduce some notations about these $K$ settings. Let $(p_k)_{2 \leq k\leq K}$ be $(K-1)$ real numbers in $[1/4, 1/2)$. Let $p_1 = 1/2$. Let us write for any $1 \leq k\leq K$, $\nu_k:=\mathcal B(p_k)$ for the Bernoulli distribution of mean $p_k$, and $\nu_k':=\mathcal B(1-p_k)$ for the Bernoulli distribution of mean $1-p_k$.
We define the product distributions $\mathcal{G}^i$ where $i \in \{1, ..., K\}$ as $\nu_1^i \otimes ... \otimes \nu_K^i$ where for $1 \leq k \leq K$, $$\nu_k^i := \nu_i \mathbf 1\{k \neq i\} + \nu_i' \mathbf 1\{k = i\}.$$
The bandit problem associated with distribution $\mathcal G^i$, and that we call ``the bandit problem $i$" is such that for any $1 \leq k \leq K$, arm $k$ has distribution $\nu_k^i$, i.e.~all arms have distribution $\nu_k$ except arm $i$ that has distribution $\nu_i'$. We write for any $1 \leq i \leq K$, $\mathbb P_{i}:=\mathbb P_{(\mathcal G^i)^{\otimes T}}$ for the probability distribution of the bandit problem $i$ according to all the samples that a strategy could possibly collect up to horizon $T$, i.e.~according to the samples $(X_{k,s})_{1 \leq k \leq K, 1 \leq s \leq T} \sim (\mathcal{G}^i)^{\otimes T}$.
We define for any $1 \leq k \leq K$ the quantities $d_k := 1/2 - p_k$.
Set also for any $i \in \{1, ..., K\}$ and any $k \in \{1, ..., K\}$
$$\Delta_k^i = d_i + d_k,~~~\mathrm{if}~~ k \neq i~~~~~~\mathrm{and}~~~~~~\Delta_i^i = d_i.$$
In the bandit problem $i$, as \textit{the arm with the best mean is $i$} (and its mean is $1-p_i = 1/2 + d_i$), one can easily see that the $(\Delta_k^i)_k$ are the arm gaps of the bandit problem $i$.
We also define for any $1 \leq i \leq K$ the quantit
$$H(i) := \sum_{1 \leq k \leq K, k \neq i} (\Delta_k^i)^{-2},$
with $H(1) = \max_{1\leq i \leq K}H(i)$. The quantities $H(i)$ correspond to the complexity $H$ computed for the bandit problem $i$ and introduced in Equation~\eqref{eq:comp} (first quantity). We finally define the quantity
$$h^* = \sum_{K \geq k \geq 2} \frac{1}{d_i^{2} H(i) }.$$
We can now state our main technical theorem - \textit{we remind that there is only one arm with highest mean in the bandit problem $i$, and that this arm is arm $i$, so $\mathbb P_{i} (\hat k_T \neq i)$ is the probability under bandit $i$ of not identifying the best arm and recommending a sub-optimal arm.}
\begin{theorem}\label{thm:main}
For any bandit strategy that returns the arm $\hat k_T$ at time $T$, it holds that
$$\max_{1 \leq i \leq K} \mathbb P_{i} (\hat k_T \neq i) \geq \frac{1}{6}\exp\Big(-60\frac{T}{H(1)} -2 \ \sqrt[]{T\log(6TK)}\Big),$$
where we remind that $H(1) = \max_i H(i)$ and also
$$\max_{1 \leq i \leq K}\Bigg[ \mathbb P_{i} (\hat k_T \neq i) \times \exp\Big(60\frac{T}{{H(i)}h^*}
+ 2 \ \sqrt[]{T\log(6TK)}\Big) \Bigg]\geq 1/6.$$
\end{theorem}
The proof of this result is different from the proof of other lower bounds for best arm identification in the fixed budget setting as in~\cite{audibert2010best}. Its construction is not based on a permutation of the arms, but on a flipping of each arm around the second best arm - see Subsection~\ref{sec:proof}. A similar construction can be found in~\cite{kaufmann2014complexity}. However, similarly to~\cite{audibert2010best}, in this paper, a single complexity $H$ is used in the proof, while our proof involves a range of complexities. The idea of the proof is that for any bandit strategy there is at least one bandit problem $i$ among the $K$ described where an arm will be pulled less than it should according to the optimal allocation of the problem $i$ - and when this happens, the algorithm makes a mistake with probability that is too high with respect to the complexity $H(i)$ of the problem. This Theorem is a stronger version of Theorem~\ref{thm:main3} since it states than even if the learner knows that the bandit problem he faces is one of $K$ problems fully described to him, he will nevertheless make an error with probability lower bounded by problem dependent quantities that are much larger than the ones in~\cite{audibert2010best,kaufmann2014complexity}.
A version of this theorem that is easier to read and that holds for $T$ large enough, is as follows.
\begin{corollary}\label{thm:main2}
Assume that $T\geq \max\Big(H(1), H(i)h^*\Big)^2 4\log(6TK)/ (60)^2$. For any bandit strategy that returns the arm $\hat k_T$ at time $T$, it holds that
$$\max_{1 \leq i \leq K} \mathbb P_{i} (\hat k_T \neq i) \geq \frac{1}{6}\exp\Big(-120\frac{T}{H(1)} \Big) = \frac{1}{6}\exp\Big(-120\frac{T}{\max_iH(i)} \Big),$$
and also
$$\max_{1 \leq i \leq K}\Bigg[ \mathbb P_{i} (\hat k_T \neq i) \times \exp\Big(120\frac{T}{{H(i)}h^*} \Big) \Bigg]\geq 1/6.$$
\end{corollary}
Note that both Theorems~\ref{thm:main} and Corollary~\ref{thm:main2} hold for any $p_2, \ldots, p_k$ that belong to $[1/4, 1/2)$ and are therefore quite general.
\section{Proof of the theorems}\label{sec:proofg}
\subsection{Proof of Theorem~\ref{thm:main}}\label{sec:proof}
\paragraph{Step 1: Definition of a high probability event where empirical KL divergences concentrate} For two distributions $\nu,\nu'$ defined on $\mathbb R$ and that are such that $\nu$ is absolutely continuous with respect to $\nu'$, we write
$$\text{KL}(\nu,\nu') = \int_{\mathbb R} \log\Big(\frac{d\nu(x)}{d\nu'(x)}\Big)d\nu(x),$$
for the Kullback leibler divergence between distribution $\nu$ and $\nu'$.
Let $k \in \{1, ..., K\}$. Let us write
$$ \text{KL}_k := \text{KL}(\nu_k', \nu_k) = \text{KL}(\nu_k, \nu_k') = (1 - 2p_k)\log\big(\frac{1 - p_k}{p_k}\big),$$
for the Kullback-Leibler divergence between two Bernoulli distributions $\nu_k$ and $\nu_k'$ of parameter $p_k$ and $1-p_k$. Since $p_k \in [1/4,1/2)$, the following inequality holds:
\begin{equation}\label{eq:gapKL}
\text{KL}_k \leq 10d_k^2.
\end{equation}
Let $1\leq t\leq T$. We define the quantity:
\begin{align*}
\widehat{\text{KL}}_{k,t} &= \frac{1}{t} \sum_{s=1}^t \log(\frac{d \nu_k}{d \nu_k'}(X_{k,s}))\\
&= \frac{1}{t} \sum_{s=1}^t \mathbf 1\{X_{k,s} = 1\} \log(\frac{p_i}{1-p_i}) + \mathbf 1\{X_{k,s} = 0\}\log(\frac{1-p_i}{p_i}),
\end{align*}
where by definition for any $s \leq t$, $X_{k,s}\sim_{i.i.d} \nu_k^i$.
Let us define the event
\begin{align*}
\xi &= \Big\{\forall 1 \leq k\leq K, \forall 1 \leq t\leq T, |\widehat{\text{KL}}_{k,t}| - \text{KL}_k\leq 2 \ \sqrt[]{\frac{\log(6TK)}{t}} \Big\}.
\end{align*}
We now state the following lemma, i.e.~a concentration bound for $|\widehat{\text{KL}}_{k,t}|$ that holds for all bandit $i$ with $1\leq i\leq K$.
\begin{lemma}\label{xi}
It holds that
$$\mathbb P_{i}(\xi) \geq 5/6.$$
\end{lemma}
\begin{proof}
If $k \neq i$ (and thus $\nu_k^i = \nu_k$) then $\mathbb E_{\mathcal G^i} \widehat{\text{KL}}_{k,t} = \mathrm{KL}_k$ and if $k = i$ (and thus $\nu_k^i = \nu_k'$) then $\mathbb E_{\mathcal G^i} \widehat{\text{KL}}_{k,t} = -\mathrm{KL}_k$. Moreover note that since $p_k \in [1/4,1/2)$
$$|\log(\frac{d \nu_k}{d \nu_k'}(X_{k,s}))| = |\mathbf 1\{X_{k,s} = 1\} \log(\frac{p_i}{1-p_i}) + \mathbf 1\{X_{k,s} = 0\}\log(\frac{1-p_i}{p_i})| \leq \log(3).$$
Therefore, $\widehat{\text{KL}}_{k,t}$ is a sum of i.i.d.~samples that are bounded by $\log(3)$, and whose mean is $\pm \mathrm{KL}_k$ depending on the value of $i$. We can apply Hoeffding's inequality to this quantity and we have that with probability larger than $1-(6KT)^{-1}$
$$|\widehat{\text{KL}}_{k,t}| - \text{KL}_k\leq \sqrt{2}\log(3) \ \sqrt[]{\frac{\log(6TK)}{t}}.$$
This assertion and an union bound over all $1 \leq k \leq K$ and $1 \leq t \leq T$ implies that $\mathbb P_{\mathcal G^i}(\xi) \geq 5/6$, as we have $\sqrt{2}\log(3) < 2$.
\end{proof}
\paragraph{Step 2: A change of measure}
Let now $\mathcal{A}lg$ denote the active strategy of the learner, that returns some arm $\hat k_T$ at the end of the budget $T$. Let $(T_k)_{1 \leq k \leq K}$ denote the numbers of samples collected by $\mathcal{A}lg$ on each arm of the bandits. These quantities are stochastic but it holds that $\sum_{1 \leq k \leq K} T_k =T$ by definition of the fixed budget setting. Let us write for any $0 \leq k \leq K$
$$t_k = \mathbb E_{1} T_k.$$
It holds also that $\sum_{1 \leq k \leq K} t_k =T$
We recall the change of measure identity (see e.g.~\cite{audibert2010best}) which states that for any measurable event $\mathcal{E}$ and for any $2 \leq i \leq K$ :
\begin{equation}\label{cm}
\mathbb{P}_{i}(\mathcal{E}) = \mathbb{E}_{1}\Big[\mathbf{1}\{\mathcal{E}\}\exp\big(-T_i\widehat{\text{KL}}_{i,T_i}\big)\Big],
\end{equation}
as the product distributions $\mathcal G^i$ and $\mathcal G^1$ only differ in $i$ and as the active strategy only explored the samples $(X_{k,s})_{k \leq K, s \leq T_k}$.
Let $2\leq i \leq K$. Consider now the event
$$\mathcal{E}_i = \{\hat k_T = 1\} \cap \{\xi\} \cap \{T_i \leq 6 t_i\},$$
i.e.~the event where the algorithm outputs arm $1$ at the end, where $\xi$ holds, and where the number of times arm $i$ was pulled is smaller than $6t_i$. We have by Equation~\eqref{cm} that
\begin{align}
\mathbb{P}_{i}(\mathcal{E}_i) & =\mathbb{E}_{1}\Big[\mathbf{1}\{\mathcal{E}_i\}\exp\big(-T_i\widehat{\text{KL}}_{i,T_i}\big)\Big]\nonumber\\
& \geq \mathbb{E}_{1}\Big[\mathbf{1}\{\mathcal E_i\}\exp\Big(-T_i\text{KL}_i -2 \ \sqrt[]{T_i\log(6TK)}\Big)\Big] \nonumber\\
& \geq \mathbb{E}_{1}\Big[\mathbf{1}\{\mathcal E_i\}\exp\Big(-6t_i\text{KL}_i -2 \ \sqrt[]{T\log(6TK)}\Big)\Big]\nonumber\\
& \geq \exp\Big(-6t_i\text{KL}_i -2 \ \sqrt[]{T\log(6TK)}\Big) \mathbb{P}_{1}(\mathcal E_i),\label{eq:event}
\end{align}
since on $\mathcal E_i$, we have that $\xi$ holds and that $T_i \leq 6t_i$, and since $\mathbb E_{1} \widehat{\text{KL}}_{i,t} = \text{KL}_i$ for any $t \leq T$
\paragraph{Step 3 : Lower bound on $\mathbb{P}_{1}(\mathcal E_i)$ for any reasonable algorithm} Assume that for the algorithm $\mathcal Alg$ that we consider
\begin{equation}\label{eq:bras0}
\mathbb E_{1}(\hat k_T \neq 1) \leq 1/2,
\end{equation}
i.e.~that the probability that $\mathcal Alg$ makes a mistake on problem $1$ is less than $1/2$. Note that if $\mathcal Alg$ does not satisfy that, it performs badly on problem $1$ and its probability of success is not larger than $1/2$ uniformly on the $K$ bandit problems we defined.
For any $2 \leq k \leq K$ it holds by Markov's inequality that
\begin{align}\label{eq:ma}
\mathbb P_{1} (T_k \geq 6 t_k) \leq \frac{\mathbb E_1 T_k}{6t_k} = 1/6,
\end{align}
since $\mathbb E_{1} T_k = t_k$ for algorithm $\mathcal Alg$,
So by combining Equations~\eqref{eq:bras0},~\eqref{eq:ma} and Lemma~\ref{xi}, it holds by an union bound that for any $2 \leq i \leq K$
$$\mathbb P_{1}(\mathcal E_i) \geq 1 - (1/6+1/2 +1/6) = 1/6.$$
This fact combined with Equation~\eqref{eq:event} and the fact that for any $2 \leq i \leq K$ $\mathbb{P}_{i}(\hat k_T \neq i) \geq \mathbb{P}_{i}(\mathcal{E}_i)$ implies that for any $2 \leq i \leq K$
\begin{align}
\mathbb{P}_{i}(\hat k_T \neq i) &\geq \frac{1}{6}\exp\Big(-6t_i\text{KL}_i -2 \ \ \sqrt[]{T\log(6TK)}\Big)\nonumber\\
&\geq \frac{1}{6}\exp\Big(-60t_id_i^2 -2 \ \ \sqrt[]{T\log(6TK)}\Big),\label{eq:event2}
\end{align}
where we use Equation~\eqref{eq:gapKL} for the last step.
\paragraph{Step 4 : Conclusions.}
Since $\sum_{2 \leq k \leq K} d_k^{-2} = H(1)$, and since $\sum_{1 \leq k \leq K} t_k = T$, then there exists $2 \leq i \leq K$ such that
$$t_i \leq \frac{T}{H(1)d_i^2},$$
as the contraposition yields an immediate contradiction.
For this $i$, it holds by Equation~\eqref{eq:event2} that
\begin{align*}
\mathbb{P}_{i}(\hat k_T \neq i) \geq\frac{1}{6}\exp\Big(-60\frac{T}{H(0)} - 2 \ \sqrt[]{T\log(6TK)}\Big).
\end{align*}
This concludes the proof of the first part of the theorem (note that $H(1) = \max_i H(i)$).
Since $h^* = \sum_{2 \leq k \leq K} \frac{1}{d_k^{2} H(k) }$ and since $\sum_{1 \leq k \leq K} t_k = T$, then there exists $2 \leq i \leq K$ such that
$$t_i \leq \frac{T}{h^* d_i^{2} H(i)}.$$
For this $i$, it holds by Equation~\eqref{eq:event2} that
\begin{equation}
\mathbb{P}_{i}(\hat k_T \neq i) \geq \frac{1}{6}\exp\Big(-\frac{60T}{h^*H(i)} -2 \ \sqrt[]{T\log(6TK)}\Big).\nonumber\\
\end{equation}
This concludes the proof of the second part of the theorem.
\subsection{Proof of Theorem~\ref{thm:main3}}\label{sec:proof2}
The proof of the first equation in this theorem follows immediately from Corollary~\ref{thm:main2} since $H(1) = \max_i H(i)$.
The proof of the first equation in this theorem follows as well from Corollary~\ref{thm:main2} by taking $d_k = \frac{1}{4}(k/K)$ for $k \geq 2$ (and therefore $p_k = 1/2 - \frac{1}{4}(k/K) \in [1/4, 1/2)$). Note first that this problem belongs to $\mathbb B_a$ with $a = 11K^2$, since $H(i) \leq H(1) \leq 11K^2$. In this case, for any $1\leq i\leq K$, we have
$$d_i^2H(i) = d_i^2 \sum_{k\neq i} \frac{1}{(d_i+d_k)^2} \leq d_i^2\Big( \frac{i}{d_i^2} + \sum_{k> i} \frac{1}{d_k^2} \Big) \leq i + i^2\sum_{K \geq k \geq i} \frac{1}{k^2} \leq i +i^2 (\frac{1}{i} - \frac{1}{K})\leq 2i.$$
This implies that
$$h^* \geq \sum_{k = 2}^K \frac{1}{2i}\geq \frac{1}{2} (\log(K+1)-\log(2)) \geq \frac{3}{10}\log(K).$$
This concludes the proof.
\section{An $\alpha-$parametrization}
Building on the ideas exposed in the very last part of the proof, we now consider $d_k^\alpha = \frac{1}{4}(k/K)^\alpha$ for $k \geq 2$, $\alpha \geq 0$. A such construction was already considered for the fixed confidence setting in~\cite{jamieson2013finding}. First, let us state that for any $\alpha$, we have the following inequalities: $H(1) \geq H(i) \geq H(K)$, with $H(K)$ (the easiest problem) of order $K$ for all $\alpha$. The hardest problem on the other hand, has complexity of order
$$
H(1) \simeq
\left\{
\begin{array}{ll}
\frac{1}{1-2\alpha}K, \text{for} \ \alpha < 1/2 \\
\log(K)K, \text{for} \ \alpha = 1/2\\
\frac{1}{2\alpha-1}K^{2\alpha}, \text{for} \ \alpha > 1/2
\end{array}
\right. .
$$
For $\alpha < 1/2$, both the easiest and hardest problems in our restricted problem class have a similar complexity up to a constant. On the other hand, for $\alpha >1/2$, we have $H(1)$ of order $H(K)^{2\alpha}$, spanning a range of problems with varying complexities. One can easily check that for $\alpha > 1/2$, we have $h^*$ of order at least $\log(K)$ (as we did for $\alpha = 1$ in the previous section). On the other hand, for $\alpha < 1/2$, we can upper bound $h^*$ as follows:
$$
h^* = \sum_{i=2}^K \frac{1}{d_i^2H(i)} \leq \frac{1}{H(K)}\sum_{i=2}^K \frac{1}{d_i^2} = \frac{H(i)}{H(K)},\\
$$
and this ratio is upper bounded by a constant, as both terms are of order $K$. As such, this construction does not imply that a $\log(K)$ adaptation price is unavoidable in all cases, and the question remains open on whether there exists an algorithm that can effectively adapt to these easier problems.
\section*{Conclusion}
In this paper, our main result states that for the problem of best arm identification in the fixed budget setting, if one does not want to assume too tight bounds on the complexity $H$ of the bandit problem, then any bandit strategy makes an error on some bandit problem $\mathcal G$ of complexity $H(\mathcal G)$ with probability at least of order
$$\exp(-\frac{T}{\log(K)H(\mathcal G)}).$$
This result formally disproves the general belief (coming from results in the fixed confidence setting) that there must exist an algorithm for this problem that, for any problem of complexity $H$, makes an error of at most $$\exp(-\frac{T}{H}).$$
This highlights the interesting fact that for this fixed budget problem and \textit{unlike what holds in the fixed confidence setting}, there is a price to pay for adaptation to the problem complexity $H$. This kind of ``adaptation price phenomenon" can be observed in many model selection problems as e.g.~sparse regression, functional estimation, etc, see~\cite{lepski1997optimal, bunea2007sparsity} for illustrations in these settings where such a phenomenon is well known.
This also proves that strategies based on the Successive Rejection of the arms as the Successive Reject of~\cite{audibert2010best}, are optimal. Our proofs are simple and we believe that our result is an important one, since this closes a gap that had been open since the introduction of the fixed confidence best arm identification problem by~\cite{audibert2010best}.
\paragraph{Acknowledgement} This work is supported by the DFG's Emmy Noether grant MuSyAD (CA 1488/1-1).
|
\section{Introduction}
Quantum systems have dynamics that appear to be beyond the capacity of classical computers to simulate as the size of the system increases. However, a controlled quantum simulator may enable an understanding of such systems that eludes classical description, as the emulation of one system by another can take full advantage of the underlying quantum evolution~\cite{feynman82,georgescu2014quantum}.
One promising avenue for quantum simulation uses massless bosons -- typical photons -- as the constituent particles and examines the phases of matter that can arise with the inclusion of interactions between these particles \cite{aspuru2012photonic,noh2016quantum,
heidemann2007evidence,peyronel2012quantum}. Perhaps the most dramatic possibilities arise in circuit quantum electrodynamics (QED) \cite{houck2012chip}, where the Josephson effect provides a strong microwave nonlinearity, though similar improvements are now becoming available in semiconductor, molecular, and atomic nonlinearities in small optical domain cavities \cite{Hennessy2007,kim.2013.373--377,RYDBERG}. These photonic systems are particularly interesting given our ability to control the dispersion relation of the particles, including, e.g., the creation of effective mass~\cite{kasprzak2006bose,klaers2010bose} or synthetic gauge fields \cite{Raghu2008,Hafezi2011,Fang2012} as well as the character of their interaction.
As a starting point, Bose-Einstein condensation of photons has been observed in recent experiments using cavity polaritons \cite{kasprzak2006bose,balili2007bose,deng2010exciton,tosi.2012.190--194} or with dye microcavities \cite{klaers2010bose} using these ideas.
Unfortunately, the vacuum is the typical ground state for such systems, and thus efforts for quantum simulation with light have focused on driving systems far from equilibrium to provide sufficient numbers of photons. This makes predicting the dynamics and steady state behavior an outstanding challenge~\cite{keeling.2011.131--151,nissen.2012.}.
On the other hand, electronic transport theory, pioneered in the works of Landauer \cite{landauer1957spatial,landauer1970electrical,imry1999conductance}, B\"{u}ttiker \cite{buttiker1986four}, and Imry \cite{imry1999conductance}, has successfully dealt with a different problem: What is the quantum version of Ohm's law, i.e., the relationship between chemical potential difference (voltage) and particle flux (current), for describing the motion of electrons in mesoscopic systems \cite{landauer1957spatial,landauer1970electrical,buttiker1986four,imry1999conductance,caroli1971direct,meir1992landauer,haug2008quantum}? Of particular use have been mathematical tools such as non-equilibirum Green's function methods \cite{keldysh1965diagram,langreth1976linear,meir1992landauer,haug2008quantum}, which enable predictions for systems even at large voltage bias and with strong interactions.
In this article we consider whether a photonic version of the Landauer-type transport exists, and find that for parametrically coupled semi-infinite leads (transmission lines), a natural photonic voltage arises with an associated Ohm's law-type behavior for the photon flux. Our results rely upon the most recent of several approaches for developing a photonic equivalent to this voltage-bias \cite{wurfel1982chemical,herrmann2005light,
schmitt2014observation,ries1991chemical}, including equilibriation of light coupled to electrons flowing in a diode \cite{ries1991chemical,smestad1992luminescence,berdahl1985radiant} and, more recently, parametrically coupled photonic systems \cite{hafezi2015chemical}.
Specifically, we derive the non-equilibrium transport of light under the parametric coupling scheme using non-equilibrium Green's function (NEGF) formalism. We study the photon flux as the equivalent of a current through a parametrically driven mesoscopic region, and show that the photon flux formula can be understood in the Landauer sense, as a transport from a chemical potential imbalance from the parametric coupling, with the addition of an anomalous particle-nonconserving squeezing term. Intuitively, our result connects the photon flow between a low frequency bath and an optical bath as mediated by a mesoscopic, interacting region. Thus we provide a rigorous framework for studying such near-equilibrium photonic systems without resorting to \textit{ad hoc} tools for steady-state dynamics. Furthermore, our result predicts a quantitative link between the photon flux and the Green's function, which provides a possible testing ground for photonic quantum simulations even without particle number conservation.
\section{Photon Transport through a trivial scatterer}
\begin{figure}
\includegraphics[width=2.8in]{fig1}
\caption{(a) Our conceptually simplest system of two semi-infinite leads, with a time-dependent coupling between them and a photodetector connecting to the right lead. (b) A potential physical implementation with a Josephson parametric coupler, driven with a flux bias line, between two transmission lines.}
\label{interface}
\end{figure}
We start by developing our photonic analog to voltage bias. Consider a photonic (optical or microwave) system coupled to two baths: one associated with the typical decay of excitations into other modes via, e.g., imperfect mirrors, while the other is associated with a second bath coupled time-dependently with fast sinusoidal variation of the coupling constant at angular frequency $\omega_p$. In particular, in Ref.~\cite{hafezi2015chemical}, one of us showed that a time-dependent bath coupling can lead to the equilibration of a small system best described by a grand canonical ensemble distribution, i.e., a system of photons with a chemical potential. However, in that work crucial questions -- such as what happens when coupled to two baths -- were largely detailed heuristically. Here we focus on building a formalism, analogous to the finite-bias Green's function approach for electronic transport.
In particular, we describe the two baths as semi-infinite transmission lines for our purposes, with the parametrically coupled bath being the `left' lead, and the natural bath corresponding to photon loss being the `right' lead, which could correspond to an outgoing optical signal to be measured with a photodetector (Fig.~\ref{interface}). This is now analogous to electronic transport at finite voltage bias, where the voltage is equivalent to the chemical potential $\hbar \omega_p$.
As a toy model, and to help develop the formalism, we start with the simplest setup in which the scatterer is trivial -- a section of transmission line -- and the problem now reduces to the case with left and right semi-infinite leads coupled parametrically (see Fig.~\ref{interface}). The Hamiltonian of the the system is $H=H_L+H_R+H_T(t)$, with
\begin{align}
&H_L=\sum_{\alpha} \epsilon_{\alpha} a^{\dagger}_{\alpha} a_{\alpha}, \, H_R=\sum_{\beta} \epsilon_{\beta} b^{\dagger}_{\beta} b_{\beta}, \notag\\ &H_T(t)=\cos(\omega_p t) \sum_{\alpha,\beta} \lambda_{\alpha \beta}u_{\alpha}u_{\beta}.
\label{Hsub}
\end{align}
Here $H_L$ and $H_R$ are Hamiltonians of left and right transmission lines respectively, and $H_T(t)$ is the time-dependent tunneling coupling between the two subsystems. The summation index $\alpha$ labels the states in the left transmission line with energies $\hbar \omega_{\alpha}=\epsilon_{\alpha}$, photon annihilation operators $a_{\alpha}$, displacement operators $u_{\alpha}=\sqrt{\frac{\hbar}{2 \omega_\alpha}}\left(a^{\dagger}_{\alpha}+a_{\alpha}\right)$, and momentum operators $p_{\alpha}=i\sqrt{\frac{\hbar \omega_\alpha}{2 }}\left(a^{\dagger}_{\alpha}-a_{\alpha} \right)$, while $\beta$, $\hbar \omega_{\beta}=\epsilon_{\beta}$, $b_{\beta}$, $u_{\beta}$, and $p_{\beta}$ represent states in the right. $\lambda_{\alpha,\beta}$ are the coupling constants.
We may organize the Hamiltonian in a matrix form
\begin{align}
H=\frac{1}{2}\vec{p}^{\text{T}}\vec{p}+\frac{1}{2}\vec{u}^{\text{T}}\mathbf{K} \vec{u}
\label{Hmatrix}
\end{align}
by introducing displacement and momentum vectors
\begin{align}
\vec{u}=\begin{pmatrix} \vec{u}_L\\ \vec{u}_R \end{pmatrix}, \vec{p}=\begin{pmatrix} \vec{p}_L\\ \vec{p}_R \end{pmatrix},
\label{up}
\end{align}
with elements $(u_{L(R)})_{\alpha(\beta)}\equiv u_{\alpha(\beta)}$ and $(p_{L(R)})_{\alpha(\beta)}\equiv p_{\alpha(\beta)}$.
$\vec{u}$ and $\vec{p}$ follow the equal time commutation relation
\begin{align}
\left[ \vec{u}(t),\vec{p}^{\text{T}}(t) \right]=i\hbar \mathbf{I}.
\label{commutation}
\end{align}
Here $\mathbf{I}$ is the identity matrix, and T denotes the matrix transpose. $\mathbf{K}$ is a symmetric spring's constant matrix and can be further separated into diagonal and off-diagonal parts $\mathbf{K}=\mathbf{D}+\mathbf{V}(t)=\mathbf{D}+\mathbf{V}\cos(\omega_pt)$, where
\begin{align}
\mathbf{D}=\begin{pmatrix} \mathbf{D}^L & 0\\ 0 & \mathbf{D}^R \end{pmatrix}, \mathbf{V}(t)=\begin{pmatrix} 0 &\mathbf{V}^{LR}(t)\\ \mathbf{V}^{RL}(t) & 0 \end{pmatrix}.
\label{DV}
\end{align}
Here $D^L_{\alpha \alpha '}=\omega^2_{\alpha} \delta_{\alpha \alpha '}$, $D^R_{\beta \beta '}=\omega^2_{\beta} \delta_{\beta \beta '}$, and $V^{LR}_{\alpha \beta}(t)=V^{RL}_{\beta \alpha}(t)=\cos(\omega_p t) \lambda_{\alpha \beta}.$
Assume the two subsystems were initially decoupled and in their own thermal equilibrium, and the parametric coupling is adiabatically turned on at $t = - \infty$ and turned off at $t = \infty$. Our goal is to find the photonic current transported between the two ends and express it in a Landauer-like formula in order to predict the current based on an effective chemical potential difference analogous to a voltage bias.
The current on the right at some later time $t$ is defined as the temporal change of the total number of photons in the right transmission line $N_R=\sum_{\beta} b_{\beta}^{\dagger} b_{\beta}$, which corresponds to an expected photodetector signal. We have
\begin{align}
J_R(t) \equiv \left\langle \dot{N}_R(t) \right\rangle =\left\langle \frac{d \sum_{\beta} b_{\beta}^{\dagger} b_{\beta} (t)}{dt}\right\rangle.
\label{JR}
\end{align}
The angular bracket denotes ensemble average over the initial equilibrium density of states, while the operators are in the Heisenberg picture.
According to the Heisenberg equation of motion,
\begin{align}
\dot{N}_R(t)=-\frac{1}{\hbar} \left. \left( \frac{\partial}{\partial t^{'}}\left[ \vec{u}_R^{\text{T}}(t {'}) \tilde{\mathbf{V}}^{RL}(t) \vec{u}_L(t)\right] \right) \right|_{t {'}=t},
\label{Ndot}
\end{align}
where $\tilde{V}^{RL}_{ \beta \alpha} (t) \equiv V^{RL}_{ \beta \alpha} (t) /\omega_{\beta}=\cos(\omega_p t)\lambda_{\alpha \beta}/\omega_{\beta}.$
One can connect the current expression with Keldysh Green's functions \cite{keldysh1965diagram} by introducing the non-equilibrium lesser Green's function defined as
\begin{align}
\mathbf{G}^{<}(t,t {'})\equiv -\frac{i}{\hbar} \left\langle \vec{u}(t {'}) \vec{u}^{\text{T}}(t)\right\rangle ^{\text{T}},
\label{Gless}
\end{align}
which can also be split into four blocks associated with left and right transmission lines. We can write the current using the lesser Green's function as
\begin{align}
J_R(t)=-i \left. \left( \frac{\partial}{\partial t^{'}}\text{Tr}\left[ \mathbf{G}^{<}_{LR}(t,t{'}) \tilde{\mathbf{V}}^{RL}(t)\right] \right) \right|_{t {'}=t}.
\label{JLinG}
\end{align}
The trace here means tracing over photon states $\alpha$.
We now follow the standard Keldysh formalism (NEGF formalism) \citep{keldysh1965diagram} to study the transport formula \cite{caroli1971direct,meir1992landauer,haug2008quantum,wang2014nonequilibrium}. Since we define our Green's functions on displacement operators $u$ instead of photon creation operators $a^{\dagger}$, our problem structurally resembles more the thermal transport cases \citep{wang2014nonequilibrium} than electronic ones. We remind the reader here that since the parametric coupling varies with time and allows pair production and annihilation mechanisms, many identities and tricks in previous works involving steady state or particle-conserving assumptions cannnot be applied here.
The equation of motion of the contour ordered Green's function defined on the Keldysh contour $C$ follows
\begin{align}
\frac{\partial^2}{\partial \tau^2} \mathbf{G}^c(\tau, \tau {'})+\mathbf{K} \mathbf{G}^c(\tau, \tau {'})=-\delta(\tau, \tau {'}) \mathbf{I},
\label{GEOM}
\end{align}
while the noninteracting equilibrium Green's function $\mathbf{g}^c(\tau, \tau {'})$ follows the equation of motion
\begin{align}
\frac{\partial^2}{\partial \tau^2} \mathbf{g}^c(\tau, \tau {'})+\mathbf{D} \mathbf{g}^c(\tau, \tau {'})=-\delta(\tau, \tau {'}) \mathbf{I}.
\label{gEOM}
\end{align}
One can easily verify that $\mathbf{G}^c(\tau, \tau {'})$ follows the Dyson equation
\begin{align}
\mathbf{G}^c(\tau, \tau {'})=\mathbf{g}^c(\tau, \tau {'})+\int_{C} d \tau {''}\mathbf{g}^c(\tau, \tau {''}) \mathbf{V}(\tau {''}) \mathbf{G}^c(\tau {''}, \tau {'}).
\label{Dyson}
\end{align}
Using the Langreth theorem of analytic continuation \cite{langreth1976linear}, the lesser Green's function can be expressed as an integral form on the real axis
\begin{align}
\mathbf{G}^<_{LR}(t, t {'}) & \approx \int_{-\infty}^{\infty} dt_1 \left\{ \mathbf{g}^r_L(t, t_1) \mathbf{V}^{LR}(t_1) \mathbf{g}^<_R(t_1, t {'}) \right. \notag\\
& + \left. \mathbf{g}^<_L(t, t_1) \mathbf{V}^{LR}(t_1) \mathbf{g}^a_R(t_1, t {'})\right\} +\mathcal{O}(\lambda^2)
.
\label{Glessfin}
\end{align}
Here the $r$ and $a$ superscripts stand for retarded and advanced Green's functions, and we treat $\lambda$ as a perturbation.
The equilibrium Green's functions used in the $\mathbf{G}^<_{LR}(t, t {'})$ expression are given by
\begin{align}
&(g^r_L)_{\alpha}(t, t_1)=\frac{-i}{2 \omega_{\alpha}}\theta(t-t_1) \left( e^{-i \omega_{\alpha} (t-t_1)}-e^{i \omega_{\alpha} (t-t_1)}\right),\notag\\
&(g^<_R)_{\beta}(t_1, t{'})=\frac{-i}{2 \omega_{\beta}}\left\{ n_R(\epsilon_{\beta}) e^{-i \omega_{\beta}(t_1-t{'})}+[1+n_R(\epsilon_{\beta})] e^{i \omega_{\beta}(t_1-t{'})}\right\},\notag\\
&(g^<_L)_{\alpha}(t, t_1)=\frac{-i}{2 \omega_{\alpha}}\left\{ n_L(\epsilon_{\alpha}) e^{-i \omega_{\alpha}(t-t_1)}+[1+n_L(\epsilon_{\alpha})] e^{i \omega_{\alpha}(t-t_1)}\right\},\notag\\
&(g^a_R)_{\beta}(t_1, t {'})=\frac{-i}{2 \omega_{\beta}}\theta(t{'}-t_1) \left( e^{i \omega_{\beta} (t_1-t{'})}-e^{-i \omega_{\beta} (t_1-t{'})}\right).\notag\\
\label{g}
\end{align}
Here $n_{L(R)}(\epsilon_{\alpha (\beta)})=(e^{(\epsilon_{\alpha (\beta)}-\mu_{L(R)})/k_BT}-1)^{-1}$ are the bosonic occupation number in left and right transmission lines. The chemical potentials $\mu_L=\mu_R=0$ for photons, $k_B$ is the Boltzmann constant, and $T$ is the initial temperature of the system.
By inserting the expression for $\mathbf{G}^<_{LR}(t, t {'})$, the current is now
\begin{widetext}
\begin{align}
J_R(t)=&-i \left( \left.\frac{\partial}{\partial t^{'}}\text{Tr}\left[ \int_{-\infty}^{\infty} dt_1 \left\{ \mathbf{g}^r_L(t, t_1) \mathbf{V}^{LR}(t_1) \mathbf{g}^<_R(t_1, t {'})+ \mathbf{g}^<_L(t, t_1) \mathbf{V}^{LR}(t_1) \mathbf{g}^a_R(t_1, t {'})\right \} \tilde{\mathbf{V}}^{RL}(t)\right] \right) \right|_{t {'}=t} \notag\\
=&-i \left( \left. \frac{\partial}{\partial t^{'}}\sum_{\alpha,\beta} \int_{-\infty}^{\infty} dt_1 \frac{\lambda_{\alpha \beta}^2}{\omega_{\beta}} \cos(\omega_p t_1)\cos(\omega_p t)\left\{ (g^r_L)_{\alpha}(t, t_1) (g^<_R)_{\beta}(t_1, t {'})+ (g^<_L)_{\alpha}(t, t_1) (g^a_R)_{\beta}(t_1, t {'})\right \}\right) \right|_{t {'}=t} \notag\\
=&\sum_{\alpha,\beta} \frac{\lambda_{\alpha \beta}^2}{4\omega_{\alpha}\omega_{\beta}} \int_{0}^{\infty} d \tau \left( \left\{\cos(\omega_p \tau)+\cos[\omega_p (2t-\tau)]\right\}\cos[(\omega_{\beta}-\omega_{\alpha})\tau]\left[n_L(\epsilon_{\alpha})-n_R(\epsilon_{\beta})\right] \right. \notag\\
&+ \left. \left\{\cos(\omega_p \tau)+\cos[\omega_p (2t-\tau)]\right\}\cos[(\omega_{\beta}+\omega_{\alpha})\tau]\left[ n_L(\epsilon_{\alpha})+n_R(\epsilon_{\beta})+1 \right] \right).
\label{JR1}
\end{align}
\end{widetext}
In the last equality we have used the identity $\cos(\omega_p t_1)\cos(\omega_p t)=\left\{\cos(\omega_p \tau)+\cos[\omega_p (2t-\tau)]\right\}/2$, and changed the integral variable to $\tau=t-t_1.$ The only explicit $t$ dependence arises in the $\cos[\omega_p (2t-\tau)]$ factor. Averaging over one pump cycle $2 \pi/ \omega_p$ takes this factor to zero. We thus neglect those terms with $\cos[\omega_p (2t-\tau)]$ in the spirit of the rotating wave approximation.
We assume the coupling constant only depends on the mode energy, $\lambda_{\alpha \beta}=\lambda(\epsilon_{\alpha},\epsilon_\beta)$, and take the continuum limit of energy so that $\sum_{\alpha,\beta}=\int_{0}^{\infty} d \epsilon_{\alpha} \rho_L(\epsilon_{\alpha})\int_{0}^{\infty} d \epsilon_{\beta} \rho_R(\epsilon_{\beta})$. Here $\rho_L$ and $\rho_R$ are the energy density of states in the left and right transmission lines. Note that $\int_0^{\infty} \cos[(\omega-\omega_1)\tau] d\tau=\pi \delta(\omega-\omega_1)$.
We now arrive at a current formula with three terms:
\begin{align}
\bar{J}_R=&\int_{\hbar \omega_p}^{\infty} d \epsilon T (\epsilon,\epsilon-\hbar \omega_p) \left[ n_L(\epsilon)-n_R(\epsilon-\hbar \omega_p)\right] \notag\\
+& \int_{\hbar \omega_p}^{\infty} d \epsilon T(\epsilon-\hbar \omega_p,\epsilon) \left[ n_L(\epsilon-\hbar \omega_p)-n_R(\epsilon) \right] \notag\\
+& \int_0^{\hbar \omega_p} d \epsilon T(\epsilon,\hbar \omega_p-\epsilon) \left[ n_L(\epsilon)+n_R(\hbar \omega_p-\epsilon)+1 \right]
\label{JRfinal}
\end{align}
Here $\bar{J}_R$ represents the time-averaged current under the rotating wave approximation, and $T(\epsilon_1,\epsilon_2)$ is the transmission function defined as $T(\epsilon_1,\epsilon_2)=\frac{\hbar^3 \pi}{8 \epsilon_1 \epsilon_2} \lambda^2(\epsilon_1,\epsilon_2)\rho_L(\epsilon_1)\rho_R(\epsilon_2).$ $\bar{J}_L$ can be calculated with similar formulations.
Note that the $1/\epsilon$ factor in the transmission function and the bosonic occupation numbers $n_{L(R)}(\epsilon)$ go to infinity as the photon energy approaches zero. One can ensure the convergence of our model by the choice of a three-dimensional reservoir on the low-frequency side. The presence of the nonlinear interaction in the case of an interacting mesoscopic region can regulate the problem as well \cite{hafezi2015chemical}. The power from the pump that generates the parametric coupling should be finite, and as the IR divergence is approached for lower dimensional systems, an appropriate inclusion of pump depletion will be necessary to develop a complete understanding of the problem.
The first line of Eq. (\ref{JRfinal}) can be interpreted as a Landauer-like transport with an effective chemical potential $\hbar \omega_p$ on the right transmission line, and the second line represents a Landauer-like transport with an effective chemical potential $\hbar \omega_p$ on the left. The third line is a particle-nonconserving term due to pair creation and annihilation mechanisms allowed by the oscillating $u$-$u$ type coupling. Two-mode squeezed states of light \cite{milburn1984multimode} are generated through this mechanism with the photon pairs entangled. One will expect a thermal state when tracing over the output modes on one side of such photon pairs.
Note that the current formula is consistent with Fermi's golden rule: The parametric coupling $\cos(\omega_p t)$ only allows transition with $E_f-E_i=\Delta E=\pm \hbar \omega_p$, where $E_f$ and $E_i$ are the energies of the final and initial states. For $\omega_p=0$, the current equation reduces to the usual Landauer form proportional to $n_L(\epsilon)-n_R(\epsilon),$ which is essentially zero when the two transmission lines are at the same temperature.
The particle-nonconserving nature of the problem is manifested by identifying the anomalous current $\bar{J_A}\equiv (\bar{J}_R+\bar{J}_L)/2=\int_0^{\hbar \omega_p} d \epsilon T(\epsilon,\hbar \omega_p-\epsilon) \allowbreak \left[ n_L(\epsilon)+n_R(\hbar \omega_p-\epsilon)+1 \right]$, which is only zero when $T(\epsilon,\hbar \omega_p-\epsilon)=0$ throughout the range, as is the case in Fig. \ref{center}(b). This term can also be understood in Fermi's golden rule point of view, considering the harmonic perturbation $H_T(t)$. According to Fermi's golden rule, the pair creation (annihilation) rates $R_c$ ($R_a$) are:
\begin{align}
&R_c= \int_0^{\hbar \omega_p} d \epsilon T(\epsilon,\hbar \omega_p-\epsilon) \left[ n_L(\epsilon)+1\right] \left[n_R(\hbar \omega_p-\epsilon)+1 \right], \notag\\
&R_a= \int_0^{\hbar \omega_p} d \epsilon T(\epsilon,\hbar \omega_p-\epsilon) n_L(\epsilon) n_R(\hbar \omega_p-\epsilon).
\label{RcRa}
\end{align}
The net creation rate is thus $R_c-R_a=\bar{J}_A.$
One can find the non-equilibrium transport part of the current by subtracting the anomalous squeezing (particle-nonconversing) term $\bar{J}_A$, and we are left with the normal current $\bar{J}_N\equiv (\bar{J}_R-\bar{J}_L)/2$, the first two lines of Eq. (\ref{JRfinal}). We note here that the asymmetry between right and left is necessary for the transport to occur; the first two lines of Eq. (\ref{JRfinal}) will cancel each other otherwise.
To focus on the transport mechanism only, we consider an energy gap on the right transmission line [see Fig.~\ref{center}(b)] such that $\forall \alpha, \beta, \epsilon_{\beta}>\hbar \omega_p,\epsilon_{\beta}>\epsilon_{\alpha}$. This gap setup prevents the pair creation and annihilation mechanisms, leaves us with a conserved current and permits a direct photonic analog to electronic transport. The system follows the transport formula $\bar{J}_R=\int_{\epsilon_{\beta,\rm min}}^{\epsilon_{\alpha,\rm max}+\hbar \omega_p} d \epsilon T (\epsilon-\hbar \omega_p,\epsilon) \left[ n_L(\epsilon-\hbar \omega_p)-n_R(\epsilon)\right]$, which is equivalent to a non-equilibirum transport current under a chemical potential imbalance $\mu_L=\hbar \omega_p, \mu_R=0.$
One can see the resemblance between our gapped transport equation and the I-V (current-voltage) characteristic of an ideal light emitting diode (LED) \cite{ries1991chemical} by relating the chemical potential $\hbar \omega_p$ to $qV$, and the gap energy $\epsilon_{\beta,\rm min}$ to the photon energy threshold $e_g$, and working under the region $\epsilon_{\alpha,\rm max}+\hbar\omega_p \gg k_B T$. However, we cannot yet make a direct connection mathematically with the somewhat different problem of electron transport through a diode combined with emission of photon into an interacting region.
\section{Photon Transport through a mesoscopic central region}
\begin{figure}
\includegraphics[width=3.0in]{fig2}
\caption{(a) The generic mesoscopic scenario, with two semi-infinite leads coupled parametrically to an intermediate mesoscopic region $C$ provides a photonic equivalent to a voltage bias and a (nonlinear) impedance provided by the region $C$. (b) A schematic diagram of the parametric coupling mechanism, in which a low energy photonic mode on the left side is up-converted via the pump photon with energy $\hbar \omega_p$ to a higher energy mode on the right. The case where $L$ and $R$ leads have a high (left) and low (right) frequency cutoff is shown, to prevent anomalous squeezing terms.}
\label{center}
\end{figure}
Now we consider a more generic case with a center mesoscopic region placed between the transmission lines, with parametric coupling between the center region and the left transmission line (see Fig.~\ref{center}). We replace the time-dependent barrier $H_T(t)$ with $H_C+H_{CL}(t)+H_{CR}$, and the Hamiltonian becomes
\begin{align}
H=H_L+H_C+H_R+H_{CL}(t)+H_{CR},
\label{CH}
\end{align}
\begin{align}
& H_L=\sum_{\alpha} \epsilon_{\alpha} a^{\dagger}_{\alpha} a_{\alpha}, \, H_R=\sum_{\beta} \epsilon_{\beta} b^{\dagger}_{\beta} b_{\beta},\notag \\
&H_C= \sum_{\gamma,\gamma {'}} t_{\gamma \gamma {'}} c^{\dagger}_{\gamma} c_{\gamma '}+H_{\rm int}, \, \notag \\
&H_{CL}(t)=\cos(\omega_p t) \sum_{\alpha,\gamma} \lambda_{\alpha \gamma}u_{\alpha}u_{\gamma}, \notag\\
&H_{CR}=\sum_{\beta,\gamma} \breve{\lambda}_{\beta \gamma}u_{\beta}u_{\gamma}.
\label{CHsub}
\end{align}
The summation index $\gamma$ labels the states in the center with energies $\hbar \omega_{\gamma}=\epsilon_{\gamma}$ and photon annihilation operators $c_{\gamma}$. Note that we place all time-dependence in the center to left coupling $H_{CL}(t)$. The central region can contain some nonlinear interacting term $H_{\rm int}$ as well as non-trivial single-particle potential effects from $t_{\gamma\gamma'}$.
We again assume the subsystems were initially in their own thermal equilibrium before the parametric coupling adiabatically turned on at $t = - \infty$.
The current on the right can be expressed as
\begin{align}
J_R(t)=-i \left( \left. \frac{\partial}{\partial t^{'}}\text{Tr}\left[ \mathbf{G}^{<}_{CR}(t,t{'}) \tilde{\mathbf{V}}^{RC}\right]\right) \right|_{t {'}=t}.
\label{CJRinG}
\end{align}
Here $\tilde{V}^{RC}_{\beta \gamma}=V^{RC}_{\beta \gamma}/ \omega_{\beta}=\breve{\lambda}_{\beta \gamma} / \omega_{\beta}.$
Accordingly to the Dyson equation, $\mathbf{G}^c_{CR}(\tau, \tau {'})=\int_{C} \mathbf{G}_{CC}^c(\tau, \tau {''}) \mathbf{V}^{CR}\mathbf{g}^c_R(\tau {''}, \tau {'}).$ Using the Langreth theorem of analytic continuation and defining $\mathbf{\tilde{\Sigma}}_R(t_1,t{'})=\mathbf{V}^{CR}\mathbf{g}_R(t_1,t{'}) \mathbf{\tilde{V}}^{RC},$ we have
\begin{align}
J_R(t)=-i \int_{-\infty}^{\infty} dt_1 \Big( {\frac{\partial}{\partial t '}}\text{Tr} & \left[ \mathbf{G}^r_{CC}(t, t_1) \mathbf{\tilde{\Sigma}}^<_R(t_1, t {'})+ \right. \notag\\
& \left. \left. \left. \mathbf{G}^<_{CC}(t, t_1) \mathbf{\tilde{\Sigma}}^a_R(t_1, t {'})\right] \right) \right|_{t {'}=t}.
\label{CJRgeneral}
\end{align}
This is our generic current expression analogous to the Meir-Wingreen equation in the electronic transport theory \cite{meir1992landauer}.
One can reformulate the current through Fourier transform
\begin{align}
J_R(t)=\int_{-\infty}^{\infty} \frac{\omega d \omega}{2 \pi} \text{Tr}\left[ \mathbf{G}^r_{CC}(t, \omega) \mathbf{\tilde{\Sigma}}^<_R(\omega)+ \mathbf{G}^<_{CC}(t, \omega) \mathbf{\tilde{\Sigma}}^a_R(\omega)\right].
\label{CJRFT}
\end{align}
Here $\mathbf{G}_{CC}(t, t{'}) \equiv \int_{-\infty}^{\infty} \frac{d \omega}{2 \pi} e^{-i \omega (t-t{'})} \mathbf{G}_{CC}(t,\omega )$, $\mathbf{G}_{CC} (t, \omega ) \equiv \int_{-\infty}^{\infty} d t {'} e^{ i \omega (t-t{'})} \mathbf{G}_{CC} (t, t{'})$.
We note here that due to the time-dependent coupling, the center Green's function is not in a steady state, and its Fourier transform function therefore depends on the initial (or ending) time index.
We now examine the simplified case of non-interacting center, and the main result of this work follows through. Non-interacting mesoscopic transport theory provides interesting phenomena such as weak localization, ballistic-to-diffusive transitions, and weak-antilocalization \cite{lee1985disordered,imry2002introduction,akkermans2007mesoscopic}. For $H_{\rm int}=0$, the center Green's function follows the Dyson equation
\begin{align}
\mathbf{G}^c_{CC}(\tau, \tau {'})=\mathbf{g}^c_C(\tau, \tau {'})+\int_{C} \mathbf{g}_{C}^c(\tau, \tau_1) \mathbf{\Sigma}^c_{\rm tot}(\tau_1,\tau_2) \mathbf{G}_{CC}^c(\tau_2, \tau {'}).
\label{CDysonBallistic}
\end{align}
Here $\mathbf{\Sigma}^c_{\rm tot}(\tau_1,\tau_2)$ is the total self-energy of the center,
\begin{align}
&\mathbf{\Sigma}^c_{\rm tot}(\tau_1,\tau_2)=\mathbf{\Sigma}^c_L(\tau_1,\tau_2)+\mathbf{\Sigma}^c_R(\tau_1,\tau_2), \notag\\
&\mathbf{\Sigma}^c_L(\tau_1,\tau_2)\equiv \mathbf{V}^{CL}\cos(\omega_p \tau_1)\mathbf{g}^c_L(\tau_1,\tau_2)\cos(\omega_p \tau_2)\mathbf{V}^{LC},\notag\\
&\mathbf{\Sigma}^c_R(\tau_1,\tau_2)\equiv \mathbf{V}^{CR}\mathbf{g}^c_R(\tau_1,\tau_2)\mathbf{V}^{RC}.
\label{CSigma}
\end{align}
Note that for the case of interacting center, there will be additional contribution to the self-energy depending on the details of $H_{\rm int}$.
Specifically, the center greater and lesser Green's function follows
\begin{align}
\mathbf{G}^{\stackrel{>}{<}}_{CC}(t, t {'})=\int dt_1 dt_2 \mathbf{G}^r_{CC}(t, t_1) \mathbf{\Sigma}^{\stackrel{>}{<}}_{\rm tot}(t_1,t_2) \mathbf{G}^a_{CC}(t_2, t {'}),
\end{align}
\label{CG><}
which allows us to further simplify the current expression.
The center Green's function under parametric coupling can be expanded with the harmonics of the coupling frequency $\omega_p$. Specifically, $\mathbf{G}_{CC} (t, \omega )=\sum_n\mathbf{G}_{CC,n} (\omega) e^{2 n i \omega_p t}, n \in \mathbb{Z}$. Under the rotating wave approximation, we neglect fast oscillating terms with $n \neq 0$ and keep only the $n=0$ steady part of the current. The time averaged current is now
\begin{align}
\bar{J}_R=\int_{-\infty}^{\infty} \frac{d \omega}{2 \pi} \omega \text{Tr}\left[ \mathbf{G}^r_{CC,0}(\omega) \mathbf{\tilde{\Sigma}}^<_R(\omega)+ \mathbf{G}^<_{CC,0}(\omega) \mathbf{\tilde{\Sigma}}^a_R(\omega)\right].
\label{CJRRWA}
\end{align}
Since the current is real, $\bar{J}_R=(\bar{J}_R+\bar{J}^*_R)/2$. Using the general identity $G^>-G^<=G^r-G^a$ and identities for steady state Green's functions in the frequency domain $[G^r(\omega)]^{\dagger}=G^a(\omega)$, $[G^<(\omega)]^{\dagger}=-G^<(\omega)$,
\begin{align}
\bar{J}_R=\int_{-\infty}^{\infty} \frac{d \omega}{4 \pi} \omega \text{Tr} & \left\{ \left[\mathbf{G}^>_{CC,0}(\omega)-\mathbf{G}^<_{CC,0}(\omega)\right] \mathbf{\tilde{\Sigma}}^<_R(\omega) \right. \notag\\
& \left.-\mathbf{G}^<_{CC,0}(\omega) \left[\mathbf{\tilde{\Sigma}}^r_R(\omega)-\mathbf{\tilde{\Sigma}}^a_R(\omega)\right]\right\}.
\label{CJRRWA2}
\end{align}
Expanding the non-interacting center greater and lesser Green's functions to the leading order term yields:
\begin{align}
\mathbf{G}^{\stackrel{>}{<}}_{CC}(t, t {'}) \approx& \int dt_1 dt_2 \mathbf{g}^r_{CC}(t, t_1) \mathbf{\Sigma}^{\stackrel{>}{<}}_{\rm tot}(t_1,t_2) \mathbf{g}^a_{CC}(t_2, t {'})\notag\\
&+\mathcal{O}(\lambda^3), \notag\\
\mathbf{G}^{\stackrel{>}{<}}_{CC,0}(\omega) \approx &\mathbf{g}^r_{C}(\omega) \mathbf{\Sigma}^{\stackrel{>}{<}}_{\rm tot,0}(\omega) \mathbf{g}^a_{C}(\omega) \notag\\
=& \mathbf{g}^r_{C}(\omega) \left( \mathbf{V}^{CR} \mathbf{g}^{\stackrel{>}{<}}_R(\omega)\mathbf{V}^{RC}\right. \notag\\
&+\frac{1}{4}\mathbf{V}^{CL} \mathbf{g}^{\stackrel{>}{<}}_L(\omega+\omega_p)\mathbf{V}^{LC}\notag\\
&+\left.\frac{1}{4}\mathbf{V}^{CL} \mathbf{g}^{\stackrel{>}{<}}_L(\omega-\omega_p)\mathbf{V}^{LC} \right) \mathbf{g}^a_{C}(\omega).
\label{CG><p}
\end{align}
By inserting the equilibrium Green's function for left and right transmission lines $g_L$ and $g_R$, we arrive at a formula similar to the trivial scatterer problem
\begin{align}
\bar{J}_R=&\int_{\hbar \omega_p}^{\infty} d \epsilon T_C (\epsilon,\epsilon-\hbar \omega_p) \left[ n_L(\epsilon)-n_R(\epsilon-\hbar \omega_p)\right] \notag\\
+& \int_{\hbar \omega_p}^{\infty} d \epsilon T_C(\epsilon-\hbar \omega_p,\epsilon) \left[ n_L(\epsilon-\hbar \omega_p)-n_R(\epsilon) \right] \notag\\
+& \int_0^{\hbar \omega_p} d \epsilon T_C(\epsilon,\hbar \omega_p-\epsilon) \left[ n_L(\epsilon)+n_R(\hbar \omega_p-\epsilon)+1 \right] .
\label{CJRfinal}
\end{align}
where the center transmission function is
\begin{align}
&T_C(\epsilon_{\alpha},\epsilon_{\beta})=\frac{\pi \hbar^3}{8}\text{Tr}\left[ \mathbf{g}_c^r(\epsilon_{\beta})\mathbf{\Lambda}_R(\epsilon_{\beta})\mathbf{g}_c^a(\epsilon_{\beta})\mathbf{\Lambda}_L(\epsilon_{\alpha})\right], \notag\\
&[\Lambda_L(\epsilon_{\alpha})]_{\gamma_1,\gamma_2}=\rho_L(\epsilon_{\alpha})\lambda_{\gamma_1}(\epsilon_{\alpha})\lambda_{\gamma_2}(\epsilon_{\alpha})/\epsilon_{\alpha},\notag \\ &[\Lambda_R(\epsilon_{\beta})]_{\gamma_1,\gamma_2}=\rho_R(\epsilon_{\beta})\breve{\lambda}_{\gamma_1}(\epsilon_{\beta})\breve{\lambda}_{\gamma_2}(\epsilon_{\beta})/\epsilon_{\beta}.
\end{align}
\label{CTC}
We again have the first two lines of Eq. (\ref{CJRfinal}) as the Landauer-like transport terms, and the last line is the particle-nonconserving part due to the oscillating $u$-$u$ type coupling.
The system will undergo non-equilibirum transport with an effective chemical potential imbalance $\hbar \omega_p$ under specific gap setups, and the current expression resembles the I-V characteristic of an ideal light-emitting diode.
\section{Conclusions}
We have derived the photonic flux between different baths parametrically-coupled to an intermediate system and found a Landauer-like transport formula for non-interacting centers. However, we also have another regime, with a particle-nonconserving term, which we can interpret as a two-mode squeezing output. The consequences of this latter regime for observation and even application remain to be explored, and are beyond the scope of the present work. We have also shown a potential extension of these techniques at the formal level to the interacting case, but suggest that applying these results, e.g., to photon-blockaded systems to see the non-classical light output would be an intriguing next step.
\begin{acknowledgments}
Funding is provided by NSF Physics Frontier Center at the JQI, and from the Princeton Center for Complex Materials, a MRSEC supported by NSF Grant No. DMR 1420541.
\end{acknowledgments}
|
\section{Introduction}
Wireless sensor nodes are small computer devices with low production costs, equipped with a radio antenna and sensors capable of sensing one or more environmental parameters~\cite{Akyildiz2002}. As sensor nodes are designed to have low production costs, their computational and energy resources are several orders of magnitude smaller than those of typical workstations. This constraint does not prevent Wireless Sensor Networks (WSNs) to be composed of hundreds of wireless sensor nodes deployed to monitor the environment and engineering structures, track objects, detect meteorological phenomena, among others.
Apart from being used for traditional monitoring tasks, in the last years, WSNs have been incorporated into Internet of Things (IoT) applications, where connected cloud computing services analyze the collected data and consequently trigger reactions~\cite{Bellavista2013}. As sensor nodes usually cannot compute complex algorithms that require a long runtime or significant memory resources, a data analysis performed in sensor nodes may not be as accurate or as fast as those computed in multi-core workstations with high storage and memory capacity.
As a side note, the IoT has a broader scope than traditional WSNs, because it is composed of more powerful devices that can compute complex algorithms, interact with humans and also provide machine-to-machine communication. Indeed, the power of the IoT is not only concentrated in devices and their applications but also, among others, in potential data analytics and interactions between different device types. Therefore, while early (traditional) WSNs performed simple data collection tasks and were merely considered as data providers, sensor nodes can be nowadays benefited by intelligent data applications, such as the optimization of the sensors' sampling interval based on data predictions~\cite{Dias2016}. By way of illustration, in the Entomatic project\footnote{http://entomatic.upf.edu}, wireless sensor nodes periodically report information on pest population density and environmental parameters, such as temperature and relative humidity. The data analysis results can be used to spray pesticides intelligently, i.e., only when--and where--necessary.
Our contribution is an architecture that relies on a robust entity to exploit the asymmetry in WSN devices' capacities when computing data algorithms. First, it overcomes a common restriction imposed in other platforms for WSNs~\cite{turon2005mote,ruzzelli2008octopus}, where WSN owners have to upload new applications manually whenever they decide to take decisions based on sophisticated data processing techniques computed outside the WSNs. At the same time, our architecture completes other IoT platforms that integrate a user interface to visualize and analyze the collected data~\cite{nimbits2016,realtimeio2016}, because it offers means for WSNs improving their operation according to the knowledge acquired in data analysis. In this paper, we consider a set of use cases to illustrate these features.
\section{Proposed Architecture}
\label{section:architecture}
We propose an architecture that exploits the sensor nodes hardware limitations and the physical distance from data origin to the central server. Our platform fits the main principles of data science related with sensed data: its collection; description; storage; maintenance; discovery; visualization; and analysis. As the platform stores and publishes collected sensed data, it is possible to visualize measurements and other collected values, as well as reproduce IoT scenarios to optimize data acquisition and its further analysis.
The data analysis can rely on complex computations to extract knowledge from the collected data and provide reliable services at a low cost~\cite{7123563}. In this architecture, they can be delivered as a service to WSNs, such that sensor nodes will be able to apply the knowledge in their favor, e.g., changing their operation to report measurements more often and detail the variations in the environment.
\subsection{Components}
\begin{figure}[t]
\centering
\includegraphics[width=0.45\textwidth]{dashboard-architecture}
\caption{System Architecture}
\label{fig:system-architecture}
\end{figure}
We consider a typical IoT environment composed by several WSNs with ordinary wireless sensor nodes reporting to predetermined sinks, called as \emph{Gateways} (\gateway{s}) in Figure~\ref{fig:system-architecture}. The proposed architecture is centered in the \emph{Data Analytics for Sensors
Dashboard} (\dashboard{}) and consists of interconnected components that can exchange information with trusted entities, eventually from different domains. The components of this architecture are:
\subsubsection{Wireless sensor nodes}
Taking advantage of their proximity to the data origin, they perform default sensing tasks in their deployment area and transmit their measurements via radio to a \gateway{}.
\subsubsection{Gateways}
They forward the gathered information to the central server and disseminate occasional instructions and updates to wireless sensor nodes. \gateway{s} are the link between ordinary wireless sensor nodes and the \dashboard{}, using a point-to-point connection over the Internet or a local network.
\subsubsection{\dashboard{}}
The central component of this architecture has three primary responsibilities: collecting, storing and publishing data transmitted by wireless sensor nodes. The ability to collect data requires a direct communication with the WSN and is fundamental for the other two responsibilities. Storing the collected data in a database (DB) allows further access to historical information, besides providing data visualization to network owners and other users. Finally, communicating data to other systems allows the dashboard to outsource data processing, such as filtering and analyzing the collected data and, especially, predicting future measurements.
\subsubsection{Data Analytics Server}
The Data Analytics Server can process computationally expensive real-time analysis over data. To do that, it can rely on external data resources, such as public services and other databases via the Internet. Indeed, tasks processed by the Data Analysis Server could not be assigned to sensor nodes, due to their constrained hardware and limited communication with external networks.
\subsection{Communication with external components}
The communication between \dashboard{} and any external component follows a standard: external data (coming from \gateway{s} and the Data Analytics Server) is received via APIs (GET or POST HTTP requests); and outbound data is announced in the form of events via socket connections to previously registered services that keep listening for updates. If connected to the Internet, the \dashboard{} may make data publicly available to remote access via an online web-based platform. Hence, besides storing data and managing users access, the system architecture supports data analytics algorithms and allows the \dashboard{} to change sensors' operations according to the data they have previously reported. For example, sensor nodes reconfiguration may be done via peer-to-peer or multicast communication, based on the analysis published by the \dashboard{}.
\section{Self-Management Demo}
In this demonstration, we show that our architecture can exploit the link between the \dashboard{} and a WSN, by analyzing the measured data and adjusting the operation of sensor nodes according to the analysis outcomes. Our WSN is composed by one \gateway{} and four sensor nodes that work independently of each other, representing use cases in which sensor nodes monitor parameters from indoor and outdoor environments.
The \dashboard{} is deployed in a well-dimensioned machine without energy nor performance constraints, with reliable Internet connection and direct access to the other architecture components. A Web User Interface provides the user a means of visualizing the collected data, the sensor nodes' location in a map, occasional failures that may be reported and results from the data analysis, performed by an external server. The Data Analytics Server is developed in R programming language and may run two types of analysis: \emph{data relevance assessment} and \emph{forecasting}. In the following, we describe each analysis procedure, illustrating the interconnection between a WSN, the \dashboard{} and the Data Analytics Server in a real deployment.
\subsection{Data relevance assessment}
In this analysis, data is used to calculate the importance of a sensor in the measurements reported by the whole WSN. To achieve that, it may be possible to use external information, such as the time of the day and the temperature reported by an online weather service. As a result, if the data analysis is about the relevance of the data generated by sensor nodes, the output is a suggestion about updating the sensor nodes' operation to sample more (or less) often.
\subsubsection{Single rule demo}
\begin{figure}[t]
\centering
\includegraphics[width=0.38\textwidth]{results-print2}
\caption{In this example, the sampling interval is updated every $12$ minutes, and it varies between $60$, $120$ and $240$ seconds.}
\label{fig:adaptive-sampling}
\end{figure}
One example of this type of data analysis is a single rule defining the sampling interval according to the time of the day. For instance, if it is a working hour, sensor nodes placed in an office have to measure the temperature once every $5$ minutes; otherwise, they can be set to measure it less often (e.g., every $30$ minutes). Hence, the WSN would report more detailed information when people are in its surroundings because their presence may impact the temperature and also because other systems (such as the air conditioning and personal computers) work only during working hours, provoking higher variations and sharper changes. Figure~\ref{fig:adaptive-sampling} depicts an image from the \dashboard{}, illustrating a rule that changes the sampling interval every $12$ minutes.
\subsubsection{External information demo}
Another example is the use of the Internet access to compare the data reported by sensor nodes with the temperature values reported by a weather service. If the values coincide sufficiently, the \dashboard{} stores in the database the forecast values from the weather service and reduces the sensor nodes' sampling interval to save their battery. Otherwise, sensor nodes keep measuring as often as possible to inform the temperature values accurately to the WSN owner. The support of a reliable external forecasting service can be justified by the high complexity of forecasts provided by its supercomputers, which are unfeasible to be locally performed without additional--and significant--costs.
\subsection{Forecasting}
The strategy known as Dual Prediction Scheme (DPS) relies on forecasts of values that will be measured by the wireless sensor nodes~\cite{Dias2016b}. After predicting the future measurements, the \dashboard{} informs the sensor nodes which specific values are expected to be measured, as illustrated in Figure~\ref{fig:timeline-dps-ch}. If the actual measurements match to the forecasts (or do not differ by more than an accepted threshold), they are not transmitted, saving sensor nodes' batteries and reducing the wireless medium occupancy.
\begin{figure}[t]
\centering
\includegraphics[width=0.45\textwidth]{timeline-dps-ch}
\caption{In a DPS, a measurement is transmitted only if its
forecast is inaccurate. The \dashboard{} may be responsible for
transmitting new prediction models every time interval after the
initialization phase.}
\label{fig:timeline-dps-ch}
\end{figure}
\subsubsection{Autoregressive Integrate Moving Average demo}
The AutoRegressive Integrate Moving Average (ARIMA) is a technique used to forecast values which can accurately predict up to $20$ future measurements and reduce up to $50\%$ of the number of transmissions~\cite{Dias2016}. Due to its high computational complexity, an ARIMA model computation might waste sensor nodes' battery without providing enough savings to compensate it. Therefore, even though the ARIMA method is not suitable to the simplest wireless sensor nodes, in this architecture we can exploit the computational asymmetry of WSN devices and avoid unnecessary data transmissions.
\section{Conclusion}
In this paper, we present a system architecture that enables the integration of device types with different computing powers and capacities. To demonstrate its application in a real-world use case, we apply the obtained results in real-time data analysis to optimize a WSN's operation and visualize its evolution in a Web User Interface.
\section*{Acknowledgment}
This work has been partially supported by the Spanish Government through the
project TEC2012-32354 (Plan Nacional I+D), by the Catalan Government
through the project SGR-2014-1173 and by the European Union through the
project FP7-SME-2013-605073-ENTOMATIC.
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Modern radio interferometric arrays deliver large volumes of data, in order to reach higher sensitivities yielding new science. To reach the full potential of such arrays, estimation of systematic errors in the data and correction for such errors (also called as calibration) is essential. This is not a trivial task for an array with hundreds of receivers that collect data over many hours and at thousands of different frequencies. A case in point being the square kilometre array (SKA), which is in the planning phase. Thus, there is an urgent need for computationally efficient and robust algorithms. On the other hand, there is a surge in research related to large scale and distributed data processing algorithms (also called as big-data), which we can exploit to solve some of these problems.
Our recent work \cite{DCAL} introduced distributed-calibration as a way of distributing the computational burden over a network of computers while at the same time improving the quality of calibration. We essentially exploited the continuity of systematic errors over frequency to enforce an additional constraint onto calibration. This reduces calibration to a consensus optimization \cite{boyd2011} problem and we used alternating direction method of multipliers (ADMM) \cite{BT} as the underlying algorithm in the proposed distributed calibration scheme.
Consensus optimization, practically implemented with ADMM, has been extensively studied and is deployed in a wide variety of application areas (some recent examples are \cite{Chang2014,Wei2012,Erseghe12}). In addition, similar work is beginning to appear in radio astronomical imaging \cite{Ferrari2014,PURIFY,Onose}. However, compared with other users of ADMM, we observe several unique properties of the calibration problem that we face. First, the cost function used in calibration is non-linear and non-convex. The systematic errors are mainly caused by directional effects such as the ionosphere and the receiver beam shape. Although we know the general properties of such errors, building an entirely accurate model (for instance for their variation with frequency) is not feasible. Hence, we enforce consensus only by using an approximate model, and this is clearly different and also more involved from most other applications. Indeed, other applications such as consensus averaging, where consensus is enforced on a constant value, use a perfect model. Furthermore, most other applications use complicated network topologies (that in turn affect the performance of ADMM) and on the other hand, in our case, we have a much simpler (and fully connected) network with one fusion center.
Of particular interest is the convergence rate of ADMM, which depends on many factors including the penalty parameter and the network topology \cite{nishihara2015general}. In most cases, the penalty parameter is selected by trial and error, following some general guidelines \cite{BT}. However, for specific problems, better methods to select the penalty have been proposed \cite{nishihara2015general,Teix2016,Ghadimi2015}. Recent work \cite{Hong15} has suggested to select the penalty parameter as large as possible to make the objective function strongly convex. Hence for our problem, we study the Hessian of the cost function to select appropriate values for the penalty parameter. For calibration along multiple directions in the sky, we can select different penalty values along each direction. Intuitively, we select a large penalty along directions with higher signal where we have more confidence in our model. These directions are mostly close to the center of the field of view. On the other hand, for directions far away from the center, we select a smaller penalty.
The rest of the paper is organized as follows: In section \ref{sec:calib} we give an overview of radio interferometric calibration. Next, in section \ref{sec:dist}, we present distributed calibration based on consensus optimization. We also present a scheme based on the Hessian of the cost function to select the penalty parameter. Simulation results are presented in section \ref{sec:results} where we demonstrate the improved performance with a refined penalty parameter. Finally, we draw our conclusions in section \ref{sec:conclusions}.
Notation: Matrices and vectors are denoted by bold upper and lower case letters as ${\bf J}$ and ${\bf v}$, respectively. The transpose and the Hermitian transpose are given by $(.)^T$ and $(.)^H$. The matrix Frobenius norm is given by $\|.\|$. The set of real and complex numbers are denoted by ${\mathbb R}$ and ${\mathbb C}$. The identity matrix is given by $\bf I$. The matrix trace operator is given by $\rm{trace}(.)$.
\section{Radio Interferometric Calibration}\label{sec:calib}
Consider a radio interferometric array with $N$ receivers. The sky is composed of many discrete sources and we consider calibration along $K$ directions in the sky. The observed data at a baseline formed by two receivers, $p$ and $q$ is given by
\cite{HBS}
\begin{equation} \label{ME}
{\bf V}_{pq}=\sum_{k=1}^K{\bf J}_{pk} {\bf C}_{pqk} {\bf J}_{qk}^H + {\bf N}_{pq}
\end{equation}
where ${\bf V}_{pq}$ ($\in \mathbb{C}^{2\times 2}$) is the observed {\em visibility} matrix (or the cross correlations). The systematic errors that need to be calibrated for station $p$ and $q$ are given by the Jones matrices ${\bf J}_{pk},{\bf J}_{qk}$ ($\in \mathbb{C}^{2\times 2}$), respectively. Note that since $K$ directions are calibrated, for each station, there are $K$ Jones matrices (so $KN$ in total). The sky signal (or {\em coherency}) along the $k$-th direction is given by ${\bf C}_{pqk}$ ($\in \mathbb{C}^{2\times 2}$) and is known a priori. The values of ${\bf J}_{pk},{\bf J}_{qk}$ and ${\bf C}_{pqk}$ in (\ref{ME}) are implicitly dependent on sampling time and frequency of the observation. The noise matrix ${\bf N}_{pq}$ ($\in \mathbb{C}^{2\times 2}$) is assumed to have complex, zero mean, circular Gaussian elements.
Estimating the Jones matrices in (\ref{ME}) can be further simplified by using the space alternating generalized expectation maximization (SAGE) algorithm \cite{Fess94,Kaz2}. In a nutshell, using SAGE algorithm, we can simplify calibration along $K$ directions to $K$ single direction calibration subproblems (see \cite{Kaz2} for details). Calibration along the $k$-th direction is done by using the effective observed data
\begin{equation} \label{ME1}
{\bf V}_{pqk} = {\bf V}_{pq} - \sum_{l=1,l\ne k}^K\widehat{\bf J}_{pl} {\bf C}_{pql} \widehat{\bf J}_{ql}^H
\end{equation}
using current estimates $\widehat{\bf J}_{pl}$ and $\widehat{\bf J}_{ql}$ and for an array with $N$ receivers, we can form at most $N(N-1)/2$ baselines that collect visibilities as in (\ref{ME1}), for any given time and frequency sample. We define our objective function (for the $k$-th direction) under a Gaussian noise model as
\begin{equation} \label{cost1}
g_{k}({\bf J}_{1k},{\bf J}_{2k},\ldots)= \sum_{p,q}\| {\bf V}_{pqk} - {\bf J}_{pk} {\bf C}_{pqk} {\bf J}_{qk}^H \|^2
\end{equation}
where the summation is over the baselines $pq$ that have data. By increasing the time and frequency interval within which data are collected, this summation can be expanded (thus improving the signal to noise ratio).
By defining ${\bf J}$ ($\in \mathbb{C}^{2N\times 2}$) as the augmented matrix of Jones matrices of all stations along the $k$-th direction,
\begin{equation}
{\bf J}\buildrel\triangle\over=[{\bf J}_{1k}^T,{\bf J}_{2k}^T,\ldots,{\bf J}_{Nk}^T]^T,
\end{equation}
and ${\bf A}_p$ ($\in \mathbb{R}^{2\times 2N}$) (and ${\bf A}_q$ likewise) as the canonical selection matrix
\begin{equation} \label{Ap}
{\bf A}_p \buildrel\triangle\over=[{\bf 0},{\bf 0},\ldots,{\bf I},\ldots,{\bf 0}],
\end{equation}
(only the $p$-th block of (\ref{Ap}) is an identity matrix) we can rewrite (\ref{cost1}) as
\begin{equation} \label{cost2}
g_{k}({\bf J})= \sum_{p,q}\| {\bf V}_{pqk} - {\bf A}_p{\bf J} {\bf C}_{pqk} ({\bf A}_q{\bf J})^H \|^2.
\end{equation}
Calibration along the $k$-th direction is the estimation of ${\bf J}$ by minimizing (\ref{cost2}). Note that (\ref{cost2}) has to be minimized for each direction $k=1\ldots K$ and updated values of (\ref{ME1}) are re-used until convergence is reached in the SAGE algorithm. We also note that (\ref{cost2}) only gives solutions for one frequency and time interval, and to calibrate the full dataset, many such solutions are obtained for data observed at different time and frequency intervals.
\section{Distributed Calibration}\label{sec:dist}
We have introduced calibration along $K$ directions, but only working on a single frequency and time sample in section \ref{sec:calib}. In this section, we consider calibrating data observed at $P$ different frequencies, but only along $1$ direction, because this can easily be extended to $K$ directions using the SAGE algorithm. We impose an additional constraint that tries to preserve continuity of ${\bf J}$ in (\ref{cost2}) over frequency. To solve this, we introduced the use of consensus optimization in \cite{DCAL}, where the objective function is modified into an augmented Lagrangian
\begin{equation} \label{aug}
L_f({\bf J}_f,{\bf Z},{\bf Y}_f)=g_{f}({\bf J}_f) + \|{\bf Y}_f^H({\bf J}_f-{\bf B}_f {\bf Z})\| + \frac{\rho}{2} \|{\bf J}_f-{\bf B}_f {\bf Z}\|^2
\end{equation}
where the subscript $(.)_f$ denotes data (and parameters) at frequency $f$. In (\ref{aug}), $g_{f}({\bf J}_f)$ is the original cost function as in (\ref{cost2}), except that the subscripts denote frequency $f$. The Lagrange multiplier is given by ${\bf Y}_f$ ($\in \mathbb{C}^{2N\times 2}$). The calibration parameters are given by ${\bf J}_f$ ($\in \mathbb{C}^{2N\times 2}$). The continuity in frequency is enforced by the frequency model given by ${\bf B}_f$ ($\in \mathbb{R}^{2N\times 2NF}$), which is essentially a set of basis functions in frequency, evaluated at $f$. The global variable ${\bf Z}$ ($\in \mathbb{C}^{2NF\times 2}$) is shared by data at all $P$ frequencies.
The ADMM iterations for solving (\ref{aug}) are given as
\begin{eqnarray} \label{step1}
({\bf J}_f)^{n+1}= \underset{{\bf J}}{\argmin}\ \ L_f({\bf J},({\bf Z})^n,({\bf Y}_f)^n)\\ \label{step2}
({\bf Z})^{n+1}= \underset{{\bf Z}}{\argmin}\ \ \sum_f L_f(({\bf J}_f)^{n+1},{\bf Z},({\bf Y}_f)^n)\\ \label{step3}
({\bf Y}_f)^{n+1}=({\bf Y}_f)^n + \rho\left( ({\bf J}_f)^{n+1}-{\bf B}_f ({\bf Z})^{n+1} \right)
\end{eqnarray}
where we use the superscript $(.)^n$ to denote the $n$-th iteration. The steps (\ref{step1}) and (\ref{step3}) are done for each $f$ in parallel. The update of the global variable (\ref{step2}) is done at the fusion center. More details of these steps can be found in \cite{DCAL}.
In this paper, we study strategies for selecting the penalty parameter $\rho$ to get faster convergence and accurate results. In order to do this, we use the Hessian operator of the cost function (\ref{cost2}), which is given as \cite{DCAL,ICASSP13},
\begin{eqnarray}\label{Hess}
\lefteqn{\mathrm{Hess}_f\left(g_{f}({\bf {J}}),{\bf {J}},{\bmath \eta}\right)}\\\nonumber
&=&\sum_{p,q}\left( {\bf {A}}_p^T \left( ({\bf {V}}_{pqf}-{\bf {A}}_p{\bf {J}}{\bf {C}}_{pqf}{\bf {J}}^H{\bf {A}}_q^T) {\bf {A}}_q {\bmath \eta}\right.\right.\\\nonumber
&& \left.\left.- {\bf {A}}_p({\bf {J}}{\bf {C}}_{pqf} {\bmath \eta}^H + {\bmath \eta}{\bf {C}}_{pqf}{\bf {J}}^H) {\bf {A}}_q^T{\bf {A}}_q{\bf {J}}\right) {\bf {C}}_{pqf}^H\right. \\\nonumber
&&\left. + {\bf {A}}_q^T \left( ({\bf {V}}_{pqf}-{\bf {A}}_p{\bf {J}}{\bf {C}}_{pqf}{\bf {J}}^H{\bf {A}}_q^T)^H {\bf {A}}_p {\bmath \eta}\right.\right.\\\nonumber
&& \left.\left.- {\bf {A}}_q({\bf {J}}{\bf {C}}_{pqf} {\bmath \eta}^H + {\bmath \eta}{\bf {C}}_{pqf}{\bf {J}}^H)^H {\bf {A}}_p^T{\bf {A}}_p{\bf {J}}\right) {\bf {C}}_{pqf}\right) \\\nonumber
\end{eqnarray}
where ${\bmath \eta}\in \mathbb{C}^{2N\times 2}$.
For convexity, we need a positive definite Hessian. Since we have a Hessian operator (instead of a matrix), we need to find the smallest eigenvalue of the Hessian, and for convexity, this should be positive. In order to find this, we define a cost function as
\begin{eqnarray} \label{hcost}
\lefteqn{h({\bmath \eta})\buildrel\triangle\over= \frac{1}{2}\mathrm{trace}\left({\bmath \eta}^H \mathrm{Hess}_f\left(g_{f}({\bf {J}}),{\bf {J}},{\bmath \eta}\right)\right.}\\\nonumber
&&+\left.\mathrm{Hess}_f^H\left(g_{f}({\bf {J}}),{\bf {J}},{\bmath \eta}\right) {\bmath \eta}\right)
\end{eqnarray}
and we find the smallest eigenvalue $\lambda$ by solving
\begin{eqnarray} \label{eig}
&&\lambda=\underset{{\bmath \eta}}{\argmin}\ \ \ \ h({\bmath \eta})\\\nonumber
&&{\mathrm{subject\ to}}\ \ {\bmath \eta}^H{\bmath \eta}={\bf I}.
\end{eqnarray}
The constraint ${\bmath \eta}^H {\bmath \eta}={\bf I}$ makes the minimization of (\ref{hcost}) restricted onto a complex Stiefel manifold \cite{AMS}, which can be easily solved by using the Riemannian trust region method \cite{RTR,manopt}. In order to do this, we require the gradient and Hessian of $h({\bmath \eta})$, which are given as
\begin{equation}
\mathrm{grad}\left(h({\bmath \eta}),{\bmath \eta}\right)= \mathrm{Hess}_f\left(g_{f}({\bf {J}}),{\bf {J}},{\bmath \eta}\right)
\end{equation}
and
\begin{equation}
\mathrm{Hess}\left(h({\bmath \eta}),{\bmath \eta},{\bmath \zeta}\right)= \mathrm{Hess}_f\left(g_{f}({\bf {J}}),{\bf {J}},{\bmath \zeta}\right),
\end{equation}
where ${\bmath \zeta} \in \mathbb{C}^{2N\times 2}$.
After obtaining $\lambda$ from (\ref{eig}), our strategy is to select $\rho$ such that $\rho+\lambda\ge 0$ so that the Hessian of the augmented Lagrangian (\ref{aug}) is positive semi-definite \cite{Hong15}. In order to do this, we need an estimate for ${\bf J}$ in (\ref{hcost}). We can find this by initial calibration with a pre-determined value of $\rho$ (say $\rho=0$). Once we obtain $\widehat{\bf J}$, we use (\ref{eig}) to find $\lambda$ and afterwards we update $\rho$. Note that $\lambda$ is dependent on $f$, but we ignore the frequency dependence of $\lambda$ and use one value of $f$ (typically the middle) to estimate it.
So far, we have considered calibration along one direction only. The next question that we must answer is how to select $\rho$ for calibration along $K$ directions in the sky. For each direction, ${\bf {C}}_{pqf}$ in (\ref{Hess}) will influence the value of $\lambda$. If the centroid of the source (cluster) \cite{Kazemi3} is along $l,m$ direction in the sky and if its effective (unpolarized) intensity is $\alpha$, we have
\begin{equation} \label{coh}
{\bf {C}}_{pqf} \approx \exp\left(\jmath \phi(l,m,p,q)\right) \alpha {\bf I}
\end{equation}
where $\phi(l,m,p,q)$ is the phase contribution and ${\bf I}$ is a $2\times 2$ identity matrix. Hence ${\bf {C}}_{pqf}$ is a diagonal scalar matrix. If $\widehat{\bf J}$ is close to the true solution, the term ${\bf {V}}_{pqf}-{\bf {A}}_p{\bf {J}}{\bf {C}}_{pqf}{\bf {J}}^H{\bf {A}}_q^T$ becomes negligible compared with the other terms in (\ref{Hess}). The remaining terms have a product ${\bf {C}}_{pqf} {\bf {C}}_{pqf}^H$ and the phase term in (\ref{coh}) cancel out. Therefore, for different clusters, the value for $\lambda$ obtained by (\ref{eig}) is mainly determined by the squared effective intensity $\alpha^2$ of each source. Hence, once we have determined a suitable value for $\rho$ for one direction, the corresponding values for other directions can be determined by scaling by the squared effective intensity.
\section{Simulation Results}\label{sec:results}
We simulate an array of $N=47$ receivers that calibrate along $K=5$ directions in the sky. The matrices ${\bf J}_{pk},{\bf J}_{qk}$ in (\ref{ME}) are generated with their elements having values drawn from a complex uniform distribution in $[0,1]$, multiplied by a frequency dependence given by a random $7$-th order polynomial. The intensities of the $K=5$ sources are randomly generated in the range $[1,5]$ and their positions are randomly chosen in a field of view of about $7\times 7$ square degrees. The variation of intensities with frequency is given by a power law with randomly generated exponent in $[-1,1]$. The noise matrices ${\bf N}_{pq}$ in (\ref{ME}) are simulated to have complex circular Gaussian random variables. The variance of the noise is changed according to the signal to noise ratio ($\rm{SNR}=10$)
\begin{equation}
\mathrm{SNR}\buildrel\triangle\over=\frac{\sum_{p,q} \|{\bf V}_{pq}\|^2}{\sum_{p,q} \| {\bf N}_{pq}\|^2}.
\end{equation}
With this setup, we generate data for $P=8$ frequency channels in the range $115$ to $185$ MHz. For calibration, we setup a $3$-rd order polynomial model ($F=4$), using Bernstein basis functions \cite{Farouki} for the matrix ${\bf B}_f$ in (\ref{aug}). Note that we intentionally use a lower order frequency dependence than what is actually present in the data to create a realistic scenario when the exact model is not known. During calibration, initial values for the parameters are always set as ${\bf J}_p={\bf I}$ for $p\in[1,N]$. Unless stated otherwise, all directions have the same value of $\rho$. We use $50$ ADMM iterations, and after the $1$-st iteration, we solve (\ref{eig}) to estimate $\lambda$, and we get a typical value of $\lambda=-150$ for a source with unit amplitude. Regardless, we perform calibration with various values of $\rho$ to compare performance.
We find the normalized (averaged over all directions) mean squared error (NMSE) between true ${\bf J}_f$ and its estimate as
\begin{equation}\label{nmse}
\mathrm{NMSE}\buildrel\triangle\over=\frac{1}{\sqrt{2KN}}\sqrt{\sum_k \|{\bf {J}}_f-\widehat{\bf {J}}_f {\bf {U}}\|^2}
\end{equation}
to measure the accuracy of calibration. In (\ref{nmse}), ${\bf U}$ is a unitary matrix that removes the unitary ambiguity in the estimated $\widehat{\bf J}_f$ \cite{interpolation}.
In Fig. \ref{fignmse}, we show the NMSE for various values of $\rho$, with increasing number of ADMM iterations. We see that for $\rho+\lambda>0$ ($\rho=200$) we get the best performance, but increasing $\rho$ too much beyond this value ($\rho=1000$) shows no additional improvement. A notable behavior of the NMSE is the enhancement of the error at the edges (especially at low ADMM iterations), which we attribute to Runge's phenomenon \cite{Runge} in polynomial interpolation.
\begin{figure}[htbp]
\begin{minipage}[b]{0.98\linewidth}
\begin{minipage}[b]{0.48\linewidth}
\centering \centerline{\epsfig{figure=eusipco_figures/nmse_rho5_1.eps,width=4.2cm}}
\vspace{0.2cm}\centerline{$\ \rho=5$}
\end{minipage}
\begin{minipage}[b]{0.48\linewidth}
\centering \centerline{\epsfig{figure=eusipco_figures/nmse_rho50_1.eps,width=4.2cm}}
\vspace{0.2cm}\centerline{$\rho=50$}
\end{minipage}
\begin{minipage}[b]{0.48\linewidth}
\centering \centerline{\epsfig{figure=eusipco_figures/nmse_rho200_1.eps,width=4.2cm}}
\vspace{0.2cm}\centerline{$\rho=200$}
\end{minipage}
\begin{minipage}[b]{0.48\linewidth}
\centering \centerline{\epsfig{figure=eusipco_figures/nmse_rho1000_1.eps,width=4.2cm}}
\vspace{0.2cm}\centerline{$\rho=1000$}
\end{minipage}
\end{minipage}
\caption{NMSE for various $\rho$ with increasing ADMM iterations.}
\label{fignmse}
\end{figure}
In Fig. \ref{fignmseall}, we show the final NMSE for 50 ADMM iterations, which once again shows that $\rho=200$ gives the best result.
\begin{figure}[htbp]
\begin{minipage}[b]{0.98\linewidth}
\centering
\centerline{\epsfig{figure=eusipco_figures/nmse_all_1.eps,width=5.4cm}}
\end{minipage}
\caption{NMSE for various $\rho$ after $50$ ADMM iterations.} \label{fignmseall}
\end{figure}
\begin{figure}[htbp]
\begin{minipage}[b]{0.98\linewidth}
\centering
\centerline{\epsfig{figure=eusipco_figures/nmse_rhovar_1.eps,width=5.4cm}}
\end{minipage}
\caption{NMSE after $50$ ADMM iterations with fixed $\rho$ along all directions and varying $\rho$ according to squared intensity.} \label{fignmsevar}
\end{figure}
In Fig. \ref{fignmsevar}, we show NMSE for a simulation with intensities at mid frequency $5,3,3,2$ and $1.5$ along the $K=5$ directions. In one calibration, we use regularization $\rho=400$ for all directions and in the other, we use $\rho$ equal to $400,144,144,64$ and $36$ respectively. We see that varying $\rho$ in proportion to the squared intensity gives the better NMSE.
\section{Conclusions}\label{sec:conclusions}
We have investigated refining the performance of distributed calibration based on consensus optimization in this paper. We have used the Hessian of the cost function to appropriately select the penalty parameter such that the augmented Lagrangian becomes convex. Furthermore, in a multi-directional calibration scheme, we have proposed to scale the penalty parameter proportional to the squared intensity along each direction. According to our simulations, such fine-tuning of parameters gives superior performance in terms of accuracy and convergence of the distributed calibration scheme.
\bibliographystyle{IEEE}
|
\section{Introduction}
The search for time-reversal violation (TV) has long been a subject
of considerable experimental and theoretical interest. It is
partially motivated by the need for CP-violation beyond that encoded
in the Standard Model (SM) Cabibbo-Kobayashi-Maskawa (CKM) matrix to
explain the cosmic baryon asymmetry\cite{Sakharov:1967dj}. Assuming
CPT is a good symmetry of nature, searches for TV provide a probe of
this possible CP-violation. Experimentally, the parity (P)- and
T-violating (PVTV) sector is being probed with great sensitivity
through electric dipole moment (EDM) searches, with the three most
stringent limits having been obtained for the $^{199}$Hg
atom\cite{Graner:2016ses}, the electron (extracted from the ThO
molecule)\cite{Baron:2013eja}, and the
neutron\cite{Afach:2015sja,Baker:2006ts}.
hand, the P-conserving and T-violating (PCTV) sector ( equivalent to
C- and CP-violation assuming CPT) has received considerably less
attention. Experimental efforts in the sector include measurement of
anomalous $\eta$-decay channels such as $\eta\rightarrow
2\pi^0\gamma$, $3\pi^0\gamma$, $3\gamma$ \cite{Beringer:1900zz} and
the $D$-coefficient in the $\beta$-decay of polarized neutrons
\cite{Mumm:2011nd} and $^{19}$Ne \cite{Hallin:1984mr}. These
processes are sensitive probes of \lq\lq new physics\rq\rq\, because
the Standard Model (SM) contributions are usually small
\cite{Dicus:1975cz,Herczeg:1997se}. There is a SM final-state
interaction that could mimic a non-zero $D$-coefficient in
$\beta$-decay at order $10^{-5}$ for neutron \cite{Callan:1967zz}
and $10^{-4}$ for $^{19}$Ne \cite{Hallin:1984mr} but the application
of heavy baryon effective field theory allows a precise computation
of this contribution (up to 1\% accuracy in the case of neutron
\cite{Ando:2009jk}).
Theoretically, the effect of PCTV physics due to beyond Standard
Model (BSM) interactions can be studied in a model-independent way
using effective field theory (EFT). In this approach, one has
integrated out the BSM heavy degrees of freedom (DOF). In this
context, it was observed in Ref. \cite{Conti:1992xn} that any EDM
limits imply severe bounds on PCTV observables since a PCTV
interaction in the presence of P-violating SM radiative corrections
will induce an EDM. While special exceptions to this argument may
occur \cite{RamseyMusolf:1999nk,Kurylov:2000ub}, the question
remains as to the prospective impact of, and motivation for,
improved probes of flavor-conserving PCTV observables. Recently, the
authors of Ref. \cite{Ng:2011ui} addressed this question in the EFT
context, studying the contribution of the \lq\lq left-right four
fermion\rq\rq\, (LR4F) operator to the $D$-coefficient of the
neutron $\beta$-decay (defined below). They find that the neutron
EDM sets an indirect bound on the $D$-coefficient that is three
orders of magnitude more stringent than its direct experimental
bound.
In this paper, we observe that there exists a set of dimension-six
four-fermion operators involving right-handed neutrinos that (a)
contribute to the $D$-coefficient and (b) are relatively
unconstrained by EDM limits. Because the SM charge changing weak
interaction involves purely left-handed leptons, this contribution
to neutron decay does not interfere linearly with the SM
contribution, resulting in a quadratic, rather than linear,
dependence on the operator Wilson coefficients. Nonetheless, present
limits on $D$ probe the TeV mass scale. We also show that while a
subset of these operators generate hadronic EDMs, their effects are
suppressed by loop factors as well as $\Lambda_{\chi}/v$ where
$\Lambda_{\chi}\sim 1$ GeV is the chiral symmetry breaking scale and
$v=246$ GeV is the Higgs vacuum expectation value (VEV). The
resulting neutron EDM sensitivity to $\Lambda$ is also at the TeV
scale and does not preclude a non-zero result in a next generation
$D$-coefficient probe.
Interestingly, indirect constraints from T-conserving observables
may be more severe. These observables include Large Hadron Collider
(LHC) results for the process $pp\to e + X + MET$ (missing
transverse energy) and neutrino mass. The latter constraints also
rely on naturalness considerations, a somewhat subjective criteria.
The former imply that an order of magnitude improvement in
$D$-coefficient sensitivity would be required in order to discover
evidence for PCTV right-handed neutrino interactions.
Our analysis leading to these conclusions is organized as follows.
In Sec. \ref{sec:vR} we introduce the relevant set of dimension-6
operators and discuss the experimental $D$-coefficient constraint on
their Wilson coefficients. We then compare this constraint to those
implied by LHC data, hadronic EDMs as well as neutrino mass and
naturalness considerations. For comparison, we perform in Sec.
\ref{sec:novR} a similar analysis of other dimension-6 operators
that do not involve right-handed neutrinos. We show that any attempt
to evade current EDM constraints and yet keep the size of the
$D$-coefficient experimentally accessible would involve fine tuning
at the $10^{-11}$ level. We conclude in Sec. \ref{sec:conclusion}.
\section{\label{sec:vR}Dimension-six operators with right-handed neutrinos}
The PCTV observable of interest in $\beta$-decay involves a triple
correlation of the spin of the decaying particle and the momenta of
the outgoing leptons that enters the differential $\beta$-decay
rate. In what follows, we focus on neutron, for which the
experimental bound on the PCTV triple correlation is the most
stringent. However, the discussion below can be easily generalized
to other cases. The differential decay rate for a polarized neutron
is given by:
\begin{eqnarray}
\nonumber
\frac{d\Gamma}{dE_ed\Omega_ed\Omega_\nu}=\frac{G_F^2V_\mathrm{ud}^2}{(2\pi)^5}(g_V^2+3g_A^2)|\vec{p}_e|E_eE_\nu^2\\
\times
\left[1+a\frac{\vec{p}_e\cdot\vec{p}_\nu}{E_eE_\nu}+\hat{s}\cdot(A\frac{\vec{p}_e}{E_e}+B\frac{\vec{p}_\nu}{E_\nu}+D\frac{\vec{p}_e\times\vec{p}_\nu}{E_eE_\nu})\right]
\end{eqnarray}
where $\hat{s}$ is the unit polarization vector of the neutron;
$\vec{p}_e$ and $\vec{p}_\nu$ are the electron and anti-neutrino
momenta, respectively, with corresponding energies $E_{e(\nu)}$;
$G_F$ is the Fermi constant; and $V_\mathrm{ud}$ is the first
generation element of the Cabibbo-Kobayashi-Maskawa (CKM) matrix.
The most stringent experimental limit on the $D$-coefficient is
given by $D=(-0.96\pm1.89\pm1.01)\times10^{-4}$ \cite{Mumm:2011nd}
which translates into an upper bound of $|D|<4\times 10^{-4}$ at
90\% CL \cite{Vos:2015eba}.
Theoretically, a non-vanishing contribution can be generated by the
interference of amplitudes involving a small set of dimension $d=6$
effective operators. Considering only first generation SM fermions
and requiring SM gauge invariance, one finds a limited set of such
$d=6$ TV operators (see Ref. \cite{Grzadkowski:2010es} for a
complete list of gauge-invariant $d=6$ operators involving SM
fields). As we discuss in Section \ref{sec:dipole}, EDM constraints
imply severe bounds on the contribution of these operators to $D$.
Extending the set of fields to include right-handed (RH) neutrinos,
one finds an additional set of four-fermion operators that
contribute to $D$ at tree-level and that are relatively immune to
EDM constraints \cite{Cirigliano:2012ab}:
\begin{eqnarray}\label{eq:Oi}
\hat{O}_1&=&\frac{c_1(\mu)}{\Lambda^2}\bar{L}^i\nu_R\bar{u}_RQ^i+h.c.\nonumber\\
\hat{O}_2&=&\frac{c_2(\mu)}{\Lambda^2}\varepsilon^{ij}\bar{L}^i\nu_R\bar{Q}^jd_R+h.c.\nonumber\\
\hat{O}_3&=&\frac{c_3(\mu)}{\Lambda^2}\varepsilon^{ij}\bar{L}^i\sigma^{\mu\nu}\nu_R\bar{Q}^j\sigma_{\mu\nu}d_R+h.c.
\end{eqnarray}
where $\Lambda$ is the BSM mass scale and $\mu$ is the
renormalization scale. These operators are analogous to the
semi-leptonic four-fermion operators of type $(\bar{L}R)(\bar{L}R)$
and $(\bar{L}R)(\bar{R}L)$ in Ref. \cite{Grzadkowski:2010es}. Also
notice that the Wilson coefficients $c_1-c_3$ are functions of the
renormalization scale $\mu$, which as to be taken as the hadronic
scale when we discuss the bounds of the Wilson coefficients from
low-energy experiments.
It is straightforward to compute the contributions of
$\hat{O}_{1-3}$ to the $D$-coefficient. The dominant affect is
quadratic in the $c_i/\Lambda^2$, as the linear interference term is
suppressed by the neutrino mass.
Following Ref.~\cite{Jackson:1957zz}, we obtain, to leading
non-trivial order in $\{c_i\}$,
\begin{equation}
\label{eq:Dcoeff}
D=-\frac{g_Sg_T}{\Lambda^4}\frac{1}{G_F^2V_{ud}^2(g_V^2+3g_A^2)}\mathrm{Im}[(c_1-c_2)c_3^*]\Big\vert_{\mu=\mu_h}
\end{equation}
where $g_S$ and $g_T$ are the nucleon scalar and tensor charges,
respectively, and $\mu_h\approx1$ GeV is the hadronic scale.
Even though one pays a price in BSM sensitivity owing to a quadratic
rather than linear dependence on the $c_i/\Lambda^2$, the gain
achieved by avoiding EDM constraints is considerable (see Section
\ref{sec:dipole}). Taking the updated lattice calculation of
$g_S=0.97(12)(6)$ and $g_T=0.987(51)(20)$
\cite{Bhattacharya:2016zcn}, we obtain:
\begin{equation}\label{eq:bound}
|\frac{\mathrm{Im}[(c_1-c_2)c_3^*]}{\Lambda^4}|\Big\vert_{\mu=\mu_h}<3\times10^{-1}\
\mathrm{TeV}^{-4}
\end{equation}
If we take $\{c_i\}\sim c$ without distinguishing the real and
imaginary part, then this inequality implies that existing
$D$-coefficient studies probe BSM T-violating interactions with RH
neutrinos with a sensitivity of $(v/\Lambda)^2c\sim3\times 10^{-2}$
at $\mu=\mu_h$. One could estimate the sensitivity to $\Lambda$ by
assuming that $c_i\sim1$ at $\mu\approx\Lambda$. QCD running in
$\overline{\mathrm{MS}}$ scheme gives
$c_{1,2}(\Lambda)\approx0.56c_{1,2}(\mu_h)$ and $c_3(\Lambda)\approx1.2c_3(\mu_h)$ for $\Lambda>m_W$ where $m_W$ is the mass of the W-boson (see,
e.g. Ref. \cite{Broadhurst:1994se}). Then, the current bound implies $\Lambda\buildrel > \over {_\sim} 1$
TeV.
The operators ${\hat O}_{1-3}$ can induce hadronic EDMs at one-loop order, but their contributions also scale
quadratically with the $c_i/\Lambda^2$.
In particular, the combination of $\hat{O}_1$ and $\hat{O}_2$ may
induce the CP-odd four-quark operator \cite{Engel:2013lsa}
\begin{equation}
\frac{C_{quqd}^{(1)}(\mu)}{\Lambda^2}
\varepsilon^{ij}\bar{Q}^iu_R\bar{Q}^jd_R+h.c.
\end{equation}
via the one-loop graph of Fig. \ref{fig:4q}. Contributions from loop
momenta $k < \Lambda$ vanish, as seen explicitly in dimensional
regularization (DR), because the amplitude involves a
quadratically-divergent integral with massless propagators and
because it is infrared finite. Non-vanishing contributions result
from $k\buildrel > \over {_\sim} \Lambda$ that are associated with matching onto the
{\em a priori} unknown ultraviolet complete theory that generates
the non-vanishing $c_i$. Estimating these matching contributions
using a cut-off regulator \cite{Erwin:2006uc} yields\footnote{It is
possible that in the full theory, a symmetry implies vanishing
matching contributions, but we will be more general here.}
\begin{equation}\label{eq:4fto4q}
\frac{C_{quqd}^{(1)}(\Lambda)}{\Lambda^2}\sim\frac{\Lambda^2}{16\pi^2}\frac{c_1^*c_2}{\Lambda^4}\Big\vert_{\mu=\Lambda}=\frac{c_1^*c_2}{16\pi^2\Lambda^2}\Big\vert_{\mu=\Lambda}
\end{equation}
This four-quark operator will in turn induce a neutron EDM $d_n$. To
evaluate this contribution, one must first evolve $C_{quqd}^{(1)}$
from $\mu=\Lambda$ down to $\mu=\mu_h$. In principle, this can only
be done if one knows the exact value of $\Lambda$. However, since
the evolution depends only logarithmically on $\Lambda$, it is
reasonable to take $\Lambda\sim1$ TeV as an illustration, giving
$C_{quqd}^{(1)}(\mu_h)=7.2C_{quqd}^{(1)}(\Lambda)$
\cite{Dekens:2013zca}.
Next, in order to find the relation between $C_{quqd}^{(1)}(\mu_h)$
and the induced hadronic EDMs one needs to compute corresponding
hadronic matrix elements. First-principle calculations of such
matrix elements are challenging, and presently only exist for simple
systems such as $\rho$-meson (see, {\em e.g.} Ref.
\cite{Pitschmann:2012by} and references therein). The results of
such calculations are generally consistent with the
order-of-magnitude estimation based on na{\"i}ve dimensional
analysis (NDA) \cite{Manohar:1983md,Weinberg:1989dx,deVries:2012ab},
so here we shall also provide an NDA estimation of $d_n$:
\begin{eqnarray}\label{eq:NDA} d_n&\sim&
e\frac{\Lambda_\chi}{16\pi^2}\frac{\mathrm{Im}C_{quqd}^{(1)}(\mu_h)}{\Lambda^2}\nonumber\\
&\approx&
e\frac{\Lambda_\chi}{16\pi^2}\times7.2\frac{\mathrm{Im}C_{quqd}^{(1)}(\Lambda)}{\Lambda^2}\nonumber\\
&\approx& 9.4\times
10^{-23}(\frac{v}{\Lambda})^2\mathrm{Im}\{c_1^*c_2\}|_{\mu=\Lambda}e\:\mathrm{cm}.
\end{eqnarray}
This EDM is suppressed by
$1/(16\pi^2)^2$ as well as $\Lambda_\chi/v$. Given the current upper
bound $d_n<3.0\times10^{-26}e\:\mathrm{cm}$ at 90\% CL
\cite{Afach:2015sja} we see that the existing neutron EDM limits are
probing $(v/\Lambda)^2c^2\sim3\times10^{-4}$ at $\mu=\Lambda$ which
implies $\Lambda\buildrel > \over {_\sim} 10$TeV if $c(\Lambda)\sim1$.
At first glance, the neutron EDM sensitivity to $\Lambda$ is slightly tighter than that of the
$D$-coefficient. However, since both estimations made in Eq.
\eqref{eq:4fto4q} and \eqref{eq:NDA} allow an error within an order
of magnitude, one may reasonably conclude that the sensitivities of
$d_n$ and the $D$-coefficient are
comparable. Furthermore, hadronic and atomic EDMs depend only on
$c_1^*c_2$ and provide no direct constraint on the contribution from
$c_jc_3^\ast$ ($j=1,2$) in Eq.~(\ref{eq:Dcoeff}).
We now consider constraints from T conserving observables. First, we
note that LHC studies of the process $pp\rightarrow e+X+MET$ place
stringent bounds on the operators in \eqref{eq:Oi}\footnote{We thank
M. Gonzales-Alonso for pointing out these constraints.}.
Following Ref. \cite{Cirigliano:2012ab}, one may define two
dimensionless quantities:
\begin{eqnarray}
\tilde{\epsilon}_S&=&-\frac{c_1-c_2}{2\sqrt{2}G_F
V_{ud}\Lambda^2}\nonumber\\
\tilde{\epsilon}_T&=&\frac{c_3}{2\sqrt{2}G_F V_{ud}\Lambda^2}.
\end{eqnarray} The contribution from $\hat{O}_{1,2,3}$ to the total cross-section $\sigma_{tot}$ of the $pp\rightarrow e+X+MET$ process measured by LHC can be written as $\sigma_{tot}=\sigma_S|\tilde{\epsilon}_S|^2+\sigma_T|\tilde{\epsilon}_T|^2$. Therefore LHC is sensitive to $|\tilde{\epsilon}_S|$ and $|\tilde{\epsilon}_T|$
while the
$D$-coefficient probes the combination of products
$\mathrm{Re}\tilde{\epsilon}_T\mathrm{Im}\tilde{\epsilon}_S-\mathrm{Re}\tilde{\epsilon}_S\mathrm{Im}\tilde{\epsilon}_T$.
The bounds on the ${\tilde\epsilon}$ parameters obtained in Ref.
\cite{Cirigliano:2012ab} assume contributions from one operator at a
time. However, when comparing with the $D$-coefficient sensitivity,
one must take both $\tilde{\epsilon}_S$ and $\tilde{\epsilon}_T$,
since the $D$-coefficient probes products of the two. Recasting the
analysis of Ref.~\cite{Cirigliano:2012ab} is nevertheless
straightforward because $\sigma_S$ and $\sigma_T$ are known. The
constraint equation in Ref.~\cite{Cirigliano:2012ab} then implies an
elliptical bound in the $|\tilde{\epsilon}_S|-|\tilde{\epsilon}_T|$
plane. One should also remember that the LHC constraints should be
run down to $\mu=\mu_h$ for a fair comparison with the
$D$-coefficient.
Since there are four real parameters in the problem (the Re and Im
parts of $\tilde\epsilon_{S,T}$) , it us useful to make simplifying
assumptions in order to compare the LHC and $D$-coefficient
sensitivities. To that end, we will
assume for the moment that
$\mathrm{Re}\tilde{\epsilon}_S=\mathrm{Im}\tilde{\epsilon}_T=0$ so
both the LHC and the neutron $D$-coefficient results set
constraints on $\mathrm{Re}\tilde{\epsilon}_T$ and
$\mathrm{Im}\tilde{\epsilon}_S$ (see Fig. \ref{fig:LHCvsD}). In this
case, one sees that the sensitivity of neutron decay to the
$D$-coefficient has to be improved by roughly a factor of 15 in
order to match the sensitivity of the 7-TeV LHC results. Results
at $\sqrt{s}=8$ TeV for the same channel at are also available. As
there is no significant deviation from SM prediction
\cite{ATLAS:2014pna,Khachatryan:2014tva}, the LHC bound on
$|\tilde{\epsilon}_S|$ and $|\tilde{\epsilon}_T|$ will be even more
stringent than quoted above, although a detailed analysis has yet
to be performed\footnote{The LHC sensitivity will, of course,
improve further with the data obtained from Run II .}.
\begin{figure}
\centering
\includegraphics[scale=0.30]{LHCvsD2}
\caption{\raggedright\label{fig:LHCvsD}(Color online)
Exclusion plot for $\mathrm{Re}\tilde{\epsilon}_{T}$ and
$\mathrm{Im}\tilde{\epsilon}_{S}$ at $\mu=\mu_h$ from the 7-TeV LHC
data (red solid line) as well as the bound from the neutron
$D$-coefficient with the current precision level (blue dotted line) and 15 times of the current precision level (green dashed line) respectively assuming
$\mathrm{Re}\tilde{\epsilon}_{S}=\mathrm{\mathrm{Im}\tilde{\epsilon}_{T}=0}.$}
\end{figure}
One may also derive interesting but less direct constraints on $\hat{O}_1$ and $\hat{O}_2$ from the scale of neutrino mass and naturalness considerations.
Above the electroweak scale, the leading contribution to $m_\nu$
comes from a one-loop diagram with a quark Yukawa insertion,
inducing the Yukawa interaction term $\bar{L}\tilde{H}v_R$, as shown
in Fig. \ref{fig:numass}. Again this contribution vanishes in DR so
we estimate it using simple dimensional analysis. After electroweak
symmetry breaking, one obtains
\begin{equation}
m_\nu\sim\frac{c_i}{\Lambda^2}\frac{\Lambda^2}{16\pi^2}m_q=\frac{c_i
m_q}{16\pi^2}
\end{equation}
where $m_q$ is the light quark mass and $i=1,2$. Taking $m_\nu<1$ eV
and $m_q\approx 5$ MeV we obtain $c_i<3\times10^{-5}$. We
stress that this bound is not airtight, as the result may
vary considerably, depending on the specific symmetry of the underlying BSM
scenario. Neutrino mass naturalness bounds also do not constrain the tensor
interaction strength $c_3$. Should a next generate $D$-coefficient measurement yield a non-vanishing result, the comparison with
neutrino mass naturalness considerations would provide interesting input for model-building.
\begin{figure}
\centering
\begin{subfigure}[b]{0.25\textwidth}
\includegraphics[width=\textwidth]{1}
\caption{\label{fig:4q}(a)}
\end{subfigure}
\begin{subfigure}[b]{0.25\textwidth}
\includegraphics[width=\textwidth]{2}
\caption{\label{fig:numass}(b)}
\end{subfigure}
\caption{\raggedright Leading loop contributions that provide
indirect bounds on $c_1$ and $c_2$. Figure (a) induces a four-quark
operator that generates hadronic EDMs. Figure (b) generates a
neutrino mass after electroweak symmetry breaking.}
\end{figure}
\section{\label{sec:novR}Operators without right-handed neutrinos}
\label{sec:dipole}
In contrast to the discussion of Section \ref{sec:vR}, we consider here $d=6$ operators that contain only
left-handed (LH) neutrino fields and show that any contributions to the $D$-coefficient are severely constrained by present EDM limits.
In the four-fermion sector, the only operator that gives a
tree-level $D$-coefficient scaling linearly with the BSM coupling
strength has the form of $\bar{u}_R\gamma^\mu d_R\bar{e}_L\gamma_\mu
\nu_L$ as discussed in Ref. \cite{Ng:2011ui}. It is actually derived
from a gauge-invariant dim-6 operator:
\begin{equation}
\label{eq:hud}
\hat{O}_{Hud}=i\frac{C_{Hud}}{\Lambda^2}(\tilde{H}^\dagger
D_\mu H)(\bar{u}_R\gamma^\mu d_R)+h.c..
\end{equation}
Below the electroweak scale, exchange of the $W$-boson contained in
the covariant derivative with the left-handed charged weak current
leads to both the semi-leptonic four-fermion operator listed above
as well as
a four-quark operator of the form
$\bar{u}_R\gamma^\mu d_R\bar{d}_L\gamma_\mu u_L$. Both operators share the
same Wilson coefficient (up to $V_{ud}$), which
is tightly constrained by the the four-quark contribution to the neutron EDM.
The three remaining semi-leptonic four-fermion operators that
contain T-odd components are the scalar and tensor operators of the
type $(\bar{L}R)(\bar{L}R)$ and $(\bar{L}R)(\bar{R}L)$
\cite{Grzadkowski:2010es,Engel:2013lsa}:
\begin{eqnarray}\label{eq:4fermion}
\hat{O}_{ledq}&=&i\frac{\mathrm{Im}C_{ledq}}{\Lambda^2}\bar{L}^ie_R\bar{d}_RQ^i+h.c.\nonumber\\
\hat{O}_{lequ}^{(1)}&=&i\frac{\mathrm{Im}C_{lequ}^{(1)}}{\Lambda^2}\varepsilon^{ij}\bar{L}^ie_R\bar{Q}^ju_R+h.c.\nonumber\\
\hat{O}_{lequ}^{(3)}&=&i\frac{\mathrm{Im}C_{lequ}^{(3)}}{\Lambda^2}\varepsilon^{ij}\bar{L}^i\sigma^{\mu\nu}e_R\bar{Q}^j\sigma_{\mu\nu}u_R+h.c.
\end{eqnarray}
Similar to the operators in Eq. \eqref{eq:Oi}, they induce a
$D$-coefficient that scales quadratically with $c_i/\Lambda^2$.
However, $\hat{O}_{ledq}$, $\hat{O}_{lequ}^{(1)}$,
$\hat{O}_{lequ}^{(3)}$ contribute linearly to EDMs of
paramagnetic atom and molecules and diamagnetic atoms at tree level
as well as hadronic and electron EDMs at the one-loop level.
In particular, the ACME limit on the EDM of ThO molecule
implies a strong constraint on
$\mathrm{Im}(C_{ledq}-C_{ledq}^{(1)})$ (see, e.g. Ref.
\cite{Chupp:2014gka}). The resulting indirect constraints on the
associated $D$-coefficient contributions are severe.
The remaining class of operators that give rise to the
$D$-coefficient at tree-level are dipole-like operators. One may
wonder whether EDM constraints to such operators may be avoided with
an appropriate choice of Wilson coefficients at low energy. We will
show, however, that this is not possible without fine-tuning at the
level of many orders of magnitude.
To simplify our discussion, let us concentrate on the dipole-like
operators in the purely leptonic sector:
\begin{eqnarray}\label{eq:op}
\hat{O}_{eB}&=&i\frac{g'\mathrm{Im}C_{eB}}{\Lambda^2}\bar{L}\sigma^{\mu\nu}H
e_RB_{\mu\nu}+h.c.\nonumber\\
\hat{O}_{eW}&=&i\frac{g\mathrm{Im}C_{eW}}{\Lambda^2}\bar{L}\sigma^{\mu\nu}\frac{\tau^i}{2}H
e_RW^i_{\mu\nu}+h.c.\nonumber\\
\hat{O}_{eH^3}&=&i\frac{\mathrm{Im}C_{eH^3}}{\Lambda^2}\bar{L}He_RH^\dagger
H+h.c.
\end{eqnarray}
The first-two operators are dipole-like while the third operator is
included as well because it mixes with the first two via electroweak
renormalization. Only $\hat{O}_{eW}$ contributes to $D$, as it is the only one containing a $W$ field. After electroweak symmetry-breaking, one finds
\begin{equation}
D=-\frac{4\sqrt{2}g_A^2}{g_V^2+3g_A^2}(\frac{m_e}{v})(\frac{v}{\Lambda})^2\mathrm{Im}C_{eW}\ \ \ .
\end{equation}
Note the presence of the $m_e/v$ suppression due to the existence of a
derivative in the operator $\hat{O}_{eW}$. The current upper bound on
the neutron $D$-coefficient implies
$(v/\Lambda)^2|\mathrm{Im}C_{eW}|<1\times 10^2$.
The same set of operators also induces an electron EDM, given by
\begin{equation}
d_e=-\frac{\sqrt{2}e}{v}(\frac{v}{\Lambda})^2(\mathrm{Im}C_{eB}-\mathrm{Im}C_{eW})
\end{equation}
The current upper bound on $d_e$ \cite{Baron:2013eja} implies
$(v/\Lambda)^2|\mathrm{Im}C_{eB}-\mathrm{Im}C_{eW}|<7.7\times10^{-13}$.
At first glance, it seems that one could simply choose
$\mathrm{Im}C_{eB}=\mathrm{Im}C_{eW}$ at low energy to avoid the EDM
constraint. We want to argue that, however, this choice is highly
unnatural because the operators in Eq. \eqref{eq:op} mix under
electroweak renormalization as
\begin{equation}
\frac{d\Theta}{d\ln\mu}=\left(\begin{array}{ccc}
\frac{151g'^{2}-27g^{2}}{192\pi^{2}} & -\frac{3gg'}{64\pi^{2}} & 0\\
-\frac{gg'}{16\pi^{2}} & \frac{-11g^{2}+3g'^{2}}{192\pi^{2}} & 0\\
-\frac{3g'(g^{2}-3g'^{2})}{16\pi^{2}} &
-\frac{9g(g^{2}-g'^{2})}{32\pi^{2}} &
-\frac{3(9g^{2}+7g'^{2})}{64\pi^{2}}\end{array}\right)\Theta
\end{equation}
where $\Theta=\left(\begin{array}{ccc} g'\mathrm{Im}C_{eB} &
g\mathrm{Im}C_{eW} & \mathrm{Im}C_{eH^3}\end{array}\right)^{T}$.
Numerically, if we assume that the bounds on the $D$-coefficient is
marginally satisfied at $\mu=m_W$ (i.e.
$(v/\Lambda)^2|\mathrm{Im}C_{eW}|=1\times 10^{2}$), then after the
electroweak renormalization we find that
$(v/\Lambda)^2|\mathrm{Im}C_{eB}-\mathrm{Im}C_{eW}|\approx 4.0$ at
$\mu=10$ TeV. However, this number has to be fine-tuned to a
precision level of $2\times10^{-11}\%$ in order to satisfy the EDM
bound at low energy, and therefore it is obviously not natural. The
dipole-like operators in the quark sector suffer from the same
problem. We conclude that, in the absence of RH neutrinos, EDM
constraints imply that the existence of an observable
$D$-coefficient is highly unlikely.
\section{\label{sec:conclusion}Conclusion}
If neutrinos are Dirac particles, implying the existence of light
$\nu_R$ in nature, then present limits on the $D$-coefficient
indicate that the mass scale of any associated PCTV interactions may
be quite significant: $\Lambda/c \buildrel > \over {_\sim} 1$ TeV, where $c$ denotes a
$d=6$ operator Wilson coefficient. The corresponding reach
associated with limits on the PCTV triple correlation in polarized
$^{19}$Ne decay are somewhat weaker, but nevertheless quite
interesting. The observation of a non-zero effect in a next
generation experiment with either the neutron or nuclei is not
precluded by constraints from EDM search null results. On the other
hand, the LHC results for $pp\to e+X+ MET$ present a greater
challenge, implying at least an order of magnitude improvement in
neutron decay PCTV correlation sensitivity would be needed for
discovery of a non-zero $D$-coefficient. Should such an observation
occur, resolving the tension between a non-zero PCTV correlation
measurement and neutrino mass naturalness considerations would
provide an interesting challenge for model building.
\section*{Acknowledgements}
The authors wish to thank Jordy de Vries and Barry Holstein for many useful
discussions. This work was supported in part under U.S. Department of Energy contract
number DE-SC0011095 (MJRM and CYS) and NSF grant number is PHY-1205896 (BEM).
\bibliographystyle{prsty}
|
\section{Introduction}
The \emph{boosting} approach to learning binary classifiers is to construct a
weighted-majority ensemble of them by incrementally adding base classifiers(\cite{SF12}).
This process is guided by a potential function that is an upper bound on the training error.
The weights assigned to the training examples communicate the gradient of the potential function to the base learner.
Like other supervised learning algorithms, boosting algorithms require a sufficiently large labeled training set in order to produce an accurate classifier.
Such algorithms do not make use of \emph{unlabeled} data, which is typically much more abundant.
On the other hand, semi-supervised learning approaches~(\cite{CSZ06})
attempt to use both labeled and unlabeled training examples. The basic
idea in many approaches is to augment the labeled set by inferring the
label of unlabeled examples from their labeled neighbors in some way.
Such inference uses the "hallucinatory labels" on the unlabeled data that tend to agree with the labeled ones (e.g. \cite{BNS06}).
In this paper, we present a semi-supervised learning approach that uses the opposite strategy.
Instead of using the unlabeled examples to \emph{enhance} the labeled examples, we use the unlabeled examples to \emph{muffle}
the effect of the labeled examples, and hallucinate labels which tend to oppose the labeled ones.
This strategy arises from a transductive inference approach,
assuming that the labeled and unlabeled examples are drawn from the same distribution.
To create labeled and unlabeled training sets,
the label of each example is either exposed (labeled example) or left hidden (unlabeled example) independently at random.
The task of learning in this scenario is to accurately predict the label of the unlabeled examples.
This task is significantly easier (\cite{V82}) than the task of standard (inductive) learning, which is to generate a rule
that will accurately predict on {\em any} as-yet-unseen examples drawn from the same distribution as the training set.
\iffalse
We focus on this issue in this paper,
applying the idea of incremental classifier aggregation to a general semi-supervised setting.
Here we are not only given a labeled set $L$,
but also an unlabeled set $U$, so that we know the ensemble classifiers' predictions on the unlabeled data.
In principle, the extra unlabeled information should be useful in
improving performance over the labeled data alone, and lead to better generalization.
\fi
In this paper, we devise algorithms which build empirically on recent work of
\cite{BF15}. That paper directly considers the \emph{test} error
based on the error rates of the ensemble classifiers (described in
Section \ref{sec:boundtesterr}) and unlabeled data, and outlines a prediction
algorithm that achieves this bound. The intuition behind the
algorithm is that the aggregated prediction on unlabeled examples
should be between $-1$ and $+1$ (see Fig. \ref{fig:predscorefunc}). The "muffling"
behavior occurs when the aggregate prediction on an unlabeled
examples is outside this range. One can represent muffling by
assigning to the unlabeled example a hallucinatory label that is the
opposite of the predicted label.
\iffalse
It shows that robust aggregated
predictors have \emph{low margin} on the given unlabeled data
generally -- that is, on most of the unlabeled examples, the
ensemble's predictions are split close to evenly between the two
labels. The idea behind this is that the predictor avoids committing
too aggressively to any of its predictions on $U$, which can lead to
very poor generalization if $U$'s actual labels are not as predicted.
\fi
We use this muffling principle to devise several simple algorithms,
notably a sequential scheme ($\algname$) that incorporates ensemble classifiers one at a
time into an aggregated classifier. At every step, it chooses a new
classifier that tends to \emph{disagree} with the current majority
opinion on examples in the unlabeled set.
The rest of the paper is organized as follows. In
Section~\ref{sec:boundtesterr}, we briefly review the theory presented
in \cite{BF15}. In Section~\ref{sec:marvin} we describe the algorithm
$\algname$. In Section~\ref{sec:mower} we describe the algorithm $\textsc{HedgeMower}$,
a way of exploiting the partitionings used by ensemble classifiers like decision trees.
Section \ref{sec:experiments} contains a comparative experimental evaluation of our
algorithm on a number of datasets.
In Section \ref{sec:disc} we draw some conclusions from the experiments,
and in Section~\ref{sec:relwork} we make connections to past and
future work.
\section{Setup: Minimizing Worst-Case Test Error}
\label{sec:boundtesterr}
In this paper, we study a learning scenario where we have two types of data drawn i.i.d. from the same distribution:
a labeled set $ L = \{ (x^L_1, y_1^L) , \dots, (x^L_m, y_m^L) \}$ and an unlabeled set $U = \{x^U_1, \dots, x^U_n \} $.
We have at our disposal an ensemble $\mathcal{H}$ of predictors,
and are tasked to classify the unlabeled data $U$ as accurately as possible.
\footnote{We will see that in our case this transductive setting, measuring performance over known $U$, is essentially equivalent to the statistical learning setting,
where $L,U$ are i.i.d. and the test data are another i.i.d. sample.}
Write $\text{clip} (x) = \min(1, \max( -1, x))$ and $[n] = \{ 1,2,\dots,n \}$, as well as $a_{n_1: n_2}$ to denote the set $\{a_{n_1}, a_{n_1 + 1}, \dots, a_{n_2} \}$.
All vector inequalities, as well as functions like $\sgn(\mathbf{v})$, are component wise.
Our setting slightly modifies that of \cite{BF15},
considering an ensemble of $p$ classifiers.
Its predictions on the unlabeled data are denoted by $\vF$:
\begin{equation}
\label{eq:defoff}
\vF =
\begin{pmatrix}
h_1(x_1^U) & \cdots & h_1 (x_n^U) \\
\vdots & \ddots & \vdots \\
h_p(x_1^U) & \cdots & h_p (x_n^U)
\end{pmatrix}
\in [-1, 1]^{p \times n}
\end{equation}
We denote the columns of $\vF$ as $\mathbf{x}_j^U =
(h_1(x_j^U), \cdots, h_p (x_j^U))^\top$, and the rows as $\mathbf{h}_i = (h_i(x_1^U), \cdots, h_i (x_n^U))^\top$, omitting the superscript $U$.
The test set has some binary labels $(y_1; \dots; y_n) \in \{-1,1\}^n$,
which are unknown to the predictor.
However, the test labels are allowed to be randomized,
represented by values in $[-1,1]$ instead of just the two values $\{ -1, 1\}$.
So it is convenient to write the labels on $U$ as $\mathbf{z} = (z_1; \dots; z_n) \in [-1,1]^n$.
The idea of \cite{BF15} is to formulate the ensemble aggregation problem as a
two-player zero-sum game between a predictor and an adversary.
In this game, the predictor is the first player,
who plays $\vg = (g_1; g_2; \dots; g_n)$,
a randomized label $g_j \in [-1,1]$ for each example $\{x_j\}_{j=1}^{n}$.
The adversary is then allowed to set the labels $\mathbf{z} \in [-1,1]^n$.
A successful predictor player corresponds to a robust learning algorithm,
able to generalize as well as possible given its lack of full label information.
The key point is that when any classifier $i$ is known to perform well to a certain degree on the test data,
its predictions $\mathbf{h}_i$ on the test data are a reasonable guide to $\mathbf{z}$,
and correspondingly give us information by constraining $\mathbf{z}$ to be "near" them.
Each classifier in the ensemble thus contributes to an intersecting set of constraints,
which interact in ways that depend on the ensemble's test predictions.
To discuss these ideas, suppose the predictor has knowledge of a \emph{correlation vector}
$\vb \in (0, 1]^p$ such that
$
\forall i \in [p] , \; \frac{1}{n} \sum_{j=1}^n h_i (x_j) z_j \geq b_i
$,
i.e. $ \frac{1}{n} \vF \mathbf{z} \geq \vb$.
These $p$ inequalities represent upper bounds on individual classifier error rates,
which can be estimated from the training set w.h.p. when the training and test data are i.i.d.,
in a standard way also used by ERM \cite{BF15}.
So in our game-theoretic formulation,
the adversary plays under ensemble error constraints defined by $\vb$.
The predictor attempts to
\emph{minimize the worst-case expected loss on the test data}
(w.r.t. the randomized labeling $\mathbf{z}$), which we write
$\ell (\mathbf{z}, \vg) := \frac{1}{n} \sum_{j=1}^{n} \frac{1}{2} (1 - z_j g_j) $.
The goal is to drive this loss down to $\approx V$, the best upper bound on error that any predictor can guarantee, given the information in $\vF$ and $\vb$:
\begin{align}
\label{eq:game1eq}
V &:= \min_{\vg \in [-1,1]^n} \; \max_{\substack{ \mathbf{z} \in [-1,1]^n , \\ \frac{1}{n} \vF \mathbf{z} \geq \vb }} \; \ell (\mathbf{z}, \vg)
\end{align}
The main result of the prior work \cite{BF15} expresses $V$ and the optimal predictor strategy
$\displaystyle \vg^* $,
which achieves the optimum of \eqref{eq:game1eq}.
To state it, define the convex \textbf{potential well} as $\Psi (x) = \max(1, \abs{x})$;
the \textbf{slack function} as
$\gamma (\sigma) := \gamma_{\vb} (\sigma) := - \ip{\vb} {\sigma} + \frac{1}{n} \sum_{j=1}^n \Psi ( \ip{\mathbf{x}^U_{j}} {\sigma} )$;
and the \textbf{optimal weights} $ \sigma^* := \argmin_{\sigma \geq 0^p} \gamma (\sigma)$.
\footnote{Here $\ip{\cdot}{\cdot}$ denotes the usual inner product in $\RR^p$ --
as we move to a setting where the ensemble grows with time $t$, it will denote an inner product over the ensemble classifiers learned so far,
and will be clear from context.}
Then the semi-supervised aggregation game of \eqref{eq:game1eq} has a conveniently expressible solution.
\begin{theorem}[\cite{BF15}]
\label{thm:gamesolngen}
The minimax value of the game \eqref{eq:game1eq} is
\begin{align*}
V = \frac{1}{2} \min_{\sigma \geq 0^p} \lrb{ - \ip{\vb}{\sigma} + \frac{1}{n} \sum_{j=1}^n \Psi ( \ip{\mathbf{x}_{j}^U}{\sigma} ) }
= \frac{1}{2} \min_{\sigma \geq 0^p} \gamma (\sigma)
:= \frac{1}{2} \gamma (\sigma^*)
\end{align*}
The minimax optimal predictions for all $j \in [n]$ are
$g_j^* := [\vg (\sigma^*)]_j := \mbox{clip} (\ip{\mathbf{x}_{j}^U}{\sigma^*})$.
\end{theorem}
This suggests that given any ensemble, we should try to play $\vg^*$ to perform well on $U$,
finding the $\sigma^*$ that minimizes the slack function $\gamma (\cdot)$ and then playing $\vg (\sigma^*)$.
We can approximately optimize to find $\sigma \approx \sigma^*$, in which case predicting with $\vg (\sigma)$ is near-optimal (\cite{BF15}).
This is a semi-supervised alternative to the common supervised learning principle of empirical risk minimization (\cite{V82}),
in which a bound on training error is minimized.
The minimax predictor takes an easily interpretable and convenient form.
For each $x_j^U \in U$, it only depends on the \textbf{score} $\ip{\mathbf{x}_{j}^U}{\sigma^*}$ of $x_j^U$
with respect to the weights $\sigma^*$.
The \textbf{margin} (defined as $\abs{\text{score}}$) can be interpreted as a notion of confidence, for which this paper provides empirical evidence.
\begin{wrapfigure}{r}{0.5\textwidth}
\begin{center}
\includegraphics[width=0.4\textwidth]{predscorefuncs.png}
\end{center}
\caption{Potential well and prediction on an unlabeled example, as a function of its score.}
\label{fig:predscorefunc}
\end{wrapfigure}
The average potential $\Psi (\cdot)$ of the unlabeled data regularizes the problem
by encouraging us to put weight on classifiers which disagree, so that the margin stays low.
Thm. \ref{thm:gamesolngen} provides a direct proof that this strategy generalizes well by directly addressing test error,
even though it contrasts starkly with \emph{max-margin} approaches known to generalize in fully supervised settings (\cite{SFBL98}).
We refer to this minimax framework (of \cite{BF15}) as \emph{muffled learning}
to emphasize its learning principle of actively distrusting overly confident predictions.
As a consequence of Thm. \ref{thm:gamesolngen},
it always performs at least as well as any single classifier when $\vb$ is estimated accurately -- in other words, unlabeled data do not hurt (\cite{BF15}).
We follow previous work (\cite{BF15, BF15b}) in emphasizing that while the transductive setting is convenient for a clean muffled formulation,
in this case it is not a restrictive assumption for high $n$, since the data are i.i.d. (see Appendix \ref{sec:minibatch} for details).
For the remainder of this paper, we investigate ways to minimize $\gamma (\cdot)$ generally (for any ensemble) and practically.
All of these are algorithms to find a weight vector $\sigma$ that leads to a good predictor, and to learn the associated ensemble.
So our algorithms always inherit the aforementioned prediction-based advantages of the muffled learning framework.
\section{An Algorithm for Incrementally Aggregating Classifiers}
\label{sec:marvin}
Directly minimizing the slack function with an ensemble generated a priori like a random forest can enjoy some practical success,
but it has been reported (\cite{BF15b}) be too conservative, because the bound $V$ on error is too loose.
However, adding more classifiers to any ensemble can only lower its $V$, because $\mathbf{z}$ is at least as constrained after the addition.
Therefore, a natural strategy to mitigate the bound's looseness is to call upon a larger ensemble.
So we elect to build our predictor incrementally, from classifiers in a possibly infinite ensemble $\mathcal{H}$.
Supervised boosting algorithms have long (\cite{F95}) done this efficiently by accessing a learning algorithm that the booster calls as needed,
to generate classifiers one at a time.
We also use such a learner as a subroutine -- it returns a classifier from $\mathcal{H}$ that approximately minimizes error on its input among $h \in \mathcal{H}$.
This is efficiently implemented for many hypothesis classes, like decision trees, linear classifiers, and other supervised learning approaches --
our method is capable of using any of these.
Our algorithm, $\algname$, repeatedly requests the classifier in $\mathcal{H}$ that minimizes error on inputs comprised of $m+n$ weighted examples:
the $m$ labeled examples in $L$, and the $n$ unlabeled examples in $U$ with purposefully \emph{hallucinated} labels that change every iteration.
\footnote{
The actual algorithm run is a minibatch version adapted slightly for the stochastic setting (Appendix \ref{sec:minibatch}).
}
\begin{algorithm}[tp]
\caption{$\algname$
\label{alg:realalg}
\begin{algorithmic}
\STATE
{\bfseries Input:} Size-$m$ labeled set $L$, size-$n$ unlabeled set $U$
\STATE
{\bfseries Initialize weights:} $\sigma_0 = \mathbf{0}$, so that $\ip{\mathbf{x}}{\sigma^{t-1}} = 0$ for all $x \in U$
\FOR{$t = 1$ {\bfseries to} $T$}
\STATE
\textbf{Hallucinate label} for each $x_j^U \in U$: \quad
$\displaystyle \tilde{y}_{j}^{t} = - \sgn (\ip{\mathbf{x}^U_j}{\sigma^{t-1}}) \;\cdot \ifn{ \abs{\ip{\mathbf{x}^U_j}{\sigma^{t-1}}} \geq 1 }$
\vspace{1mm}
\STATE
\textbf{Find a classifier $h^t \in \mathcal{H}$} that approximately minimizes weighted error over combined data:
\begin{align}
\label{eq:ermcall}
h^t = \argmax_{h \in \mathcal{H}} \;\lrb{ \frac{1}{m} \sum_{i=1}^{m} y_i^L h (x_i^L) + \frac{1}{n} \sum_{j=1}^{n} \tilde{y}_{j}^{t} h (x_j^U) }
\end{align}
\STATE
\textbf{Add $h^t$ to predictor} with positive weight $\sigma^{t}$ found by line search (e.g. Appendix \ref{sec:gss})
\STATE
\emph{Optional, $\algname$-C:} Total correction -- minimize the slack function over the ensemble so far: $h^1, \dots, h^t$
\STATE
\emph{Optional, $\algname$-D:} If $h^t$ is a decision tree, add all internal nodes of $h^t$ too before performing total correction (see Sec. \ref{sec:mower}).
\ENDFOR
\STATE
{\bfseries Output:} Predictor $g_T (\mathbf{x}) = \mbox{clip}(\ip{\mathbf{x}}{\sigma^T})$
\end{algorithmic}
\end{algorithm}
$\algname$ is straightforward to specify (Alg. \ref{alg:realalg}), with no parameters to tune.
It ignores all currently hedged unlabeled examples because they are already minimizing $\Psi$,
and sends every clipped unlabeled example to the error-minimizing oracle with a hallucinated label of the minority prediction,
to encourage its margin towards zero.
Labeled examples are sent to the oracle unchanged,
and the data are weighted so that $L$ and $U$ have equal weights when no unlabeled data are hedged (see \eqref{eq:ermcall}).
\iffalse
\toinsert{The new algorithm generalizes in a unique manner, conservatively minimizing the margins of unlabeled data.
A useful way to visualize this is the data's score distribution, seen in Figure \ref{fig:genplots} on $L, U$, and a holdout set.
On $L$, the data eventually have scores whose signs match their labels, because $L$ is the unchanging part of the oracle input every iteration.
If the algorithm is keeping scores on $U$ in check, then most examples in $U$ will be hedged for a significant number of iterations,
causing $L$ to dominate in relative weight and influence.
In contrast, supervised boosting tends to herd labeled scores in the positive direction (\cite{SF12}) much less strongly. }
\toinsert{The scores on $U$ are clearly being forced into $[-1,1]$,
and they concentrate at the edges of this interval;
$\algname$ does not keep them from meandering out of the hedged region $[-1,1]$,
but it forces them back to the edge of the region if they do.
Also, in spite of continually hallucinating a label of opposite sign to the ensemble,
notice that the clusters around scores $\pm 1$ roughly correspond to data with classes $\pm 1$,
because of the influence of the labeled data.
Finally, $\algname$'s holdout set scores are very different from Adaboost's;
the latter are indistinguishable from scores on $U$, as for any supervised algorithm.
The combined effect of learning the true labels on $L$ and purposefully hallucinating often-incorrect labels on $U$
is that our holdout score distribution generalizes well, as shown in Figure \ref{fig:genplots}. }
\fi
The $\algname$ update can be seen as greedy coordinate descent on the slack function in the high-dimensional space spanned by $\mathcal{H}$ (Appendix \ref{sec:marvinderiv}).
This dimensionality is a thorny theoretical issue, so that even though the slack function is convex,
a practical step size schedule is not easily understood using optimization-based analysis of coordinate descent.
Even the step size's proportionality constant is of great importance in ensuring that the method converges quickly,
and a good choice depends on the interactions between dimensions (ensemble classifiers) in complex ways.
All these considerations motivate us to use line search to find the appropriate step size;
this is crucial to achieving quick convergence and enabling our total correction results, with details in the appendices.
Another way to improve performance that we experiment with follows the example of \emph{totally corrective} algorithms for supervised boosting (\cite{WLR06}).
After adding each new ensemble classifier, this approach minimizes the objective function over the entire cumulative ensemble so far.
It is especially appropriate in our case because the slack function is convex, so efficient optimization methods are guaranteed to make progress.
We return to this idea in Sec. \ref{sec:experiments}, where we implement $\algname$-C, the totally corrective version of $\algname$.
\section{Maximizing the Performance of an Ensemble of Trees}
\label{sec:mower}
$\algname$ addresses the central issue of learning an ensemble to aggregate, while simultaneously learning the aggregation function also by minimizing the slack function $\gamma(\sigma)$.
To date, the only other work that has attempted to empirically minimize the slack function is the recent paper \cite{BF15b},
whose idea applies to ensembles of decision trees or other partitioning classifiers.
They augment the ensemble with \emph{specialists} constructed from the leaves of the trees
each of which predicts only on the data falling into it, contributing local information about the true labels.
The work \cite{BF15b} optimizes using standard gradient descent without line search,
give evidence that the benefits of such partitioning specialists may complement sequential boosting-type procedures,
and ultimately pose the fusion of the two approaches as an open problem.
\begin{figure*}[tp]
\vskip -0.1in
\begin{center}
\centerline{\includegraphics[width=\linewidth,height=115pt]{wilson_tree.png}}
\label{icml-historical}
\end{center}
\vskip -0.25in
\caption{Effect of Wilson score interval (Sec. \ref{sec:mower}) for measuring $\vb$, on a decision tree with 9 internal nodes (left). In the middle, using 100 labeled data: for each internal node, green bar is fraction of labeled data falling into that node; red bar is the plugin estimate of $\vb$; blue bar is Wilson lower bound, calculated from values of green and red bars. Right: as middle, but with 400 labeled data.}
\label{fig:wilson_tree}
\end{figure*}
We address this problem, extending the idea of \cite{BF15b} to handle all internal nodes, not just the potentially prohibitively small leaves.
So we need calculate the components of $\vb$, i.e. the errors of all the specialists representing internal nodes, simultaneously.
These errors are close to their estimates from the labeled data with high probability --
this \emph{uniform convergence} of the estimates is deeply studied in learning theory (\cite{V82}).
The theoretical uniform bounds are too loose for direct use, though,
so we upper-bound each individually with some very high confidence (e.g. 99.9\%);
by uniform convergence, this probably constitutes a valid uniform bound on the vector $\vb$.
For each node, we are estimating a binomial proportion (say $p$, using an estimate $\hat{p}$);
the natural option for this is Wald's confidence interval with width $\sqrt{\frac{\hat{p} (1 - \hat{p})}{m}}$.
However, this fails to provide adequate coverage in two regimes of interest in decision tree partitionings:
a small number of labeled data falling into a leaf,
and very skewed leaves ($\hat{p} \approx 0$ or $1$).
To maintain coverage of our interval in these situations, we calculate $\vb$ using the lower bound provided by Wilson's score interval (\cite{W27}) for each node of the tree.
This follows accepted practice for estimating $p$ in the aftermentioned regimes of interest (\cite{BCD01}).
Figure \ref{fig:wilson_tree} depicts the effect of using Wilson's interval, even on small pure leaves -- it implements nonuniform shrinkage of all errors towards $1/2$
to ensure good coverage,
in keeping with the conservatism of muffling.
The only parameter here is the confidence level (i.e. the allowed probability of failure) --
a higher such probability makes the prediction more aggressive,
resulting in most of the internal nodes getting "mowed down" to a Wilson lower-bound of $0$, as seen in Figure \ref{fig:wilson_tree}.
We call the resulting algorithm $\textsc{HedgeMower}$, minimizing the slack function over the random forest trees combined with all their internal nodes.
See Sec. \ref{sec:genbeh} for a full specification of the algorithm and of Wilson's score interval.
We find that line search and Wilson's interval are crucially important to our empirical performance.
Line search results in significant improvements over SGD with a stepsize schedule, even without any additional specialist nodes,
surprising in light of the reports of this strategy's ineffectiveness in \cite{BF15b}.
To highlight this, in addition to $\textsc{HedgeMower}$ we implement a simplified algorithm we call $\textsc{HedgeMower}$-1, which simply minimizes the slack function using the complete random forest trees, without augmenting with any specialists.
Following the specialist formulation of \cite{BF15b}, we are adding specialists of many various sizes representing variation at many scales,
represented by a different scaling factor for each column of $\vF$, so that dimensions are "unnormalized" by design.
An open problem of \cite{BF15b} is to use second-order or other convex optimization methods to continue to make progress despite such multiscale issues;
but we believe our approach of first-order line search works satisfactorily here, and it is very efficient (Appendix \ref{sec:gss}).
Fig. \ref{fig:genplots} shows the effect of $\textsc{HedgeMower}$ and $\textsc{HedgeMower}$-1 for a couple of datasets;
the muffling effect of $\Psi$ (recall Fig. \ref{fig:predscorefunc}) is readily apparent,
particularly when specialist knowledge is incorporated ($\textsc{HedgeMower}$, right column of Fig. \ref{fig:genplots}).
\begin{figure*}[tp]
\vskip -0.1in
\begin{center}
\centerline{\includegraphics[width=\linewidth,height=140pt]{mower_score_dists.png}}
\label{icml-historical}
\end{center}
\vskip -0.25in
\caption{Score distributions of four different classifiers in the columns from left to right -- respectively logistic regression, random forests, $\textsc{HedgeMower}$-1, and $\textsc{HedgeMower}$.
Top row: covtype dataset. Bottom row: ssl-secstr. Blue dashed lines are at scores of $\pm 1$, the inflection points of $\Psi$. }
\label{fig:genplots}
\end{figure*}
\section{Empirical Results}
\label{sec:experiments}
We now turn to implementing $\algname$ and $\textsc{HedgeMower}$ on a variety of datasets.
We summarize the algorithms we implement; the first four have each been described in previous sections,
and $\algname$-D combines the two ideas.
\begin{itemize}
\item
$\textsc{HedgeMower}$-$1$ -- Minimize $\gamma (\sigma)$ using just whole RF trees, and none of their specialists.
\item
$\textsc{HedgeMower}$ -- Add all internal nodes of RF trees as specialists, minimize slack function.
\item
$\algname$ -- Add one (non-specialist) tree at a time.
\item
$\algname$-C -- Like $\algname$, but with total correction each timestep.
\item
$\algname$-D -- Similar to $\algname$-C (with total correction), but adding a (non-specialist) tree \emph{and its internal nodes} each timestep, like $\textsc{HedgeMower}$.
\end{itemize}
Our implementations of our new algorithms are in Python, and use the open-source package \texttt{scikit-learn}.
We restrict the labeled data available to the algorithm by various orders of magnitude when feasible, to explore its effect.
Unused labeled examples are combined with the test examples (and the extra unlabeled set, if any is provided)
to form the set of unlabeled data used by the algorithm. All algorithms are run with 100 base unregularized decision trees as ensemble classifiers where applicable.
Class-imbalanced and noisy datasets are included, so that AUC is an appropriate measure of performance.
Results and 95\% confidence intervals are given from 20 Monte Carlo trials (details in appendix), each starting with a different random subsample of labeled data.
Further information\footnote{Our code is available at \url{https://github.com/aikanor/marvin}.} on the sources of the datasets can be found in the appendices (and in \cite{BF15b},
where many of them were used).
\vspace{0.5em}
\begin{table}[tp]
\centering
\rotatebox{90}{
\scriptsize
\setlength\tabcolsep{1.5pt}
\begin{tabular}
{|
M{0.09\linewidth} || M{0.05\linewidth}
|| M{0.065\linewidth} || M{0.065\linewidth} | M{0.065\linewidth} || M{0.065\linewidth} || M{0.065\linewidth} || M{0.065\linewidth} |||| M{0.065\linewidth} || M{0.065\linewidth} || M{0.065\linewidth} || M{0.065\linewidth}
|}
\hline
Dataset & \# labeled & \textsc{Hedge}-\textsc{Mower}-$1$ & \textsc{Hedge}-\textsc{Mower} & \# relevant nodes / \# total nodes per tree & $\algname$ & $\algname$-C & $\algname$-D & RF &Ada-Boost & \hspace{-0em}LogitBoost & LR \\ \hline \hline
\multirow{3}{*}{\texttt{kagg-prot}}
& 100 & 0.58 $\pm$ 0.12 & 0.66 $\pm$ 0.08 & 2 / 12 & 0.65 $\pm$ 0.09 & 0.66 $\pm$ 0.08 & 0.65 $\pm$ 0.12 & 0.64 $\pm$ 0.04 & 0.65 $\pm$ 0.05 & 0.66 $\pm$ 0.06 & 0.64 $\pm$ 0.04 \\ \cline{2-12}
& 500 & \textbf{0.77 $\pm$ 0.05} & 0.76 $\pm$ 0.05 & 13 / 50 & \textbf{0.80 $\pm$ 0.02} & \textbf{0.83 $\pm$ 0.03} & 0.71 $\pm$ 0.04 & 0.72 $\pm$ 0.03 & 0.72 $\pm$ 0.02 & 0.74 $\pm$ 0.01 & 0.70 $\pm$ 0.02 \\ \cline{2-12}
& 2500 & 0.83 $\pm$ 0.03 & 0.83 $\pm$ 0.02 & 61 / 242 & 0.79 $\pm$ 0.04 & 0.83 $\pm$ 0.04 & 0.81 $\pm$ 0.05 & 0.80 $\pm$ 0.04 & 0.77 $\pm$ 0.06 & 0.79 $\pm$ 0.06 & 0.75 $\pm$ 0.04 \\ \hline \hline
\multirow{3}{*}{\texttt{covtype}}
& 1K & 0.81 $\pm$ 0.02 & \textbf{0.82 $\pm$ 0.02} & 23 / 131 & \textbf{0.82 $\pm$ 0.02} & \textbf{0.84 $\pm$ 0.05} & \textbf{0.88 $\pm$ 0.04} & 0.766 $\pm$ 0.014 & 0.729 $\pm$ 0.015 & 0.762 $\pm$ 0.014 & 0.715 $\pm$ 0.015 \\ \cline{2-12}
& 10K & 0.871 $\pm$ 0.005 & 0.870 $\pm$ 0.015 & 254 / 1103 & 0.942 $\pm$ 0.004 & \textbf{0.969 $\pm$ 0.005} & 0.956 $\pm$ 0.018 & 0.848 $\pm$ 0.006 & 0.738 $\pm$ 0.01 & 0.794 $\pm$ 0.006 & 0.734 $\pm$ 0.009 \\ \cline{2-12}
& 100K & 0.941 $\pm$ 0.004 & 0.891 $\pm$ 0.005 & 1963 / 8543 & 0.961 $\pm$ 0.005 & \textbf{0.979 $\pm$ 0.003} & \textbf{0.974 $\pm$ 0.003} & 0.926 $\pm$ 0.006 & 0.749 $\pm$ 0.001 & 0.796 $\pm$ 0.002 & 0.769 $\pm$ 0.002 \\ \hline \hline
\multirow{3}{*}{\texttt{ssl-secstr}}
& 100 & 0.58 $\pm$ 0.06 & 0.55 $\pm$ 0.11 & 2 / 18 & 0.61 $\pm$ 0.05 & 0.58 $\pm$ 0.03 & 0.64 $\pm$ 0.06 & 0.55 $\pm$ 0.03 & 0.55 $\pm$ 0.03 & 0.55 $\pm$ 0.04 & 0.56 $\pm$ 0.02 \\ \cline{2-12}
& 1K & \textbf{0.70 $\pm$ 0.04} & \textbf{0.68 $\pm$ 0.01} & 14 / 164 & 0.64 $\pm$ 0.02 & 0.64 $\pm$ 0.02 & \textbf{0.66 $\pm$ 0.03} & 0.64 $\pm$ 0.02 & 0.63 $\pm$ 0.02 & 0.64 $\pm$ 0.01 & 0.64 $\pm$ 0.01 \\ \cline{2-12}
& 10K & 0.729 $\pm$ 0.006 & \textbf{0.741 $\pm$ 0.006} & 173 / 1630 & 0.680 $\pm$ 0.005 & 0.70 $\pm$ 0.01 & 0.673 $\pm$ 0.010 & 0.673 $\pm$ 0.004 & 0.638 $\pm$ 0.009 & 0.673 $\pm$ 0.006 & 0.701 $\pm$ 0.002 \\ \hline \hline
\multirow{2}{*}{\texttt{adult}}
& 100 & \textbf{0.78 $\pm$ 0.08} & \textbf{0.81 $\pm$ 0.03} & 4 / 16 & 0.73 $\pm$ 0.03 & \textbf{0.77 $\pm$ 0.02} & 0.72 $\pm$ 0.05 & 0.65 $\pm$ 0.09 & 0.68 $\pm$ 0.06 & 0.66 $\pm$ 0.07 & 0.68 $\pm$ 0.08 \\ \cline{2-12}
& 1K & \textbf{0.84 $\pm$ 0.01} & \textbf{0.85 $\pm$ 0.02} & 29 / 126 & 0.75 $\pm$ 0.02 & 0.78 $\pm$ 0.03 & 0.78 $\pm$ 0.04 & 0.72 $\pm$ 0.03 & 0.73 $\pm$ 0.02 & 0.73 $\pm$ 0.03 & 0.75 $\pm$ 0.03 \\ \hline \hline
\multirow{2}{*}{\texttt{ssl-text}}
& 100 & 0.68 $\pm$ 0.05 & 0.65 $\pm$ 0.09 & 5 / 24 & 0.68 $\pm$ 0.02 & 0.73 $\pm$ 0.08 & 0.69 $\pm$ 0.08 & 0.64 $\pm$ 0.09 & 0.65 $\pm$ 0.05 & 0.64 $\pm$ 0.06 & 0.66 $\pm$ 0.16 \\ \cline{2-12}
& 1K & 0.85 $\pm$ 0.02 & 0.88 $\pm$ 0.05 & 54 / 237 & 0.81 $\pm$ 0.01 & 0.86 $\pm$ 0.05 & 0.89 $\pm$ 0.06 & 0.83 $\pm$ 0.07 & 0.80 $\pm$ 0.04 & 0.82 $\pm$ 0.07 & 0.86 $\pm$ 0.03 \\ \hline \hline
\multirow{3}{*}{\texttt{kagg-cred}}
& 1K & \textbf{0.74 $\pm$ 0.03} & \textbf{0.75 $\pm$ 0.04} & 21 / 49 & 0.59 $\pm$ 0.02 & \textbf{0.66 $\pm$ 0.06} & \textbf{0.70 $\pm$ 0.10} & 0.56 $\pm$ 0.03 & 0.59 $\pm$ 0.02 & 0.59 $\pm$ 0.03 & 0.53 $\pm$ 0.05 \\ \cline{2-12}
& 10K & \textbf{0.75 $\pm$ 0.04} & \textbf{0.73 $\pm$ 0.02} & 144 / 356 & 0.62 $\pm$ 0.02 & \textbf{0.75 $\pm$ 0.02} & \textbf{0.72 $\pm$ 0.03} & 0.581 $\pm$ 0.014 & 0.588 $\pm$ 0.015 & 0.590 $\pm$ 0.013 & 0.52 $\pm$ 0.02 \\ \cline{2-12}
& 100K & \textbf{0.74 $\pm$ 0.01} & \textbf{0.74 $\pm$ 0.09} & 1404 / 4015 & 0.62 $\pm$ 0.02 & \textbf{0.76 $\pm$ 0.02} & \textbf{0.74 $\pm$ 0.04} & 0.588 $\pm$ 0.005 & 0.591 $\pm$ 0.007 & 0.595 $\pm$ 0.005 & 0.519 $\pm$ 0.008 \\ \hline \hline
\multirow{2}{*}{\texttt{cod-rna}}
& 1K & 0.943 $\pm$ 0.015 & \textbf{0.969 $\pm$ 0.004} & 31 / 80 & 0.929 $\pm$ 0.008 & 0.94 $\pm$ 0.02 & \textbf{0.95 $\pm$ 0.02} & 0.87 $\pm$ 0.03 & 0.90 $\pm$ 0.02 & 0.90 $\pm$ 0.02 & 0.87 $\pm$ 0.02 \\ \cline{2-12}
& 10K & \textbf{0.973 $\pm$ 0.006} & \textbf{0.967 $\pm$ 0.008} & 205 / 532 & 0.96 $\pm$ 0.01 & \textbf{0.979 $\pm$ 0.004} & \textbf{0.973 $\pm$ 0.007} & 0.93 $\pm$ 0.02 & 0.91 $\pm$ 0.02 & 0.935 $\pm$ 0.016 & 0.92 $\pm$ 0.02 \\ \hline \hline
\multirow{3}{*}{\texttt{SUSY}}
& 1K & \textbf{0.82 $\pm$ 0.02} & \textbf{0.830 $\pm$ 0.015} & 25 / 82 & 0.78 $\pm$ 0.02 & 0.76 $\pm$ 0.04 & 0.78 $\pm$ 0.02 & 0.771 $\pm$ 0.006 & 0.769 $\pm$ 0.009 & 0.771 $\pm$ 0.005 & 0.775 $\pm$ 0.006 \\ \cline{2-12}
& 10K & \textbf{0.837 $\pm$ 0.005} & \textbf{0.829 $\pm$ 0.009} & 187 / 717 & 0.791 $\pm$ 0.010 & 0.818 $\pm$ 0.005 & 0.801 $\pm$ 0.005 & 0.784 $\pm$ 0.003 & 0.777 $\pm$ 0.006 & 0.788 $\pm$ 0.003 & 0.779 $\pm$ 0.002 \\ \cline{2-12}
& 100K & \textbf{0.849 $\pm$ 0.004} & 0.833 $\pm$ 0.007 & 1547 / 7185 & 0.816 $\pm$ 0.009 & \textbf{0.84 $\pm$ 0.02} & \textbf{0.82 $\pm$ 0.04} & 0.797 $\pm$ 0.003 & 0.797 $\pm$ 0.005 & 0.791 $\pm$ 0.003 & 0.779 $\pm$ 0.002 \\ \hline \hline
\end{tabular}
}
\caption{Area under ROC curve for $\algname$ and $\textsc{HedgeMower}$ variants, and supervised ensemble algorithms,
all run with/to 100 ensemble classifiers unless otherwise stated.
All $\algname$ variants are run with unregularized decision tree weak learners.
The "\# relevant nodes..." column refers to the average number of internal nodes not mowed down by Wilson's interval
when running the "$\textsc{HedgeMower}$" column,
as a fraction of the total number of internal nodes.
95\% confidence intervals indicated using 20 Monte Carlo trials, with best algorithm(s) for each row in bold. }
\label{tab:allauc}
\end{table}
We compare our algorithms' performance to that of standard supervised ensemble algorithms under the same conditions -- AdaBoost and LogitBoost,
random forests (100 trees, default parameters) as a high-performance supervised ensemble algorithm, and logistic regression.
We find that one or more of our new algorithms is sufficient to achieve significant improvements over the baselines in all cases.
We further discuss this, and Table 1, in Sec. \ref{sec:disc}.
Many of our datasets are large enough that $U$ will not fit in memory,
making the batch boosting method impractical.
However, there is a fairly straightforward minibatch remedy:
store only a fixed-size minibatch of unlabeled examples, and periodically replace this batch
(or similar, e.g. in a streaming setting, replace a randomly selected example in the batch with each new example that arrives).
This is explained in Appendix \ref{sec:minibatch}.
The only tuning done for the new algorithms is of the Wilson failure probability (details in Appendix \ref{sec:genbeh}).
This applies to all our new algorithms, because all use line search, which requires $\vb$
(just one component at a time for $\algname$, and many at once for the other four algorithms).
The situation is particularly complex for $\algname$-D, which needs enough labeled data to estimate $\vb$ for a growing ensemble including many specialists.
It is certainly possible that further parameter tuning will lead to better performance in future, but we aim to highlight the approaches'
simplicity and generality in this initial evaluation.
\section{Discussion}
\label{sec:disc}
The results of Table 1 show that we achieve significant improvement over the baselines in all cases.
The situation is more unclear when choosing between the new algorithms, an exciting source of future open problems.
However, we can still deduce some statements and recommendations from Table 1.
For a combination of simplicity, speed, and good performance with low variance, we recommend $\algname$-C.
It only adds one new classifier per iteration,
and its results dominate $\algname$ across experiments; total correction with line search is effective on the convex slack function.
As the culmination of our ideas in this paper, $\algname$-D might be expected to perform best overall.
This is possibly true for many datasets, but we cannot conclude this in general, because $\algname$-D often has high variance.
We believe this has to do with the complex way in which it uses labeled data, both online and to estimate specialist errors.
Further exploration appears warranted, because such optimization was out of our scope here.
We find that the algorithms here converge quickly in a number of ways.
Our results typically can be achieved with a small fraction of the unlabeled data available;
beyond this point, we believe that there is a statistical bottleneck in estimating the first term of the slack function, involving $\vb$.
In addition, the table makes clear the profound effect of Wilson's interval on $\textsc{HedgeMower}$, which uses all internal nodes.
Other heuristics, like selecting just the top $k$ nodes by Wilson score for some $k$, perform almost as well (not shown).
All this makes the final decision rule of our algorithms essentially a thresholded linear combination of a few white-box tree learners,
which has the advantageous side effect of being nicely interpretable - the score is just an additive combination of specialist rules, each of which can be written as a decision rule
(involving both the asleep/awake status and predictions when awake), similar to alternating decision trees \cite{FM99} or similar tree ensembles.
\section{Related and Future Work}
\label{sec:relwork}
Semi-supervised learning has been an active area of research over the last decade (\cite{CSZ06}),
mostly involving graph-based methods like label propagation that operate pairwise on the data, and also including the transductive SVM \cite{J99} and other algorithms \cite{ZG09}.
These generally try to locate the decision boundary at low-density regions of the unlabeled data (\cite{CZ05});
when formulated as max-margin methods, they stand in stark contrast to the muffled min-margin idea for generalization.
The labels typically agree with the labeled data and some type of unlabeled cluster structure used as a regularizer (\cite{BNS06}),
while our regularizer is in the same spirit but encourages the opposite muffling behavior.
Other more coarse-grained methods using discriminative statistics \cite{CSZ06, QSCL09} are more in the spirit of our algorithms.
Semi-supervised algorithms for boosting have previously drawn some attention for their applications,
notably in \cite{GLB08, KJJL09}, which also hallucinate labels over the unlabeled data;
but they do not use the muffling framework.
The only practical work which does is the aforementioned method of \cite{BF15b},
and we directly address a main open problem posed in that paper, about combining specialist information with the incremental aggregation idea of boosting.
Another fascinating open problem, building further on these ideas,
is how to target areas of the space with specialist classifiers as part of the incremental process,
rather than just using the specialists provided by decision trees.
This paper is related to the significant existing supervised boosting literature \cite{SF12}.
Such algorithms concern the incremental classifier aggregation idea,
and generally attempt to minimize some convex upper bound on error on $L$.
It would be of interest to incorporate other notions from boosting theory, like weak learnability, into our framework,
or investigate if they are even necessary.
Our understanding of the generalization behavior of the muffling framework is still just beginning, though it is clear that estimation of $\vb$ is heavily involved.
There is already a theoretical connection established between this estimation, generalization, and classifier complexity as measured by $\vnorm{\sigma}_1$,
which relates to $L_{\infty}$ norm constraints on the adversary in the minimax formulation (\cite{BF15c}),
and building on this could yield fruitful practical insights.
Finally, we plan to explore practical applications at larger scale to investigate the space of ensembles that can be aggregated --
for instance, decision trees can be inappropriate in high dimension, and efficient linear classifiers could be used instead.
Deep learning of features is another possibility we would like to explore with the muffling framework,
especially in light of the profound and rapidly expanding set of connections between deep and semi-supervised learning (\cite{B09}).
\newpage
\small
|
\section{Introduction}
Matveev~\cite{matveev_2-3_connectivity_paper, matveev} and (independently) Piergallini~\cite{piergallini} show that the set of triangulations of a three-manifold is connected under 2-3 and 3-2 moves, as long as we ignore the small number of triangulations of manifolds that consist of only a single tetrahedron. For these, no 2-3 or 3-2 move can be applied. Said another way, we consider the \emph{Pachner graph} $\mathfrak{T}(M)$, whose vertices are the triangulations of a given manifold $M$, where two vertices are connected by an edge if the corresponding triangulations are related by a 2-3 move. Matveev's and Piergallini's result then says that $\mathfrak{T}(M)$ is connected, again if we ignore the one-tetrahedron triangulations.
A natural question to ask is whether this remains true when we impose conditions on the triangulations. In other words we consider the connectivity of the subgraph of $\mathfrak{T}(M)$ corresponding to a given property. For example, we could consider the subgraph of geometric triangulations. Hoffman noted that the subgraph of geometric triangulations of the figure 8 knot complement is not connected; in fact there are isolated geometric triangulations (this observation is mentioned in \cite{dadd_duan}). However, geometricity is a very strong property. There are many weaker properties, corresponding to larger subgraphs, which may be connected. Of particular interest are the properties of \emph{1-efficiency} and of \emph{having essential edges}. The \emph{3D index} is a recent quantum object mapping ideal triangulations to Laurent series, introduced by Dimofte, Gukov, and Gaiotto \cite{dimofte_gaiotto_gukov_1, dimofte_gaiotto_gukov_2}. Garoufalidis \cite{garoufalidis_3D_index} showed that the 3D index is invariant under 2-3 and 3-2 moves, provided that it is defined on both sides of the move. In \cite{ghrs_index_structures}, Garoufalidis, Hodgson, Rubinstein and the author show that the 3D index is defined only on 1-efficient triangulations. Therefore, if we knew that the subgraph of $\mathfrak{T}(M)$ corresponding to 1-efficient triangulations were connected, we would have an invariant of the manifold. Unfortunately this is not known. Also in \cite{ghrs_index_structures}, we define an extremely small subgraph of the 1-efficient triangulations that we can prove is connected, and since the subgraph is determined purely by the topology of the manifold, we have an invariant. However, this is a somewhat unsatisfying workaround. A very similar story holds for the \emph{1-loop invariant}, as defined by Dimofte and Garoufalidis \cite{dimofte_garoufalidis}. In this case, the triangulations are required to have solutions to Thurston's gluing equations corresponding to the complete hyperbolic structure, which is implied by the triangulation having essential edges. Again it is not known if the triangulations with essential edges form a connected subgraph of $\mathfrak{T}(M)$.
In this paper, we answer the connectivity question in the affirmative for the subgraph of triangulations with the property of \emph{having no degree one edges}.
\begin{thm}
Let $M$ be an oriented three-manifold other than the lens space L(4,1). If $M$ is closed, let $\mathfrak{T}(M)$ denote the set of single vertex triangulations of $M$. If $M$ has boundary, let $\mathfrak{T}(M)$ denote the set of ideal triangulations of $M$. Exclude from $\mathfrak{T}(M)$ any triangulations consisting of a single tetrahedron. Then the subgraph of $\mathfrak{T}(M)$ consisting of triangulations without degree one edges is connected under 2-3 and 3-2 moves.
\label{main theorem}
\end{thm}
The restrictions on the manifold being orientable, and the triangulation being either ideal or having only a single vertex are likely not very serious. Most probably, a few other special cases would need to be ruled out. For brevity however, in this paper we restrict to these cases. The restriction ruling out L(4,1) is a special case, similar to the cases of single-tetrahedron triangulations. Most of the triangulations of the manifold L(4,1) that do not have degree one edges are in one large connected component, but there is also an infinite family of isolated triangulations without degree one edges, for which any 2-3 move would introduce a degree one edge. These triangulations have no degree three edges, and so there are no 3-2 moves to consider. We discuss this family in Remark \ref{L41_triangs}.
The property of having no degree one edges is very weak, but is a prerequisite for a triangulation to have essential edges, or to be 1-efficient, or for most other properties of triangulations that have been investigated. We might hope that it will be possible to build up to connectivity of these stronger properties from weaker properties such as having no degree one edges.
\section{Definitions and preliminaries}\label{defns_and_prelims}
\begin{defn}
A \emph{model triangulation} is a set of
identical oriented $3$--simplices together with a set of
orientation-reversing face pairings. A face cannot be glued to itself, and each face must be paired with another face. We refer to the quotient space after making the identifications given by the face pairings as a \emph{triangulation}.
The simplices prior to identification, and their vertices, edges, and so on, are
called \emph{model cells}. The map of a model cell to its image in
the triangulation need not be a homeomorphism -- this occurs whenever a cell is glued to itself in some fashion.
An \emph{ideal triangulation} is the result of removing the vertices from a triangulation, producing a manifold with boundary. For a closed manifold $M$, we define $\mathfrak{T}(M)$ to be the set of single-vertex triangulations. For a manifold $M$ with boundary, we define $\mathfrak{T}(M)$ to be the set of ideal triangulations of $M$, and we assume that $M$ has no spherical boundary components.
\end{defn}
\begin{rmk}
A triangulation as defined above (i.e. with material vertices, rather than an ideal triangulation) is topologically a manifold if and only if the \emph{link} of each vertex (i.e. the boundary of a small closed regular neighbourhood of the vertex) is a sphere.
\end{rmk}
\begin{defn}
The \emph{degree} of an edge $e$, written $\deg(e)$, is the number of model edges that are identified to form $e$.
\end{defn}
In this paper, we will translate freely back and forth between a triangulation and its dual \emph{special spine}.
\begin{defn}
A \emph{spine} of a manifold $M$ with non-empty boundary is a compact subpolyhedron $P$, such that $M$ collapses to $P$. For a closed manifold $M$, a spine is a spine of the complement of an open ball in $M$. If we are given a triangulation of a manifold, we obtain the dual (special) spine by inserting into each tetrahedron a \emph{butterfly}. See Figure \ref{butterfly}. A spine is \emph{simple} if every point on a spine has a neighbourhood homeomorphic to one of the pictures shown in Figure \ref{simple_spine_points}.
\begin{figure}[htb]
\centering
\subfloat[A butterfly inside of a tetrahedron.]{
\includegraphics[width=0.2\textwidth]{Figures/butterfly}
\label{butterfly}
}
\quad
\subfloat[Three different types of point on a simple spine. From left to right: a non-singular point, a triple point, and a true vertex.]{
\includegraphics[width=0.65\textwidth]{Figures/point_types_on_spine}
\label{simple_spine_points}
}
\caption{}
\label{spine_definitions}
\end{figure}
A connected component of the set of non-singular points is a \emph{2-stratum}. A connected component of the set of triple points is a \emph{1-stratum}. A simple spine is \emph{special} if each 2-stratum is an open disk and each 1-stratum is an open interval.
\end{defn}
For further details on special spines and the moves on them described in this paper, see Matveev's book, \emph{Algorithmic topology and classification of 3-manifolds}~\cite{matveev}.
\begin{defn}
Let $\mathcal{T}$ be a triangulation with at least two tetrahedra. A \emph{2-3 move} can be performed on any pair of distinct tetrahedra of $\mathcal{T}$ that share a triangular face $\bigtriangleup$. We remove $\bigtriangleup$ and the two tetrahedra, and replace them with three tetrahedra arranged around a new edge, which has endpoints the two vertices not on $\bigtriangleup$. See Figure \ref{2-3}. A \emph{3-2 move} is the reverse of a 2-3 move, and can be performed on any triangulation with a degree three edge $e$, where the three tetrahedra incident to $e$ are distinct (i.e. the three model edges that are identified to form $e$ are edges of distinct model tetrahedra).
\end{defn}
\begin{figure}[htb]
\centering
\subfloat[The 2-3 and 3-2 moves.]{
\labellist
\small\hair 2pt
\pinlabel 2-3 at 380 395
\pinlabel 3-2 at 380 314
\endlabellist
\includegraphics[width=0.44\textwidth]{Figures/2-3_move_tet_and_spine}
\label{2-3}}
\qquad
\subfloat[The 0-2 and 2-0 moves.]{
\labellist
\small\hair 2pt
\pinlabel 0-2 at 339 539.5
\pinlabel 2-0 at 339 452.5
\endlabellist
\includegraphics[width=0.39\textwidth]{Figures/0-2_move_tet_and_spine}
\label{0-2}
}
\caption{Moves on (topological) triangulations, and their dual special spines.}
\label{2-3 and 0-2 moves}
\end{figure}
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $e$ at 121 145
\pinlabel 0-2 at 274 147
\endlabellist
\includegraphics[width=0.5\textwidth]{Figures/0-2_move_cross-section2}
\caption{The 0-2 move, shown acting on a cross-section through the book of tetrahedra around the edge $e$. Edges in the plane of the cross-section are drawn with a thick line, while intersections with triangular faces are drawn with a thin line. The intersections with the two triangular faces the move is performed on are drawn with a dashed line. One half book is on the left, the other on the right.}
\label{0-2_cross-section}
\end{figure}
The focus of this paper is on connectivity of triangulations under the 2-3 and 3-2 moves. In order to organise the argument we will also consider a number of other moves, including the 0-2 and 2-0 moves, and (later) the V-move.
\begin{defn}
\label{0-2_move}
Let $\mathcal{T}$ be a triangulation. A \emph{0-2 move} can be performed on any pair of distinct triangular faces of $\mathcal{T}$ that share an edge $e$. Around the edge $e$, the tetrahedra of $\mathcal{T}$ are arranged in a cyclic sequence, which we call a \emph{book of tetrahedra}. (Note that tetrahedra may appear more than once in the book.) The two triangles and $e$ separate the book into two \emph{half--books}. We unglue the tetrahedra that are identified across the two triangles, duplicating the triangles and also duplicating $e$. We glue into the resulting hole a pair of tetrahedra glued to each other in such a way that there is a degree two edge between them. See Figure \ref{0-2}. In Figure \ref{0-2_cross-section} we see the effect of the 0-2 move on the book of tetrahedra around $e$. The figure shows a cross-section through the book, perpendicular to $e$.
A \emph{2-0 move} is the reverse of a 0-2 move, and can be performed on any triangulation with a degree two edge, where the two tetrahedra incident to that edge are distinct, and the two edges opposite the degree two edge are not identified.
\end{defn}
\begin{rmk}
The 0-2 move is also called the {\em lune} move in the dual language of special spines.
\end{rmk}
\begin{thm}[Matveev~\cite{matveev, matveev_2-3_connectivity_paper}, Piergallini~\cite{piergallini}] \label{connectivity}
Let $\mathcal{T}$ and $\mathcal{T}'$ be two triangulations of a manifold, both of which have at least two tetrahedra, and which have the same number of vertices as each other. Then $\mathcal{T}$ and $\mathcal{T}'$ are connected by a sequence of 2-3 and 3-2 moves.
\end{thm}
Our goal is to modify a path of triangulations given by the above theorem to avoid degree one edges.
\section{Anatomy of a degree one edge}
By definition, a degree one edge is formed from a single model edge $e$. The model tetrahedron that has $e$ as an edge has the two model faces incident to $e$ paired with each other. See Figure \ref{make_deg_1_edge}. Note that only one triangle of a triangulation containing a degree one edge is incident to that edge.
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $e$ at 251 39
\pinlabel $e'$ at 273.5 65
\endlabellist
\includegraphics[width=0.8\textwidth]{Figures/make_degree_1_edge}
\caption{A degree one edge is formed by gluing two faces of a tetrahedron to each other by ``closing the book'' around the edge. Edges of the tetrahedron are drawn with a thick line, while ``horizon'' lines of a triangular face that curves away from us are drawn with a thin line.}
\label{make_deg_1_edge}
\end{figure}
In order to avoid introducing degree one edges when performing 2-3 and 3-2 moves, it will be useful to know under what conditions a degree one edge can arise. A 2-3 move alters the degrees of \emph{at most} nine edges of the triangulation: the nine edges of the two tetrahedra shown in the upper left of Figure \ref{2-3}. When a 2-3 move is performed, the degrees of the three ``equatorial'' edges go down by one, and the degrees of the other six edges go up by one. We say \emph{``at most} nine edges'' here, because there may be identifications among the edges, due to gluings not shown in Figure \ref{2-3}. Thus, in general, a 2-3 move can reduce the degree of an edge of the triangulation by up to three (if the three equatorial edges are identified with each other), while a 3-2 move can reduce the degree of an edge of the triangulation by up to six. However, any such large jumps down in degree cannot result in a degree one edge, as we will see in Corollary \ref{must_go_from_2_to_1}, which is derived from the following lemma.
\begin{lemma}\label{ways_to_create_deg_one}
There are two ways in which a degree one edge can be created, either via a 3-2 move or a 2-3 move. The two possibilities are shown in Figure \ref{deg2_to_deg1_edge}. In both cases, all of the tetrahedra shown are distinct (i.e. each tetrahedron shown corresponds to a distinct model tetrahedron). The triangulations before and after these moves may or may not have identifications amongst the boundary faces of the collections of tetrahedra shown.
\end{lemma}
\begin{figure}[htb]
\centering
\subfloat[The unique way in which a degree one edge can be formed via a 2-3 move.]{
\labellist
\small\hair 2pt
\pinlabel 2-3 at 110 51
\endlabellist
\includegraphics[width=0.45\textwidth]{Figures/deg2_to_deg1_2-3}
\label{deg2_to_deg1_2-3}
}
\subfloat[The unique way in which a degree one edge can be formed via a 3-2 move.]{
\labellist
\small\hair 2pt
\pinlabel 3-2 at 110 51
\endlabellist
\includegraphics[width=0.45\textwidth]{Figures/deg2_to_deg1_3-2}
\label{deg2_to_deg1_3-2}
}
\caption{}
\label{deg2_to_deg1_edge}
\end{figure}
\begin{proof}
We consider the reverse move: starting with a tetrahedron $T$ incident to a degree one edge $e$, and considering which 2-3 or 3-2 moves can be applied to remove the degree one edge.
First, consider the possible 3-2 moves that involve $T$ (and therefore can affect $e$). In Figure \ref{make_deg_1_edge}, $e$ is labelled, as is its opposite edge $e'$. A 3-2 move can only be applied to a degree three edge, all of whose incident tetrahedra are distinct (i.e. three distinct model tetrahedra have model edges identified to form the degree three edge). Therefore we cannot apply a 3-2 move to $e$, since its degree is incorrect, or to the other two edges incident to our tetrahedron, since the model tetrahedron has two of its model edges identified at each of these two edges. So the only possible edge that a 3-2 move can be applied to is $e'$. This move is shown in reverse in Figure \ref{deg2_to_deg1_2-3}. A 3-2 move results in two distinct tetrahedra, so we must start with two distinct tetrahedra when producing a degree one edge as in Figure \ref{deg2_to_deg1_2-3}.
Second, consider the possible 2-3 moves that involve $T$ (and therefore can affect $e$). Such a move must be performed on a triangle of the triangulation, whose two model triangles are on distinct model tetrahedra. Therefore we cannot use the one triangle incident to $e$, and must use one of the other two faces of $T$. Up to symmetry these give the same configuration. The reverse move is shown in Figure \ref{deg2_to_deg1_3-2}. The result of the 2-3 move consists of three distinct tetrahedra, so we must start with three distinct tetrahedra when producing a degree one edge as in Figure \ref{deg2_to_deg1_3-2}.
\end{proof}
\begin{cor}\label{must_go_from_2_to_1}
A degree one edge comes from reducing the degree of a degree two edge by one. No single 2-3 or 3-2 move can convert any edge of degree higher than two into a degree one edge.
\end{cor}
\begin{proof}
Lemma \ref{ways_to_create_deg_one} lists the two ways to produce a degree one edge, and in both cases the edge in question has degree two beforehand.
\end{proof}
\begin{rmk}
It is possible that a 3-2 move can make two degree one edges at once. A 2-3 move can also make more than one degree one edge, but only in a triangulation with multiple material vertices, or both ideal and material vertices.
\end{rmk}
\section{Proof of the main theorem}
In this section, we prove Theorem \ref{main theorem}. We first describe the general strategy of the proof, with details to come in the subsequent pages.
We are given two triangulations $\mathcal{T}$ and $\mathcal{T}'$ of a manifold $M$, neither of which has a degree one edge, and we must find a path of 2-3 and 3-2 moves which connects from one to the other, and which does not pass through a triangulation with a degree one edge. Theorem \ref{connectivity} gives us the existence of a path $\gamma = (\mathcal{T} = \mathcal{T}_1, \ldots, \mathcal{T}_N = \mathcal{T}')$ connected by 2-3 and 3-2 moves, although some of the intermediate triangulations $\mathcal{T}_i, 1<i<N$ may have degree one edges. Our plan is to modify the path to detour around any such triangulations.
First, observe that the tetrahedron $T$ incident to a degree one edge $e$ is unchanged for the entire lifetime of the degree one edge. Any 2-3 or 3-2 move that alters $T$ is one of the moves shown in Figure \ref{deg2_to_deg1_edge}, as we saw in the proof of Lemma \ref{ways_to_create_deg_one}. This means in particular that the triangle $\bigtriangleup$ incident to $e$ is also unchanged for the lifetime of the degree one edge.
Suppose at first that there are no degree one edges in the path $\gamma$ from $\mathcal{T}_1$ to $\mathcal{T}_i$. Suppose that a single degree one edge $e$ is introduced in the triangulation $\mathcal{T}_{i+1}$, persists until $\mathcal{T}_{j-1}$, and then is removed in $\mathcal{T}_j$. Let the single triangle incident to $e$ be called $\bigtriangleup$. We take a detour from $\mathcal{T}_i$, performing a sequence of moves (which do not go through any triangulations with degree one edges) that results in a triangulation $\overline{\mathcal{T}_i}$. This triangulation is identical to $\mathcal{T}_i$, other than that the triangle $\bigtriangleup$ has been unglued, and a triangulation of a three-ball with boundary consisting of two triangles is glued onto the two revealed faces of the triangulation. Said another way, we unglue the triangle $\bigtriangleup$, and insert a \emph{triangular pillow}. Our triangular pillow must not have any degree one edges inside of it. The act of gluing it in where $\bigtriangleup$ was increases the degree of the three edges incident to $\bigtriangleup$.
We now continue the path $\gamma$, with $\overline{\mathcal{T}_i}$ in place of $\mathcal{T}_i$. The next move no longer produces a degree one edge, since the degree of $e$ has been increased by the insertion of the triangular pillow. As we continue this parallel path to the original $\gamma$, no moves alter the triangular pillow, by our previous observation about the fact that $\bigtriangleup$ is unchanged for the lifetime of the degree one edge.
We continue, up until we get to the triangulation $\overline{\mathcal{T}_j}$, which is the same as $\mathcal{T}_j$ with $\bigtriangleup$ unglued and the triangular pillow inserted into the resulting hole. After this move, we remove the triangular pillow by another sequence of moves, that converts $\overline{\mathcal{T}_j}$ back into $\mathcal{T}_j$ (this is precisely the reverse process to that of inserting the triangular pillow). Having completed our detour, we continue with the path $\gamma$, until we reach $\mathcal{T}_N = \mathcal{T}'$. The resulting path then has no triangulations with degree one edges.
If we can perform the above detour, then we can similarly deal with paths with multiple degree one edges, even multiple produced on the same move. All we need to do is apply the move of inserting a triangular pillow for each degree one edge in turn, leaving it there for the lifetime of the degree one edge, then remove it immediately afterwards.
Given this strategy, all we need to do is describe our triangular pillow, and show how it can be inserted into the triangle incident to an edge $e$ just before the degree of $e$ is to be reduced to one, all without introducing any degree one edges in the process.
\subsection{The triangular pillow}
\label{triangular_pillow}
First, we describe the triangulation we use for our triangular pillow. Figure \ref{make_bird_beak} shows a \emph{bird beak}, which is two tetrahedra arranged around a degree two edge. On the right we draw the bird beak in a suggestive manner -- rotating the upper half down so that it looks more like a real-life bird's beak. In Figure \ref{combine_two_beaks}, we glue together two bird beaks, interleaving the mandibles of each beak. The resulting triangulation of the three-ball has two triangular faces on its boundary, and has no degree one edges in its interior. Ungluing a triangle $\bigtriangleup$ of a triangulation and inserting this triangular pillow adds three, three, and eight to the degrees of the edges incident to $\bigtriangleup$.
\begin{figure}[htb]
\centering
\includegraphics[width=0.5\textwidth]{Figures/make_bird_beak}
\caption{A bird beak, is two tetrahedra arranged around a degree two edge.}
\label{make_bird_beak}
\end{figure}
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $3$ at 200 36
\pinlabel $3$ at 238 36
\pinlabel $8$ at 219 9
\endlabellist
\includegraphics[width=0.65\textwidth]{Figures/combine_two_beaks}
\caption{Our triangulated triangular pillow is formed from two bird beaks, interleaved with each other. On the right, the contributions to the degrees of the three edges incident to the triangular pillow are shown. The four edges internal to the triangular pillow have degrees two, two, three and three.}
\label{combine_two_beaks}
\end{figure}
\subsection{The V-move}
We must insert our triangular pillow using 2-3 and 3-2 moves. As a first step, we introduce a bird beak. We use Matveev's \emph{V-move}~\cite[Figure 1.13]{matveev}. The effect of the V-move is shown in Figure \ref{triangulation_V-move}: it wraps a bird beak around two faces of a tetrahedron. Note that there is a symmetry in the resulting three tetrahedra: we can also see this as wrapping a bird beak around the opposite two faces of the same tetrahedron. There are three different possible V-moves to apply in a tetrahedron, because there are three pairs of opposite edges.
\begin{figure}[htb]
\centering
\includegraphics[width=0.5\textwidth]{Figures/triangulation_V-move}
\caption{The V-move wraps a bird beak around two faces of a tetrahedron. }
\label{triangulation_V-move}
\end{figure}
\begin{figure}[htb]
\centering
\subfloat[Before the V-move.]{
\includegraphics[width=0.35\textwidth]{Figures/tet_triang_and_spine}
}
\qquad
\subfloat[After the V-move.]{
\includegraphics[width=0.35\textwidth]{Figures/V-move_triang_and_spine}
}
\caption{The V-move, with both the triangulation and dual special spine shown.}
\label{triangulation_and_spine_V-move}
\end{figure}
The V-move can be implemented using 2-3 and 3-2 moves, assuming that the triangulation has more than one tetrahedron. The process is easier to understand in the dual setting of special spines. Figure \ref{triangulation_and_spine_V-move} shows the V-move again, with both the triangulation and the special spine shown. Vertices, edges and faces of the dual spine correspond to tetrahedra, faces and edges of the triangulation.
\begin{lemma}
Let $\bigtriangleup$ be a face of a triangulation which has no degree one edges. Suppose that $\bigtriangleup$ is incident to two distinct tetrahedra $T_1, T_2$, and suppose that the edges of $\bigtriangleup$ (which may not be distinct) all have degree at least three. Then we can perform any of the three V-moves in either $T_1$ or $T_2$ by sequences of 2-3 and 3-2 moves, none of which introduce a degree one edge at any point.
\label{V-move_lemma}
\end{lemma}
\begin{proof}
Figure \ref{V-move}, adapted from \cite[Figure 1.15]{matveev} shows the process of implementing the V-move with 2-3 and 3-2 moves, using the dual spine to the triangulation. By inspection, we can check that only the three faces marked with a $*$ have their degree reduced below their starting value during the process, and only by one, during the first 2-3 move. (In the dual picture, these three faces correspond to the three edges of the triangle $\bigtriangleup$.) Of the new faces added, the smallest degree we see is two. Moreover, by Corollary \ref{must_go_from_2_to_1}, even if there were identifications between the faces marked $*$, the only way that such a face could become degree one is if it were degree two at the start. By assumption, this is not the case.
\end{proof}
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $*$ at 65 205
\pinlabel $*$ at 56 187
\pinlabel $*$ at 63 166
\pinlabel 2-3 at 345 201
\pinlabel 2-3 at 530 190
\pinlabel 2-3 at 530 64
\pinlabel 3-2 at 345 50
\endlabellist
\includegraphics[width=0.9\textwidth]{Figures/V-move}
\caption{The V-move is a composition of 2-3 and 3-2 moves.}
\label{V-move}
\end{figure}
\subsection{Rotating the mandibles of a bird beak}
With the V-move, we introduce a bird beak, wrapped around two faces and across their common edge $e$ of a tetrahedron. We can think of the introduction of the bird beak as a 0-2 move (as in Definition \ref{0-2_move}), and so it splits the book of tetrahedra around $e$ into two half-books, one of which contains only the one tetrahedron we wrapped the bird beak around.
We will also need to be able to rotate the mandibles of the bird beak around in the split book of tetrahedra, effectively moving tetrahedra from one half-book to the other. This will let us close the mandibles on each other and make our triangular pillow. The move shown in Figure \ref{move_beak} achieves this goal, allowing us to move one mandible of the beak from one side of a tetrahedron to the other, assuming that this third tetrahedron is distinct from the two tetrahedra of the bird beak. The move is shown in the dual picture in Figure \ref{move_beak_spine}, which is adapted from \cite[Fig. 1.18]{matveev}.
\begin{figure}[htb]
\subfloat[The move can be performed by applying a 2-3 move followed by a 3-2 move.]{
\centering
\labellist
\small\hair 2pt
\pinlabel 2-3 at 297 110
\pinlabel 3-2 at 631 110
\endlabellist
\includegraphics[width=0.9\textwidth]{Figures/rotate_beak}
}
\subfloat[The move shown in cross-section view. The lower mandible of the beak rotates to the right within the book of tetrahedra, moving the tetrahedron $T$ from the right half-book to the left half-book.]{
\centering
\labellist
\small\hair 2pt
\pinlabel $T$ at 153 55
\pinlabel $T$ at 440 50
\endlabellist
\includegraphics[width=0.5\textwidth]{Figures/rotate_beak_cross-section2}
}
\caption{Rotating a mandible of a bird beak past a tetrahedron.}
\label{move_beak}
\end{figure}
\begin{rmk}
The move shown in Figure \ref{move_beak} does not allow us to rotate the two mandibles of a single bird beak \emph{past each other}. Here the configuration to consider involves a model triangle $\bigtriangleup$, two of whose edges are identified to form a single edge $e$ of the triangulation. If we perform a 0-2 move, inserting a bird beak and splitting apart the book of tetrahedra around $e$, it can happen that the two mandibles of the beak are glued to each other across $\bigtriangleup$, and we may want to swap their order. The move shown in \cite[Fig. 1.19]{matveev} allows us to rotate mandibles past each other in such a configuration, but we will not require such a move in this paper.
\label{move_beak_past_self}
\end{rmk}
Note that we do not want to rotate one mandible to close onto the other mandible of a single bird beak, as this would result in a degree one edge where the beak is folded over onto itself.
\begin{lemma}
Assume that we have a triangulation with a bird beak, and that collapsing the bird beak via a 2-0 move would not result in a triangulation with a degree one edge. Then
we can perform the move of rotating the bird beak past a tetrahedron that is not one of the two tetrahedra of the bird beak, by performing a 2-3 followed by a 3-2 move, neither of which introduce a degree one edge, unless this move results in a bird beak folded over onto itself.
\label{rotate_bird_beak}
\end{lemma}
\begin{proof}
See Figure \ref{move_beak_spine}. The faces marked $*$ have degree at least two, since otherwise we have a bird beak folded over onto itself. All other faces in the sequence above have degree at least two, either by directly counting vertices in the diagram, or because the corresponding face after performing the 2-0 move has degree at least two, by assumption.
\end{proof}
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel 2-3 at 341 458
\pinlabel 3-2 at 723 458
\pinlabel 2-0 at 356 265
\pinlabel 2-0 at 710 265
\pinlabel $*$ at 41 432
\pinlabel $*$ at 908 341
\endlabellist
\includegraphics[width=0.9\textwidth]{Figures/rotate_bird_beak_spine2}
\caption{The dual picture of rotating a bird beak past a tetrahedron.
\label{move_beak_spine}
\end{figure}
If we start with a triangulation with no degree one edges and then perform a V-move, this lemma allows us to rotate the mandibles of the resulting bird beak to any position we wish without introducing any degree one edges, as long as we move the mandibles of the beak through a part of the book of tetrahedra for which all tetrahedra are distinct (to avoid the issues discussed in Remark \ref{move_beak_past_self}), and we do not close the beak onto itself.
\subsection{Inserting a triangular pillow}
We have our triangulation $\mathcal{T}_i$, which has a triangle $\bigtriangleup$, with incident edge $e$ which is degree two, and which in $\mathcal{T}_{i+1}$ becomes degree one. Let the other two edges of $\bigtriangleup$ be called $e'$ and $e''$, and suppose that the two tetrahedra incident to $e$ are also glued along the edges $\bar{e}'$ and $\bar{e}''$. See Figure \ref{near_tri_and_degree_2}.
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $\bigtriangleup$ at 41 74
\pinlabel $e$ [t] at 45 58
\pinlabel $e'$ [bl] at 95 87
\pinlabel $\bar{e}'$ [tl] at 95 28
\pinlabel $f$ [l] at 69 34
\pinlabel $e''$ [br] at 35 87
\pinlabel $\bar{e}''$ [tr] at 35 28
\pinlabel $v$ [l] at 127 58
\endlabellist
\includegraphics[width=0.3\textwidth]{Figures/near_tri_and_degree_2}
\caption{The vicinity of the triangle $\bigtriangleup$, with edge $e$ of degree two.}
\label{near_tri_and_degree_2}
\end{figure}
Our plan is to use two V-moves, on tetrahedra incident to each of $e'$ and $e''$, to introduce bird beaks that split the books of tetrahedra around $e'$ and $e''$. We then use Lemma \ref{rotate_bird_beak} to rotate the mandibles of the two beaks around until they form our triangular pillow.
In order to apply Lemma \ref{V-move_lemma} and so apply the V-move, splitting the book of tetrahedra around $e'$ say, we need to find a triangle incident to $e'$, all of whose edges are degree at least three. In particular, we require that $e'$ has degree at least three -- we first check that this is the case.
Suppose not, and assume that $e'$ has degree two. Then considering Figure \ref{near_tri_and_degree_2}, we see that the triangulation would have a vertex $v$ with spherical link, with only the two tetrahedra incident to $e$ incident to it. But there must then be another vertex of the triangulation, since not all vertices of the two tetrahedra are at $v$. In the closed case this contradicts the assumption in Theorem \ref{main theorem} that the triangulation has only a single vertex. In the case that $M$ has boundary, the spherical link of $v$ gives another contradiction, since we assume that $M$ has no spherical boundary components.
So we know that $e'$ has degree at least three, as do $e'', \bar{e}',$ and $\bar{e}''$ by the same argument. Next, we need to find a triangle incident to $e'$, all of whose edges have degree at least three. Consider Figure \ref{near_tri_and_degree_2} again. We know that the edge $\bar{e}'$ has degree at least three. If the edge $f$ also has degree at least three then we are done: we have found a triangle incident to $e'$ with three edges of degree at least three. If not, then $f$ must have degree two (since there are no degree one edges), and $f$ is incident to another tetrahedron, $T$, say, which is incident to $e', \bar{e}', e'', \bar{e}''$ and one other edge, say $f_1$, which is opposite $f$ in $T$. The tetrahedron $T$ has two triangles that are incident to $e'$, one of which we have already checked, and the other, which has edges $e', e'',$ and $f_1$. We saw above that $\bar{e}''$ has degree at least three, and so if $f_1$ also has degree three, then we are again done. If not, then again $f_1$ must have degree two, and we continue with another tetrahedron incident to $e', \bar{e}', e'', \bar{e}''$ and one other edge, say $f_2$, which is opposite $f_1$ in the new tetrahedron.
We continue in this fashion, building a stack of tetrahedra, until we either find a triangle incident to $e'$ and two other edges of degree at least three, or the stack wraps around to join onto the back side of the two tetrahedra we started with, and with the final degree two edge being the edge opposite $e$ in the back tetrahedron shown in Figure \ref{near_tri_and_degree_2}. Note that this is the only way in which the stack of tetrahedra can glue to itself.
Since the manifold is orientable, there are four cases to consider, depending on the angle by which the front of the stack of tetrahedra we have built is glued onto the back.
\begin{itemize}
\item If the stack is glued with no rotation (and so there must be an even number of tetrahedra in the stack), then the four vertices of the tetrahedra are distinct after gluing with spherical vertex links, which again contradicts the hypotheses given in Theorem \ref{main theorem}.
\item If the stack is glued with a half-turn rotation (and so again there must be an even number of tetrahedra in the stack), then there are two distinct vertices after gluing, each with spherical vertex links, and again we contradict the hypotheses given in Theorem \ref{main theorem}.
\item If the stack is glued with either a quarter turn or a three quarters turn, then we have a single-vertex triangulation of the manifold L(4,1), which is the exception listed in Theorem \ref{main theorem}.
\end{itemize}
\begin{rmk}
The above argument constructs the triangulations of L(4,1) mentioned in the introduction. Each is formed from a stack of an odd number of tetrahedra, with degree two edges between each pair of tetrahedra in the stack. The two triangles at the top of the stack are glued to the two triangles at the bottom of the stack, with a quarter (or equivalently, three-quarter) turn. No edge has degree three, so no 3-2 move is possible. Moreover, every possible 2-3 move introduces a degree one edge.
\label{L41_triangs}
\end{rmk}
So, we find a triangle incident to $e'$ for which all edges have degree at least three. Then using Lemma \ref{V-move_lemma}, we perform a V-move that results in a bird beak that splits apart the book of tetrahedra around $e'$. Figure \ref{first_V-move} shows the result in the case that the edge $f$ has degree at least three -- we get a bird beak wrapped around one of the two tetrahedra incident to $e$. Otherwise, we obtain a bird beak wrapped around a tetrahedron somewhere further up the stack. By using Lemma \ref{rotate_bird_beak}, we can rotate the mandibles of such a beak one by one down the stack (noting that all tetrahedra of the stack are distinct), until it is in the position shown in Figure \ref{first_V-move}. By moving the lower mandible before the upper one, we can ensure that we do not close the bird beak onto itself when rotating its mandibles, which would produce a degree one edge.
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $e$ [t] at 45 58
\pinlabel $\bar{e}'$ [tl] at 95 28
\pinlabel $f$ [l] at 69 34
\pinlabel $e''$ [br] at 35 87
\pinlabel $\bar{e}''$ [tr] at 35 28
\endlabellist
\subfloat[After performing the V-move, assuming that $f$ has degree at least three.]{
\includegraphics[width=0.4\textwidth]{Figures/first_V-move}
\label{first_V-move}
}
\qquad
\subfloat[After performing the second V-move.]{
\labellist
\small\hair 2pt
\pinlabel $e$ [t] at 45 58
\pinlabel $\bar{e}'$ [tl] at 95 28
\pinlabel $f$ [l] at 69 34
\pinlabel $\bar{e}''$ [tr] at 35 28
\endlabellist
\includegraphics[width=0.4\textwidth]{Figures/second_V-move}
\label{second_V-move}
}
\caption{V-moves used to make two bird beaks.}
\label{V-move_in_place}
\end{figure}
Next, we need to add the other bird beak needed to make our triangular pillow. As before, we know that edges $e''$ and $\bar{e}''$ have degree at least three, and after our previous moves, $f$ now has degree at least three as well. So we can apply Lemma \ref{V-move_lemma} here, splitting apart the book of tetrahedra around $e''$ and obtaining Figure \ref{second_V-move}. Finally, using Lemma \ref{rotate_bird_beak}, we close the two bird beaks around each others mandibles, producing our triangular pillow. See Figure \ref{close_bird_beaks}.
\begin{figure}[htb]
\centering
\labellist
\small\hair 2pt
\pinlabel $e$ [t] at 45 58
\pinlabel $\bar{e}'$ [tl] at 95 28
\pinlabel $f$ [l] at 69 34
\pinlabel $\bar{e}''$ [tr] at 35 28
\endlabellist
\subfloat[After closing the first bird beak.]{
\includegraphics[width=0.4\textwidth]{Figures/close_first_beak}
\label{close_first_beak}
}
\qquad
\subfloat[After closing the second bird beak.]{
\labellist
\small\hair 2pt
\pinlabel $e$ [t] at 45 58
\pinlabel $\bar{e}'$ [tl] at 95 28
\pinlabel $f$ [l] at 69 34
\pinlabel $\bar{e}''$ [tr] at 35 28
\endlabellist
\includegraphics[width=0.4\textwidth]{Figures/close_second_beak}
\label{close_second_beak}
}
\caption{Closing the bird beaks.}
\label{close_bird_beaks}
\end{figure}
To summarise, in this section we described a procedure to introduce a triangular pillow without introducing a degree one edge at any stage. This then completes the proof of Theorem \ref{main theorem}.
\bibliographystyle{../hamsplain}
|
\section{Introduction}
The core concepts of computer security have remained mostly the same since they were developed and articulated in the 1970’s and late 1960’s, e.g., in the seminal papers on protection by Lampson and by Saltzer and Schroeder~\cite{Protection,SaltzerAndSchroeder}.
However, since then, surprisingly little progress has been made in increasing the practical security of computer users, especially compared to the exponential improvements in other aspects of computing during the same time.
Indeed, by most measures, the problems of computer security have grown progressively worse over those decades, even though a plethora of computer security mechanisms has been developed and deployed, and the use of technologies such as advanced cryptography has become ubiquitous.
Today’s computer users suffer every year some inconvenience or damage as a result of computer security issues, as their personal information is stolen from their devices or vulnerable databases, as their computers are infected with unwanted software or harvested into botnets, or simply as their credit cards or vehicles are recalled because of a security breach or vulnerability.
It’s therefore worthwhile to consider approaches to computer security based on new foundations---although any such effort must acknowledge that security is too intricate and pervasive a problem to admit a panacea.
This short position paper presents a \emph{data-driven software security model} that is founded on an abstraction of \emph{empirical programs}, which pairs software with data on all security-relevant events in all of its execution traces.
Enforcing this data-driven model will, for example,
naturally
disallow network access by the Microsoft Windows Solitaire game~\cite{Solitaire}
and also disallow messages that trigger the Heartbleed vulnerability~\cite{Heartbleed}.
In the model, security policies are automatically derived from historical evidence,
and the Solitaire game doesn't use the Windows networking libraries, even though it includes them,
and no naturally-occurring TLS heartbeat messages with huge payloads have ever existed.
This paper further motivates and describes this data-driven model of software security, and gives examples of practical, useful methods for its application.
\subsection{Security Models and the Difficulty of Setting Policies}
It is not a new observation that we have made disappointingly little progress in secure computing.
In a series of talks and papers around the turn of the millennium, Butler Lampson, one of the founding figures in computer security, gave a good overview of the field, its aspirations and promise, and its failures in practice~\cite{LampsonRealWorld}.
In this work,
Lampson convincingly explains the difficulties of computer security, e.g., how it is even harder than traditional security because universal networking has created an unbounded set of would-be attackers, and because computers’ precise, faithful execution implies that exploits can only be stopped by fully guaranteeing the absence of vulnerabilities.
\begin{figure}[t]
\centering
\begin{tabular}{lcl}
Security policy & $=$ & Functional specification \\
Security mechanism & $=$ & Software implementation \\
Security assurance & $=$ & Program correctness \\
\emph{Security model} & $=$ & \emph{Programming methodology}
\end{tabular}
\caption{The correspondence between aspects of computer security and computer software, the last one an addition to those identified by Lampson~\cite{LampsonRealWorld}.}\label{fig:terminology}
\end{figure}
\begin{figure*}[t]
\centering
\fbox{\begin{varwidth}{\dimexpr\textwidth-2\fboxsep-2\fboxrule\relax}
{\em Permit only low-level executions that programmers intended to be possible, \\
unless given explicit, special permission.}
\end{varwidth}}
\caption{The general maxim of the widely-used programmer-intent software security model, as applied to security-relevant events during software execution.}\label{fig:piss}
\end{figure*}
Most importantly, Lampson draws the correspondence, shown in Figure~\ref{fig:terminology}, between security policies, mechanisms, and assurance and the more general pursuit of program correctness (i.e., establishing that software correctly implements a specification).
To his list, Figure~\ref{fig:terminology} adds a correspondence between security models, such as access-control lists, capabilities, or information-flow tracking, and different programming methodologies, such as imperative or functional programming, logical or declarative formalisms, or reactive software modules.
Like programming paradigms,
security models can be best seen as alternative means towards the same end
(at least, if we follow Lampson and ignore general non-interference, etc.~\cite{hyperprop}).
Tasks and approaches easy in one Turing-complete language
may be difficult in another,
and the same holds true for security models,
which, in general,
can restrict executions only to the same set
of enforceable security policies~\cite{Enforceable}.
Lampson's correspondence highlights how the definition of the right security policy---which can be enforced with assurance by the selected mechanisms---is the key obstacle to creating secure software.
Software developers are famously loath to fully specify their intended functionality, and infamously likely to get such specifications wrong, when forced to create them; furthermore, only recently has it become practically feasible to guarantee program correctness for simple software~\cite{SeL4}.
Figure~\ref{fig:terminology} identifies security policy as just a form of specification,
albeit one that is particularly crucial and hard to get right.
It can only be in vain hope that we ask software developers, administrators, or users to define security policy and select enforcement mechanisms,
when the more fundamental, and much better understood, task of creating software that correctly performs its task has proven to be such an insurmountable challenge.
Depending on the task at hand, and the guarantees needed, certain security models and mechanisms may be particularly well-suited to provide assurance, just as software may be best implemented using a certain methodology.
Security policies can range widely in their intent and granularity,
like any other form of software specification,
and for some purposes it can be simple to derive and set the intended policy.
For example,
simple prohibitions of permitted information-flow can comprehensively protect secrecy,
just as declaratively implementing functionality in Datalog can fully ensure decidability.
Security models that enable simple, useful security policies for large classes of software are clearly advantageous, even though their simplicity will prevent such models from addressing many real-world security concerns.
A particularly important
security model that admits simple policies is one that encompasses mechanisms to thwart the exploits of low-level software vulnerabilities.
\subsection{Automatically-derived Policies for Low-level Software}
Over the last couple of decades, stack-based buffer overflows and memory-corruption vulnerabilities have become a primary exploit vector and a critical software security issue.
In defending against such attacks, the security model outlined in Figure~\ref{fig:piss} has been particularly effective at defining simple, useful security policies that successfully prevent exploits.
In this security model---here termed ``programmer-intent software security''---security policies are automatically derived from software source code (or binaries) by identifying simple program properties that are obviously true, based on the programming-language abstractions and semantics and the clear intent of the programmers.
This security model has been instantiated multiple times, for example by placing canaries or cookies on the execution stack to preserve the programmer-intended integrity of function return-address values, or by introducing artificial heterogeneity and randomness to capture an intent that programs be insensitive to the concrete representation of values like pointers~\cite{lowLevelSoftwareSecurity}.
Those instantiations
have adhered to the maxim of Figure~\ref{fig:piss}
and have had to explicitly permit only a handful of special
cases, e.g., for dynamic loading and libraries and signal delivery.
Unfortunately, many of those instantiations tie policy and mechanism too closely and intricately together for the underlying model to be clearly identifiable.
The clearest examples of this security model is the work to enforce the programmer's intended control and data flow that is often termed Control-Flow Integrity (CFI) and Data-Flow Integrity (DFI)~\cite{programShepherd,CFI,SafeCODE,DFI,WIT,fwdCFI}.
In that work, it is clear how different security policies of varying granularity can be automatically derived from the software itself, and how those policies can (greatly) constrain the attacker from exploiting low-level vulnerabilities.
Those policies can
operate at different levels of abstraction and
aim to enforce a detailed abstract model of the program with full precision,
or make only binary, coarse-grained distinctions between code or data,
or even leave certain activity completely unconstrained.
For example, some CFI mechanisms apply only to C++ VTables, but do so precisely, whereas others apply to all indirect control flow in a very coarse manner;
similarly, data-flow integrity mechanisms differ in their abstractions, with some based on data allocation but others on read/write patterns, etc.
\begin{figure*}[t]
\centering
\fbox{\begin{varwidth}{\dimexpr\textwidth-2\fboxsep-2\fboxrule\relax}
{\em Permit only executions that historical evidence shows to be common enough,\\
unless given explicit, special permission.}
\end{varwidth}}
\caption{The general maxim of the data-driven software security model of this paper, as applied to security-relevant events during software execution.}\label{fig:ddss}
\end{figure*}
The great advantage of these control-flow and data-flow security mechanisms
is that their policy is dictated by the already-written software program.
In no case
is the user required to specify policy details: they must simply choose between security policies with different enforcement mechanisms, and different levels of assurance.
Of course, that choice may still be challenging, because a wide variety of mechanisms exists---ranging from binary-translation emulators to highly-optimized compiler-inserted inline guards---with greatly differing software-engineering and performance characteristics.
Furthermore, the parameters of each mechanism are likely tweakable, e.g., to change the level of precision or make other tradeoffs.
Finally, only some of those mechanisms will attempt to provide high assurance, such as those that verify the static CFI properties of the final binary~\cite{Strato}.
However, these are easily-made, commonplace engineering tradeoffs, compared to the primary obstacle of writing security policies or program specifications identified by Lampson.
Mechanisms that embody this programmer-intent security model are now used near universally, in one form or another, undoubtedly due to their lack of user-specified security policies.
However, at best, application of this model removes just one set of vulnerabilities---approximately the ones eliminated by a rigorously type-safe programming language---and leaves unremedied other vulnerabilities, such as actual logic errors made by programmers.
It is worthwhile to consider other ways in which useful, practical security policies can be defined without user specification, e.g., by automatically basing such policies on the experience gathered from absolutely all executions of a software program.
\section{A Data-driven Software Security Model}
Let's posit that software programs could be accompanied by a comprehensive summary of \emph{all} executions of that program, detailing the program's historical behavior in every instance.
Indeed, let's define a new abstraction of an \emph{empirical program} as the static software program (e.g., its source code or executable text) coupled with the multiset of \emph{all} of its execution traces.
Those traces would be captured at some level of granularity
and,
for the purposes of security,
may include only security-relevant execution events.
Since modern software is generally online---to receive security updates, if nothing else---the practical implementation of this empirical program abstraction is not a farfetched idea.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth]{ac_dds}\vspace*{-1.4em}
\caption{Data-driven software security tied to access control, as per~\cite{LampsonRealWorld}.}\label{fig:ddsacls}
\end{figure}
Such an abstraction would naturally support the security model outlined in Figure~\ref{fig:ddss}, which simply prohibits all novel, security-relevant behavior, unless especially permitted.
This model could, by default, prevent many software attacks,
such as privilege-escalation exploits
of the vulnerabilities
regularly discovered in esoteric operating system services.
Most recently,
this model's enforcement would have blocked exploits of the CVE-2016-0728 vulnerability
by prohibiting use of the Linux \texttt{keyctl} system call\label{ref:keyctl}
in commonly-used applications,
since historical evidence
would have shown that this software never used \texttt{keyctl} or kernel keyrings.
Of course, to apply this model in a practical security enforcement mechanism, there are a number of obstacles and naturally-arising questions.
For example, what security policy should be enforced on the first program execution?
But, whatever the obstacles,
the data-driven security model of Figure~\ref{fig:ddss}
is attractive for many reasons.
In particular,
it seems like a natural basis for software security enforcement
to consider how that software has behaved in the past, overall
(cf., state-based security enforcement
that considers only the current execution~\cite{hbac}).
Also,
as shown in Figures~\ref{fig:ddsacls} and~\ref{fig:ddsif},
this model can be naturally combined with existing security models,
by simply ensuring that operations proceed and information flows
in accordance with historical audit logs
(thereby raising the importance of audit to parity with
the other aspects of Lampson's Gold Standard~\cite{LampsonRealWorld}).
\begin{figure}[t]
\vspace*{0.43in}
\centering
\includegraphics[width=\columnwidth]{if_dds}\vspace*{-1em}
\caption{Data-driven software security tied to information flow, as per~\cite{LampsonRealWorld}.}\label{fig:ddsif}
\end{figure}
Crucially, this data-driven security shares with programmer-intent security
the great advantage
that policy is primarily dictated by the already-written (and -executed) empirical programs.
Therefore,
policies can be simple and easy to specify.
The primary parameters of data-driven security policies would be the event
abstraction used in execution traces---for example,
network service requests, API or system calls, function calls, or simply security privileges---as
well as
the historical frequency by which
security-relevant events must have been seen,
to be permitted in the current execution.
Although historical evidence may be interpreted in myriad ways
(e.g., using elaborate machine learning),
efficiency and robustness concerns are likely to favor simpler methods
(e.g., utilizing the set of observed events, not their sequences).
For clarity,
this position paper considers
only security-related events, such as system calls,
which can be easily summarized and combined using set-theoretic operations.
An event is considered to be supported by
historical evidence
if it has occurred at least once (or $k$ times for some fixed, low threshold $k$)
in the execution traces;
otherwise, it is prohibited by the security policy.
Security policies
that such sets of permitted operations to reduce software attack surface
have proven to be of great practical value,
e.g., in network firewalls and operating-system sandboxing~\cite{ChromeSandbox}.
Simple, data-driven software security policies
can be particularly well suited
to large-scale, popular software applications,
which are typically
bloated in size,
composed of innumerable platforms, modules, and libraries,
and full of arcane or unused functionality.
In such commonly-used software
there exist many dusty corners with lurking vulnerabilities,
as well as embedded interpreters, dynamic-library loaders, and reflection APIs
that attackers can exploit to perform arbitrary behavior.
Simply disallowing previously-unseen security-relevant events
can effectively thwart such attacks,
and for widely-used software
that can be a strong basis of historical evidence
from an abundance of execution traces.
Of course, data-driven security is not a panacea,
and its limitations, obstacles, and open questions must be carefully considered.
However,
as outlined above,
data-driven security enforcement can provide significant benefits,
at least for the commonly-used software that most affects users.
\subsection{Differences from Anomaly and Intrusion Detection}
The most prominent past attempts
at automatically deriving security policies from execution traces
have centered on the
techniques of anomaly detection and software intrusion detection.
Although the details differ,
the common idea in this work is
to use a set of traces collected from benign training runs (or trial runs)
to determine what constitutes ``normal'' executions,
and enforce compliance to this security policy
during a later operational phase~\cite{Denning,AnomalySurveyTwo}.
Because
software tends to have a long tail of possible behaviors
and
these techniques use only a partial set of training traces
(e.g., from a special training phase),
they invariably suffer from a great number of false reports
when deployed in practice.
As a result,
they have seen limited practical use,
except as voluminous but helpful warning notifications for manual operators.
\paragraph*{Comprehensiveness}
This paper's data-driven software security model
relies on the empirical program abstraction to
avoid such falsely-reported security violations.
By definition, empirical programs include \emph{all} execution traces,
not just those from training runs.
Those traces should also include all executions
performed during the software's development and testing,
which should ensure that any latent, actual software feature will be represented,
even for the first use of unpopular software.
The long-tail behavior of commonly-used software
should be particularly transparent---or so it would seem---since
their policies could summarize
billions of execution traces.
\paragraph*{Level of abstraction}
For the above to hold true,
the abstractions and methods used for data-driven security policies
must be chosen carefully.
They cannot, in particular, be so fine-grained
that they include user-specific behavior that is orthogonal to the software's semantics.
For example,
security policies
based on the content of user input to text-editing software,
or the URL bar of a Web browser,
would be likely guaranteed to trigger false errors in the future.
Therefore,
data-driven security is best based on simpler, coarser abstractions,
such as sets of system calls used in this paper,
perhaps augmented with invariants on parameter values and their magnitudes.
Techniques for selecting such abstractions
can be
built on the foundation of inferring convergent software invariants~\cite{Daikon,Diduce}.
\paragraph*{Development-process integration}
Also,
data-driven security techniques must be
at least partially integrated into engineering processes
and preferably used throughout the software development lifecycle.
For example,
test coverage analysis could be used to determine
when
security policies are comprehensive and permissive enough to
allow wide software deployment.
Fuzzing, concolic execution, or other automated testing techniques
could be used to increase coverage (e.g., as in~\cite{MiningSandboxes}).
However,
care would have to be taken, or software
with embedded interpreters or other
general-purpose modules
might be found to exhibit all possible behaviors.
Such software-engineering integration
would also be invaluable in maintaining security policies
as software is updated for security, stability, or behavior
(which can be frequent in modern software),
or is given the occasional major upgrade.
(Of course,
a proper accounting of software updates
would require redefining
the empirical program abstraction
to account for differences in versions over time.)
Most importantly,
the maintenance of data-driven security policies would have to be
integrated into software development,
with the deployment of new, summarized security policy
treated as a form of software security update.
\paragraph*{Real-world execution traces}
No matter how software development is performed,
there is no substitute for
the evidence from real-world execution traces
in the crafting of data-driven security policies.
Testing is at best partial,
and software is often used for unanticipated purposes
in unexpected configurations;
no testing framework can hope to
replicate all end-user activity and the environments
that may affect software execution.
It is inevitable that
some execution paths
will be seen first during real-world execution
(e.g., it's not hard to imagine that
this might be true for
software-emulated denormal floating-point computations).
Furthermore,
it may be important
that security policy disallow some behavior
that is common during software development,
such as debugger- or automation-driven inputs;
otherwise,
data-driven security policies
might, by default, open the door to
cases like the well-known sendmail Debug script vulnerability, CVE-1999-0095.
\paragraph*{Context-specificity}
To increase the precision of security policies,
execution traces can
include information about environmental factors,
such as time, locale, user preferences, etc.,
which may affect software behavior in myriad ways.
For example,
software execution may be highly dependent on time:
the Y2K bug and its effect comes to mind,
as does end-of-year processing in some business software.
In this case,
one could hope this behavior had been
captured in execution traces during functionality or integration testing.
More problematically, some software may contain functionality ``easter eggs''
that were hidden even during software development.
In this latter case,
one can ask whether software behavior that has never been seen before, even during testing,
should not simply be counted as a bug, resulting in program termination;
certainly, software crashes often---without good reason---and here at least there would be a good reason.
\paragraph*{Remediation options}
More generally,
a range of remedial actions may be taken
upon violation of a data-driven security policy.
Fail-stop enforcement that halts execution
may be practical in many cases,
e.g., for non-critical software that can be restarted by the user,
and where data loss is not a concern,
as long as the software can be reset to avoid an infinite loop of halting.
Also,
it may sometimes be practical to simply ask the user whether to proceed,
and summarize their decisions to set future policy;
a variant of this approach is used for security-relevant permissions on some mobile phone and Web platforms.
Finally,
even if execution is silently allowed to proceed,
both system administrators and software developers
could make good use of high-accuracy reports of software exploits,
e.g., to set firewall rules or develop security updates.
Remediation by silent alerts would blur the distinction
with traditional anomaly detection,
except that data-driven security enforcement
should trigger vastly fewer alerts.
\paragraph*{Deployment bootstrapping}
Even when using silent alerts,
there should be no enforcement of
data-driven security policies
until
those policies have converged and stabilized,
through the summarization of sufficient real-world execution traces.
This is a key
distinction between
the data-driven software security model
and traditional anomaly detection.
By integrating with the software-development lifecycle,
including initial real-world use of the software,
the process of converging to stable security policies
can be monitored
and at some point---after careful consideration using domain knowledge---the
switch must be flipped to start enforcing those policies.
Thus,
data-driven enforcement won't likely bring any security benefits
during software development, testing, and early deployments;
its benefits will accrue primarily after software has become frequently-used enough,
which, fortunately, is also when its improved security will impact the most users.
\subsection{Open Questions and Formal Modeling}
There still remain open questions
about this data-driven software security model that cannot be answered here.
For example, if security policy is driven by real-world execution traces,
it is not difficult to imagine that attackers might try to get malicious behavior
classified as benign, e.g., by using a Sybil attack to facilitate Mimicry attacks~\cite{Sybil,Mimicry}.
In other contexts, such as crowd-sourced restaurant reviews,
those attacks are largely prevented by risk analysis of registered user accounts.
However, those attacks might be devastating in a software context,
especially where there is no well-defined, accountable registry of users.
Even so,
in some domains, such as datacenter computing,
this obstacle can clearly be overcome,
e.g., by eliminating Sybils or managing their number.
There also remain more formal questions
about the model,
especially as regards to its empirical program abstraction.
From a formal-language perspective,
programs can be seen as language recognizers,
and from this viewpoint
the insecurity of
modern software partially stems from it recognizing too large a set of inputs.
An empirical program restricts the set of recognized inputs
by disallowing some events in execution traces,
at some levels of abstraction,
while the static text of the software itself
implicitly defines a subset of allowed events.
Clearly,
it would be helpful in applying the data-driven software security model
if there was a sound basis for formal reasoning about empirical programs.
However,
it is first necessary to establish
that comprehensive software execution tracing can be done, efficiently enough, in practice,
and used to derive data-driven policies that provide useful security benefits.
\section{Methods for Data-driven Software Security}
Despite its relative simplicity and other attractive qualities,
a data-driven software security model
is not likely to be straightforward to apply, in most domains.
For example,
both
the crucial integration with software-engineering processes
and the
deployment of mechanisms to construct and maintain security policy
are likely to be major challenges, by themselves.
Also,
it is still more of an art than science to
select the abstractions, granularity, and thresholds of empirical programs,
and to determine when trace data has converged into useful security policies---requiring
the skills of hard-to-find artisan security engineers
that are domain experts.
Even so,
over the last few years,
at Google
we have constructed
several data-driven security mechanisms, and have utilized them in various ways,
as part of software products and production infrastructure.
Many of those mechanisms have been experimental, but some have seen significant deployment.
This section describes three of those mechanisms,
which establish that
execution traces can be collected
with low-enough overhead and in a way that protects end-user privacy,
and that those traces can be used to protect users' security and privacy in novel ways.
\subsection{Efficient Monitoring of Software Execution Details}
At Google,
in work led by Michael Vrable,
we have considered
data-driven security policies about system-call behavior
for production software that runs in our datacenters.
For such software, the empirical program abstraction can
be relatively easily realized
by integrating with Google's test-driven development
and by collecting execution-trace summaries
from thousands of process instances.
Also,
fail-stop enforcement can be particularly
well-suited to fault-tolerant software,
which is designed to gracefully tolerate process failures
and automatically discard requests that trigger such failures.
Our experience has shown that system-call-trace-based security policies
can be efficiently collected, summarized, and enforced
using standard technologies,
such as \texttt{ptrace} and \texttt{seccomp\_bpf} on Linux---neither
of which incurs any significant per-system-call cost,
even if applied holistically as might by done with a system-wide profiler~\cite{DCPI}.
We have developed further techniques for efficient tracing
at the level of functions, library routines, or network messages,
based on reordering executable-binary code (including message-marshaling code)
such that execution of supposedly-unused code
can be blocked
using operating-system memory protection.
Finally, we have created mechanisms to robustly handle
abrupt, unexpected changes in software behavior,
and the resulting storms of events and other disruptions.
Therefore,
the data-driven security model can be realized efficiently, in practice,
at least for policies that involve sets of system calls or sets of other support routines and services.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth]{syscallno}
\includegraphics[width=\columnwidth,trim=4 0 0 0]{syscalluse}
\caption{
A representation of the system calls in an empirical program of a large, feature-rich software service,
drawn from concrete data on the first 200 seconds after startup.
Above, the cumulative number of system calls used converges to fewer than 80, out of the 300+ in Linux.
Below,
system call use varies between initialization phases,
with very few calls used in the final production-service phase
(only the first 9 system calls are represented).}\label{fig:chips}
\end{figure}
Furthermore, our work has established
that enforcement of such policies can significantly reduce the attack surface
that is exposed to potentially-vulnerable software.
When viewed as an empirical program,
even the most feature-rich software
makes a relatively modest use of underlying system services---although
the software may include code, modules, and libraries for
all possible functionality, as well as several kitchen sinks.
For example,
the upper half of
Figure~\ref{fig:chips}
shows how, empirically, one such program uses less than a third of the Linux system calls.
The lower half of Figure~\ref{fig:chips} also shows
how such software can exhibit phase-specific behavior
that can be used to improve security,
e.g., by further constraining policy after initialization,
much as is done by the Chrome Web browser~\cite{ChromeSandbox}.
Notably,
in this case,
the software made no use of the \texttt{keyctl} call,
so an automatically-derived security policy
would have prevented exploits of the CVE-2016-0728 vulnerability discussed on page~\pageref{ref:keyctl}.
\subsection{Privacy-preserving Learning of Software Execution Data}
Software monitoring can have privacy consequences for users,
even if it is not immediately apparent how the data is privacy sensitive.
Simply knowing that a software feature has been active can have significant consequences---as
is effectively demonstrated
in television courtroom dramas,
where perpetrators are frequently found guilty
based on evidence from log records captured for benign purposes.
There are many possible methods
that can protect the privacy and anonymity of software users
participating in the collection of execution traces for empirical programs.
For example,
data can be collected and combined using elaborate cryptographic methods---such
such as partially-homomorphic Paillier encryption---or users can simply
rely on Tor-like network anonymization
when providing trace data.
At Google,
we have developed and deployed
some particularly attractive
privacy-preserving technologies for monitoring client data
based on
the the ideas of randomized response~\cite{rappor}.
These technologies are available in the open-source RAPPOR project,
to be found at \texttt{\url{https://github.com/google/rappor}}~\cite{rappor}.
RAPPOR is already used, extensively,
to gather statistics about client-side values
such as user-provided URL domains
in the Chrome Web browser.
RAPPOR
can be easily applied to collect execution trace data,
while preserving both the privacy and anonymity of users.
In particular,
for reports about the frequency of system calls used by empirical programs
it is sufficient to utilize ``basic RAPPOR''---a simple, binary form of randomized response.
Such reports would be low overhead,
both in terms of computation and their size,
and can be easily aggregated into data-driven security policies.
\subsection{Matching User Expectations and Software Permissions}
At Google,
we have in the last few years
developed techniques for estimating
how users expect software to behave,
in different aspects.
This effort has been performed in the context of online software markets,
such as those that exist for mobile phone and Web platforms.
The goal has been to improve the
use of security-related permissions by software in thosse markets
by creating ``peer groups'' of similar software~\cite{applesAndOranges}.
Our premise has been that users will expect
similar behavior
from software that they perceive to offer similar functionality.
Recently,
in work led by Martin Pelikan,
we have gotten very good results
finding apparently-similar software
using modern machine-learning techniques---in
particular,
by using \texttt{word2vec} skip-gram models
on the data available about different software,
such as its descriptions and how users' interact with the software in online markets.
As a result,
we have been able to craft quite accurate software peer groups
of software that provides similar functionality,
and thereby have established
a good basis of comparison
from which to estimate users' expectation.
For the purposes of security,
we have found that
knowing how users expect software to behave
(or a good estimate thereof)
can provide many benefits.
In particular,
it allows actual, concrete software behavior---such as
that which might be provided by an empirical program---to
be compared against users' expectation.
If there is a (large) discrepancy,
a number of remedial approaches can be taken.
Most notably, the software developers can be notified,
and asked to remedy the situation;
other alternatives
include
further automated or manual review of the software,
different handling of the software in the market,
or even asking users their opinion.
Some of those remedial approaches
have been found suitable for practical application,
when measurements have proven them to have clear benefits to users.
\section{Conclusion}
When deciding whether software should be permitted to perform a security-relevant action,
it seems like a good idea to consider the historical evidence
of what actions that software has performed in the past.
For popular, widely-used software,
there are literally billions of executions from which to draw such historical evidence,
thereby allowing a very accurate view of what constitutes ``normal'' software execution to be established.
Furthermore,
for the most part,
this same software is already networked,
and could provide execution trace data to online services that aggregated such evidence.
Motivated by the above,
a distinct model for data-driven software security can be established.
This data-driven model is different from the traditional approaches of anomaly and intrusion detection,
e.g., in its comprehensiveness and integration with software-development processes.
This model
immediately raises concerns about efficiency, privacy, and practical utility,
but these concerns
can be positively addressed,
using existing techniques and mechanisms.
While many questions remain about the model's general applicability and deployment,
its enforcement could already provide substantial security benefits
to software that runs in some important domains---by reducing its attack surface,
and
thereby protecting the
software in the same manner as firewalls
have very successfully protected networks.
\section*{Acknowledgment}
The ideas in this position paper have benefited from the thoughts and work of many.
In alphabetical order, thanks go to
Mart{\'\i}n Abadi,
Blake Bassett,
Ludovico Cavedon,
Andy Chu,
Giulia Fanti,
Iulia Ion,
Suman Jana,
Noah Johnson,
Aleksandra Korolova,
Karen Lees,
Ilya Mironov,
Martin Pelikan,
Vasyl Pihur,
Ananth Raghunathan,
Sooel Son,
Dawn Song,
Michael Vrable,
Moti Yung,
and
Yinqian Zhang.
Thanks also to Alessandra Gorla, Florian Gross, and Andreas Zeller
for collaboration based on their related work in~\cite{Chabada}
and discussion of their most recent work on mining sandboxes~\cite{MiningSandboxes}.
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Clusters of galaxies are the largest gravitationally bounded structures in the Universe and play a fundamental role
in testing cosmological models, investigating the formation and growth of structures of the Universe and properties of dark matter (DM).
The combination of different independent techniques (stellar kinematics in the Brightest Cluster Galaxy, X-ray, weak and strong lensing, stellar kinematics, etc.) allows to robustly constrain the DM density profile of galaxy clusters from the inner region up to large radial distances ($\rm \sim 5\, Mpc$) \citep[see][]{Sand2002,Sand2004,Newman2009,Newman2011,Umetsu2012,Umetsu2016,Biviano2013,Balestra2015}.
Gravitational lensing offers a unique technique to investigate the mass distribution of galaxy clusters, since it is generated by the total mass of the lens, i.e. both the baryonic and dark matter components. In the strong lensing regime, when giant arcs and multiple images of background sources are generated, the mass distribution of the inner region of galaxy clusters can be reconstructed in great detail \citep[e.g. see][]{Eichner2013, Monna2015, Grillo2014}.
In this context the Cluster Lensing And Supernovae survey with Hubble (CLASH, \citealt{Postman2012a}) and the CLASH-VLT survey, provide high resolution photometry and spectroscopic dataset to measure cluster mass profiles by combining strong and weak lensing analyses (see \citealt{Umetsu2012}, \citealt{Coe2012}, \citealt{Medezinski2013}). In particular, the total mass profile of inner regions ($\rm R < 100-200 kpc$) of galaxy clusters can be derived with extremely high precision of few percentage using deep high resolution HST photometry, as that collected by the CLASH and the deeper Hubble Frontier Fields (HFF, P.I. J. Lotz) surveys \citep[e.g., see ][]{ Jauzac2015, Grillo2014}.
\\
In this paper we present the strong lensing reconstruction of the core of the galaxy cluster MACS J2129.4-0741 (\citealt{Ebeling2007}, hereafter MACS\,2129) using the high resolution photometry from CLASH and new spectroscopy from the CLASH-VLT survey.
This is a massive galaxy cluster ($M_{2500}=4.7\pm1.7\times10^{14}h_{70}^{-1}M_{\odot}$, \citealt{Donahue2014}), selected within the CLASH survey for its lensing strength.
A peculiar bright red galaxy is sextuply lensed in its core, and the deflection of this multiple system is strongly affected by the gravitational action of the central galaxies. Six further multiple images systems are identified in the cluster core. These allow us to put strong constraints on the mass distribution in the innermost region of the cluster. \citet{Zitrin2011a} present a first strong lensing model of the cluster, which is subsequently refined using the CLASH photometry in \citealt{Zitrin2015}, reproducing the multiple images with rms of $\sim2\arcsec$. Here we perform a more detailed analysis of the cluster core and take advantage of the new CLASH-VLT spectroscopic data to trace the cluster mass distribution with higher accuracy.
\begin{figure*}
\centering
\includegraphics[width=18cm]{macs2129_fov.png}
\caption{\small Color composite image of the core of MACS\,2129, generated using the CLASH HST dataset: Blue=F435W+F475W; Green=F555W+F606W+F625+F775W+F814W+F850LP; Red=F105W+F110W+F125W+F140W+F160W.
We label with different colors and numbers the seven multiple images systems used in the SL analysis. All systems have spectroscopic measurements except system 5 and 7. System 8 is a candidate quadruple lensed source. Image 8.1 was serendipitously discovered within our CLASH-VLT survey and has redshift $z_{sp}=4.4$. We selected three candidate counter images based on photometry, which are labelled as 8.2, 8.3 and 8.4.
We also label the galaxy GR used as reference for the galaxy luminosity scaling relation and the two cluster members G1 and G2 which are individually optimized in our model.
\label{fig:fov}
\end{figure*}
The paper is organized as follows.
In Section~\ref{sec:dataset} we present the photometric and spectroscopic datataset.
In Section~\ref{sec:lensing} we introduce the cluster mass components and the lensed systems used in the analysis. In Section~\ref{sec:point_model} we describe the performed lensing analysis and in Section~\ref{sec:results} we provide and discuss the results.
Summary and conclusions are given in Section~\ref{sec:conclusions}. Finally in the Appendix we summarize the CLASH spectroscopy for the background sources and cluster members, and provide the 2D spectra and spectral energy distribution (SED) fit of the multiple images used in the lensing analysis. \\
Through the paper we assume a cosmological model with Hubble constant $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$,
and density parameters $\Omega_{\rm m} = 0.3$ and $\Omega_{\Lambda}=0.7$. Magnitudes are given in the AB system. At the redshift of the cluster ($\rm z_{cl}=0.589$) 1\arcsec corresponds to 6.63 kpc.
\section{Photometric and Spectroscopic Dataset}
\label{sec:dataset}
\subsection{ HST Photometry}
MACS\,2129 was observed between May and August 2011 as part of the CLASH survey with the HST Advanced Camera for Surveys (ACS) and the HST Wide Field Camera 3 (WFC3) UVIS and IR cameras. These observations provide high resolution photometry in 16 filters covering the UV, optical and NIR range . In addition, photometry in the ACS/WFC/F555W filter is also available from the public HST archive.
The photometric data set (available at http://archive.stsci.edu/prepds/clash, see \citet[][]{Postman2012}) is composed by HST mosaic drizzled 65mas/pixel images generated
with the \texttt{Mosaicdrizzle} pipeline \citep[see][]{Koekemoer2011}. These cover a field of view (FOV) of $\sim 3.5'\times3.5'$ in the ACS images and $\sim2'\times2'$ in the WFC3IR images.
In Tab.~\ref{tab:phot} we list the complete filter list of our photometric dataset with the respective total exposure time and $5\sigma$ detection limit. \\
Using these data,
we generate a multi-band photometric catalog of fluxes extracted within $0.6\arcsec$ (9 pixels) diameter aperture, using \texttt{SExtractor} 2.8.6 \citep{Bertin1996} in dual image mode. As detection image we use the weighted sum of all the WFC3IR images.
\begin{table}
\caption{Photometric Dataset summary for MACS\,2129: column (1) filters, column (2) total observation time in seconds, column (3) HST instrument, column (4) the $5\sigma$ magnitude depth within $0.6\arcsec $.}
\centering
\footnotesize
\begin{tabular}{|c|c|c|c|}
\hline
\hline
Filter&Total time (s)&Instrument&5$\sigma$ Depth\\
\hline
f225w & 6934 & WFC3/UVIS & 25.37\\
f275w & 7243 & WFC3/UVIS & 25.42 \\
f336w & 4580 & WFC3/UVIS & 25.78 \\
f390w & 4563 & WFC3/UVIS & 26.47 \\
f435w & 3728 & ACS/WFC & 26.22 \\
f475w & 4040 & ACS/WFC & 26.64 \\
f555w & 8880 & ACS/WFC& 26.97 \\
f606w & 3728 & ACS/WFC & 26.86 \\
f625w & 3846 & ACS/WFC & 26.37 \\
f775w & 4048 & ACS/WFC & 26.24 \\
f814w & 13396 & ACS/WFC & 27.14 \\
f850lp & 7808 & ACS/WFC & 25.91 \\
f105w & 1006 & WFC3/IR & 26.21 \\
f110w & 1409 & WFC3/IR & 26.72 \\
f125w & 1409 & WFC3/IR & 26.41 \\
f140w & 2312 & WFC3/IR & 26.87 \\
f160w & 3620 & WFC3/IR & 26.66 \\
\hline
\end{tabular}
\label{tab:phot}
\end{table}
\subsection{VLT Spectroscopy}
MACS\,2129 was observed with the Visible Multi-Object Spectrograph \citep[VIMOS,][]{LeFevre2003} at ESO VLT as a part of the CLASH-VLT ESO Program 186.A-0798 \citep[P.I. Rosati P., see][]{Rosati2014}. The cluster is partially obscured in the north-east side by a Galactic cirrus, thus only the central and South-western portion of the cluster were targeted by our VIMOS observations, covering a field of $16^\prime\times18^\prime$. Only two 4h-pointings were performed, one using the low resolution (LR) blue grism (covering the spectral range of 3700-6700\,\AA\, with a resolution of 180) and one using the medium resolution (MR) red grism (which covers the 4800-10000\,\AA\, with resolution of 580). The spectra were reduced with the Vimos Interactive Pipeline Graphical Interface \citep[VIPGI][]{Scodeggio2005} and redshifts were measured following the procedure described in \citealt{Balestra2015}. A quality flag (QF) is assigned to each spectroscopic redshift depending if it is ``secure" (QF=3), ``likely" (QF=2) , ``insecure" (QF=1) or if it is based on a strong emission line (QF=9). With a number of slits significantly reduced ($\sim1/10$) compared to other clusters observed in the CLASH-VLT campaign, we obtained a total of 281 redshifts, including 48 cluster members and 19 lensed sources. In the core of the cluster ($\rm R<1.5^\prime$), we have secured redshifts for 4 cluster members, 4 multiple images systems and 15 additional lensed background sources. In the appendix we provide the complete list of spectroscopic redshifts of cluster members and lensing features with their respective QF (see Tab.\,\ref{tab:galaxy_vlt} and \,\ref{tab:lensed_images}). \\
We also collected additional spectroscopic data available from literature \citep{Stern2010} which provide spectroscopic measurements of 6 further cluster members in the cluster core, observed with LRIS instrument on the Keck I telescope.
Recently new spectroscopic measurements for cluster members and lensing features of MACS\,2129 has been released from the Grism Lens-Amplified Survey from Space \citep[GLASS, ][]{Treu2015}. However it provides no additional cluster member with robust $z_{sp}$ measurement to our galaxy sample.
\section{Strong Lensing Ingredients }
\label{sec:lensing}
We perform the strong lensing analysis in the core of MACS\,2129, using the parametric mass modeling
software \texttt{"Gravitational Lensing Efficient Explorer" (GLEE)}, developed by S. H. Suyu and A. Halkola \citep{Suyu2010,Suyu2012}.
We adopt parametric mass profiles to describe the smooth large scale cluster dark halo (DH) and the cluster members. The positions of the observed multiple images are used as constraints to estimate the mass profiles parameters.
Through a simulated annealing minimization in the image plane we derive the best fitting model which reproduces the observed images. Then we perform a Monte Carlo Markov Chain (MCMC) sampling to find the most probable parameters and uncertainties for the cluster mass components.
\subsection{Cluster dark matter halo}
The large scale smooth DH of the cluster is described through a Pseudo Isothermal Elliptical Mass Distribution (PIEMD) profile \citep{Kassiola1993}, with projected surface mass density given by
\begin{equation}
\Sigma(R)= \frac{\sigma^2}{2 {\rm G}}(r_{\rm c}^2+R^2)^{-0.5}
\end{equation}
where $\sigma$ is the halo velocity dispersion and $r\rm_c$ is the core radius. The projected radius R is given by $R^2=x^2/(1+e)^2+y^2/(1-e)^2$ , where $e=(1-b/a)/(1+b/a)$ is the halo ellipticity, and $b/a$ is the halo axial ratio. For $(b/a\rightarrow1,r_c\rightarrow0)$, the asymptotic Einstein radius $\theta_E$ is given by
\begin{equation}
\theta_E=4\pi\left(\frac{\sigma}{c}\right)^{2}\frac{D_{ds}}{D_s}=\Theta_E\frac{D_{ds}}{D_s}
\label{eq:theta}
\end{equation}
where $D_{s}$ is the source distance and $D_{ds}$ is the distance between the lens and the source.
In Eq.\,\ref{eq:theta} we introduce the Einstein parameter $\Theta_E$, which is the Einstein ring when the ratio $D_{ds} /D_{s}$ is 1. In our analysis the amplitude of the DH mass component is described through the Einstein parameter $\Theta_E$, which is optimized in the range $[5\arcsec,45\arcsec]$. The core radius $r_c$ is optimized within [0\arcsec,25\arcsec], while position angle PA and axis ratio vary within $[-90^{\circ},90^{\circ}]$ and [0,1], respectively.
The position of the DH centre is optimized within $\pm3\arcsec$ around the central brightest cluster member (BCG). We adopt flat priors on all the DH parameters.
\subsection{Cluster members}
\label{sec:cmembers}
\begin{table}
\centering
\footnotesize
\caption{List of the spectroscopically confirmed cluster members in the core of MACS\,2129: column (1) id, columns (2) and (3) RA and DEC in degrees, column (4) auto magnitude extracted with \texttt{SExtractor} in the F814w filter, column (5) spectroscopic redshift.}
\begin{threeparttable}
\begin{tabular}{|c|c|c|c|c|}
\hline
\hline
ID& RA&DEC & $\rm F814W_{auto}$&z\\
\hline
873 & 322.35712& $-$7.68909& $22.19\pm 0.01 $ & 0.583\tnote{a} \\
945 & 322.35495& $-$7.69178& $20.16\pm 0.01 $ & 0.596\tnote{a} \\
1298& 322.35095& $-$7.69715& $20.92\pm 0.01 $ & 0.593\tnote{a} \\
1747& 322.35324& $-$7.70795& $21.15\pm 0.01 $ & 0.590\tnote{a} \\
BCG & 322.3588 & $-$7.69105& $19.42\pm 0.01 $ & 0.589\tnote{b} \\
884 & 322.34265& $-$7.68953& $20.47\pm 0.01 $ & 0.590\tnote{b} \\
947 & 322.35574& $-$7.69188& $20.41\pm 0.01 $ & 0.586\tnote{b} \\
984 & 322.3631 & $-$7.69129& $20.92\pm 0.01 $ & 0.579\tnote{b} \\
1167& 322.35013& $-$7.69443& $21.10\pm 0.01 $ & 0.586\tnote{b} \\
1676& 322.3504 & $-$7.70633& $19.68\pm 0.01 $ & 0.596\tnote{b} \\
\hline
\end{tabular}
\begin{tablenotes}
\item[a] redshift measurement from CLASH-VLT survey
\item[b] redshift measurement from \cite{Stern2010}
\end{tablenotes}
\end{threeparttable}
\label{tab:vlt_membrs}
\end{table}
In addition to the 10 spectroscopic members ( see Tab.\,\ref{tab:vlt_membrs}), we select further cluster members in the core of the cluster ($\rm R<1.5^\prime$) using the CLASH photometry. These are bright ($\rm f814w_{auto}<24$) galaxies which lie on the cluster red sequence ($\rm f625w-f814w\leq 1\pm0.3$), see Fig.\,\ref{fig:col_mag}. In addition we require that these candidate cluster members have photometric redshift $\rm z_{ph}$ within $\rm z_{cl}\pm\Delta z$, with $\rm \Delta z= 0.04(1+z_{cl})$.
Photometric redshifts are computed based on the CLASH photometry using the spectral energy distribution (SED) fitting code \texttt{LePhare}\footnote{http://www.cfht.hawaii.edu/$\sim$arnouts/lephare.html} \citep{Arnouts1999,Ilbert2006}.
To estimate photometric redshifts we use galaxy SED templates from the COSMOS \citep{Ilbert2009} template set.
\begin{figure}
\centering
\includegraphics[width=9cm]{color_diagr_cl_new1.png}
\caption{\small Color magnitude diagram for sources extracted in the WFC3IR FOV of the cluster MACS\,2129 at $z=0.589$.
We plot the color from the aperture magnitudes f625w and f814w, versus the \texttt{SExtractor} mag$\_$best in the f814w filter.
Blue circles are all the sources extracted in the WFC3IR FOV;
Red circles are the spectroscopically confirmed cluster members ;
Green circles are cluster member candidates with $z\rm_{ph}\in[0.52,0.65]$ and $\rm f814w\_mag\_best<24$.
}
\label{fig:col_mag}
\end{figure}
To account for the interstellar medium extinction we apply the Calzetti extinction law \citep{Calzetti2000} to the starburst templates, and the SMC Prevot law \citep{Prevot1984} to the Sc and Sd templates.\\
We select 83 bright cluster member candidates in the cluster core based on the photometry. Two further galaxies are included in the sample after visual inspection, which have bluer color ($\rm f625w-f814w\sim0.6$) due to photometric contamination by nearby blue sources. We have a total of 85 candidate cluster members which add to the 10 spectroscopically confirmed ones.\\
These 95 galaxies are included in the cluster SL mass model and
are described with dual pseudo isothermal elliptical profiles (dPIE) \citep{Elisa2007}. The projected surface mass density for this profile is
\begin{equation}
\Sigma(R)= \frac{\sigma^2}{2 {\rm G} R}\frac{r_{tr}^2}{r_{tr}^2-r_c^2} \left(\frac{1}{\sqrt{1+r_c^2/R^2}}-\frac{1}{\sqrt{1+ r_{tr}^2/R^2}}\right)
\label{eq:bbs}
\end{equation}
where $R^2=x^2/(1+e)^2+y^2/(1-e)^2$ as for the PIEMD mass profile, $\rm r_c$ is the core radius, $\rm r_{tr}$ is the so-called truncation radius which marks the region where the density slope changes from $\rho\propto r^{-2}$ to $\rho\propto r^{-4}$.
The total mass for this profile is given by
\begin{equation}
M_{tot}=\frac{\pi\sigma^2}{G}\frac{r_{tr}^2}{r_{tr}+r_{c}}\,\,.
\label{eq:m_bbs}
\end{equation}
We use vanishing core radii for the galaxies, thus, for each galaxy mass profile we have 2 free parameters, i.e. the velocity dispersion $ \sigma$ and the truncation radius $r_t$.\\
We adopt as reference galaxy (GR) one of the candidate cluster members, and then derive the velocity dispersion and size of all the other cluster members through the following luminosity scaling relations, according to the Faber-Jacksons and the Fundamental Plane relations:
\begin{equation}
\sigma=\sigma_{GR}\left(\frac{L}{L_{GR}}\right)^{\delta}
\label{eq:F_J}
\end{equation}
\begin{equation}
r_{\rm tr}=r_{\rm tr,GR}\left(\frac{L}{L_{GR}}\right)^\alpha=r_{\rm tr,GR}\left(\frac{\sigma}{\sigma_{GR}}\right)^\frac{\alpha}{\delta}
\label{eq:r_tr}
\end{equation}
Following our previous work, we adopt as exponent of the scaling relations $\delta=0.3$ and $\alpha=0.4$, assuming a constant mass to light ratio. See \cite{ Eichner2013, Monna2015} for a detailed derivation of these values.\\
The reference galaxy GR (Ra=322.3626, Dec=-7.69238 degrees, see Fig.\,\ref{fig:fov}) is a cluster member with magnitude $\rm F814w\_auto=21.3$ and axis ratio $\rm b/a=0.95$, which lies $\sim14\arcsec$ from the BCG. The velocity dispersion and truncation radius of GR are optimized with flat priors within the ranges of [100,300] km/s and [1,100] kpc, respectively. The mass parameters $\sigma$ and $\rm r_{tr}$ for all the other cluster members are then derived through the previous scaling relation.\\
The three central galaxies (the BCG and the two galaxies G1 and G2, lying on its sides, see Fig.\,\ref{fig:fov}) clearly affect the distortion of the radial multiple images of system 1, thus we individually optimized their parameters through the analysis.
The truncation radii for these galaxies vary within [1,100]kpc, the velocity dispersions are optimized in [100,300] km/s, except for the BCG for which we use a larger range of [100,410] km/s. In addition also their axis ratio $b/a$ and position angle PA are optimized, using the value extracted from the f814w image with gaussian priors.\\ For all the other galaxies $b/a$ and PA are fixed to the value extracted with Sextractor in the filter f814w.
\subsection{Lensed images}
\label{sec:multiple_images}
\begin{table*}
\caption{Summary of the multiple images identified in the core of MACS\,2129: column (1) ID, column (2) and (3) RA and DEC in degrees, column (4) source redshift. Column (5) and (6) give the redshift of the lensed sources and the offset $\delta\theta$ between the observed and predicted multiple images as derived from our SL model, see Section \ref{sec:results}. All the systems have spectroscopic measurements, except for system 5 and 7, for which we provide the photometric redshift measurements from the online public CLASH archive, with their 68\% ranges.}
\centering
\footnotesize
\begin{tabular}{|c|c|c|c|c|c|}
\hline
\hline
ID& RA& DEC& $\rm z_s$& $z_{sl}$& $\delta\theta (\arcsec)$\\
\hline
1.1 & 322.35797 & -7.68588 & 1.36&1.36 &0.3 \\
1.2 & 322.35965 & -7.69082 & " &"&0.2 \\
1.3 & 322.35925 & -7.69095 & " &"&0.1 \\
1.4 & 322.35712 & -7.69109 & " &"&0.2 \\
1.5 & 322.35764 & -7.69115 & " &"&0.4 \\
1.6 & 322.35861 & -7.69489 & " &"&0.6 \\
2.1 & 322.35483 & -7.6907 & 1.04&1.04 &0.2 \\
2.2 & 322.35477 & -7.6916 & " &" &0.1 \\
2.3 & 322.35538 & -7.69332 & " &"&0.4 \\
3.1 & 322.35022 & -7.68886 & 2.24&2.24 &1.8 \\
3.2 & 322.35004 & -7.68982 & " &"&0.3 \\
3.3 & 322.35095 & -7.69577 & " &"&1.7 \\
4.1a & 322.36642 & -7.68674 & 2.24&" &0.4 \\
4.1b & 322.36651 & -7.68689 & " &"&0.3 \\
4.2a & 322.36693 & -7.68831 & " &"&0.4 \\
4.2b & 322.36695 & -7.68820 & " &"&0.2 \\
4.3a & 322.36679 & -7.69497 & " &"&0.3 \\
4.3b & 322.36666 & -7.69525& " &"&1.3 \\
5.1& 322.36422& -7.69387& 1.8 [1.4, 2.0]&{$1.71\pm0.06$}&0.6\\
5.2& 322.36460& -7.69137& 2.0 [1.8, 2.3]&"&0.3\\
5.3& 322.36243& -7.68493& 1.8 [1.4, 2.2]&"&0.5\\
6.1& 322.35094& -7.69333& 6.85 &6.85&1.4\\
6.2& 322.35324& -7.69744& " &"&1.7\\
6.3& 322.35394& -7.68164& " &"&1.9\\
7.1& 322.35714& -7.69425& 1.4 [1.2, 1.5]&{$1.33\pm0.05$}&0.4\\
7.2& 322.35625& -7.69172& - &" &0.7\\
7.3& 322.35670& -7.68554& 1.4 [1.1, 1.5]&"&0.5\\
8.1 & 322.35698 &-7.68924 &4.41 & 4.41& 0.5\\
8.2 & 322.36167 &-7.68808 &" & "& 1.4\\
8.3 & 322.35860 &-7.68491 &" & "& 1.2\\
8.4 & 322.36035 &-7.70094 &" & "& 1.4\\
\hline
\end{tabular}
\label{tab:multiple_images}
\end{table*}
Eight systems of multiple images are clearly identified in the HST images of MACS\,2129, see Fig.\,\ref{fig:fov}. Seven of these were presented in \citet{Zitrin2015} (system 1 to 7), while system 8 was serendipitously discovered as a $\rm z_{sp}=4.4$ lensed source in the cluster core by our VLT observations. Within the CLASH-VLT survey, we targeted 21 lensed sources, including some belonging to the multiple images systems from \citet{Zitrin2015}, which were bright enough for VIMOS follow up. In Tab.\,\ref{tab:lensed_images} in the Appendix we list the redshift measurement for all the lensed features: 10 objects have secure redshift measurement (QF=3), nine sources have likely redshift estimate (QF=2) and for two objects the $\rm z_{sp}$ are based on a strong emission line (QF=9). We obtain spectroscopic measurement for 5 systems of multiple images (systems 1 to 4 and system 8). In addition system 6 has been spectroscopically confirmed and analyzed in detail in \citet{Huang2016}. These spectroscopic measurements will be used as constraints in our strong lensing analysis.
Two further systems of multiple images are used in the SL modeling (system 5 and 7): these are identified by \citet{Zitrin2015} and have no spectroscopic measurement. The photometric redshifts of the multiple images associated to these systems are derived using the CLASH\footnote{available at http://archive.stsci.edu/prepds/clash/} photometry with the SED fitting code \texttt{Bayesian Photometric Redshift, BPZ} (see Tab.\,\ref{tab:lensed_images} and Fig. \ref{fig:dan_phot}). In the lensing analysis the mean $\rm z_{ph}$ for each system is used as starting value for the source redshift $\rm z_s$, and it is optimized with gaussian prior using as width the 68\% uncertainties on the photometric redshift estimates. \\
In the following we give a short description of each lensed system. \\
\textbf{System 1.} A peculiar six times lensed red galaxy appears in the cluster core. Four multiple images are located on the sides of the BCG, and other two tangential images are $\sim16$\arcsec northern and southern the BCG. Spectroscopic measurements of the image 1.1 and 1.5 place this galaxy at $\rm z_{sp}=1.36$. In Fig.\,\ref{fig:spectra1} the 1D and 2D spectra for images 1.1 and 1.5 are shown. The spectrum of this early type galaxy shows a prominent 4000\,\AA\, break and no O[II] line in emission.\\%
\textbf{System 2.}
On the west side of the cluster, at a distance of $\sim14$\arcsec, a tangential arc and its counter image are found.
The arc has a spectroscopic redshift of 1.04 with QF=2. The spectrum for this system is shown in Fig.\,\ref{fig:spectra1} \\
\textbf{System 3.}
At larger distance on the west side, there is a second long arc with respective counter image, which have likely spectroscopic redshift $\rm z_{sp}=2.24$ (QF=2). In Fig.\,\ref{fig:spectra2} we show the 1D and 2D spectra for images 3.1 and 3.3.\\
\textbf{System 4.}
A faint triply lensed source is identified on the eastern side of the cluster, at a distance of $\sim30$\arcsec. The three multiple images are clearly identified being composed by two bright knots. The image 4.3 was targeted with VIMOS, providing a likely source redshift of 2.24 (QF=2), see Fig.\,\ref{fig:spectra2}. \\
\textbf{System 5.}
This is a triply lensed system lying on the east side of the cluster core, at a distance of $\sim20\arcsec$ from the BCG. Photometric redshift estimate place this lensed galaxy at $\rm z_{ph}\sim1.5-2$.\\
\textbf{System 6.}
This system is composed by three optical dropouts with photometric redshift $\rm z_{ph}\sim6.5$. \citealt{Huang2016} have recently published the spectroscopic confirmation for this system using Keck/DEIMOS and HST/WFC3/G102 grism data. A clear Lyman$-\alpha$ emission line places this triply imaged system at $z_{sp}=6.85$. \\
\textbf{System 7.}
This system is another triply lensed galaxy which has photometric redshift estimation of $\rm z_{ph}\sim1.4$\\
In Fig.\,\ref{fig:dan_phot} in the Appendix we show the optical and IR color cutout for the systems 5, 6 and 7, with their respective SED fit and photometric redshift, as available from the public CLASH archive$^{2}$.\\
\textbf{System 8.}
This is a candidate quadruply lensed source at $\rm z_{sp}=4.4$. Image 8.1 was serendipitously discovered in the vicinity of a cluster member targeted with our VIMOS observations. We obtained a spectroscopic redshift of $\rm z_{sp}=4.4$, based on a single emission line, identified as Lyman$-\alpha$ (see Fig.\,\ref{fig:spectra_z4}). We photometrically identified 3 possible counter images of image 8.1 in the HST imaging which are supported by our lensing model, and are thus included in the model itself as constraints.\\
All the systems are summarized in Tab.\,\ref{tab:multiple_images}. In total we have 31 multiple images which provide constraints on our lensing model.
We adopt uncertainties of $1\arcsec$ on the position of all the multiple images (except for system 1, see Sec.\,\ref{sec:point_model} for details), to take into account lens substructures and line of sight structures not considered in our model, which may introduce an offset of 0.5\arcsec-2.5\arcsec on multiple images prediction \citep{Host2012,Jullo2010}.
\section{Strong Lensing Analysis}
\label{sec:point_model}
Given the mass components and the multiple images systems presented in the previous section, we perform the SL analysis of MACS\,2129 by minimizing the distance between the observed and predicted multiple images. This is performed through a standard $\chi^2$ minimization, where the $\chi^2$ is defined as
\begin{equation}
\chi^2_{tot}=\sum\left(\frac{|\vec{\theta}^{pred}_i-\vec{\theta}^{obs}_{i}|}{\sigma^{pos}_i}\right)^2 \,,
\label{eq:chi_q}
\end{equation}
where ${\vec{\theta}^{pred}}_i$ is the predicted position of the $i-$th multiple image, $\vec{\theta}^{obs}_{i}$ is the observed multiple image position and $\sigma^{pos}_i$ is its uncertainty.
We carry out the SL modeling through subsequent iterations, in which we include step by step the lensed systems in the model. \\
First, we perform the minimization analysis using as constraints only the two inner multiple systems with measured spectroscopic redshift (systems 1 and 2). Through this preliminary modeling we keep fixed the b/a and PA of the BCG, G1 and G2. The position angle and axial ratio of these three galaxies are fixed to the values extracted with Sextractor in the filter f814w. The two lensed systems 1 and 2 provide 18 constraints through the positions of the respective multiple images. The free parameters are 18 in total: 6 for the DH, 2 for each of the individually optimized galaxy (GR, BCG, G1 and G2) and 4 for the x- and y- position of the lensed sources. Once we reach a preliminary model with good prediction for the multiple images,
we add the outer spectroscopically confirmed systems 3, 4a and 4b. This increases the number of constraints to 36, while the free parameters are 24.
The inclusion of these outer multiple images affects the accuracy in reproducing the central system 1: the minimization is mainly driven by the tangential multiple images of systems 2, 3 and 4, which place constraints on the large scale DH, to the detriment of the central radial images of system 1. In order to improve the mass model in the very inner region, then we reduce the positional error of system 1 to $\sigma^{pos}_{sys1}=0.5\arcsec$, which gives a higher weight to this system in the $\chi^2_{tot}$ minimization. As a result of this approach, we reach preliminary good prediction for all the spectroscopic confirmed systems, with $\rm \chi^2=16.9/12\,dof$ and median rms 0.5\arcsec.
In the next step we include the multiple systems 5 to 8 with subsequent iterations and repeat the $\chi^2$ minimization. The source redshift for system 5 and 7 is unknown, thus we use their photometric redshift as starting value, and optimize them using a gaussian prior with width given by their respective $\rm z_{ph}$ uncertainties. At this stage we release the b/a and PA of BCG, G1 and G2 and optimize within the range described in the previous section. In addition we also allow the core radii of these three galaxies to vary within [0,3.5] kpc. The total number of degrees of freedom is 27, having finally 62 constraints and 41 free parameters.
Besides the minimization analysis, we run MCMC chains to obtain the model which reproduces with best accuracy the observed multiple images and the uncertainties on our free parameters, i.e. the mass components parameters and the source redshift for systems 5 and 7.
\section{Results and Discussion}
\label{sec:results}
After the minimization and MCMC analyses, our final best mass model of the cluster has $\chi^2_{tot}=29$ having 21 degrees of freedom. It reproduces the multiple images with a median accuracy of $0.4^{\prime\prime}$.
In Fig.\ref{fig:rms} we show the histogram of the offsets $\delta\theta$ between the observed and predicted images, which are listed in Tab.\ref{tab:multiple_images}.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{rms_histo.png}
\caption{\small Histogram of the offsets $\delta\theta$ between the observed position of the multiple images and the predicted ones from our best SL model. Units are arcseconds. The vertical dashed line mark the median value of $\delta\theta$.}
\label{fig:rms}
\end{figure}
In the inner region, inside the Einstein parameter of the cluster, the lensed systems 1, 2, 5 and 7 are reproduced with high accuracy, having median $\delta\theta=0.4\arcsec$. Only the central system 8, which is a candidate quadruply lensed galaxy, has median $\delta\theta=1.3\arcsec$. The images 8.1, which is spectroscopically confirmed at redshift $\rm z_{sp}=4.4$, is reproduced with a small offset of 0.5\arcsec, whereas the three candidate counter images, photometrically selected, are predicted with accuracy of 1.3\arcsec. At larger radii ($\rm r\gtrsim30\arcsec$), we find that the two systems on the west of the cluster (system 3 and 6) are reproduced with median $\delta\theta=1.7\arcsec$, while on the east side, System 4 is reproduced with high accuracy of $0.4\arcsec$.
Based on the symmetry of the model (see Fig.\,\ref{fig:cl_sys1}), we can exclude that the particular shallowness of the DH profile is affecting the lensing prediction at large radii. Indeed, if this is the case, we would expect large $\delta\theta$ also for System 4. Thus we conclude that the larger offset $\delta\theta=1.7\arcsec$ for the prediction of system 3 and 6 is likely related to lens substructures or density inhomogeneities along the line of sight, which are not taken into account in our mass model. However our photometric dataset does not reveal any apparent intervening structure in this region.\\
\begin{figure}
\centering
\includegraphics[width=9cm]{1.png}
\caption{\small RGB color composite image of MACS\,2129 with overplotted the critical lines (in blue) for a source at the redshift $z_s=1.36$. The multiple images used to constrain the cluster mass profile are labeled with squares in different colors (see Tab.\ref{tab:lensed_images}). The axis are in arcseconds with respect to the BCG position }
\label{fig:cl_sys1}
\end{figure}
In Tab.\,\ref{tab:mass_param} we provide the final values obtained for the cluster mass components parameters. \\
The cluster DH has central position very close to the BGC, with $\rm x_{DH}=-2.0\arcsec\pm0.7\arcsec$ and $\rm y_{DH}=1.5\arcsec\pm0.3\arcsec$ with respect to the BCG. It is highly elongated along the WE direction, with axis ratio $b/a=0.32\pm0.03$ and position angle PA$=-7^{\circ}\pm1^{\circ}$, measured counterclockwise from the west direction. The Einstein parameter is $\rm \Theta_e=29\pm4$ arcseconds, which corresponds to an Einstein radius of $\rm \theta_{E,sys1}=14\pm2$ arcseconds for a source at the redshift of system 1, $\rm z_{sys1}=1.36$. The DH has a quite large core radius of $\rm r_c= 101_{-11}^{+13}$ kpc.
Prediction of large DH core radii for galaxy clusters is not an unusual result in strong lensing analyses \citep[see e.g. ][]{ Richard2010a, Johnson2014, Jauzac2014, Jauzac2015, Grillo2014}.
In the innermost region of the cluster, lensing probes the total mass of the cluster core. This is given by the superposition of the PIEMD profile describing the large scale DH and the dPIE profile which describes the baryonic mass component of the BCG. Obviously the core of the large scale DH profile correlates with the baryonic mass distribution of the BCG. The amplitude of the BCG stellar mass profile is described by a \textit{central velocity dispersion $\rm\sigma_{0,BCG}^{dPIE}$} parameter, for the dPIE profile. This does not correspond to the spectroscopic measured central stellar velocity dispersion of the galaxy $\rm\sigma_{sp}^{*}$, since in this case $\rm\sigma_{0,BCG}^{dPIE}$ simply provide an estimate of the mass amplitude of the BCG stellar component, and is not related to the dynamical velocity dispersion of the galaxy. This can be seen in \citet{Newman2013a, Newman2013b} where, by combining lensing and kinematic analyses, they derive the amplitude $\rm\sigma_{0,BCG}^{dPIE}$ of the BCG stellar mass profile for several galaxy clusters, and the results indeed depart from the spectroscopic measurement of the central velocity dispersion of the BCGs. E.g., for the BCG of Abell\,611 they get $\rm\sigma_{0,BCG}^{dPIE}=164\pm33\,km/s$ for the baryonic dPIE profile, whereas the spectroscopic measurement of the central velocity dispersion provides $\rm\sigma_{BCG, sp}^{*}=317\pm20\,km/s$.
\citealt{Sand2002, Sand2004} and \citealt{Newman2013a, Newman2013b} show the utility to combine the BCG stellar kinematic analysis with strong lensing to disentangle the DH and the stellar mass components in the innermost region of galaxy clusters.
\citealt{Newman2013b} separate the baryonic and dark matter components in seven galaxy clusters. They find that the size and mass of the BCG correlate with the core radii of the DH mass profile, and that larger BCG are hosted by clusters with larger core radii. However they obtain typical DH core radii $\rm r_c\approx 10-20 kpc$ for the clusters of their sample. Thus the large core radius resulting in our analysis may reflect an overestimate of the mass associated to the BCG component.
Indeed from our analysis we get $\rm\sigma_{0,BCG}^{dPIE}=370\pm25\,km/s$ as amplitude of the BCG stellar mass profile, which is rather high when compared to the results from \citet{Newman2013a, Newman2013b} for the BCG stellar profiles. In order to proper disentangle the large scale DH and the BCG baryonic mass components in the core of the cluster, the lensing analysis alone is not enough, and additional constraints, as kinematics, are needed. In Fig. \ref{fig:dh_bcg} we plot the DH core radius versus the dPIE velocity dispersion of the BCG stellar mass profile for our mass model to show the correlation between this two parameters. In Fig.\,\ref{fig:mass_prof} we also show the mass profile for both this mass components. Given the total mass profile of the cluster robustly probed by the lensing analysis, we can see that a higher BCG stellar mass profile would imply a shallower profile for the DH, i.e. a larger DH core.
In addition we verified that, if we did not use a separate mass profile to describe the BCG and the inner most cluster mass would have been modeled by a PIEMD profile alone, then we would obtain a cluster core radius as small as $\rm \approx 5 kpc$, which places a lower limit to the DH core. \\
\begin{figure}
\centering
\includegraphics[width=8cm]{dhcore_bcg.png}
\caption{\small Large scale DH core radius versus the velocity dispersion of the dPIE mass profile describing the cluster BCG. The gray scale correspond to the $1\sigma$ (black), $2\sigma$ (dark gray) and $3\sigma$ (light gray) confidence level region.}
\label{fig:dh_bcg}
\end{figure}
Given our final best model of the cluster, we extract the projected mass of the different cluster components within the Einstein parameter. The projected mass of the large scale DH is constrained with 8\% precision, with $\rm M_{DH}=(8.6\pm0.6)\times10^{13}\rm{M_\odot}$ within $\rm \Theta_{E}=29\pm4$\arcsec. The BCG has a projected baryonic mass of $\rm M_{BCG}(<\Theta_E)=(8.4\pm2)\times10^{12}\rm{M_\odot}$ .\\
The reference galaxy GR has central velocity dispersion $\sigma_{0,GR}=185_{-16}^{+ 20}$ km/s and truncation radius $\rm r_{\rm tr, GR}= 66_{- 28 }^{+ 22}$ kpc, providing the following cluster scaling relation:
\begin{equation}
\rm r_{\rm tr}=66_{- 28 }^{+ 22} kpc\left(\frac{\sigma}{185_{-16}^{+ 20}\,km/s}\right)^\frac{4}{3}\,.
\label{eq:scaling_rel}
\end{equation}
The other two galaxies individually optimized, G1 and G2, get velocity dispersion with $1\sigma$ errors of $\lesssim20\%$, but their truncation radii have very large uncertainties
The total projected mass of the other cluster galaxies is constrained at level of $12\%$, with $\rm M_{gal}=(4.4\pm0.5)\times10^{13}\rm{M_\odot}$ within the Einstein parameter $\rm \Theta_{E}$. \\
Overall the joint projected cluster mass of the large scale DH, BCG and the galaxy component is $\rm M_{tot}( <\Theta_E)=(1.35\pm0.03)\times10^{14}\rm{M_\odot}$ and it is constrained with $2\%$ precision.
In Figure \ref{fig:mass_prof} we show the 2D projected mass profiles of the DH, BCG, the galaxy component and the total mass of the cluster core up to a radial distance of $50\arcsec$ from the cluster center. \\
\citet{Zitrin2015} published a first model of MACS\,2129 using the CLASH dataset. They obtain a total projected mass of $\rm M_e=(9.2\pm0.9)\times10^{13}M_{\odot}$ within the Einstein radius $\theta_{E}=19\arcsec\pm2\arcsec$ for a source at redshift $\rm z_s=2$. Within the same radius we extract a total mass of $\rm M_{tot}(<19\arcsec)=(8.9\pm0.1)\times10^{13}M_{\odot}$, which is fully consistent with the previous analysis from \citet{Zitrin2015}. We highlight that the larger uncertainties in \citet{Zitrin2015} account also for systematics, derived by modelling the DH cluster with different analytic mass profiles.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{mass_plot1.png}
\caption{\small Projected 2D mass profile at 68\% confidence level of the core of MACS\,2129. In red we plot the BCG stellar mass profile, in blue we plot the total mass of the other cluster members, in gray the cluster DH, and in black the total mass profile.}
\label{fig:mass_prof}
\end{figure}
\begin{table}
\centering
\caption{\small Final parameters of the mass components of MACS\,2129 with their respective $1\sigma$ uncertain.
The DH position (x,y) is given in arcseconds with respect to the BCG. The core and truncation radii are in kpc, velocity dispersions are in km/s and PA are in degrees measured counterclockwise from the west direction.}
\footnotesize
\begin{tabular}{|l|c|}
\hline
\hline
DH & \\
\hline
x & $ -2.0\pm0.7$\\
y & $ 1.5 \pm0.3$\\
b/a & $ 0.32 \pm0.03$\\
PA & $ -7\pm1$\\
$\Theta_e (")$ & $ 29 \pm4$\\
$\rm r_{c}$& $ 101 _{- 11 }^{+ 13 }$\\
\hline
GR & \\
\hline
$\sigma_0$& $185_{-16}^{+20}$\\
$\rm r_{\rm tr}$ & $ 66 _{- 28 }^{+ 22 }$\\
\hline
BCG &\\
\hline
b/a & $ 0.78 \pm0.10$\\
PA & $ 2.5\pm2.8$\\
$\sigma_0$& $370\pm25$\\
$\rm r_{c}$ & $ 1.1 _{- 0.8}^{+ 1.2 }$\\
$\rm r_{\rm tr}$ & $ 87 _{- 35 }^{+ 32 }$\\
\hline
G1 & \\
\hline
b/a & $ 0.86\pm0.05$\\
PA & $ 54\pm3$\\
$\sigma_0$& $224_{-38}^{+52}$\\
$\rm r_{c}$ & $1.6 _{- 1.0 }^{+ 1.1 }$\\
$\rm r_{\rm tr}$ & $ 57 _{- 37 }^{+ 30 }$\\
\hline
G2 & \\
\hline
b/a & $ 0.80\pm0.05$\\
PA & $ 130\pm3$\\
$\sigma_0$& $255\pm28$\\
$\rm r_{c}$ & $ 1.3_{- 1.0}^{+ 1.3 }$\\
$\rm r_{\rm tr}$ & $ 25_{- 21 }^{+ 28 }$\\
\hline
\hline
\end{tabular}\\
\label{tab:mass_param}
\end{table}
\section{Summary and Conclusions}
\label{sec:conclusions}
In this paper we presented a precise strong lensing analysis of the massive cluster of galaxies MACS\,2129. We combine the CLASH high resolution photometry with the new VLT spectroscopy from the CLASH-VLT surveys to robustly select cluster members and multiple images needed for the SL modeling. This cluster shows a peculiar sextuply lensed red passive galaxy in its core, which has spectroscopic redshift $\rm z_{sp}=1.36$.
Six additional multiple lensed sources are clearly identified in the cluster core. Our CLASH-VLT spectroscopic data provide spectroscopic redshift confirmation for three of these lensed systems.
We use these multiple images to constrain the mass profile in the inner most region of the cluster with very high precision. Our best mass reconstruction of the cluster reproduces the observed multiple images with a median accuracy of $0.4\arcsec$.
The overall high accuracy in predicting the position of the multiple images directly translate to a high precision reconstruction of the total mass distribution of the cluster.\\
The cluster DH projected mass is constrained with 8\% precision, with $\rm M_{DH}=(8.6\pm0.6)\times10^{13}\rm M_{\odot}$ within the Einstein parameter $\Theta_E=29\arcsec\pm4\arcsec$. The DH has a large core radius $\rm r_c=101_{-11}^{+13} kpc$, similar to other galaxy cluster \citep{Richard2010, Johnson2014, Jauzac2014, Jauzac2015, Grillo2014}. However, the DH core radius correlates with the BCG mass profile. When the BCG baryonic and the cluster DH components are properly disentangled by combining lensing and kinematic analyses, typical smaller core radii $(\rm r_c\approx 10-20 kpc)$ are found for cluster DHs, see \citealt{Newman2013a, Newman2013b}. Thus the large core radius we derived likely reflects an overestimate of the mass associated to the BCG mass component. Nevertheless the final 2D projected total mass of the cluster is constrained with an accuracy of 2\%. Within the Einstein parameter it is $\rm M_{tot}=(1.35\pm0.03)\times10^{14}\rm M_{\odot}$. \\
In addition to the mass model for the core of MACS\,2129, we also provide the complete list of robust spectroscopic redshifts of cluster members and lensed features measured with VLT/VIMOS within the CLASH-VLT survey, see the Appendix \ref{sec:appendix}.
\section*{Acknowledgements}
This work is supported by the Transregional Collaborative Research Centre TRR 33 - The
Dark Universe and the DFG cluster of excellence ``Origin and Structure of the Universe".
The CLASH Multi-Cycle Treasury Program (GO-12065) is based on observations made with the NASA/ESA Hubble Space Telescope. The Space Telescope Science Institute is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract NAS 5-26555.
SHS gratefully acknowledges support from the Max Planck Society through the Max Planck Research Group.
P.R. and A. Mercurio acknowledge support
from PRIN-INAF 2014 1.05.01.94.02 (PI M. Nonino).
P.R. acknowledges the hospitality and support of the visitor program of the DFG cluster of excellence ``Origin and Structure of the Univers''.
C.G. acknowledges support by VILLUM FONDEN Young Investigator Programme through grant no. 10123.
\addcontentsline{toc}{chapter}{Bibliography}
\bibliographystyle{mn2efix}
|
\section{Introduction}
The motivation behind this paper is a basic question, due to S.-S. Chern and J. K. Moser \cite{CM74}, which roughly asks if there are compact, strictly pseudoconvex CR manifolds of dimension three that do not possess (CR) umbilical points. The answer to the question, in this generality, is well known to be 'yes'. E. Cartan had much earlier \cite{Cartan33} discovered a 1-parameter family of real hypersurfaces $\mu_\alpha\subset {\4P}^2$ that are compact, strictly pseudoconvex, homogeneous and non-spherical (hence without umbilical points). A family of 2:1-covers of these CR manifolds are embeddable in ${\4C}^3$. The universal cover of $\mu_\alpha$ is the sphere $S^3$, and by pulling back the CR structures of $\mu_\alpha$ one obtains a family of CR structures on $S^3$ that do not possess any umbilical points. The latter structures, however, are well known to not be embeddable in ${\4C}^n$ for any $n$. (See \cite{Isaev06}, \cite{Rossi65}.) A precise question that remains to be answered is then the following:
\begin{Que}\Label{CMQ} Does there exist a compact, strictly pseudoconvex CR manifold $M$ in ${\4C}^2$ that does not have any umbilical points?
\end{Que}
We point out the analogy with the classical Caratheodory Conjecture (see, e.g., \cite{Hamburger40}, \cite{Ivanov02}) regarding umbilical points on compact surfacces embedded in ${\4R}^3$, and refer the interested reader to the paper \cite{EDumb15} for a closer discussion of the analogy between the notion of (CR) umbilical points in CR geometry and that of umbilical points in the classicial geometry of surfaces in ${\4R}^3$.
Let $M=M^{2n+1}$ be a real hypersurface in ${\4C}^{n+1}$ and $p\in M$. The lowest order (local) invariant of $M$ is its Levi form at $p$, which roughly speaking is a Hermitian form on the tangent space $T^{1,0}_pM$ of $(1,0)$-vectors tangent to $M$ at $p$. A hypersurface is said to be Levi nondegenerate at $p$ if the Levi form is nondegenerate at $p$, and strictly pseudoconvex if the Levi form at $p$ is definite. Levi nondegeneracy of $M$ at $p$ can be detected by the condition $J_p\neq 0$, where $J=J(\rho)$ is
Fefferman's Monge-Ampere operator \cite{Fefferman76}
\begin{equation}\Label{FeffJ}
J(\rho):=(-1)^{n+1} \det
\begin{pmatrix}
\rho & \rho_{\bar Z}\\
\rho_{Z} &\rho_{Z\bar Z}
\end{pmatrix}
\end{equation}
applied to a local defining function $\rho$ of $M$ near $p$; i.e., $M$ is locally given by $\rho=0$, and $d\rho\neq 0$ on $M$. In \eqref{FeffJ}, the notation used is $\rho_{\bar Z}:=(\rho_{\bar Z_1},\ldots,\rho_{\bar Z_{n+1}})$, $\rho_{Z}$ is its conjugate transpose $\rho_Z=(\rho_{\bar Z})^*$, and $\rho_{Z\bar Z}$ is the $(n+1)\times(n+1)$ matrix $(\rho_{Z_k\bar Z_j})$. One of the main contributions in this paper is the introduction of a sequence of invariant determinants, for hypersurfaces in ${\4C}^2$, that can be viewed as higher order analogs of Fefferman's operator $J$ in \eqref{FeffJ}.
Chern and Moser showed \cite{CM74} that if $M$ is strictly pseudoconvex at $p$, then that there are formal holomorphic coordinates $Z=(z,w)=(z_1,\ldots,z_n,w)$ (convergent if $M$ is real-analytic), vanishing at $p$, such that $M$ can be expressed in Chern--Moser normal form. Rather than describing this normal form precisely here, we simply note that in these coordinates $M$ is expressed as a graph \begin{equation}
{\sf Im}\, w=\varphi(z,\bar z,{\sf Re}\, w),
\end{equation}
where the graphing function has the form
\begin{equation}\Label{CMnormform}
\varphi(z,\bar z,{\sf Re}\, w)=\sum_{j=1}^n|z_j|^2+R_{m}(z,\bar z)+O(m+1).
\end{equation}
Here, $R_{m}(z,\bar z)$ is a homogeneous Hermitian polynomial of degree $m$ with $m=6$ for $n=1$ and $m=4$ for $n\ge 2$, $O(k)$ signify terms of weight $\geq k$ in $(z,\bar z, {\sf Re}\, w)$, where $(z,\bar z)$ are assigned weight one and ${\sf Re}\, w$ weight two. If $n\geq 2$, then $m=4$ and $R_4(z,\bar z)$ is of bidegree $(2,2)$; $R_4(z,\bar z)$ represents the CR curvature tensor $S_{\alpha\bar\beta\nu\bar\mu}$ at $p=(0,0)$ as its sectional curvature
\begin{equation}\Label{R_4}
R_4(z,\bar z)=\sum_{\alpha,\beta,\nu,\mu}S_{\alpha\bar\beta\nu\bar\mu}z_\alpha z_\nu\overline{z_\beta z_\mu}.
\end{equation}
The CR curvature tensor $S_{\alpha\bar\beta\nu\bar\mu}$ has Hermitian curvature symmetries and zero trace. (The latter is equivalent to $R_4(z,\bar z)$ being a harmonic function of $z$.) The real dimension of the space of such curvature tensors (of Weyl--Bochner type) is $n^2(n-1)(n+3)/3$ (see \cite{sitaramayya73}). From this it follows that the condition of being umbilical at a point $p$ is equivalent to $n^2(n-1)(n+3)/3$ independent real equations on the 4-jet of the CR manifold $M^{2n+1}$ at $p$. For $n=2$, this means 5 equations on a 5-dimensional manifold, but for $n\geq 3$, this is an "overdetermined" system. More precisely, an application of Thom's Transversality Theorem \cite{GG86} (see \cite{BRZ07} for similar arguments) shows that a {\it generic} strictly pseudoconvex hypersurface $M^{2n+1}\subset {\4C}^{n+1}$ (i.e., in a dense open subset in the compact-open topology of $C^\infty$-mappings $M^{2n+1}\hookrightarrow{\4C}^{n+1}$) has no umbilical points when $n\geq 3$, and at most isolated umbilical points when $n=2$. Moreover, to further support the statement that umbilical points are rare when $n\geq2$, we mention that Webster has shown \cite{Webster00} that every non-spherical real ellipsoid in ${\4C}^{n+1}$ is free of umbilical points when $n\geq 2$.
In this paper, we shall consider the case $n=1$, i.e., strictly pseudoconvex hypersurfaces $M=M^3$ in ${\4C}^2$. In this case, the CR curvature vanishes identically, and the lowest order nontrivial invariant in the Chern--Moser normal form occurs in weight $m=6$, and $R_6(z,\bar z)$ in \eqref{CMnormform} has the form
\begin{equation}\Label{R_6}
R_6(z,\bar z)=c_{24}z^2\bar z^4+c_{42}z^4\bar z^4,\quad c_{42}=\overline{c_{24}}\in {\4C},
\end{equation}
where $c_{24}$ represents E. Cartan's "6th order tensor" $Q=Q^1{}_{\bar 1}$ at $p=(0,0)$. The hypersurface $M^3$ is umbilical at $p=(0,0)$ if $c_{24}=0$. Thus, the condition of being umbilical on $M^3$ amounts to two independent real equations in the 6-jet space of CR manifolds in ${\4C}^2$. Thus, we should expect umbilical points (if they exist!) to form real curves in $M^3$. If we consider the condition that $M^3$ is umbilical at $p$ and the rank of the differential of the two real equations $c_{24}=0$ is $\leq 1$, which amounts to the vanishing of two real determinants, then we obtain an algebraic subvariety in the 7-jet space of codimension $4$. An application of Thom's Transversality Theorem, as above, yields that a {\it generic} strictly pseudoconvex hypersurface $M^3\subset {\4C}^2$ either has no umbilical points at all or a set of umbilical points that consists of smooth real curves.
\subsection{Summary of Main Results.} The major problem addressed in this paper is that of existence of umbilical points. As is illustrated by the family of examples $\mu_\alpha\subset{\4P}^2$ and their covers, there are compact three dimensional CR manifolds without umbilical points. We shall focus here on Question \ref{CMQ} described above. Not much is known in general about this problem. It was shown by X. Huang and S. Ji \cite{HuangJi07} that every real ellipsoid in ${\4C}^2$ must have umbilical points. More recently, it was proved by the first author and S. Duong \cite{EDumb15} that every {\it circular} $M^3\subset {\4C}^2$ has umbilical points; in fact, it was proved in \cite{EDumb15} that every compact three dimensional CR manifold with a transverse free CR $U(1)$ (circle) action must have umbilical points provided that the Riemann surface $M/U(1)$ has genus $g\neq 1$. As a first step towards answering Question \ref{CMQ} more generally, we shall consider small perturbations $M_\varepsilon\subset {\4C}^2$ of the unit sphere $M_0=S^3\subset {\4C}^2$. We shall show that generic {\it almost circular} perturbations $M_\varepsilon$ must possess umbilical points. This is Theorem \ref{ACpertThm}. Precise statements and definitions are given in Section \ref{PertSec}. We also note in Remark \ref{ACrem} that real ellipsoids are not generic in the sense of Theorem \ref{ACpertThm}, but we give separately a new proof of a special case of the Huang-Ji theorem that real ellipsoids possess umbilical points; namely, we prove the existence of real curves of umbilical points in the special case where the ellipsoids are sufficiently close to spherical. This is Theorem \ref{Prop-ell}.
One of the main obstacles in investigating the existence of umbilical points on a compact real hypersurface $M^3\subset{\4C}^2$ is the lack of a global convenient representation ("formula") of Cartan's tensor $Q$. In Chern--Moser normal form at $p=(0,0)$, $Q$ is represented at $p$ by the coefficient $c_{24}$. Loboda \cite{Loboda97} discovered a "semi-global" formula that represents $Q$ for a graph $M^3\subset {\4C}^2$ provided that $M^3$ also has additional tranverse symmetry. One of the main results in this paper is a {\it global} representation of $Q$ as a nonlinear PDO acting on a defining function $\rho$ for $M^3\subset {\4C}^2$. In fact, a sequence of invariants $\det A_k(\rho)$, for $k\geq 3$, is introduced in Section \ref{SecInv} (see in particular Theorem \ref{transform}). These invariant determinants are in some sense higher order versions of Fefferman's Monge-Ampere operator $J(\rho)$. If we introduce the $(1,0)$-vector field
$$
L:=-\frac{\partial \rho}{\partial w}\frac{\partial}{\partial z}+\frac{\partial \rho}{\partial z}\frac{\partial}{\partial w}=-\rho_w\partial_z+\rho_z\partial_w,
$$
then we have, as the reader can easily check,
\begin{equation}\Label{FeffJ-2}
J(\rho)=\det
\begin{pmatrix}
\rho_z & \bar L\rho_z\\
\rho_w & \bar L \rho_w
\end{pmatrix} \mod \rho.
\end{equation}
Our invariant determinants $\det A_k(\rho)$, $k\geq 3$ are higher order analogues of the determinant on the right in \eqref{FeffJ-2}. The precise definition is given in \eqref{an}. The relationship between $\det A_k(\rho)$ and the classical Chern--Moser invariants of $M^3$ is explored in Section \ref{SecCM}. This relationship is expressed by \eqref{detAnCM}; in particular, it is shown that $\det A_3(\rho)$ represents Cartan's tensor $Q$ at every $p\in M^3$.
\section{A family of invariant determinants}\Label{SecInv}
Consider a real hypersurface $M=M^3\subset{\4C}^2$ given by $\rho(z,w,\bar z,\bar w)=0$
with $\partial\rho\ne 0$ on $M$. We use coordinates $Z=(z,w)\in {\4C}^2$.
Then the space of $(1,0)$-vectors on $M$ at every point is spanned by
\begin{equation}\Label{l}
L: = - \rho_w \partial_z + \rho_z \partial_w.
\end{equation}
We shall also use the Hessian $\rho_{Z^2}$ evaluated at $(L,L)$:
$$\rho_{Z^2}(L,L) = \rho_{zz} \rho_2^2 - 2\rho_{zw} \rho_z\rho_w + \rho_{w^2} \rho_z^2. $$
For every $n\ge 3$, consider the $(2n-1)\times(2n-1)$ matrix PDO $A_n(\rho)$ acting on the smooth function $\rho$
\begin{equation}\Label{an}
A_n=A_n(\rho):=
\begin{pmatrix}
\bar L^j (\rho_z^k \rho_w^{n-k})\cr
\bar L^j (\rho_z^s \rho_w^{n-3-s} \rho_{Z^2}(L, L))
\end{pmatrix}_{0\le j\le 2n-2,\, 0\le k\le n, \, 0\le s\le n-3},
\end{equation}
where we regard $j$ as column index and $k$ and $s$ as row indices (first followed by the second). In particular, for $n=3,4$, we obtain
\begin{equation}\Label{a3}
A_3=A_3(\rho):=
\begin{pmatrix}
\rho_w^3 & \bar L(\rho_w^3) & \cdots & \bar L^4(\rho_w^3)\cr
\rho_z\rho_w^2 & \bar L( \rho_z\rho_w^2) & \cdots & \bar L^4( \rho_z\rho_w^2)\cr
\rho_z^2\rho_w & \bar L( \rho_z^2\rho_w) & \cdots & \bar L^4( \rho_z^2\rho_w)\cr
\rho_z^3 & \bar L(\rho_z^3) & \cdots & \bar L^4(\rho_z^3)\cr
\rho_{Z^2}(L, L) & \bar L(\rho_{Z^2}(L, L)) & \cdots & \bar L^4(\rho_{Z^2}(L, L))
\end{pmatrix}
\end{equation}
and
\begin{equation}\Label{a4}
A_4=A_4(\rho):=
\begin{pmatrix}
\rho_w^4 & \bar L(\rho_w^4) & \cdots & \bar L^6(\rho_w^4)\cr
\rho_z\rho_w^3 & \bar L( \rho_z\rho_w^3) & \cdots & \bar L^6( \rho_z\rho_w^3)\cr
\rho_z^2\rho_w^2 & \bar L(\rho_z^2\rho_w^2) & \cdots & \bar L^6(\rho_z^2\rho_w^2)\cr
\rho_z^3\rho_w & \bar L( \rho_z^3\rho_w ) & \cdots & \bar L^6( \rho_z^3\rho_w )\cr
\rho_z^4 & \bar L(\rho_z^4) & \cdots & \bar L^6(\rho_z^4)\cr
\rho_w\rho_{Z^2}(L, L) & \bar L(\rho_w\rho_{Z^2}(L, L)) & \cdots & \bar L^6(\rho_w\rho_{Z^2}(L, L))\cr
\rho_z\rho_{Z^2}(L, L) & \bar L(\rho_z\rho_{Z^2}(L, L)) & \cdots & \bar L^6(\rho_z\rho_{Z^2}(L, L))
\end{pmatrix}.
\end{equation}
We also denote by $D_n=D_n(\rho)$ the upper left $(n+1)\times (n+1)$ minor of $A_n$, i.e.\
\begin{equation}\Label{dn}
D_n:=
\begin{pmatrix}
\bar L^j (\rho_z^k \rho_w^{n-k})
\end{pmatrix}_{0\le j\le n, \, 0\le k\le n}.
\end{equation}
The main interest in considering the matrices $A_n$ and $D_n$ is the following invariance property of their determinants:
\begin{Thm}\Label{transform}
For every $M$ and $n\ge 3$, the properties $\det A_n=0$ and $\det D_n=0$ at points of $M$
are independent of the choice of the defining function $\rho$ as well as of
the choice of the coordinates $Z=(z,w)\in {\4C}^2$.
More precisely, if $L^*$, $A^*_n$ and $D^*_n$ are given by \eqref{l}, \eqref{an} and \eqref{dn} respectively
with $\rho$ replaced by another defining function $\rho^*=a\rho$ (where $a$ is any nonzero real smooth function),
and $Z=(z,w)$ replaced by another (formal) holomorphic coordinate system $Z^*=(z^*,w^*)$,
we have the transformation rule
\begin{equation}
\Label{a-transform}
\delta^{n^2-1} \bar \delta^{(n-1)(2n-1)}\det A^*_n = a^{(2n-1)^2} \det A_n,
\end{equation}
\begin{equation}
\Label{d-transform}
|\delta|^{n(n+1)} \det D^*_n = a^{\frac{3n(n+1)}2} \det D_n,
\end{equation}
where $\delta$ is the Jacobian determinant of the coordinate transformation $Z^*=H(Z)$.
\end{Thm}
\begin{Rem}
It is important that $L$ and $\rho$ used in $A_n$ are related via \eqref{l}.
The invariance of the property $\det A_n=0$
does not hold for arbitrary choices of $(1,0)$ vector fields $L$.
However, given that chosen $L$, the invariance of $\det A_n$ remains
when replacing $\bar L$ by arbitrary $(0,1)$-vector fields.
\end{Rem}
To prove Theorem \ref{transform}, we require two lemmas.
\begin{Lem}\Label{a-lemma}
For any real smooth function $a=a(Z,\bar Z)$, and $L$ and $L^*$ given respectively by \eqref{l} and by the same formula with $\rho$ replaced with $\rho^*=a\rho$, we have on $M$ the identities
\begin{equation}\Label{a-transform'}
\rho^*_Z = a\rho_Z,
\quad L^* = a L,
\quad \rho^*_{Z^2}(L^*,L^*) = a^3 \rho_{Z^2}(L, L).
\end{equation}
\end{Lem}
\begin{proof}
By Leibnitz' rule on $M$ we have $\rho^*_Z = a \rho_Z $, which implies the first and second identities in \eqref{a-transform'}. By Leibnitz' rule again, for any $(1,0)$ vectors $\xi, \eta$, we also have
$$(a\rho)_{Z^2}(\xi, \eta)
= a\rho_{Z^2}
(\xi,\eta)
+ a_Z(\xi) \rho_Z(\eta)
+ a_Z(\eta) \rho_Z(\xi)
+ a_{Z^2}(\xi,\eta) \rho.
$$
Substituting $\xi=\eta=L$ and using the properties $\rho_Z(L)=0$ and $\rho=0$ on $M$,
we obtain, on $M$,
\begin{equation}\Label{rho-L}
\rho^*_{Z^2}(L, L) = a \rho_{Z^2} (L, L).
\end{equation}
Together with the second identity in \eqref{a-transform'} this yields the third identity.
\end{proof}
\begin{Lem}\Label{coord}
For any (formal) biholomorphic transformations $Z^*=H(Z)$, $\zeta^*=K(\zeta)$ of ${\4C}^2$, any complex formal power series $\rho^*(Z^*, \bar Z^*)$
and $\rho(Z,\bar Z):= \rho^*(H(Z), K(\bar Z))$,
consider $L$, $A_n$, $D_n$ and $L^*$, $A^*_n$, $D^*_n$ given by \eqref{l}, \eqref{an}, \eqref{dn} and respectively
by the same identities with $\rho$ replaced by $\rho^*$.
Then the following hold:
\begin{enumerate}
\item[(i)] The identity
\begin{equation}\Label{Z2}
\rho_{Z^2}(L,L) = (\det H_Z)^2 \rho^*_{Z^{*2}} (L^*, L^*)
\end{equation}
holds modulo a cubic homogeneous polynomial in $(\rho_z, \rho_w)$ with holomorphic coefficients in $Z$.
\item[(ii)] The determinants of $A_n$, $D_n$ and $A^*_n$, $D^*_n$ are related by
\begin{equation}
\det A_n = (\det H_Z)^{n^2-1}(\det K_{\bar Z})^{(n-1)(2n-1)} \det A^*_n,
\end{equation}
\begin{equation}
\det D_n = (\det H_Z)^{\frac{n(n+1)}2}
(\det K_{\bar Z})^{\frac{n(n+1)}2} \det D^*_n,
\end{equation}
where the matrices $A^*_n$ and $D^*_n$ are evaluated at $(Z^*,\bar Z^*)= (H(Z), K(\bar Z))$.
\end{enumerate}
\end{Lem}
\begin{proof}
We write
$$S_n:=
\begin{pmatrix}
\rho_w^n\cr
\rho_z\rho_w^{n-1}\cr
\vdots\cr
\rho_z^n
\end{pmatrix}
$$
for the standard basis of homogeneous monomials of order $n$ in $(\rho_z, \rho_w)$.
In particular, $S_n$ coincides with the
$(n+1)\times 1$ matrix (column vector) consisting of the first $n+1$ entries of the first column of $A_n$.
We also write $S^*_n$ for corresponding column of monomials in $(\rho^*_{z^*}, \rho^*_{w^*})$.
In particular,
$$
S_1=
\begin{pmatrix}
\rho_w\cr
\rho_z
\end{pmatrix},
\quad
S^*_1=
\begin{pmatrix}
\rho^*_{w^*}\cr
\rho^*_{z^*}
\end{pmatrix}.
$$
By the chain rule, we have
$$\rho_Z = \rho^*_{Z^*}\circ H_Z,$$
where $\rho^*_{Z^*}$ is evaluated at $(Z^*,\bar Z^*)=(H(Z), K(\bar Z))$.
Writing as matrix identity we obtain
$$S_1 = H_{Z} S^*_1,$$
where by abuse of notation, for $H=(f,g)$, we identify $H_Z$ with its induced matrix
\begin{equation}\Label{h-matrix}
\begin{pmatrix}
g_w & f_w\cr
g_z & f_w
\end{pmatrix}.
\end{equation}
Then viewing $n$th order homogenous monomials in $\rho_z,\rho_w$ in the $n$th tensor power of the cotangent space, we have
\begin{equation}\Label{S-rel}
S_n = (\otimes^n H_Z) S^*_n,
\end{equation}
where, by another abuse of notation, we identify the tensor power transformation $\otimes^n H_Z$
with its induced matrix in the monomial basis.
Further we have
\begin{equation}\Label{det}
\det (\otimes^n H_Z) = (\det H_Z)^{\frac{n(n+1)}2},
\end{equation}
which e.g. follows from Jordan normal form.
Since $(\otimes^n H_Z)$ is holomorphic in $Z$ and $\bar L$ is a $(0,1)$ vector field, we obtain for any $j$,
\begin{equation}\Label{first-rows}
\bar L^j S_n = (\otimes^n H_Z) \bar L^j S^*_n.
\end{equation}
In particular, we have
\begin{equation}\Label{dn-rel}
\det D_n = (\det H_Z)^{\frac{n(n+1)}2} \det E^*_n,
\end{equation}
where $E^*_n$ is the matrix obtained from $D^*_n$ with $\bar L^*$
being replaced by $\bar L$.
We next turn to the relation between $L$ and $L^*$.
By definition
$$
L=
\begin{pmatrix}
\rho_z & -\rho_w
\end{pmatrix}
\begin{pmatrix}
\partial_w\cr
\partial_z
\end{pmatrix},
\quad
L^*=
\begin{pmatrix}
\rho^*_{z^*} & -\rho^*_{w^*}
\end{pmatrix}
\begin{pmatrix}
\partial_{w^*}\cr
\partial_{z^*}
\end{pmatrix}.
$$
We further have
$$
\begin{pmatrix}
H_Z \partial_w \cr
H_Z \partial_z
\end{pmatrix}
=
\begin{pmatrix}
g_w & f_w\cr
g_z & f_z
\end{pmatrix}
\begin{pmatrix}
\partial_{w^*} \cr
\partial_{z^*}
\end{pmatrix}
= H_Z
\begin{pmatrix}
\partial_{w^*} \cr
\partial_{z^*}
\end{pmatrix},
$$
where we continue our abuse of notation by writing $H_Z$ also for its induced matrix.
Similarly,
\begin{equation}\Label{rho-rel}
\begin{pmatrix}
\rho_w & \rho_z
\end{pmatrix}
=
\begin{pmatrix}
\rho^*_{w^*} & \rho^*_{z^*}
\end{pmatrix}
\begin{pmatrix}
g_w & g_z \cr
f_w & f_z
\end{pmatrix}
=
\begin{pmatrix}
\rho^*_{w^*} & \rho^*_{z^*}
\end{pmatrix}
H_Z^t,
\end{equation}
where $H_Z^t$ is the transpose matrix.
Furthermore,
$$
\begin{pmatrix}
\rho_z & -\rho_w
\end{pmatrix}
=
\begin{pmatrix}
\rho_w & \rho_z
\end{pmatrix}
J,
\quad
J:=
\begin{pmatrix}
0 & -1\cr
1 &0
\end{pmatrix},
$$
and hence
$$
\begin{pmatrix}
\rho_z & -\rho_w
\end{pmatrix}
=
\begin{pmatrix}
\rho^*_{z^*} & -\rho^*_{w^*}
\end{pmatrix}
J^{-1} H_Z^t J.
$$
Putting everything together, we obtain
$$H_Z L =
\begin{pmatrix}
\rho_z & -\rho_w
\end{pmatrix}
\begin{pmatrix}
H_Z \partial_w \cr
H_Z \partial_z
\end{pmatrix}
=
\begin{pmatrix}
\rho_z & -\rho_w
\end{pmatrix}
H_Z
\begin{pmatrix}
\partial_{w^*} \cr
\partial_{z^*}
\end{pmatrix}
=
\begin{pmatrix}
\rho^*_{z^*} & -\rho^*_{w^*}
\end{pmatrix}
J^{-1} H_Z^t J H_Z
\begin{pmatrix}
\partial_{w^*} \cr
\partial_{z^*}
\end{pmatrix}.
$$
By direct calculation,
$$
J^{-1} H_Z^t J H_Z =
\begin{pmatrix}
0 & 1\cr
-1 &0
\end{pmatrix}
\begin{pmatrix}
g_w & g_z \cr
f_w & f_z
\end{pmatrix}
\begin{pmatrix}
0 & -1\cr
1 &0
\end{pmatrix}
\begin{pmatrix}
g_w & f_w \cr
g_z & f_z
\end{pmatrix}
= (\det H_Z)
\begin{pmatrix}
1 & 0\cr
0 &1
\end{pmatrix},
$$
and therefore
\begin{equation}\Label{LL}
H_Z L = (\det H_Z) L^*.
\end{equation}
Now, by the chain rule, for any $(1,0)$ vectors $\xi, \eta$,
$$
\rho_{Z^2}(\xi,\eta) = \rho^*_{Z^{*2}} (H_Z \xi, H_Z \eta) + \rho^*_{Z^*} (H_{Z^2}(\xi, \eta)),
$$
where we recall that $H_{Z^2}(\xi, \eta)$ is a $(1,0)$ vector.
Substituting $\xi=\eta = L$ and using \eqref{LL}, we obtain
$$
\rho_{Z^2}(L,L) = (\det H_Z)^2 \rho^*_{Z^{*2}} (L^*, L^*) + \rho^*_{Z^*}( H_{Z^2}(L,L)).
$$
Expanding the last term and using \eqref{rho-rel}, we obtain the desired identity \eqref{Z2}
modulo a homogeneous polynomial of degree 3 in $(\rho_z,\rho_w)$ with coefficients that are polynomial in the derivatives of $H$, which proves the first statement (i) of the lemma.
We now turn to the $(n-2)\times 1$ matrix (column vector) consisting of the last $n-2$ entries of the first column in $A_n$, which in the notation introduced can be expressed as $\rho_{Z^2}(L,L)S_{n-3}$.
Hence \eqref{S-rel} implies
$$\rho_{Z^2}(L,L) S_{n-3} = \rho_{Z^2}(L,L) (\otimes^{n-3} H_Z) S^*_{n-3}.$$
By the first statement (i) of the lemma, already proved, we have
$$\rho_{Z^2}(L,L) S_{n-3} = (\det H_Z)^2 (\otimes^{n-3} H_Z) \rho^*_{Z^{*2}}(L^*,L^*) S^*_{n-3}$$
modulo a homogeneous polynomial of order $n$ in $(\rho_z, \rho_w)$ with holomorphic coefficients in $Z$.
Since $\bar L$ is $(0,1)$, it commutes with those holomorphic coefficients and, consequently,
we can subtract from the last $n-2$ rows of $A_n$ suitable linear combinations of the first $n+1$ rows
to obtain a matrix with the same determinant of the form
$$
\begin{pmatrix}
\otimes^n H_Z & 0\cr
0 & (\det H_Z)^2 (\otimes^{n-3} H_Z)
\end{pmatrix}
B^*_n,
$$
where $B^*_n$ is the matrix obtained from $A^*_n$ with $\bar L^*$ (but not $L^*$!) replaced by $\bar L$.
Finally, analogously to \eqref{LL}, we have the relation
$$
K_{\bar Z}\bar L = (\det K_{\bar Z}) \bar L^*.
$$
Writing $C^*_n$ for the first column of $B^*_n$,
we obtain for any $j$,
$$(K_{\bar Z}\bar L)^j C^*_n = (\det K_{\bar Z})^j (\bar L^*)^j C^*_n$$
modulo a linear combination of the columns $(\bar L^*)^s C^*_n$ with $s<j$.
Hence, subtracting those linear combinations without changing the determinant, we obtain
$$\det B^*_n = (\det K_{\bar Z})^{(n-1)(2n-1)} \det A^*_n, $$
and similar
$$\det E^*_n = (\det K_{\bar Z})^{\frac{n(n+1)}2} \det D^*_n,$$
where $E^*_n$ was defined after \eqref{dn-rel}.
Putting everything together and using \eqref{det} we obtain the second conclusion.
\end{proof}
\begin{proof}[Proof of Theorem~$\ref{transform}$]
Clearly it suffices to prove the proposition by separately considering changes of the defining function and the coordinates. The transformation formula under a change of coordinates follows from Lemma~\ref{coord} when $\rho$ is real-analytic. However, since the matrix $A_n$
only depends on a finite order jet of $\rho$ at a reference point,
the corresponding transformation rule in \eqref{a-transform'} holds for any smooth $\rho$.
It remains to consider the change $\rho^* = a\rho$ of the defining function.
By Lemma~\ref{a-lemma}, for any $k$, $s$, $n$ as in \eqref{an}, we have
$$(\rho^*_z)^k(\rho^*_w)^{n-k} = a^n \rho_z^k \rho_w^{n-k},
\quad
(\rho^*_z)^s(\rho^*_w)^{n-3-k} \rho^*_{Z^2}(L^*, L^*)
= a^n \rho_z^k \rho_w^{n-k} \rho_{Z^2}(L, L).
$$
Then, writing
$$
C_n:=
\begin{pmatrix}
(\rho_z^k \rho_w^{n-k})\cr
(\rho_z^s \rho_w^{n-3-s} \rho_{Z^2}(L, L)
\end{pmatrix}_{0\le k\le n, \, 0\le s\le n-3}
$$
for the first column of the matrix $A_n$ given by \eqref{an}
and $C^*_n$ for the corresponding first column of $A^*_n$, we obtain
$$C^*_n = a^n C_n.$$
Then using the relation $\bar L^* = a\bar L$, we conclude for every $j$,
$$(\bar L^*)^j C^*_n = a^{n+j} \bar L^j C_n$$
modulo linear combinations of the columns $\bar L^s C_n$ with $s<j$.
Since the determinant does not change after subtracting a linear combination for columns from another column, we obtain the desired transformation rules
$$\det A^*_n = a^{(2n-1)^2} \det A_n,
\quad
\det D^*_n = a^{\frac{3n(n+1)}2} \det D_n.
$$
\end{proof}
\section{Calculation in Chern-Moser normal form}\Label{SecCM}
Note that the invariance property in Proposition~\ref{det} was obtained
for any smooth real hypersurface given by $\rho(Z,\bar Z)=0$.
If the latter is Levi-nondegenerate, we can use special defining functions
\begin{equation}\Label{normform}
\rho(z,w,\bar z,\bar w) = -{\sf Im}\, w + \varphi(z,\bar z, {\sf Re}\, w), \quad \varphi(z,\bar z, u) = \sum \varphi_{kl}(u)z^k \bar z^l,
\end{equation}
in Chern-Moser normal form (in the formal sense if $M$ is only smooth and not real-analytic) to
compute the determinant of $A_n$ at the origin.
Recall \cite{CM74} that the normal form requires $\varphi$ to satisfy
\begin{equation}\Label{CMnormal}
\varphi_{11}=1, \quad
\varphi_{0k}=\varphi_{1s}=\varphi_{22}=\varphi_{23}=\varphi_{33}=0, \quad k\ge 0, s\ge 2.
\end{equation}
In this normal form we have $(\rho_w, \rho_z)(0)=(i/2,0)$
and furthermore
$$\bar L^j \rho_w(0)= \bar L^k \rho_z(0) = 0, \quad j\geq 1,\ k\ne 1,$$
and
$$\bar L\rho_z(0) =-i/2 \ne 0. $$
Furthermore,
$$\rho_{Z^2}(L,L) = \rho_w^2 \rho_{z^2} - 2 \rho_z\rho_w \rho_{zw} + \rho_{z}^2 \rho_{w^2} $$
and
$$\rho_{\bar z^l w^s}(0) = \rho_{z \bar z^l w^s}(0) = 0, \quad l\ge 0, \quad s\ge 1,$$
imply
$$\bar L^k (\rho_{Z^2}(L,L))(0) =
(\rho_w(0))^2 \bar L^k \rho_{z^2}(0) =
(-i/2)^{k+2} \rho_{z^2\bar z^k} (0) =
(-i/2)^{k+2} \varphi_{z^2\bar z^k} (0).$$
We similarly observe that
\begin{multline}
\bar L^k (\rho_z^s\rho_w^{n-3-s}\rho_{Z^2}(L,L))(0) =\binom{k}{s}
(\rho_w(0))^{n-1-s}\bar L^s\rho_z^s (0)\bar L^{k-s} \rho_{z^2}(0) \\= (-1)^k
(i/2)^{n+k-s-1} s!\binom{k}{s}\varphi_{z^2\bar z^{k-s}} (0).
\end{multline}
Hence we obtain
\begin{equation}\Label{detAnCM}
\det A_n|_{Z=0} = c_n \det
\begin{pmatrix}
\binom{n+1}{0}\varphi_{2,n+1} & \binom{n+2}{0}\varphi_{2,n+2} & \cdots & \binom{2n-2}{0}\varphi_{2, 2n-2}\cr
\binom{n+1}{1}\varphi_{2,n} & \binom{n+2}{1}\varphi_{2,n+1} & \cdots & \binom{2n-2}{1}\varphi_{2, 2n-3}\cr
\vdots & \vdots & \ddots &\vdots\cr
\binom{n+1}{n-3}\varphi_{2,4} & \binom{n+2}{n-3}\varphi_{2, 5} & \cdots & \binom{2n-2}{n-3}\varphi_{2, n+1}\cr
\end{pmatrix},
\end{equation}
where $c_n\ne 0$ is a universal constant (independent of $\varphi$).
In particular,
\begin{equation}\Label{A3=Q}
\det A_3|_{Z=0} = c_3 \varphi_{2,4},\end{equation}
is Cartan's ``6th order tensor",
$$\det A_4|_{Z=0} = c_4 \det
\begin{pmatrix}
\varphi_{2,5} & \varphi_{2,6}\cr
5\varphi_{2, 4} & 6\varphi_{2,5}
\end{pmatrix},
$$
$$\det A_5|_{Z=0} = c_5 \det
\begin{pmatrix}
\varphi_{2,6} & \varphi_{2,7} & \varphi_{2,8}\cr
6\varphi_{2, 5} & 7\varphi_{2,6} & 8\varphi_{2,7}\cr
15\varphi_{2, 4} & 21\varphi_{2,5} & 28\varphi_{2,6}
\end{pmatrix}.
$$
The same calculations, for any hypersurface in the form \eqref{normform} with $\varphi(z,0,s)=\varphi(0,\bar z,s)\equiv 0$ as a formal power series (which can always be achieved; see \cite{BER99a}), show also
that each $\det D_n$ (given by \eqref{dn}) equals a universal constant times $(\varphi_{11})^{n}$.
Together with Theorem~\ref{transform} this yields:
\begin{Pro}\Label{dn-nonvanish} Let $M\subset {\4C}^2$ be a real smooth hypersurface
$M$ given by $\rho(Z,\bar Z)=0$. Then, $\det D_n\ne 0$ at $p\in M$ if and only if
$M$ is Levi-nondegenerate at $p$.
\end{Pro}
\section{Umbilical points on real hypersurfaces in ${\4C}^2$}
Let $M\subset {\4C}^2$ be a smooth real hypersurface and $p\in M$. Recall that $p$ is said to be an {\it umbilical point} if in Chern--Moser normal coordinates $(z,w)$ vanishing at $p$, the coefficient $\varphi_{2,4}$ in the Chern--Moser normal form (\eqref{normform} and \eqref{CMnormal}) vanishes, $\varphi_{2,4}=0$. While Chern-Moser normal coordinates and normal form are not unique, it is well known \cite{CM74} that the vanishing of $\varphi_{2,4}$ is an invariant. By Theorem \ref{transform} and \eqref{A3=Q}, we have:
\begin{Pro}\Label{Umb-A3} Let $M\subset {\4C}^2$ be defined by $\rho=0$. Then, the set $\mathcal U$ of umbilical points on $M$ is given by the equation $\det A_3(\rho)=0$.
\end{Pro}
\subsection{Umbilical indices}
For a fixed global defining equation $\rho=0$ for $M$, where $\rho$ is defined in a neighborhood of $M$, denote by $Q:=\det A_3(\rho)$, so that the set $\mathcal U\subset M$ of umbilical points is given by $Q=0$. We assume $M$ to be oriented and choose $\rho$ compatible with that orientation, i.e.\ such that the gradient of $\rho$ completes positively oriented frames in $M$ to those in ${\4C}^2$. Note that such $\rho$ always exists e.g.\ the oriented distance function. Further, $\rho$ is unique up to multiplication with a positive real function in a neighborhood of $M$ in ${\4C}^2$.
We shall say that $p\in \mathcal U\subset M$ is a {\it $1$-regular} umbilical point of $M$ if $\mathcal U$ is a smooth real curve (1-manifold) at $p$. By Thom's transversality, every hypersurface $M$
can be approximated by one having only $1$-regular umbilical points.
\begin{Def}
For every oriented closed curve $C$ in $M$ avoiding the umbilic set $\mathcal U$,
define its {\em umbilical index} to be $-1/2$ times the winding number
of $Q$ along $C$. For every $1$-regular umbilical point $p$
with chosen orientation of $\mathcal U$,
define its {\em local umbilical index}
(or umbilical index of $M$ at $p$)
to be the umbilical index of the positively oriented boundary
of any sufficiently small disk transversal to $\mathcal U$.
\end{Def}
Since $\rho$ is unique up to multiplication with positive real function,
it follows from Theorem~\ref{transform} that the index as defined
is independent of the choice of $\rho$. It further follows
from the same theorem that the umbilic index of $C$
is also independent of the choice of the ambient coordinates in ${\4C}^2$
as long as $C$ is null-homotopic.
If $\Sigma\subset M$ is an oriented surface (2-manifold) that meets $\mathcal U$ transversely at a $1$-regular point $p$,
the index of $M$ at $p$ is given by
\begin{equation}\Label{Indexp}
\iota_\Sigma(p)=-\frac{1}{2}\deg\left(\frac{Q}{|Q|}\colon \partial \Sigma_p\to S^1\right),
\end{equation}
where $\Sigma_p$ is the boundary of
a sufficiently small topological disk containing $p$
(topologically a circle $S^1$) oriented positively with respect to $\Sigma$, and where $\deg$ denotes the topological degree
(which is the same as winding number in this case).
We note that we can also express the index using an integral,
\begin{equation}\Label{Indexp}
\iota_\Sigma(p)=\frac{i}{4\pi }\int_{\partial\Sigma_p}\frac{dQ}{Q}=
\frac{i}{4\pi }\int_{\partial\Sigma_p}d\log Q.
\end{equation}
Since the 1-form $dQ/Q$ is closed away from the zeros of $Q$, we observe from \eqref{Indexp} that in fact the index at $p$ only depends on the orientation of $\Sigma$ and not on the choice of transversal $\Sigma$ itself, and that the index is constant along every component of the 1-manifold of nondegenerate umbilical points $\mathcal U_{1}$. If $p\in M$ is a $1$-regular umbilical point and the index of $M$ at $p$ (with respect to any transversal $\Sigma$) is non-zero, then we shall say that $p$ is a {\it stable} umbilical point. In view of Thom's transversality, any sufficiently small perturbation of the CR structure of $M$ near a stable umbilical point $p$ will have stable umbilical points near $p$, which motivates this terminology.
We shall use the notation $W_\gamma(R)$ for the winding number of a function $R$ on $M$ along an oriented closed curve $\gamma$ (defined only when $R$ does not vanish on $\gamma$);
\begin{equation}\Label{W(R)}
W_\gamma(R)=\frac{1}{2\pi i}\int_{\gamma}\frac{dR}{R}=
\frac{1}{2\pi i}\int_{\gamma}d\log R.
\end{equation}
Thus, in particular, if $\Sigma$ is a surface, transversal to $\mathcal U$ at $p$ and $\Sigma_p$ is its intersection with a small tubular neighborhood of $\mu$ near $p$, then by definition of the index:
$$
\iota_\Sigma(p)=-\frac{1}{2}W_{\partial\Sigma_p}(Q).
$$
If $M$ is real-analytic, then $\mathcal U$ is a real-analytic subvariety of $M$, and the set $\mathcal U_1$ of 1-regular umbilical points consist of the subset of $\mathcal U$ of regular points of dimension one. For simplicity, we shall proceed under the assumption that $M$ is real-analytic. In this case, $\mathcal U$ is either all of $M$ (we assume that $M$ is connected), in which case $M$ is locally spherical, or $\mathcal U$ is a proper subvariety. In the latter case, points of $\mathcal U$ are either 0-, 1-, or, 2-dimensional, and we decompose $\mathcal U$ accordingly, $\mathcal U=\mathcal U^0\cup \mathcal U^1\cup \mathcal U^2$; recall that the (topological) dimension of a real-analytic subvariety $\mathcal V$ at a point $p$ is the largest dimension of a nonsingular component of $\mathcal V$ with $p$ in its closure.
We note that the set of $1$-regular umbilical points $\mathcal U_1$ equals $\mathcal U^0\cup \mathcal U^1$ minus a discrete set of points. Thus, if there are no points of dimension $2$ on $\mathcal U$, then any surface $\Sigma$ that intersects $\mathcal U$ can be locally perturbed to only intersect $\mathcal U$ along the set of $1$-regular points $\mathcal U_1$. We have the following simple consequence of Stokes Theorem:
\begin{Pro}\Label{Wvsindex}
Let $M\subset {\4C}^2$ be a real-analytic hypersurface, and assume that $\mathcal U$ has no points of dimension $2$ or $3$, i.e., $\mathcal U=\mathcal U^0\cup\mathcal U^1$. Let $\gamma\subset M$ be a oriented closed curve, homologous to $0$ in $M$ and not intersecting $\mathcal U$, and $\Sigma\subset M$ an oriented surface, intersecting $\mathcal U$ transversally along $\mathcal U_1$ and with $\partial\Sigma=\gamma$. Then,
\begin{equation}
W_\gamma(Q)=-\frac{1}{2}\sum_{p\in \mathcal U_1\cap \Sigma}\iota_\Sigma(p).
\end{equation}
\end{Pro}
\begin{proof} Let $p_1,\ldots, p_k$ be the finite set of points in $\mathcal U_1\cap \Sigma$ and $\Sigma_{p_j}$ the intersection of $\Sigma$ with a sufficiently small tubular neighborhood of $\mathcal U_1$ near $p_j$ (so small that the closures of the $\Sigma_{p_j}$ are disjoint). Since $dQ/Q$ is closed in the punctured surface
$$
\Sigma':=\Sigma\setminus \bigcup_{j=1}^k \Sigma_{p_j},
$$
being locally $d\log Q$ on $\Sigma'$, we conclude by Stokes Theorem that:
\begin{equation}
W_\gamma(Q)=\frac{1}{2\pi i} \int_{\partial\Sigma}\frac{dQ}{Q}=\sum_{j=1}^k \frac{1}{2\pi i}\int_{\partial\Sigma_{p_j}} \frac{dQ}{Q}= -\frac{1}{2}\sum_{p\in \mathcal U_1\cap \Sigma}\iota_\Sigma(p).
\end{equation}
\end{proof}
\section{Umbilical points on Real Ellipsoids}\Label{EllSec}
We shall consider real ellipsoids $E\subset {\4C}^2$. A general real ellipsoid can be defined by an equation of the form
\begin{equation}
A(z^2+\bar z^2)+2(2+A)|z|^2+B(w^2+\bar w^2)+2(2+B)|w|^2=4,\quad A,B\geq 0.
\end{equation}
We shall fix $A,B\geq 0$, and assume that at least one of these, say $A$, is nonzero (so that the ellipsoid does not degenerate to a sphere). We consider a 1-parameter family of ellipsoids $E_\epsilon$, defined by $\rho_\epsilon=0$, where
\begin{equation}
\rho_\epsilon:=
\epsilon A(z^2+\bar z^2)+2(2+\epsilon A)|z|^2
+\epsilon B(w^2+\bar w^2)+2(2+\epsilon B)|w|^2
-4
\end{equation}
\begin{equation}
= -4 + 4(|z|^2 + |w|^2) + \epsilon (
A (z^2 + \bar z^2 + 2|z|^2 )
+ B (z^2 + \bar z^2 + 2|z|^2
),
\quad \epsilon >0.
\end{equation}
Note that $E_0$ is the unit sphere. We shall mainly be concerned with small perturbations of the sphere, and shall thus consider small $\epsilon>0$. Our aim is to prove the following result:
\begin{Thm}\Label{Prop-ell} For $\epsilon>0$ sufficiently small, the subset of umbilical points on $E_\epsilon$ either contains points of dimension at least $2$ or contains a curve of umbilical points.
\end{Thm}
To this end, we shall compute the matrix $A_3=A_3(\rho_\epsilon)$ on $\rho=\rho_\epsilon=0$. Since we will be concerned with small perturbations it suffices, as we shall see, to compute $A_3$ mod $O(\epsilon^3)$. We note first that
\begin{equation}\label{L-ell}
\rho_z=4\bar z+2\epsilon A (\bar z +z),\quad \rho_w=4\bar w+2\epsilon B(\bar w+ w)
\end{equation}
and
\begin{equation}
\rho_{z^2} = 2\epsilon A
,\quad \rho_{zw} = 0
,\quad \rho_{w^2} = 2\epsilon B.
\end{equation}
Therefore, we have
\begin{equation}\label{rhoZ2}
\begin{aligned}
\rho_{Z^2}(L,L)=&(\rho_w)^2\rho_{z^2}-2\rho_z\rho_w\rho_{zw}+(\rho_z)^2\rho_{w^2}\\
=&8\epsilon \left(A(2\bar w+\epsilon B(\bar w+ w))^2+B(2\bar z+\epsilon A (\bar z +z))^2\right)\\
=&32\epsilon (A\bar w^2+B\bar z^2)
+ 32\epsilon^2AB(|z|^2+|w|^2+\bar w^2+\bar z^2)\\
& + 8\epsilon^3AB\left (A(z+\bar z)^2+B(w+\bar w)^2\right).
\end{aligned}
\end{equation}
To calculate $A_3(\rho)$, we shall repeatedly apply $\bar L$, where
\begin{equation}\label{barL}
\begin{aligned}
\bar L &=
-\rho_{\bar w} \frac{\partial}{\partial \bar z}
+ \rho_{\bar z}\frac{\partial}{\partial \bar w}\ \\
&=-2(2w+\epsilon B(\bar w+ w))\frac{\partial}{\partial \bar z}+2(2z+\epsilon A(\bar z+z)\frac{\partial}{\partial \bar w}\\
&=4\left(-w\frac{\partial}{\partial \bar z}+z\frac{\partial}{\partial \bar w}\right )+2\epsilon\left(-B(w+\bar w)\frac{\partial}{\partial \bar z}+A(z+\bar z)\frac{\partial}{\partial \bar w}\right)\\
&=\bar L_0+\epsilon \bar L_1,
\end{aligned}
\end{equation}
to $\rho_{Z^2}(L,L)$,
and subsequently evaluate at $w=0$. Since $\bar L$ only involves differentiation in $\bar z$ and $\bar w$,
the result for $w=0$ will not change when replacing with $0$
all occurences of
$w$ (but not $\bar w$). Thus for $w=0$, we obtain
\begin{equation}
\begin{aligned}
& \frac1{2^k \cdot 32} \bar L^k \rho_{Z^2}(L,L) \\
& = \left(
2z \frac\partial{\partial\bar w}
+ \epsilon \left(
- B\bar w \frac\partial{\partial\bar z}
+ A(z + \bar z) \frac\partial{\partial\bar w}
\right)
\right)^k
\left(
\epsilon (A \bar w^2 + B \bar z^2)
+ \epsilon^2 AB(|z|^2 + \bar z^2 + \bar w^2).
\right)
\end{aligned}
\end{equation}
Then we obtain for $w=0$,
\begin{equation}
\bar L\rho_{Z^2}(L,L)
= \bar L^3 \rho_{Z^2}(L,L)
= \bar L^4 \rho_{Z^2}(L,L)
= O(\epsilon^2),
\quad
\frac1{2^2 \cdot 32} \bar L^2\rho_{Z^2}(L,L)
= 2^3 z^2 A \epsilon + O(\epsilon^2)
\end{equation}
\subsection{Terms of the form $\bar L^k\rho_z^3$; first row.} We note that
\begin{equation}
\rho_z^3=(4\bar z+2\epsilon A (\bar z +z))^3=8(2\bar z+\varepsilon A(\bar z+z))^3.
\end{equation}
We shall be interested in $A_3$ mod $O(\epsilon^3)$, and since all terms in the last row (computed above) are already $O(\varepsilon)$, we shall compute $\bar L^k(\rho_z)^3$ mod $O(\varepsilon^2)$. Thus, we have
\begin{equation}
\rho_z^3=8(8\bar z^3+12\varepsilon A\bar z^2(\bar z+z))+O(\varepsilon^2)
=
2^5 (2\bar z^3+3\varepsilon A(\bar z^3+z\bar z^2))+O(\varepsilon^2).
\end{equation}
We obtain for $w=0$,
\begin{equation}
\frac1{2^k}
\bar L^k \rho_z^3
=
\left(
2z \frac\partial{\partial\bar w}
+ \epsilon \left(
- B\bar w \frac\partial{\partial\bar z}
+ A(z + \bar z) \frac\partial{\partial\bar w}
\right)
\right)^k
\rho_z^3
\end{equation}
and since $\rho_z$ is independent of $w$,
\begin{equation}
\bar L\rho_z^3 = \bar L^3\rho_z^4 = \bar L^4\rho_z^3 = O(\varepsilon^2),
\end{equation}
and
\begin{equation}
\frac14 \bar L^2 \rho_z^3
= -2z \epsilon B \frac\partial{\partial \bar z} \rho_z^3
= -2^7\cdot 3\epsilon B z \bar z^2 + O(\varepsilon^2)
\end{equation}
\subsection{Terms of the form $\bar L^k\rho_z^2\rho_w$; second row.}
As before, replacing occurences of $w$ (but not $\bar w$) with $0$ (written $\cong$),
we obtain
\begin{equation}
\begin{aligned}
\rho_z^2\rho_w &\cong
(4\bar z+2\varepsilon A(\bar z+z))^2(4 + 2\varepsilon B) \bar w\\
&=2^5(2\bar z^2\bar w+\varepsilon((2A+B)\bar z^2\bar w+2Az\bar z\bar w)+O(\varepsilon^2),
\end{aligned}
\end{equation}
We compute, mod $O(\varepsilon^2)$
\begin{multline}
\frac1{2^{k+5}}
\bar L^{k} (\rho_z^2 \rho_w) \\
\cong \left(
2z \frac\partial{\partial\bar w}
+ \epsilon \left(
- B\bar w \frac\partial{\partial\bar z}
+ A(z + \bar z) \frac\partial{\partial\bar w}
\right)
\right)^k
\left(
(2\bar z^2\bar w+\varepsilon((2A+B)\bar z^2\bar w+2Az\bar z\bar w)
\right),
\end{multline}
and hence for $w=0$,
\begin{equation}
\bar L^{k} (\rho_z^2 \rho_w) = O(\epsilon),
\ k\ne 1;\quad \bar L^{4} (\rho_z^2 \rho_w) = O(\epsilon^2),
\end{equation}
and
\begin{equation}
\bar L (\rho_z^2 \rho_w) = 2^8 z\bar z^2 + O(\epsilon).
\end{equation}
\subsection{Terms of the form $\bar L^k\rho_z\rho_w^2$; third row.} We can obtain the formulas in this case by considering the previous subsection and interchanging the roles of $z$ and $w$. We obtain
\begin{equation}
\begin{aligned}
\rho_z\rho_w^2 &\cong
(4\bar z+2\varepsilon A(\bar z+z))(4 + 2\varepsilon B)^2 \bar w^2\\
&=2^5
\left(
2\bar z
+\varepsilon((2B+A)\bar z+Az)
\right)\bar w^2
+O(\varepsilon^2),
\end{aligned}
\end{equation}
As before we obtain mod $O(\varepsilon^2)$,
\begin{multline}
\frac1{2^{k+5}}
\bar L^{k} (\rho_z \rho_w^2) \\
\cong \left(
2z \frac\partial{\partial\bar w}
+ \epsilon \left(
- B\bar w \frac\partial{\partial\bar z}
+ A(z + \bar z) \frac\partial{\partial\bar w}
\right)
\right)^k
2^5\left(
(2\bar z
+\varepsilon((2B+A)\bar z+Az)\right)\bar w^2,
\end{multline}
from where as before, for $w=0$,
\begin{equation}
\bar L^k ( \rho_z\rho_w^2 ) = O(\epsilon), \quad k\ne 2,
\end{equation}
\begin{equation}
\bar L^2 ( \rho_z\rho_w^2 ) = 2^{11}z^2\bar z + O(\epsilon),
\end{equation}
and
\begin{equation}
\bar L^4 ( \rho_z\rho_w^2 )
= 2^7
{4 \choose {2}}
\epsilon
\left(
2z \frac\partial{\partial\bar w}
- \epsilon B\bar w \frac\partial{\partial\bar z}
\right)^2
(4\bar z z^2)
+ O(\epsilon^2)
= 2^{11}
{4 \choose {2}}
\epsilon
B z^3
+ O(\epsilon^2).
\end{equation}
\subsection{Terms of the form $\bar L^k\rho_w^3$; fourth row.} Following the same strategy as above,
we obtain
\begin{equation}
\rho_w^3 \cong
(4 + 2\varepsilon B)^3 \bar w^3 =
2^3 (8 + 12\varepsilon B) \bar w^3
+O(\varepsilon^2),
\end{equation}
and for $w=0$,
\begin{equation}
\bar L^k ( \rho_w^3 ) = O(\epsilon^2), \quad k\ne 3,
\end{equation}
\begin{equation}
\bar L^3 (\rho_w^3 ) = 2^{9} \cdot 6 z^3 + O(\epsilon).
\end{equation}
\subsection{Calculation of $\varepsilon^2$-term of $A_3(\rho_\varepsilon)$ along $w=0$}
From our calculations in the subsections above we obtain for $w=0$:
\begin{equation}
A_3(z,\bar z)=
\begin{pmatrix}
2^6\bar z^3+O(\epsilon) &0& O(\epsilon) &0& 0\\
O(\epsilon)& 8z\bar z^2 + O(\epsilon) & O(\epsilon)& O(\epsilon)& 0 \\
O(\epsilon)& O(\epsilon)& 2^{11} z^2\bar z & O(\epsilon)& 2^{11}
{4 \choose {2}} \epsilon B z^3\\
0& 0& O(\epsilon)& 2^{10}\cdot 3 z^3 & 0\\
2^5 \epsilon B\bar z^2 & 0& 2^{10} \epsilon Az^2 & 0& 0
\end{pmatrix}
+ O(\epsilon^2).
\end{equation}
Since the last row as well as the last column each has a factor $\varepsilon$,
we conclude
\begin{equation}\Label{det-a3}
\det A_3(\rho_\varepsilon)|_{w=0}=\varepsilon^2\Delta_2 +O(\varepsilon^3),
\quad
\Delta_2 = N AB z^9\bar z^5,
\end{equation}
where $N$ is a large positive integer.
\subsection{Umbilical points on the ellipsoids $E_\varepsilon$.} Let $S_\varepsilon$ be the ellipse (in the $z$-plane) obtained by intersecting $E_\varepsilon$ with the complex line $w=0$. If both $A, B>0$, we easily conclude that the winding number $W_{S_0}(\Delta_2)$ of $\Delta_2(z,\bar z)$ around the circle $S_0$, traversed in the positive direction, equals $4$; recall that the winding number is defined by \eqref{W(R)} from which $W_{S_0}(\Delta_2)=4$ follows immediately.
\begin{proof}[Proof of Theorem $\ref{Prop-ell}$]
Let us first assume that both $A,B>0$. Since, in this case, $\Delta_2$ does not vanish on $S_0$, it follows that $Q_\varepsilon|_{w=0}$ does not vanish on $S_\varepsilon$ for $\varepsilon>0$ sufficiently small, where
$Q_\varepsilon=\det A_3(\rho_\varepsilon)$.
It is also clear by continuity that, for sufficiently small $\varepsilon>0$, the winding number of $Q_\varepsilon$ around $S_0$ coincides with that of $\Delta_2$ around $S_0$, and also around $S_\varepsilon$. We conclude that
\begin{equation}\Label{W(Qeps)}
W_{S_\varepsilon}(Q_\varepsilon)=4,
\end{equation}
for sufficiently small $\varepsilon>0$. Now, either the set of umbilical points $\mathcal U\subset E_\varepsilon$ contains points of dimension at least 2, or there is a surface $\Sigma^\varepsilon$ in $E_\varepsilon$ that is bounded by $S_\varepsilon$ and meets the subset of 1-regular umbilical points $\mathcal U_1$ transversally; indeed, we can always find even a simply connected $\Sigma^\varepsilon$ in $E_\varepsilon$ with $\partial\Sigma^\varepsilon=S_\varepsilon$, and if $\mathcal U$ has only components of dimension 0 and 1, then small local deformations of $\Sigma^\varepsilon$ along the intersection will result in only transversal intersections along $\mathcal U_1$. It now follows from \eqref{W(Qeps)} and Proposition \ref{Wvsindex} that
\begin{equation}
\sum_{p\in \Sigma^\varepsilon\cap \mathcal U_1}\iota_{\Sigma^\varepsilon_p}(p)=-2.
\end{equation}
In particular, either $\mathcal U$ has points of dimension at least 2 or contains at least one curve of stable umbilical points when both $A,B>0$.
In the remaining case where, say, $B=0$, the ellipsoid is invariant under the circle action $(z,w)\mapsto (z,e^{it}w)$ and therefore has umbilical points along the curve of fixed points $(z,0)$, in view of the special Chern--Moser normalization at non-umbilical points \cite{CM74}, pp.\, 246--247.
\end{proof}
\section{Umbilical points on perturbations of the sphere}\Label{PertSec}
We shall consider perturbations $M_\varepsilon\subset {\4C}^2$ of the unit sphere
given by $\rho=\rho^\epsilon=0$, where
\begin{equation}\Label{pert}
\rho^\varepsilon:= \rho^0 + \varepsilon \rho',
\quad \rho^0 := -1 + z\bar z + w\bar w,
\end{equation}
$\rho'$ is a smooth real-valued function, and $\varepsilon$ is a small real parameter.
For $\varepsilon=0$ we recover the unit sphere $S^3=M_0$ and hence $\det A_3=0$.
For $\varepsilon\neq 0$, we shall consider the power series expansion of $\det A_3$ in $\varepsilon$. In that expansion, we shall compute the linear term in $\varepsilon$.
Since the expansion of $\rho_{Z^2}$ begins with a linear term in $\varepsilon$,
the only nonzero contribution to the linear term in $\varepsilon$ in the determinant \eqref{a3} will come
from $0$th order terms (in $\varepsilon$) in the first $4$ rows and $1$st order terms in the last row.
Furthermore, only $0$th order terms in the expansion of $L$ will contribute.
Thus for our computation, we only need use the terms with
$$(\rho^0_w, \rho^0_z)=(\bar w, \bar z), \quad L_0 = -\bar w \partial_z + \bar z \partial_w,$$
and hence the desired coefficient of $\varepsilon$ is
\begin{equation}\label{epscoeff}
\det
\begin{pmatrix}
D^0_3 &0\cr
* & \bar L_0^4 (\rho'_{Z^2}(L_0,L_0))
\end{pmatrix}
= (\det D^0_3) \bar L_0^4 (\rho'_{Z^2}(L_0,L_0)),
\end{equation}
where $D^0_3$ is calculated using $\rho^0$. By Proposition \ref{dn-nonvanish}, we conclude:
\begin{Pro}\Label{Lin-eps} For a perturbation of the form \eqref{pert},
\begin{equation}\Label{Lin-epsterm}
\det A_3(\rho^\varepsilon)=c_0\bar L_0^4 (\rho'_{Z^2}(L_0,L_0))\varepsilon+O(\varepsilon^2),
\end{equation}
where $c_0$ is a universal polynomial that does not vanish on the unit sphere $\rho^0=0$.
\end{Pro}
We note that
\begin{equation}\Label{rhoZ2}
\rho'_{Z^2}(L_0,L_0)= (-\bar w)^2 \rho'_{z^2} - 2\bar z\bar w \rho'_{zw} + \bar z^2 \rho'_{w^2},
\end{equation}
and observe that the coefficients in $\bar L_0$ are holomorphic, and hence repeated applications of $\bar L_0$ will not result in any differentiations of the coefficients, and we obtain
\begin{equation}\Label{l4}
\bar L^4_0=(-w\partial_{\bar z}+z\partial_{\bar w})^4=w^4\partial^4_{\bar z}-4zw^3\partial^3_{\bar z}\partial_{\bar w}+6z^2w^2\partial^2_{\bar z}\partial^2_{\bar w}-4z^3w\partial_{\bar z}\partial^3_{\bar w}+z^4\partial^4_{\bar w}.
\end{equation}
We shall consider polynomial perturbations of the form $\rho'=\sum_{k=2}^m\rho'_{k}$, where $\rho'_k$ are homogeneous polynomials of degree $k$ in $Z=(z,w)$ and $\bar Z$. We may decompose $\rho'_k$ further into bidegree, $\rho'_k=\sum_{p+q=k}\rho'_{p,q}$, where each $\rho_{p,q}$ is of bidegree $(p,q)$. Since our perturbations $\rho'$ are real-valued, we must have $\rho'_{q,p}=\overline{\rho'_{p,q}}$.
We shall use the notation $\mathcal H_k$ for the space of homogeneous polynomials of degree $k$, and $\mathcal H_{p,q}$ for those of bidegree $(p,q)$. We note that if $R\in \mathcal H_{p,q}$, then $\bar L_0^4 (R_{Z^2}(L_0,L_0))\in \mathcal H_{p+2,q-2}$. We also note that in this case $R_{Z^2}(L_0,L_0)\in \mathcal H_{p-2,q+2}$, and we conclude that $\bar L_0^4 (R_{Z^2}(L_0,L_0))=0$ unless both $p$ and $q$ satisfy $p,q\geq 2$. Let us for brevity use the notation
\begin{equation}\Label{Q0}
Q^0(R):=\bar L_0^4 (R_{Z^2}(L_0,L_0)),
\end{equation}
so that
\begin{equation}\Label{Q}
Q=Q(\rho^\varepsilon):=\det A_3(\rho^\varepsilon)=c_0Q^0(\rho')\varepsilon+O(\varepsilon^2).
\end{equation}
We may then summarize the discussion above as follows.
\begin{Pro}\Label{Q0decomp} For a real-valued polynomial $\rho'$ of degree $m$, decomposed into homogeneous components $\rho'_k$ and further decomposed into bidegree $\rho'_{p,q}$, of the form
\begin{equation}\Label{rho'decomp}
\rho'=\sum_{k=2}^m\rho'_k=\sum_{k=2}^m\sum_{p+q=k}\rho'_{p,q},\quad \rho'_{p,q}=\overline{\rho'_{q,p}},
\end{equation}
it holds that
\begin{equation}\Label{Q0decomp}
Q^0(\rho')=\sum_{k=4}^m\sum_{l=4}^k Q^0_{l,k-l},\quad Q^0_{l,k-l}=Q^0(\rho'_{l-2,k-l+2}).
\end{equation}
\end{Pro}
We have the following technical result.
\begin{Pro}\Label{gengenpert}
Let $\rho'$ be a real-valued polynomial of degree $m$, and decompose $Q^0(\rho')$ as in \eqref{Q0decomp}. Assume that:
\begin{itemize}
\item[(i)] The real-algebraic variety $\mathcal V:=\{Q^0(\rho')=0\}\cap S^3$ in $S^3$ has no points of dimension $\geq 2$.
\item[(ii)] $Q^0_{l,k-l}=0$ for
$4\leq l\leq k/2$.
\end{itemize}
Then, for sufficiently small $\varepsilon>0$, the set of umbilical points $\mathcal U$ on the perturbation $M_\varepsilon$ contains either points of dimension $\geq 2$ or a curve of stable umbilical points.
\end{Pro}
\begin{Rem}\Label{gengenrem} We make a few observations:
\begin{itemize}
\item If the degree $m\leq 7$, then condition (ii) is vacuous, and hence only (i) is required in this case.
\item If the degree $m\leq 3$, then condition (i) is never satisfied, since $Q^0(\rho')$ vanishes completely. In particular, as noted in the previous section, for real ellipsoid perturbations $E_\varepsilon$ we have $Q^0(\rho')=0$. Nevertheless, as is proved in Theorem \ref{Prop-ell}, real ellipsoids do have umbilical points.
\end{itemize}
\end{Rem}
To prove Proposition \ref{gengenpert}, we need the following lemma:
\begin{Lem}\Label{Excircle}
Let $\rho'$ be a real-valued polynomial such that condition {\rm (i)} in Proposition $\ref{gengenpert}$ holds. Then, there is point $Z_0=(z_0,w_0)\in S^3$ such that $Q^0(\rho')$ does not vanish on the circle $S_0\colon t\mapsto e^{it}Z_0$ in $S^3$.
\end{Lem}
\begin{proof} For $Z_1:=(z_1,w_1)\in S^3$, consider the circle $S_1$ in $S^3$ parametrized by $t\mapsto e^{it}(z_1,w_1)$. Let $\Sigma\subset S^3$ be a germ at $Z_1$ of an open, real-analytic surface, transverse to $S_1$ at this point. Consider the real-analytic map $\Gamma\colon \Sigma\times S^1\to S^3$, given by $\Gamma(z,w,t)=e^{it}(z,w)$ in local coordinates $t\to e^{it}$ on $S^1$. This map realizes an open subset $\Omega$ of $S^3$ as an $S^1$-fibration over $\Sigma$. If we let $\pi\colon \Omega\to \Sigma$ be the projection, then we can consider $\pi(\mathcal V)\subset \Sigma$, where $\mathcal V$ is the zero locus of $Q^0(\rho')$ as in Proposition \ref{gengenpert}. Since $\mathcal V$, by condition (i), has no points of dimension 2 or 3, $\pi(\mathcal V)$ is a proper sub-analytic subset of the open surface $\Sigma$. Thus, by choosing $Z_0=(z_0,w_0)$ in $\Sigma$ outside this projection, we find the desired oriented circle $S_0$, parametrized by $t\to \Gamma(z_0,w_0,t)$.
\end{proof}
\begin{Rem} We may parametrize all great circles on $S^3$ by blowing up the origin in ${\4C}^2$. In this way, $S^3$ becomes the unit circle in the universal line bundle $O(-1)$ over ${\4P}^2$. The corresponding projection $\pi \colon O(-1)\to {\4P}^2$ is algebraic and then the set $\pi(\mathcal V)$ of unit circles to avoid is a closed semialgebraic set in ${\4P}^2$.
\end{Rem}
We now proceed with the proof of Proposition \ref{gengenpert}. Let $S_0\colon t\mapsto e^{it}Z_0$, with $Z_0=(z_0,w_0)\in S^3$, be the circle provided by Lemma \ref{Excircle}, and define a polynomial $P(\zeta,\bar \zeta)$ of degree $m$ in the variable $\zeta\in{\4C}$ by
\begin{equation}\Label{PfromQ}
P(\zeta,\bar \zeta):=Q^0(\rho')(\zeta Z_0,\overline{\zeta Z_0}).
\end{equation}
By construction of $S_0$, $P$ does not vanish on the unit circle. The decomposition of $Q^0(\rho')$ into bidegree, given by \eqref{Q0decomp}, yields a decomposition of $P$ into bidegree:
\begin{equation}
P(\zeta,\bar\zeta)=\sum_{k=4}^m\sum_{l=4}^kp_{l,k-l}
\zeta^l\bar \zeta^{k-l},\quad p_{l,k-l}
:=Q_{l,k-l}(\zeta Z_0,\overline{\zeta Z_0})/\zeta^l\bar \zeta^{k-l}.
\end{equation}
On the unit circle $\zeta=e^{it}$, $P(\zeta,\bar \zeta)$ coincides with a rational function $R(\zeta)$,
\begin{equation}
R(\zeta)=P(\zeta,1/\zeta)=\frac{\sum_{k=4}^m\sum_{l=4}^k
p_{l,k-l}\zeta^{2l+m-k}}{\zeta^m}.
\end{equation}
If we define
\begin{equation}
p(\zeta):=\sum_{k=4}^m\sum_{l=4}^k
p_{l,k-l}\zeta^{2l+m-k},
\end{equation}
then by the construction of $p(\zeta)$ and the argument principle we conclude:
\begin{Lem}\Label{ArgP}
Let $n$ denote the number of zeros (counted with multiplicities) of $p(\zeta)$ in the unit disk $|\zeta|<1$. Then
\begin{equation}
W_{S_0}(Q^0)=n-m,
\end{equation}
where $Q^0=Q^0(\rho')$ and $S_0$ is the circle in the construction of $p(\zeta)$ above.
\end{Lem}
We may now complete the proof of Proposition \ref{gengenpert}.
\begin{proof}[Proof of Proposition $\ref{gengenpert}$] Recall that the set of umbilical points $\mathcal U$ on $M_\varepsilon$ is given by $Q=0$, where $Q$ is as in \eqref{Q}. Let $S_0$ the circle on $S^3$ as above, and let $S_\varepsilon$ be the perturbed oriented curve on $M_\varepsilon$ obtained as the intersection between the complex subspace through $Z_0=(z_0,w_0)\in S^3$ and $M_\varepsilon$. As in the proof of Theorem \ref{Prop-ell}, for sufficiently small $\varepsilon>0$, we have
\begin{equation}
W_{S_\varepsilon}(Q)=W_{S_0}(Q^0)=n-m,
\end{equation}
where $n$ and $m$ are as in Lemma \ref{ArgP}. We claim that $n-m\neq 0$. Indeed, by condition (ii), the coefficients $p_{l,k-l}=0$ for $l\leq k/2$. If we let $r$ denote the minimum integer $r=2l+m-k$ for which $p_{l,k-l}\neq 0$, then $p(\zeta)$ is divisible by $\zeta^{r}$ and since $r>m$, we conclude that $n>m$. The proof of Proposition \ref{gengenpert} is now completed in the same way as the proof of Theorem \ref{Prop-ell}.
\end{proof}
\subsection{The sphere $S^3$ as a circle bundle over ${\4P}^1$} We recall here the idea of realizing the sphere as the unit circle in the universal bundle
$\pi\colon J:=O(-1)\to {\4P}^1$; this idea has been extended to more general three-dimensional CR manifolds with a CR circle action by Bland--Duchamp \cite{BlandDuchamp91} and Epstein \cite{Epstein92}. Recall that $J$
naturally embeds into ${\4P}^1\times {\4C}^2$ in such a way that the fiber $J_Z$ over a point $Z$ in homogeneous coordinates, $Z=[z\colon w]\in {\4P}^1$, is the complex line through $(z,w)\in {\4C}^2$; $J_Z$ is parametrized by $\zeta\mapsto \zeta(z,w)$. The standard metric $|\cdot|$ on $J$ is the one induced by the Euclidian metric on ${\4C}^2$; if $s(Z)$ is a non-vanishing local section in $J$ and we write $s(Z)=(u,v)\in {\4C}^2$, then $|s|^2:=|u|^2+|v|^2$. The unit circle bundle $\tilde S^3:=\{\lambda\in J\colon |\lambda|^2=1\}$ is CR isomorphic to the unit sphere $S^3$ in ${\4C}^2$. Indeed, if we view the total space $J$ as the blow-up of the origin in ${\4C}^2$, then the CR isomorphism $\tilde \pi|_{\tilde S^3}\colon \tilde S^3\to S^3$ is the blow-down map $\tilde \pi\colon J\to {\4C}^2$ restricted to $\tilde S^3$. For convenience, we shall simply identify $S^3$ with $\tilde S^2$ via this isomorphism; in this identification, the fibers $\pi^{-1}(Z)$ in $\tilde S^3$ correspond to the great circles $t\mapsto e^{it}Z$.
We note that if $\rho'$ is a real-valued polynomial, then the projection $\pi(\mathcal V)\subset {\4P}^1$ of the real-algebraic subvariety $\mathcal V\subset S^3\cong \tilde S^3$, defined to be the zero locus of $Q^0=Q^0(\rho')$ as in Proposition \ref{gengenpert}, is a closed semialgebraic subset. An inspection of the proof of Proposition \ref{gengenpert} reveals immediately that condition (i) in the assumptions of this proposition can be replaced by the assumption that $\pi(\mathcal V)\neq {\4P}^1$. As is shown in Lemma \ref{Excircle}, condition (i) implies $\pi(\mathcal V)\neq {\4P}^1$, and the latter property is the only one used in the proof of Proposition \ref{gengenpert}. For convenience, we state the result here.
\begin{Pro}\Label{gengenpert'}
Let $\rho'$ be a real-valued polynomial of degree $m$, and decompose $Q^0(\rho')$ as in \eqref{Q0decomp}. Assume that:
\begin{itemize}
\item[(i$'$)] ${\4P}^1\setminus \pi(\mathcal V)\neq \emptyset$, where $\mathcal V\subset S^3\cong\tilde S^3$ and $\pi\colon \tilde S^3\to {\4P}^1$ are as above.
\item[(ii)] $Q^0_{l,k-l}=0$ for $4\leq l\leq k/2$.
\end{itemize}
Then, for sufficiently small $\varepsilon>0$, the set of umbilical points $\mathcal U$ on the perturbation $M_\varepsilon$ contains either points of dimension $\geq 2$ or a curve of stable umbilical points.
\end{Pro}
\subsection{Generic perturbations of the sphere} We shall denote by $\mathcal P_m$ the space of all polynomials in $Z=(z,w)$ and $\bar Z$ of degree at most $m$, and by $\mathcal P^{\bR}_m$ the real subspace of those that are real-valued. Thus, we have $$\mathcal P_m=\bigoplus_{p+q\leq m}\mathcal H_{p,q}$$ and $\rho'\in\mathcal P_m$ belongs to $\mathcal P^{\bR}_m$ when $\rho'_{p,q}=\overline{\rho_{q,p}}$ for all $p,q$. We shall show that condition (i$'$) in Proposition \ref{gengenpert'} is generic. More precisely, we shall prove the following:
\begin{Pro} \Label{gencondi} The set $\Pi_m$ of polynomials $\rho'$ in $\mathcal P^{\bR}_m$ such that $\pi(\mathcal V)={\4P}^1$, where $\mathcal V\subset S^3\cong\tilde S^3$ and $\pi\colon \tilde S^3\to {\4P}^1$ are as in Proposition $\ref{gengenpert'}$, is a real-analytic subvariety in $\mathcal P^{\bR}_m$. Moreover, if $\mathcal A\subset \mathcal P^{\bR}_m$ is any real subspace containing $\mathcal A_p:=\{ez^p\bar w^p+\bar e w^p\bar z^p\colon e\in {\4C}\}$ for some $2\leq p\leq m/2$, then $\Pi_m\cap \mathcal A$ has strictly smaller dimension than $\mathcal A$.
\end{Pro}
\begin{proof} Let $\tilde z=z/w$ be a local coordinate in the chart $U_0=\{[z\colon w]\in {\4P}^1\colon w\neq 0\}$ in ${\4P}^1$ and $\tilde Q^0=\tilde Q^0(\tilde z,\bar{\tilde z};\zeta,\bar\zeta)$ the polynomial $Q^0=Q^0(\rho')$ for some $\rho'\in \mathcal P^{\bR}_m$ in the local trivialization
$$U_0\times{\4C}={\4C}\times{\4C}\cong J|_{U_0}\subset U_0\times{\4C}^2,$$
given by
$$
(\tilde z,\zeta)\mapsto (\tilde z;\tilde\pi(\tilde z,\zeta)),\ \tilde\pi(\tilde z,\zeta):=\zeta(\tilde z,1).
$$
In other words, $\tilde Q^0=Q^0\circ \tilde \pi$; we shall denote by $\tilde Q^0_{p,q}=Q^0_{p,q}\circ\tilde \pi$, so that we have the decomposition (see Proposition \ref{Q0decomp})
\begin{equation}
\tilde Q^0=\sum_{k=4}^m\sum_{l=4}^k\tilde Q^0_{l,k-l},
\end{equation}
where each component $\tilde Q^0_{p,q}$ takes the form
\begin{equation}
\tilde Q^0_{p,q}(\tilde z,\bar{\tilde z};\zeta,\bar\zeta)=q_{p,q}(\tilde z,\bar{\tilde z})\zeta^p\bar \zeta^q=q_{p,q}(\tilde z,\bar{\tilde z})\zeta^{p-q}|\zeta|^{2q},
\end{equation}
with
\begin{equation}
q_{p,q}(\tilde z,\bar{\tilde z})=\tilde Q^0_{p,q}((\tilde z,1),(\bar{\tilde z},1))=
\left(\sum_{\alpha\leq p,\, \gamma\leq q}c_{pq;\alpha\bar \gamma}\tilde z^{\alpha}\bar{\tilde z}^{\gamma}\right).
\end{equation}
for suitable coefficients $c_{pq;\alpha\bar\gamma}$.
Recall that $\tilde S^3\subset J$ is given in these coordinates by
\begin{equation}\Label{tildeS^3}
|\zeta|^2(1+|\tilde z|^2)=1.
\end{equation}
Consequently, each $\tilde Q^0_{p,q}$ coincides on $\tilde S^3$ with the function
\begin{equation}
R_{p,q}(\tilde z,\bar{\tilde z};\zeta,\bar\zeta)=\frac{q_{p,q}(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^q}\,\zeta^{p-q},
\end{equation}
and $\tilde Q^0$ coincides with $R$, where
\begin{equation}\Label{R1}
R=\sum_{k=4}^m\sum_{l=4}^kR_{l,k-l}=\sum_{k=4}^m\sum_{l=4}^k\frac{q_{l,k-l}(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{k-l}}\,\zeta^{2l-k},
\end{equation}
a rational function in $\zeta$ with coefficients that are rational functions in $\tilde z$ and $\bar{\tilde z}$. Note that the powers of $\zeta$ range from $8-m$ to $m$. Let us collect terms of equal powers in $\zeta$ and rewrite $R$ in \eqref{R1} in the form
\begin{equation}\Label{R2}
R=\sum_{r=8-m}^m \frac{b_r(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{s_r}}\,\zeta^r
=\frac{1}{\zeta^{m-8}}\sum_{r=0}^{2m-8} \frac{b_{r+8-m}(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{s_{r+8-m}}}\,\zeta^r,
\end{equation}
where the $s_r$ are (easily computable but not important) positive integers, and the $b_r$ are polynomials in $(\tilde z,\bar{\tilde z})$.
Now, by definition of the set $\Pi_m$, we have $\rho'\in \Pi_m$ precisely when $R$ as a rational function in $\zeta$ has at least one root on the circle \eqref{tildeS^3} for every $\tilde z\in U_0\subset {\4P}^1$. Observe that that set $B_k$ of coefficients $a=(a_0,\ldots,a_k)\in {\4C}^{k+1}$ such that the polynomial $a_0+a_1\zeta+\ldots+a_k\zeta^k$ has a root on the unit circle forms a real-algebraic, Levi flat (singular) hypersurface. Thus, $\rho'\in \Pi_m$ translates into the condition that
\begin{equation}\Label{br1}
\left(\frac{b_{8-m}(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{s_{m-8}}},\frac{b_{9-m}(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{s_m+1/2}},\ldots,
\frac{b_m(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{s_m+(2m-8)/2}}\right )\in B_{2m-8},\quad \forall \tilde z\in U_0.
\end{equation}
By unraveling the construction of $R$, we note that if we expand $\rho'$ in the monomial basis $Z^I=z^\alpha w^\beta$ of $\mathcal P_m$, i.e.,
\begin{equation}
\rho'=\sum_{|I|+|J|\leq m}e_{I\bar J}Z^I\bar Z^J,\quad e_{I\bar J}=\overline{e_{J\bar I}},
\end{equation}
then the components in \eqref{br1}
$$
\frac{b_r(\tilde z,\bar{\tilde z})}{(1+|\tilde z|^2)^{s_r+r/2}}
$$
are linear in $e_{I\bar J}$ and $\overline{e_{I\bar J}}$. Consequently, we deduce from the above discussion and \eqref{br1} that $\Pi_m$ is a real-algebraic subvariety in $\mathcal P^{\bR}_m$.
To complete the proof of Proposition \ref{gencondi}, we must show that if $\mathcal A$ is as in the statement of the proposition, then the dimension of $\Pi_m\cap \mathcal A$ is strictly less than that of $\mathcal A$. For this, it suffices to show that $\Pi_m\cap\mathcal A\neq \mathcal A$. To this end, we compute $Q^0(z^p\bar w^p)$, for $p\geq 2$,
\begin{equation}
Q^0(z^p\bar w^p)=\bar L_0^4(p(p-1)z^{p-2}\bar w^{p+2})=(p+2)(p+1)p^2(p-1)^2z^{p+2}\bar w^{p-2},
\end{equation}
and similarly,
\begin{equation}
Q^0(w^p\bar z^p)=\bar L_0^4(p(p-1)w^{p-2}\bar z^{p+2})=(p+2)(p+1)p^2(p-1)^2w^{p+2}\bar z^{p-2}.
\end{equation}
Thus, if $\rho'$ is any polynomial in $\mathcal A$, resulting in the polynomial $R$ as in \eqref{R2}, then $\rho'+ez^p\bar w^p+\bar ew^p\bar z^p$, which is also in $\mathcal A$ for all $e\in {\4C}$, results in
\begin{equation}
R'=R+\frac{(e\tilde z^{p+2}+\bar e\bar {\tilde z}^{p-2})}{(1+|\tilde z|^2)^{p-2}}\, \zeta^4.
\end{equation}
From this we easily deduce that if $\rho'\in \Pi_m$, then $\rho'+ez^p\bar w^p+\bar ew^p\bar z^p$ will not be in $\Pi_m$ for $e\neq 0$; indeed, since
\begin{equation}
\tilde z\mapsto \frac{(e\tilde z^{p+2}+\bar e\bar {\tilde z}^{p-2})}{(1+|\tilde z|^2)^{p-2}},\quad e\neq 0,
\end{equation}
maps onto an open neighborhood of $0$ in ${\4C}$, this statement follows from the following simple observation:
\begin{Lem} If $p(\zeta)=\zeta^n+a_{n-1}\zeta^{n-1}+\ldots +a_0$ has a root on the unit circle, then the set of $b\in {\4C}$ such that $p(\zeta)+b\zeta^k$ has a root on the unit circle is a real-algebraic, possibly singular curve (real-algebraic variety of dimension one).
\end{Lem}
\begin{proof} Consider the (symmetric) finite polynomial mapping $\Phi\colon {\4C}^{n}\to {\4C}^{n}$ sending a collection of roots $\tau=(\tau_1,\ldots,\tau_{n})$ to the collection of coefficients $a=(a_0,\ldots, a_{n-1})$ of the polynomial
\begin{equation}
p(\zeta)=\zeta^n+a_{n-1}\zeta^{n-1}+\ldots +a_0:=(\zeta-\tau_1)\ldots(\zeta-\tau_n).
\end{equation}
Pick $p_0(\zeta)$ such that one its roots is on the unit circle, i.e., $a^0=\Phi(\tau^0)$ with $\tau^0$ in the Levi flat (singular) hypersurface $H=\cup_{j=1}^n H_j$, with $H_j:=\{\tau\colon |\tau_j|=1\}$. The polynomials $p_e(\zeta):=p_0(\zeta)+b\zeta^k$ correspond to points $a^b=a^0+(0,\ldots,b,\ldots,0)$ (with $b$ in the $(k+1)$th component) and hence their roots $\tau^b$ belong to the complex 1-dimensional subvariety $\Phi^{-1}(X_k)$, where $X_k$ denotes the complex curve $b\mapsto a^0+(0,\ldots,b,\ldots,0)$. We claim that $\Phi^{-1}(X_k)$ is not contained in $H$, which will prove the conclusion of the lemma. Indeed, $\Phi^{-1}(X_k)$ could only be contained in the Levi flat $H$ if it were contained in one of its leaves $\tau_j=c$, with $c$ constant, which is clearly impossible.
\end{proof}
As mentioned above, we have now shown that the real-algebraic subvariety $\Pi_m$ satisfies $\Pi_m\cap\mathcal A\neq \mathcal A$, which completes the proof of Proposition \ref{gencondi}.
\end{proof}
\subsection{Generic perturbations of almost circular type} Recall that a real hypersurface $M\subset {\4C}^{2}$ is called {\it circular} if $Z\in M$ implies $e^{it}Z\in M$ for all $e^{it}\in S^1$. For perturbations $M_\varepsilon$ of the sphere, as in \eqref{pert}, it is straightforward to verify that the $M_\varepsilon$ are circular for all sufficiently small $\varepsilon>0$ if and only if in the decomposition \eqref{rho'decomp} we have $\rho'_{p,q}=0$ for $|p-q|\neq 0$. It was shown in \cite{EDumb15} that compact, circular real hypersurfaces in ${\4C}^2$ always have umbilical points. Here we shall consider perturbations $M_\varepsilon$ that are {\it almost circular}, which we define to be those for which, in the decomposition \eqref{rho'decomp} of $\rho'$, we have $\rho'_{p,q}=0$ when $|p-q|\geq 4$; we also say that such $\rho'$ are almost circular. We easily observe that a polynomial $P=P(Z,\bar Z)$ is almost circular if and only if its Fourier coefficients $\hat P_k$ vanish for $|k|\geq 4$:
$$
\hat P_{k}(Z,\bar Z):=\frac{1}{2\pi}\int_{0}^{2\pi}P(e^{it}Z,e^{-it}\bar Z)e^{-ikt}dt=0,\quad |k|\geq 4.
$$
Recall that $\mathcal P_m$ denotes the space of all polynomials in $Z=(z,w)$ and $\bar Z$ of degree at most $m$. We shall denote by $\mathcal A\mathcal C_m$ the real subspace of those that are real-valued and almost circular. Thus, $\rho'\in\mathcal P_m$ belongs to $\mathcal A\mathcal C_m$ when $\rho'$ is real-valued (i.e., $\rho'\in \mathcal P^{\bR}_m$) and $\rho'_{p,q}=0$ for $|p-q|\geq 4$. We note that $\mathcal A=\mathcal A\mathcal C_m$ satisfies the hypothesis in Proposition \ref{gencondi} for all $m\geq 2$ and with any $2\leq p\leq m/2$.
\begin{Thm}\Label{ACpertThm}
For $m\geq 4$, there is a real-algebraic subvariety $\Xi_m\subset \mathcal A\mathcal C_m$ of dimension strictly less than that of $\mathcal A\mathcal C_m$ such that if $\rho'\in \mathcal A\mathcal C_m\setminus\Xi_m$, then, for sufficiently small $\varepsilon>0$, the set of umbilical points $\mathcal U$ on the perturbation $M_\varepsilon$, given by \eqref{pert}, contains either points of dimension $\geq 2$ or a curve of stable umbilical points.
\end{Thm}
\begin{proof} We shall let $\Xi_m$ be $\Xi_m:=\Pi_m\cap \mathcal A\mathcal C_m$, where $\Pi_m$ is as defined in Proposition \ref{gencondi}. As noted above, $\mathcal A=\mathcal A\mathcal C_m$ satisfies the hypotheses in Proposition \ref{gencondi} and, hence, we conclude that $\Xi_m$ is a real-algebraic subvariety of strictly lower dimension that $\mathcal A\mathcal C_m$. The conclusion of Theorem \ref{ACpertThm} now follows from Proposition \ref{gengenpert'}, since $\rho'\in \mathcal A\mathcal C_m$ clearly guarantees that condition (ii) in that proposition holds; indeed, for $\rho'_{p,q}$, we have $Q^0(\rho'_{p,q})=Q^0_{p+2,q-2}$ and if $|p-q|\leq 3$, then $l=p+2\geq(p+q+1)/2>k/2$, which is the requirement in condition (ii).
\end{proof}
\begin{Rem}\Label{ACrem}
\begin{itemize}
\item
Recall that if, for example, $m=2p$ and
$$
\rho'=\rho'_{p-1,p+1}+\rho_{p,p}+\rho'_{p+1,p-1},\quad \rho'_{p+1,p-1}=\overline{\rho'_{p-1,p+1}}
$$
($\implies \rho'\in \mathcal A\mathcal C_m$), then
$$
Q^0=Q^0(\rho')=Q^0_{p+1,p-1}+Q^0_{p+2,p-2}+Q^0_{p+3,p-3}.
$$
We note that there are plenty of polynomials of this form,
$$
Q=Q_{p+1,p-1}+Q_{p+2,p-2}+Q_{p+3,p-3},
$$
such that $\pi(\mathcal V)={\4P}^1$, where $\mathcal V$ denotes the zero locus of $Q$ in $\tilde S^3$. For example, any $Q$ of the form
$$
Q=(z+\bar z)(Q'_{p-1,p}+Q'_{p,p-1})
$$
will satisfy this, as the reader can easily verify. However, we do not know any non-trivial examples of such $Q$ that are also in the image of the linear map $Q^0$, i.e., of the form $Q=Q^0(\rho')$ with $\rho'\in\mathcal A\mathcal C_m$.
\item It is clear from the calculations in Section \ref{EllSec} that the real ellipsoids $E_\varepsilon$ are not generic in the sense of Theorem \ref{ACpertThm}, i.e., these belong to $\Pi_m$.
\end{itemize}
\end{Rem}
\def$'${$'$}
|
\section{Introduction}\label{sec:intro}
Soil moisture is a key factor in climate systems, which has a significant impact on hydrological processes, runoff generations and drought developments. To understand its spatial variability and predict values at unsampled locations, Gaussian process models are widely used \citep{Stein:1999}, where likelihood based methods are appropriate for model fitting. However, it generally requires $O(n^3)$ computations and $O(n^2)$ memory for $n$ irregularly spaced locations \citep{Sun:Stein:2014}. Similar to other climate variables, many satellite-based or numerical model generated soil moisture datasets have nearly a global coverage with high spatial resolutions, so that the exact computation of Gaussian likelihood becomes prohibitive. There are various existing methods, many of which were discussed by \citet{Sun:Li:Genton:2012}. For example,
covariance tapering \citep{Furrer:Genton:Nychka:2006,Kaufman:Schervish:Nychka:2008,Sang:Huang:2012} assumes a compactly supported covariance function, which leads to a sparse covariance matrix; low rank models, including
space-time Kalman filtering \citep{Wikle:Cressie:1999},
low rank splines \citep{Lin:Wahba:Xiang:Gao:Klein:Klein:2000},
moving averages \citep{Hoef:Cressie:Barry:2004},
predictive processes \citep{Banerjee:Gelfand:Finley:Sang:2008} and
fixed rank kriging \citep{Cressie:Johannesson:2008}, make use of a latent process with a lower dimension where the resulting covariance matrix has a low rank representation; and Markov random field models \citep{Cressie:1993,Rue:Tjelmeland:2002,Rue:Held:2005,Lindgren:Rue:Lindstrom:2011} exploit fast-approximated conditional distributions assuming conditional independence with the precision matrix being sparse. These methods use models that may allow exact computations to reduce computations and/or storage, and each has its strength and weakness. For instance, \citet{Stein:2013} studied the properties of the covariance tapers and showed that covariance tapering sometimes performs even worse than assuming independent blocks in the covariance; \citet{Stein:2014} discussed the limitations on the low rank approximations; and Markov models depend on the observation locations, and realignment to a much finer grid with missing values is required for irregular locations \citep{Sun:Stein:2014}. Recently developed methods include the
nearest-neighbor Gaussian process model \citep{Datta:Banerjee:Finley:Gelfand:2015}, which
is used as a sparsity-inducing prior within a Bayesian hierarchical modeling framework,
the multiresolution Gaussian process model \citep{Nychka:Bandyopadhyay:Hammerling:Lindgren:Sain:2015}, which constructs basis functions using compactly supported correlation function on different level of grids,
equivalent kriging \citep{Kleiber:Nychka:2015}, which uses an equivalent kernel to approximate the kriging weight function when a nontrivial nugget exists, and multi-level restricted Gaussian maximum likelihood method \citep{Candas:Genton:Yokota}, for estimating the covariance function parameters using contrasts.
An alternative way to reduce computations is via likelihood and score equation approximations. \citet{Vecchia:1988} first proposed to approximate the likelihood using the composite likelihood method, where the conditional densities were calculated by choosing only a subset of the complete conditioning set. \citet{Stein:Chi:Wetly:2004} adapted this method for restricted maximum likelihoods approximation. Instead of approximating the likelihood itself, \citet{Sun:Stein:2014} proposed new unbiased estimating equations for score equation approximation, where the sparse precision matrix approximation is constructed by a similar method. In these approximation methods, the exact likelihood and the score equations can be obtained by using the complete conditioning set to calculate each conditional density. It was shown that the approximation quality or the statistical efficiency depends on the selected size of the subset. It is common that the subset is still inadequate by considering the largest possible number of nearest neighbors, which motivates this work.
In this paper, we propose a generalized hierarchical low rank method for likelihood approximation. The proposed method utilizes low rank approximations hierarchically, which does not lead to a low rank covariance matrix approximation. Therefore, it is different from the predictive process method \citep{Banerjee:Gelfand:Finley:Sang:2008}, where the covariance matrix is approximated by a low rank representation. Furthermore, the proposed method contains the independent blocks \citep{Stein:2013} and nearest neighbors \citep{Sun:Stein:2014} approaches as special cases. The improvement of the proposed method is explored theoretically and the performance is investigated by numerical and simulation studies. We show that the hierarchical low rank approximation significantly improves the statistical efficiency of the most commonly used methods while retaining the computational efficiency, especially when the size of conditional subsets is restricted by the computational capacity, which is always the case for real datasets. For illustrations, our method is applied to a large real-world spatial dataset of soil moisture in the Mississippi River basin, U.S.A., to facilitate a better understanding of the hydrological process and climate variability. Our method is able to fit a Gaussian process model to 2 million measurements with fast computations, making it practical and attractive for very large datasets.
\section{Methodology}\label{sec:meth}
\subsection{Approximating likelihoods}\label{subsec:likeli}
Let $\{z(s):s\in
D\subset\mathbb{R}^d\}$ be a stationary isotropic Gaussian Process in a domain $D$
in the $d$-dimensional Euclidean space, and typically $d=2$.
We assume the mean of the process is zero for simplicity and the covariance function has a parametric form $C(h;\theta)=\hbox{cov}\{z(s),z(s')\}$, where $h=\|s-s'\|$ and $\theta$ is the parameter vector of length $p$.
Suppose that data are observed at $n$ irregularly spaced locations $s_1,\ldots,s_n$, then,
\[
Z=(z_1,\ldots,z_n)^{ \mathrm{\scriptscriptstyle T} }\sim N(0,\Sigma(\theta)),
\]
where $z_i=z(s_i), i=1,\ldots,n$, and $\Sigma(\theta)$ is the variance-covariance matrix with the $(i,j)^\textrm{th}$ element $C(\|s_i-s_j\|;\theta).$ For simplicity, $\theta$ is omitted in notations hereinafter unless clarification is needed.
The maximum likelihood estimate can be obtained by maximizing the log-likelihood,
\[
\ell(\theta\mid Z)=\hbox{log}\{f(Z\mid\theta)\}=-\frac{1}{2}\hbox{log}(|\Sigma|)
-\frac{1}{2}Z^{ \mathrm{\scriptscriptstyle T} }\Sigma^{-1}Z-\frac{n}{2}\hbox{log}(2\pi),
\]
where $f$ is the multivariate normal density. In practice, if the mean of $Z$ is a vector that depends linearly on unknown parameters, the restricted maximum likelihood estimate should be employed \citep{Stein:Chi:Wetly:2004}.
When computations become prohibitive, one way to approximate the likelihood is through log-conditional densities,
\[
\ell(\theta\mid Z)=\hbox{log}\{f(z_1\mid\theta)\}+\sum_{j=1}^{n-1}\hbox{log}\{
f(z_{j+1}\mid Z_j,\theta)\},
\]
where $Z_{j}=(z_1,\ldots,z_{j})^{ \mathrm{\scriptscriptstyle T} },$ for $1\leq j\leq n-1,$ indicating all the ``past" observations of $z_{j+1}$.
Since,
\[
\hbox{cov}\Biggl(
\begin{array}{c}
Z_{j} \\
z_{j+1} \\
\end{array}
\Biggr)=\Biggl(
\begin{array}{cc}
\Sigma_{jj} & \sigma_{j} \\
\sigma_{j}^{ \mathrm{\scriptscriptstyle T} } & \sigma_{j+1,j+1} \\
\end{array}
\Biggr),
\]
it is easy to show that for $j=1,\ldots,n-1$,
\begin{equation}\label{eq:cond}
\hbox{log} \{f(z_{j+1}|Z_{j})\}=-\frac{1}{2}\Biggl\{\frac{\bigl(z_{j+1}-\sigma^{ \mathrm{\scriptscriptstyle T} }_{j}\Sigma_{jj}^{-1}Z_{j}\bigr)^2}{\sigma_{j+1,j+1}-\sigma^{ \mathrm{\scriptscriptstyle T} }_{j}\Sigma_{jj}^{-1}\sigma_{j}}
+\hbox{log}\bigl(\sigma_{j+1,j+1}-\sigma^{ \mathrm{\scriptscriptstyle T} }_{j}\Sigma_{jj}^{-1}\sigma_{j}\bigr)
+\hbox{log}(2\pi)\Biggr\},
\end{equation}
which is the log-density of $w_j=b_j^{ \mathrm{\scriptscriptstyle T} } Z,$ where $b_j=(-\sigma_{j}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}^{-1},1,0,\ldots,0)^{ \mathrm{\scriptscriptstyle T} }.$
It can be shown that $w_j$s are independent and $w_j\sim N(0,v_j),$ where $v_j=b_j^{ \mathrm{\scriptscriptstyle T} }\Sigma b_j$ \citep{Stein:Chi:Wetly:2004}. \citet{Sun:Stein:2014} further showed that the precision matrix is $\Sigma^{-1}=\sum_{j=0}^{n-1}b_jb_j^{ \mathrm{\scriptscriptstyle T} }/v_j$, where $b_0=(1,0,\ldots,0)^{ \mathrm{\scriptscriptstyle T} }$ and $v_0=b_0^{ \mathrm{\scriptscriptstyle T} }\Sigma b_0$.
More generally, $z_{j+1}$ can be defined as a vector which is usually more computationally efficient, and the corresponding $b_j=(-\sigma_{j}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}^{-1},I,0,\ldots,0)^{ \mathrm{\scriptscriptstyle T} },$ where $I$ is an identity matrix of size $j$.
However, for a large $j$, it is computationally expensive to evaluate $\Sigma_{jj}^{-1}\sigma_{j}.$ \citet{Vecchia:1988} proposed approximating each conditional density by only conditioning on a subset $z_{j+1}$ consisting of $r\ll j$ nearest neighbors. The same approach is used by \citet{Stein:Chi:Wetly:2004}
for approximating the restricted maximum likelihood estimate. \citet{Sun:Stein:2014} also used the subset of nearest neighbors to approximate the precision matrix for score equation approximation.
In this paper, we propose a generalized framework that allows to approximate these conditional densities hierarchically using a low rank representation. Although we implement our algorithm for application in \S\ref{sec:app} with $z_{j+1}$ being a vector, we present and illustrate our methodology assuming $z_{j+1}$ is scalar for simplicity.
\subsection{Hierarchical low rank representation}\label{subsec:approx}
Motivated by the nearest neighbors method, where only $r\ll j$ nearest neighbors are selected to approximate $\Sigma^{-1}_{jj}\sigma_{j}$ for a large $j$ in equation~\eqref{eq:cond}, we propose a general approximation framework for $j>r$ using a low rank representation.
Denote $\Sigma^{-1}_{jj}\sigma_{j}$ by $x_j,$ or $\Sigma_{jj}x_j=\sigma_j$.
We propose to approximate $x_j$ by a low rank representation $\hat x_j=A_{j,r}\tilde x_j,$ where $\tilde x_j$ is a vector of length $r$ and $A_{j,r}$ is a $j\times r$ matrix. Then, instead of solving $\Sigma_jx_j=\sigma_j$, we minimize the norm $\|\Sigma_{jj}A_{j,r}\tilde x_j-\sigma_j\|_{\Sigma{_{jj}^{-1}}}=(\Sigma_{jj}A_{j,r}\tilde x_j-\sigma_j)^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}^{-1}(\Sigma_{jj}A_{j,r}\tilde x_j-\sigma_j)$ or equivalently solve $A_{j,r}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,r}\tilde x_j=A_{j,r}^{ \mathrm{\scriptscriptstyle T} }\sigma_j.$ Therefore, $x_j$ is approximated by,
\begin{equation}\label{eq:approx}
\hat x_j=A_{j,r}\tilde x_j=A_{j,r}(A_{j,r}^{ \mathrm{\scriptscriptstyle T} } \Sigma_{jj}A_{j,r})^{-1}A_{j,r}^{ \mathrm{\scriptscriptstyle T} }\sigma_{j},
\end{equation}
which only involves a linear solve of dimension $r$.
In this framework, we approximate $x_j$ for each $j>r$ hierarchically by a low rank representation,
which includes many commonly used strategies as special cases with different choices of $A_{j,r}$. The following are some examples:
\begin{example}
Independent blocks method (IND). In this
method, no correlation between ``past" points and the ``current" point is
considered. Namely, $A_{j,r}$ is a $0$ matrix; however, $z_{j+1}$ is a vector of length $r$ here for fair comparison to other methods in terms of computation.
\end{example}
\begin{example}
Nearest neighbors method (NN). Choose $r$ nearest neighbors of $z_{j+1}$ from $Z_j$. The corresponding $A_{j,r}$ is of $j\times r$ dimensions, where each column consists of only one element $1$ at the $k$-th row if $z_k$ is selected from $Z_j$ and zero otherwise.
\end{example}
\begin{example}
Nearest neighboring sets method (SUM). Choose $r$ nearest neighboring sets of $z_{j+1}$, where each set contains $m>1$ neighbors and a total of $mr\ll j$ neighbors are selected from $Z_j$. The matrix $A_{j,r}$ is specified as a $j\times r$ matrix with each column having $m$ elements of $1$, indicating the sum of the $m$ selected neighbors are considered. In this way, more neighbors are included while the computational cost remains the same.
\end{example}
\begin{example}
Nearest neighbors and nearest neighboring sets method (NNSUM). Combine Examples~2 and 3, where $r_1$ columns of $A_{j,r}$ are constructed as in Example~2, and $r-r_1$ are built as in Example~3. In this way, we use the exact information from the $r_1$ nearest neighbors and consider $r-r_1$ nearest neighboring sets with a total number of $r_1+m(r-r_1)$ selected nearest neighbors.
\end{example}
\subsection{Hierarchical low rank approximation method}\label{subsec:HLR}
In this section, we propose a generalized hierarchical low rank approximation method (HLR). In equation~\eqref{eq:approx}, the matrix $A_{j,r}$ is a $0$-$1$ matrix. The $r\times r$ matrix $A_{j,r}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,r}$ only extracts the corresponding rows or columns of $\Sigma_{jj}$. Now suppose we select $mr$ nearest neighbors of $z_{j+1}$, and the corresponding $A_{j,mr}$ is of size $j\times mr.$ To retain the same computational costs associated with rank $r$, we propose the following approximation,
\begin{equation}
\label{eq:nugget}
A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,mr}\approx P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }+\epsilon_j^2I_{mr},
\end{equation}
where $L_{j}$ is a positive definite matrix of dimension $r\times r,$ $P_j$ is a
$mr\times r$ matrix consisting of $r$ basis functions, $I_{mr}$ is the identity matrix of size $mr$, and $\epsilon_j^2$ is the
nugget.
By the Sherman--Morrison--Woodbury formula,
\begin{equation}\label{eq:HLR}
(P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }+\epsilon_j^2I_{mr})^{-1}= \epsilon_j^{-2}I_{mr}-\epsilon_j^{-4}P_j(L_{j}^{-1}+\epsilon_j^{-2}P_j^{ \mathrm{\scriptscriptstyle T} } P_j)^{-1}P_j^{ \mathrm{\scriptscriptstyle T} },
\end{equation}
then $(A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,mr})^{-1}$ in equation \eqref{eq:approx} can be approximated by inverting only an $r\times r$ matrix $L_{j}$. This approach shares the same spirits with the predictive process \citep{Banerjee:Gelfand:Finley:Sang:2008} and fixed rank kriging \citep{Cressie:Johannesson:2008}.
However both methods approximate the covariance function by a low rank representation while the low rank approximation is done for each $j>r$ hierarchy in our method, and the resulting approximated covariance is no longer low rank. The detailed choice for $P_j$ were discussed by \citet{Cressie:Johannesson:2008}. In this paper, we use the eigenfunctions.
\begin{figure}[b!]
\centering
\includegraphics[width=0.8\textwidth]{method.eps}
\caption{\label{fig:method}A random field where $n=5$ locations have observations}.
\end{figure}
To help comprehend, Fig.~\ref{fig:method} illustrates the methods described in \S\ref{subsec:approx} and \ref{subsec:HLR} for $n=5$ observations $Z=(z_1,\ldots,z_5)^{ \mathrm{\scriptscriptstyle T} }$. Let $r=2$, then IND considers $3$ independent blocks, and $f(Z)$ is approximated by
$f(z_5)f(z_4,z_3)f(z_2,z_1)$. For the other four methods, the conditional density is required to calculate in each hierarchy. For instance, in hierarchy $j=5$, NN approximates the conditional density $f(z_5\mid z_4,\ldots,z_1)$ by $f(z_5\mid z_4,z_3)$; SUM by $f(z_5\mid z_4+z_3,z_2+z_1)$; NNSUM by $f(z_5\mid z_4,z_3+z_2)$; and HLR by
$f(z_5\mid a_{14}z_4+a_{13}z_3+a_{12}z_2+a_{11}z_1,
a_{24}z_4+a_{23}z_3+a_{22}z_2+a_{21}z_1)$, where $a_{ij}$s are determined by the low rank approximation.
\subsection{Assessing model quality}\label{subsec:quality}
There are various ways to measure the performance of approximation methods, including the Kullback--Leibler divergence, the Godambe information matrix, and the Frobenius norm.
The Kullback--Leibler divergence computes the divergence of the approximated from the exact distributions.
For the zero-mean Gaussian process, the Kullback--Leibler divergence has the closed form,
\[
D_\textsc{K-L}(N_e\|N_a)=\frac{1}{2}\left\{\hbox{tr}(\Sigma_a^{-1}\Sigma_e)+\hbox{log}(|\Sigma_a|)
-\hbox{log}(|\Sigma_e|)-n\right\},
\]
where $N_e$ and $N_a$ stand for the exact and the approximated distributions, respectively, $\Sigma_e$ and $\Sigma_a$ are the corresponding covariance matrices, and $n$ is the dimension of the distribution.
The Godambe information matrix gives the asymptotic variances and covariances for the estimated parameters in the Gaussian process, as used by \citet{Kaufman:Schervish:Nychka:2008} and \citet{Sun:Stein:2014}.
The Frobenius norm is another way to think about this problem. However, it is a matrix norm and does not penalize the positive definiteness of a covariance matrix~\citep{Stein:2014}.
For our numerical and simulation studies in \S3, we choose the Kullback--Leibler divergence and the Godambe Information matrix to assess the quality of the approximation. Because the results in terms of showing the different performances are similar, we only present the results of Kullback--Leibler divergence. It will be shown numerically that the Kullback--Leibler divergence of the hierarchical low rank approximation method is always the smallest.
This is due to the fact that for each $j>r$, the hierarchical low rank approximation method provides a better approximation in equation~\eqref{eq:approx} by including more neighbors than the nearest neighbors method. Let $V^\textsc{N}_{jj}$ be the $r\times r$ matrix defined by $A_{j,r}^{ \mathrm{\scriptscriptstyle T} } \Sigma_{jj}A_{j,r}$ in equation~\eqref{eq:approx} using the nearest neighbors method and let $V^\textsc{H}_{jj}=P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }+\epsilon_j^2I_{mr}$ be the $mr\times mr$ matrix for approximating $A_{j,mr}^{ \mathrm{\scriptscriptstyle T} } \Sigma_{jj}A_{j,mr}$ in equation~\eqref{eq:nugget} by the hierarchical low rank approximation method, where $P_j$ consists of eigenfunctions. The following theorem shows the result that the approximation to $\Sigma_{jj}$ induced by $V^\textsc{H}_{jj}$ is better than that induced by $V^\textsc{N}_{jj}$ in terms of the Frobenius norm.
\begin{theorem}
\label{theorem}
Let $\lambda_1\geq\lambda_2\geq\cdots\geq\lambda_{mr}>0$ be the eigenvalues of $A_{j,mr}^{ \mathrm{\scriptscriptstyle T} } \Sigma_{jj}A_{j,mr}$. If $\epsilon_j^2$ in equation~\eqref{eq:nugget} satisfies
$\epsilon_j^2<(\lambda_r+\lambda_{mr})/2$,
we have,
\[
\|A_{j,mr}V^\textsc{H}_{jj}A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }-\Sigma_{jj}\|_F
\leq
\|A_{j,r}V^\textsc{N}_{jj}A_{j,r}^{ \mathrm{\scriptscriptstyle T} }-\Sigma_{jj}\|_F,
\]
where $\|\cdot\|_F$ means the Frobenius norm.
\end{theorem}
The proof is shown in the Appendix. Similar results hold for the comparison between hierarchical low rank approximation method and the nearest neighboring sets method, or the nearest neighbors and nearest neighboring sets method.
\subsection{Computational complexity and parallelization}\label{subsec:comp}
For our hierarchical low rank approximation method, we need to execute a linear solve of dimension $r$, which requires $O(\min(j,r)^3)$ computation in equation~\eqref{eq:HLR} for each hierarchy $j=1,\ldots,n-1$ assuming that the direct method is employed. Then the total computational cost is $O(r^3n)$ for likelihood approximation per value. When $r\ll n$, the computational cost is much smaller than $O(n^3)$, which is required by the Cholesky decomposition.
In practice, the computation time can be reduced further by choosing $z_{j+1}$ as a vector due to the fact that it leads to a smaller number of hierarchies that need to be evaluated.
It is also worth noting that our approach can be parallelized easily because the computation of each hierarchy is independent of each other.
\section{Numerical study}\label{sec:num}
\subsection{Design setup}
In the numerical study in this section and the following simulation study in \S\ref{sec:sim}, we focus on irregularly
spaced data with an unstructured covariance matrix \citep{Sun:Stein:2014}. The observations are generated
at the locations $n^{-1/2}(r-0.5 + X_{r\ell},\ell-0.5 +
Y_{r\ell})$
for $r,\ell\in\{1,\ldots,n^{1/2}\}$, where $n$ is the number of locations, and $X_{r\ell}$s and
$Y_{r\ell}$s are independent and identically distributed, uniform on $(-0.4, 0.4).$
The advantage of this design is that it is irregular, and we can guarantee that no two locations are too close.
Here,
we study the performances of different approximation methods proposed
in \S\ref{subsec:approx} and \S\ref{subsec:HLR} in different settings.
We consider a zero-mean Gaussian process model with Mat\'ern covariance function possibly with a nugget,
\begin{equation}\label{eq:matern}
C(h;\alpha,\beta,\nu,\tau^2)=
\alpha
\{(2\nu)^{1/2}h/\beta\}^\nu
K_\nu\{{(2\nu)^{1/2}}h/\beta\}
/
\{\Gamma(\nu)2^{\nu-1}\}+\tau^2\mathbbm{1}(h=0),
\end{equation}
where $K_\nu(\cdot)$ is the modified Bessel function of the second kind of order $\nu$, $\Gamma(\cdot)$ is the gamma function, $\mathbbm{1}(\cdot)$ is the indicator function, $h\geq 0$ is the distance between two locations, $\alpha>0$ is the sill parameter, $\beta>0$ is the range parameter, $\nu>0$ is the smoothness parameter, and $\tau^2$ is the nugget effect.
For $n$ irregularly spaced locations, the description of the five methods considered is shown in Table~\ref{tab:method}.
\begin{table}[h]
\centering
\def~{\hphantom{0}}
\caption{Description of the five methods used in the numerical study. IND, independent blocks method; NN, nearest neighbors method; SUM, nearest neighboring sets method; NNSUM, nearest neighbors and nearest neighboring sets method; HLR, the hierarchical low rank approximation method.}
\begin{tabular}[c]{l|p{.7\textwidth}}
\hline
Method&Description\\\hline
IND& Divide the locations into $\lceil n/r\rceil$ blocks, each of which contains $r$ points. $\lceil n/r \rceil$ means the largest integer that is no larger than $n/r$.
\\
NN& A number of $r$ nearest neighbors are selected to construct $A_{j,r}.$
\\
SUM& A number of $r$ nearest neighboring sets are selected and each set has 2 locations. Then a total number of $2r$ nearest neighbors are used to construct $A_{j,r}$.
\\
NNSUM& A number of $\lceil
r/2\rceil$ nearest neighbors are first selected, then the following $2(r-\lceil
r/2\rceil)$ nearest neighbors are divided into $r-\lceil
r/2\rceil$ sets of size $2$.
\\
HLR& A number of $2r$ nearest neighbors are considered, where $L_{j}$ is a $r\times r$ diagonal matrix with elements corresponding to the $r$ leading eigenvalues. $P$ consists of the $r$ corresponding eigenvectors.
\\
\hline
\end{tabular}
\label{tab:method}
\end{table}
In \S\ref{subsec:depend}--\ref{subsec:noise}, we present the Kullback--Leibler divergence calculated from different settings for the five methods with $\alpha$ fixed at $1$ and $n=900$. In \S\ref{subsec:further}, we discuss the effect of sample size $n$ and the rank $r$.
\subsection{Dependence level}\label{subsec:depend}
In the Mat\'ern model in equation~\eqref{eq:matern}, the range parameter $\beta$ controls the dependence of the process. In this section, we consider different $\beta$.
Given $\nu=0.5$, which corresponds to an exponential covariance function and $\tau^2=0.15$, the first row of Fig.~\ref{fig:numericalStudy} shows the Kullback--Leibler divergence for $\beta=0.1$, which means a weaker dependence, and $\beta=0.5$, which indicates a stronger dependence, as the rank $r$ increases from $2$ to $8$. We can see that the HLR approximation is always the best with the smallest Kullback--Leibler divergence, and SUM and NNSUM win against NN only when $r=2$ for $\beta=0.1$, while for $\beta=0.5$, the improvement of SUM and NNSUM exists up to $r=6$.
It implies that when a strong correlation is present, a small number of nearest neighbors is not adequate to
provide a good approximation of the conditional density.
It is also worth noting that the range of $r/n$ in this study is from $0.22\%$ to $0.89\%$. For very large $n$, and $r\ll n$, the improvement from HLR, SUM or NNSUM approaches can be substantial.
\subsection{Smoothness level}
In the Mat\'ern covariance function, a larger $\nu$ indicates a smoother process.
In this section, we fix $\beta=0.5$ and $\tau^2=0.$ We consider two smoothness levels with $\nu=0.5$ and $\nu=1,$ which correspond to the exponential and Whittle covariance functions, respectively.
The second row of Fig.~\ref{fig:numericalStudy} shows the Kullback--Leibler divergence.
Similarly, the HLR approach outperforms the other methods. For the rougher process, when $\nu=0.5$, SUM and NNSUM are slightly better than NN at $r=2.$ When $\nu$ increases to $1$, the improvement almost disappears and all the methods need a large $r$ to achieve similar performances as $\nu=0.5.$
\begin{figure}[p!]
\centering
\includegraphics[width=0.8\textwidth]{numericalStudy.eps}
\vskip -0.7cm
\caption{\label{fig:numericalStudy}Six panels showing the Kullback--Leibler divergence against rank with 900 locations in IND (long-dash, $---$), NN (dot-dash, $\cdot$ - $\cdot$), SUM (dashes, - - -), NNSUM (dots, $\cdot\cdot\cdot$), and HLR (solid, ---) methods. The corresponding parameters are indicated in the titles.}
\end{figure}
\subsection{Noise level}\label{subsec:noise}
The nugget effect can be viewed as measurement errors or the micro-structure in the underlying process. In this section, we consider different $\tau^2$. Given $\beta=0.25$ and $\nu=0.5$, the last row of Fig.~\ref{fig:numericalStudy} shows the Kullback--Leibler divergence for $\tau^2=0$ and $\tau^2=0.15.$ In both cases, the HLR approach still provides the best approximation, although for large $\tau^2,$ a larger $r$ is needed. If the rank $r$ is limited to a small number, we can see that SUM or NNSUM can improve NN when the process is noisy or with a larger $\tau^2.$
\subsection{Sample size and rank}\label{subsec:further}
In this section, we explore the effect of sample size given the rank $r$ or the ratio of $r/n$. Fig.~\ref{fig:range50} shows the results for a similar design as in the first row of Fig.~\ref{fig:numericalStudy} but with $n=2500$.
Comparing Fig.~\ref{fig:range50} to the first row of Fig.~\ref{fig:numericalStudy}, we can see that for a given process, a larger number of locations does require larger ranks to achieve a similar approximation quality. When $r$ is fixed, NN is often not adequate, especially for large $n$, and SUM, NNSUM, and HLR can improve the approximation by including more neighbors.
\begin{figure}[b!]
\centering
\includegraphics[width=0.8\textwidth]{range50.eps}
\caption{\label{fig:range50}Two panels showing the Kullback--Leibler divergence against rank with 2500 locations in IND (long-dash, $---$), NN (dot-dash, $\cdot$ - $\cdot$), SUM (dashes, - - -), NNSUM (dots, $\cdot\cdot\cdot$), and HLR (solid, ---) methods. The corresponding parameters are indicated in the titles.}
\end{figure}
Although it is not realistic for a large dataset, we also investigate a situation where NN is adequate to provide a good approximation at rank $r$, and then compare the Kullback--Leibler divergence for NN at $r+1$ and NNSUM with the same first $r$ nearest neighbors and one additional set containing the next $2$ nearest neighbors. We find that for $\alpha=1, \beta=0.5, \nu=0.5, \tau^2=0$ and $n=900$, NN with rank $r+1=51$ gives a Kullback--Leibler divergence as $9.4\times10^{-2}$ and NNSUM reduces Kullback--Leibler divergence by $1\%.$
\section{Simulation study}\label{sec:sim}
In \S\ref{sec:num}, we calculated the Kullback--Leibler divergence at the true parameter values. In this section, we generate $n=900$ observations with parameters $\alpha=1, \beta=0.1, \nu=0.5$ and $\tau^2=0.15.$ We run the optimization for $\alpha,\beta,\tau^2$ while fixing $\nu$ at the true value and obtain the estimates of $\alpha,\beta, \tau^2$ by maximizing the approximated likelihoods with $r=2$. We repeat the estimates procedure $500$ times and the boxplots of $\alpha$ and $\beta$ are shown in Fig.~\ref{fig:simulation}.
\begin{figure}[h!]
\centering
\includegraphics[width=0.8\textwidth]{simulation.eps}
\caption{\label{fig:simulation}Two panels showing the boxplot of parameter estimates and mean squared error times 10 (left) or 1000 (right). The solid line is a reference for the true parameter value, and the dash line is the corresponding mean squared error of the $500$ number of estimates in each method. Left: illustration for estimated $\alpha$; Right: illustration for estimated $\beta.$}
\end{figure}
We see that the estimates obtained by the hierarchical low rank approximation method have the smallest mean squared error among the 3 approximation methods and are close to the exact maximum likelihood estimation.
\section{Application}\label{sec:app}
\subsection{Dataset description}
In this section, we apply our method to modeling soil moisture, a key factor in evaluating the state of the hydrological process, including runoff generation and drought development. We consider high-resolution daily soil moisture data at the top layer of the Mississippi basin, U.S.A., on January $1^\textrm{st}$, 2014 \citep{Chaney:Metcalfe:Wood:review}.
The spatial resolution is of $0.0083$ degrees. The grid consists of $1830\times1329=2,432,070$ locations with $2,153,888$ observations and $278,182$ missing values. The illustration of the data is shown in Fig.~\ref{fig:data}.
\begin{figure}[ht]
\centering
\includegraphics[width=0.9\textwidth]{data.eps}
\caption{\label{fig:data} Soil moisture (unit: percentage) at the top layer of the Mississippi basin, U.S.A. on January $1^\textrm{st}$, 2014. }
\end{figure}
We know that a one-degree difference in latitude along any longitude line is equivalent to $111$ km; however, the distance of one-degree difference in longitude depends on the corresponding latitude. As the range of the latitude in this region is relatively small, for simplicity, we use the distance of one-degree difference in longitude at the center location of the region to represent all others, which is $87.5$ km; namely, in this region, $1^\circ$ in latitude is $111$ km and $1^\circ$ in longitude is $87.5$ km.
To understand the structure of the day's soil moisture, we fit a Gaussian process model with a Mat\'ern covariance function.
From all the locations, we randomly pick $n=2,000,000$ points, which are irregularly spaced, to train our model. To assess the quality of our model, the fitted models can be used to predict part of the left out observations.
\subsection{Estimation and prediction}\label{subsec:estimation}
To use a Gaussian process model, we first fit a linear model to the longitude and latitude as the covariates to the soil moisture. After fitting, we find the negatively skewed residuals, hence we apply a logarithm transformation with some shift.
The histogram of the transformed residual is shown in the left panel of Fig.~\ref{fig:histNvario}, which does not show strong departure from Gaussianity. To examine the isotropy of this process, we calculate the directional empirical variograms as illustrated in the right panel of Fig.~\ref{fig:histNvario}. We see the variograms on the circle with the same radius to the origin have similar values, suggesting that it is reasonable to assume an isotropic model.
\begin{figure}[ht]
\centering
\includegraphics[width=0.8\textwidth]{histNvario.eps}
\caption{\label{fig:histNvario}Left: the histogram of the transformed residuals; Right: the image plot of the empirical variogram at different distances and along different directions.}
\end{figure}
Let $z(s)$ denote the transformed residual and the region $D$ be the set of the selected locations, then the proposed Gaussian process model here is
$\{z(s):s\in
D\subset\mathbb{R}^d\}
\sim
\textsc{GP}(0,C(h;\theta)).$ We choose three different covariance functions: the exponential, which has the smoothness parameter $\nu=0.5$; the Whittle, which has $\nu=1$; and the Mat\'ern covariance function, which has an unknown $\nu$.
The formula is given in equation~\eqref{eq:matern}.
Given that the $2,000,000$ observations follow $Z\sim N(0,\Sigma(\theta))$, $\Sigma(\theta)$ is the 2 million by 2 million variance-covariance matrix, obtained from the chosen covariance function. We use nearest neighbors and hierarchical low rank approximation methods with rank $r=60$ to get the approximated likelihood and then obtain the parameter estimates. The results are shown in Table~\ref{tab:estimate}.
The Mat\'ern covariance model is more flexible by allowing to estimate $\nu$. The estimated $\nu$ in the
Mat\'ern covariance model by both methods is smaller than $0.5$, and the estimated $\beta$ has the largest value. It suggests a rougher process with a larger dependence range compared to the estimated exponential covariance model.
The last row of Table~\ref{tab:estimate} shows the values of log-likelihood per observation. For each given covariance model, the likelihood with parameters estimated by the hierarchical low rank approximation method is always larger than that by the nearest neighbors method.
Among different covariance models, the likelihood with Mat\'ern covariance is the largest.
\begin{table}[h]
\def~{\hphantom{0}}
\centering
\caption{Parameter estimation results}
\small{
\begin{tabular}{l|ccc|ccc}
\hline
& \multicolumn{3}{c|}{Nearest neighbors}& \multicolumn{3}{c}{Hierarchical low rank approximation} \\
& Exponential & Whittle & Mat\'ern & Exponential & Whittle & Mat\'ern \\\hline
Estimated~ $\alpha$
& $~~1.0073~$ & $~~0.9787~$ & $~~1.0597~$
& $~~1.0065~$ & $~~0.9789~$ & $~~1.0539~$\\
Estimated~ $\beta$ (km)
& $~21.6115~$ & $~~5.9316~$ & $222.6545~$
& $~21.2944~$ & $~~5.8216~$ & $178.2051~$\\
Estimated~ $\tau^2$
& $~~0.0107~$ & $~~0.0013~$ & $~~0.0000~$
& $~~0.0096~$ & $~~0.0012~$ & $~~0.0001~$\\
Estimated~ $\nu$
& $~~0.5000~$ & $~~1.0000~$ & $~~0.2079~$
& $~~0.5000~$ & $~~1.0000~$ & $~~0.2214~$\\
\hline
log-likelihood$/n$
& $-0.1042~$ & $-0.1417~$ & $-0.0852~$
& $-0.0941~$ & $-0.1308~$ & $-0.0761~$\\
\hline
\end{tabular}
}
\label{tab:estimate}
\end{table}
The size of the problem in this application is in the millions, a dataset which is far beyond the ability of classic analysis methods.
However, nearest neighbors and hierarchical low rank approximation methods can evaluate the approximated likelihood at each iteration in the optimization procedure within 5 and 14 minutes, respectively. The fast computation makes it highly practical for applying the proposed methods to a large real-world spatial dataset problem. The experiment is performed with the Intel Xeon E5-2680 [email protected] processor.
Next, we use the fitted Mat\'ern model by the hierarchical low rank approximation method to predict soil moisture at the $1000$ left out locations by kriging, which is known to provide the best linear unbiased prediction as well as the prediction standard errors \citep{Cressie:1993}. However, the problem here is of size $n=2,000,000$, hence kriging cannot be employed directly, because it involves a linear solve of size $n$ \citep{Furrer:Genton:Nychka:2006}.
In fact, the proposed methods in this paper can be adopted for approximating kriging equations as well. But for the purpose of validating the fitted model, we explore the exact computation method by treating the irregularly spaced data as observations on a finer regular grid with missing values. The resulting covariance matrix has a block Toeplitz Toeplitz block structure, which can be embedded in a block circulant circulant block matrix \citep{Kozintsev:1999}. Then kriging can be done by fast Fourier transformation. More details can be found in \citet{Chan:Ng:1996}. The mean squared prediction errors over the $1000$ validation locations is $4.53\times10^{-5},$ which is notably small.
\section{Discussion}\label{sec:end}
The implementation in this paper was done with a single-thread program, however as aforementioned in \S\ref{subsec:comp}, computation in each hierarchy can be paralleled, which would reduce the computation time dramatically and make applications even more practical.
The proposed method can be also extended to more complicated settings. For example, although the rank was fixed to the same in each hierarchy, it can be chosen flexibly in accordance with the number of ``past" observations that are involved in the hierarchy, which, we believe, would give a better approximation.
Moreover, for prediction problems, the proposed method can be further investigated to approximate kriging equations for large irregularly spaced spatial datasets.
\newpage
\section*{Appendix}
Recall that the dimension of $V^\textsc{H}_{jj}$ is $mr\times mr$, $V^\textsc{N}_{jj}$ is $r\times r$, $A_{j,r}$ is $j\times r$, $A_{j,mr}$ is $j\times mr$, and $\Sigma_{jj}$ is $j\times j.$ Define $B$ to be the $mr\times r$ matrix by keeping the $mr$ selected rows from $A_{j,r}$, or $B=A_{j,mr}^{ \mathrm{\scriptscriptstyle T} } A_{j,r}$. Let $M$ denote $A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,mr}$.
The proof of Theorem~\ref{theorem} is as follows.
{\bf Proof of Theorem~\ref{theorem}}
Since the equation,
\[
\begin{array}{rl}
&\|A_{j,mr}V^\textsc{H}_{jj}A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }-\Sigma_{jj}\|^2_F
-
\|A_{j,r}V^\textsc{N}_{jj}A_{j,r}^{ \mathrm{\scriptscriptstyle T} }-\Sigma_{jj}\|^2_F\\
=&
\|A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }( A_{j,mr}V^\textsc{H}_{jj}A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }-\Sigma_{jj})A_{j,mr}\|^2_F
-
\|A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }(
A_{j,r}V^\textsc{N}_{jj}A_{j,r}^{ \mathrm{\scriptscriptstyle T} }-\Sigma_{jj})A_{j,mr}\|^2_F\\
=&\|V^\textsc{H}_{jj}-A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,mr}\|^2_F
-
\|BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-A_{j,mr}^{ \mathrm{\scriptscriptstyle T} }\Sigma_{jj}A_{j,mr}\|^2_F\\
=&\|V^\textsc{H}_{jj}-M\|^2_F\
-\|BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M\|^2_F,\\
\end{array}
\]
it suffices to show,
\[\|V^\textsc{H}_{jj}-M\|_F\
\leq\|BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M\|_F.\]
Noting that $V^\textsc{H}_{jj}=P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }+\epsilon_j^2I_{mr},$ we have $\|V^\textsc{H}_{jj}-M\|_F=\|P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }-(M-\epsilon_j^2I_{mr})\|_F$.
Since $\epsilon_j^2<(\lambda_r+\lambda_{mr})/2$, we know that the eigenvalues of $M-\epsilon_j^2I_{mr}$ satisfy $\lambda_1-\epsilon_j^2\geq\lambda_2-\epsilon_j^2\geq\cdots\geq\lambda_{r}-\epsilon_j^2$ and $|\lambda_{r}-\epsilon_j^2|\geq\max^{mr}_{k=r+1}(|\lambda_k-\epsilon_j^2|)$.
Thus, $|\lambda_1-\epsilon_j^2|\geq|\lambda_2-\epsilon_j^2|\geq\cdots\geq|\lambda_{r}-\epsilon_j^2|\geq\max^{mr}_{k=r+1}(|\lambda_k-\epsilon_j^2|)$. By the construction of $P_j$ and $L_j$, and Eckart-Young-Mirsky theorem \citep{Eckart:Young:1936,Mirsky:1960}, we know,
\[
\|P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }-(M-\epsilon_j^2I_{mr})\|_F
=\inf_{\hbox{rank}(X)\leq r}\|X-(M-\epsilon_j^2I_{mr})\|_F.
\]
Noting that the rank of $BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }$ is $r$, we have $\|V^\textsc{H}_{jj}-M\|_F=\|P_jL_{j}P_j^{ \mathrm{\scriptscriptstyle T} }-(M-\epsilon_j^2I_{mr})\|_F\leq\|BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-(M-\epsilon_j^2I_{mr})\|_F=\|(BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M)-\epsilon_j^2I_{mr}\|_F.$
It is easy to observe that the diagonal elements of $BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M$ is non-positive, thus $\|(BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M)-\epsilon_j^2I_{mr}\|_F\leq\|BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M\|_F.$ Then $\|V^\textsc{H}_{jj}-M\|_F\
\leq\|BV^\textsc{N}_{jj}B^{ \mathrm{\scriptscriptstyle T} }-M\|_F.$
This completes the proof.
\bibliographystyle{agsm}
|
\section{Introduction}
Recent technological advances have made it possible to collect unprecedented amounts of data. However, extracting patterns of information from these high-dimensional massive datasets is often challenging. How do we automatically determine, among millions of measured features (variables), which are informative, and which are irrelevant or redundant? The ability to select such features from high-dimensional data is crucial for computers to recognize patterns in complex data in ways that are fast, accurate, and even human-understandable \cite{Guyon}.
An efficient method for feature selection receiving increasing attention is Column Subset Selection (CSS). CSS is a constrained low-rank-approximation problem that seeks to approximate a matrix (e.g. instances by features matrix) by projecting it onto a space spanned by only a few of its columns (features). Formally, given a matrix $A$ with $n$ columns, and a target rank $k < n$, we wish to find a size-$k$ subset $S$ of $A$'s columns such that each column $A_i$ of $A$ ($i \in \{1, \dots, n\}$) is contained as much as possible in the subspace $\text{span}(S)$, in terms of the Frobenius norm:
\begin{align*}
\text{arg max}_{S \text{ contains $k$ of $A$'s columns}} \sum_{i=1}^n \|\text{proj}(A_i \; | \; \text{span}(S))\|_2^2
\end{align*}
While similar in spirit to general low-rank approximation, some advantages with CSS include flexibility,
interpretability and efficiency during inference. CSS is an
unsupervised method and does not require labeled data, which is
especially useful when labeled data is sparse. We note, on the other
hand, unlabeled data is often very abundant and therefore scalable
methods, like the one we present, are often needed.
Furthermore, by subselecting features, as opposed to generating new
features via an arbitrary function of the input features, we keep the
semantic interpretation of the features intact. This is especially
important in applications that require interpretable models. A third
important advantage is the efficiency of applying the solution CSS
feature selection problem during inference. Compared to PCA or other
methods that require a matrix-matrix multiplication to project input
features into a reduced space during inference time, CSS only requires
selecting a subset of feature values from a new instance vector. This
is especially useful for latency sensitive applications and when the
projection matrix itself may be prohibitively large, for example in
restricted memory settings.
While there have been significant advances in CSS~\cite{Boutsidis1,Boutsidis2,Guruswami}, most of the algorithms are either impractical and not applicable in a distributed setting for large datasets, or they do not have good (multiplicative $1 - \varepsilon$) provable error bounds.
Among efficient algorithms studied for the CSS problem is the simple {\em greedy algorithm}, which iteratively selects the best column and keeps it. Recent work shows that it does well in practice and even in a distributed setting~\cite{Farahat1, Farahat2} and admits a performance guarantee \cite{Civril1}. However, the known guarantees depend on an arbitrarily large matrix-coherence parameter, which is unsatisfactory. Also, even though the algorithm is relatively fast, additional optimizations are needed to scale it to datasets with millions of features and instances.
\subsection{Our contributions}
Let $A \in \mathbb{R}^{m \times n}$ be the given matrix, and let $k$ be the target number of columns. Let $OPT_k$ denote the {\em optimal} set of columns, i.e., one that {\em covers} the maximum Frobenius mass of $A$. Our contributions are as follows.
{\em Novel analysis of Greedy.} For any $\varepsilon > 0$, we show that the natural greedy algorithm (Section~\ref{section-2}), after $r = \frac{k}{\sigma_{\min}(OPT_k) \varepsilon}$ steps, gives an objective value that is within a $(1-\varepsilon)$ factor of the optimum. We also give a matching lower bound, showing that $\frac{k}{\sigma_{\min}(OPT_k) \varepsilon}$ is tight up to a constant factor. Here $\sigma_{\min}(OPT_k)$ is the smallest squared singular value of the {\em optimal} set of columns (after scaling to unit vectors).
Our result is similar in spirit to those of~\cite{Civril1, Liberty}, but with an important difference. Their bound on $r$ depends on the {\em least} $\sigma_{\min}(S)$ over \textit{all} $S$ of size $k$, while ours depends on $\sigma_{\min}(OPT_k)$. Note that these quantities can differ significantly. For instance, if the data has even a little bit of redundancy (e.g. few columns that are near duplicates), then there exist $S$ for which $\sigma_{\min}$ is tiny, but the optimal set of columns could be reasonably well-conditioned (in fact, we would {\em expect} the optimal set of columns to be fairly well conditioned).
{\em Distributed Greedy.} We consider a natural distributed implementation of the greedy algorithm (Section~\ref{section-2}). Here, we show that an interesting phenomenon occurs: even though partitioning the input does not work in general (as in coreset based algorithms), {\em randomly} partitioning works well. This is inspired by a similar result on submodular maximization~\cite{Mirrokni}.
Further, our result implies a $2$-pass streaming algorithm for the CSS problem in the {\em random arrival} model for the columns.
We note that if the columns each have sparsity $\phi$,~\cite{Boutsidis2015} gives an algorithm with total communication of $O(\frac{sk\phi}{\varepsilon} + \frac{sk^2}{\varepsilon^4})$. Their algorithm works for ``worst case'' partitioning of the columns into machines and is much more intricate than the greedy algorithm. In constrast, our algorithm is very simple, and for a random partitioning, the communication is just the first term above, along with an extra $\sigma_{\min}(OPT)$ term. Thus depending on $\sigma_{\min}$ and $\varepsilon$, each of the bounds could be better than the other.
{\em Further optimizations.} We also present techniques to speed up the implementation of the greedy algorithm. We show that the recent result of~\cite{Mirzasoleiman} (once again, on submodular optimization) can be extended to the case of CSS, improving the running time significantly.
We then compare our algorithms (in accuracy and running times) to various well-studied CSS algorithms. (Section 6.)
\subsection{Related Work}
The CSS problem is one of the central problems related to matrix approximation. Exact solution is known to be UG-hard~\cite{Civril2}, and several approximation methods have been proposed over the years. Techniques such as importance sampling \cite{Drineas1, Frieze}, adaptive sampling \cite{Deshpande1}, volume sampling \cite{Deshpande2, Deshpande4}, leverage scores \cite{Drineas-Leverage}, and projection-cost preserving sketches \cite{Cohen} have led to a much better understanding of the problem. \cite{Guruswami} gave the optimal dependence between column sampling and low-rank approximation.
Due to the numerous applications, much work has been done on the implementation side, where adaptive sampling and leverage scores have been shown to perform well. A related, extremely simple algorithm is the greedy algorithm, which turns out to perform well and be scalable \cite{Farahat1, Farahat2}. This was first analyzed by~\cite{Civril1}, as we discussed.
There is also substantial literature about distributed algorithms for CSS \cite{Pi, Feldman, Cohen, Farahat3, Farahat4, Boutsidis2015}. In particular, \cite{Farahat3, Farahat4} present distributed versions of the greedy algorithm based on MapReduce. Although they do not provide theoretical guarantees, their experimental results are very promising.
The idea of composable coresets has been applied explicitly or implicitly to several problems~\cite{FeldmanSS13,BalcanEL13,VahabPODS2014}. Quite recently, for some problems in which coreset methods do not work in general, surprising results have shown that randomized variants of them give good approximations~\cite{BarbosaENW15,Mirrokni}. We extend this framework to the CSS problem.
\subsection{Background and Notation}
We use the following notation throughout the paper. The set of integers $\{1, \dots, n\}$ is denoted by $[n]$. For a matrix $A \in \mathbb{R}^{m \times n}$, $A_j$ denotes the $j$th column ($A_j \in \mathbb{R}^m$). Given $S \subseteq [n]$, $A[S]$ denotes the submatrix of $A$ containing columns indexed by $S$. The projection matrix $\Pi_A$ projects onto the column span of $A$.
Let $\norm{A}_F$ denote the Frobenius norm, i.e., $\sqrt{\sum_{i,j} A_{i, j}^2}$. We write $\sigma_{\min}(A)$ to denote the minimum \textit{squared} singular value, i.e., $\inf_{x:\norm{x}_2 = 1} \frac{\|Ax\|_2^2}{\|x\|_2^2}$. We abuse notation slightly, and for a set of vectors $V$, we write $\sigma_{\min}(V)$ for the $\sigma_{\min}$ of the matrix with columns $V$
\iffalse
{\bf Submodular optimization.} Given a finite set $\Omega$ and a set function $f : 2^{\Omega} \to \mathbb{R}$, define the marginal gain of adding an element $x \in \Omega$ to a set $S \subseteq \Omega$ by $\Delta(x | S) = f(S \cup \{x\}) - f(S)$. $f$ is said to be submodular if $\Delta(x | S) \geq \Delta(x | T)$ for any subsets $S \subseteq T \subseteq \Omega$ and any element $x \in \Omega \setminus T$. This is a formalization of the well-known economic principle of decreasing marginal utility. $f$ is further said to be nonnegative if $f(S) \geq 0$ for any $S \subseteq \Omega$, and monotonically nondecreasing if $f(S) \leq f(T)$ for any $S \subseteq T \subseteq \Omega$. The theory of maximizing submodular functions subject to a cardinality constraint has been well studied, and has been shown to be NP-hard [Nemhauser and
Wolsey 1978; Feige 1998]. However, it is a key result in combinatorial optimization that a simple greedy algorithm to this problem for nonnegative, monotone nondecreasing submodular functions admits a $1 - \frac{1}{e}$ constant factor approximation [Nemhauser '78].
\fi
\begin{defin}\label{defn:css-problem}
Given a matrix $A \in \mathbb{R}^{m \times n}$ and an integer $k \le n$, the \textbf{Column Subset Selection (CSS) Problem} asks to find
\[ \text{arg max}_{S \subseteq [n], |S| = k} \norm{\Pi_{A[S]}A}_F^2, \]
i.e., the set of columns that {\em best explain} the full matrix $A$.
\end{defin}
We note that it is also common to cast this as a minimization problem, with the objective being $\norm{A - \Pi_{A[S]} A}_F^2$. While the exact optimization problems are equivalent, obtaining multiplicative approximations for the minimization version could be harder when the matrix is low-rank.
For a set of vectors $V$ and a matrix $M$, we denote
\[ f_M(V) = \norm{\Pi_V M}_F^2. \]
\iffalse
In this article, instead of minimizing the unexplained (error) part $\|A - \Pi_{A[S]}A\|_F^2$ of $A$, we maximize the explained part $\|\Pi_{A[S]}A\|_F^2$ of $A$. Formally,
\begin{align}
\text{arg max}_{S \subseteq [n], |S| = k} \|\Pi_{A[S]}A\|_F^2
= \text{arg min}_{S \subseteq [n], |S| = k} \|A - \Pi_{A[S]}A\|_F^2
\end{align}
Thus, a subset of columns that maximizes explanation of $A$ will also minimized the unexplained error.
\fi
Also, the case when $M$ is a single vector will be important. For any vector $u$, and a set of vectors $V$, we write
\[ f_u(V) = \norm{\Pi_V u}_2^2. \]
\begin{remark} \label{rem:not-submodular} Note that $f_M (V)$ can be viewed as the extent to which we can {\em cover} matrix $M$ using vectors $V$. However, unlike combinatorial covering objectives, our definition is not submodular, or even subadditive.
\iffalse
\begin{defin} \label{f definition matrix}
Given $A \in \mathbb{R}^{m \times n}$, define the function: $f_A : \mathcal{P}(\mathbb{R}^m) \to \mathbb{R}$ by: $$f_A(S) = \sum_{j=1}^n f_{A_j}(S)$$ over the columns $A_j \in \mathbb{R}^m$ of $A$.
\end{defin}
\fi
As an example, consider covering the following $A$ using its own columns. Here, $f_A(\{A_1, A_2\}) = \|A\|_F^2 > f(\{A_1\}) + f(\{A_2\})$.
\[ A = \left( \begin{array}{ccc}
1 & 0 & 1 \\
1 & -1 & 0 \\
0 & 1 & 1 \end{array} \right)\]
\end{remark}
\iffalse
\subsection{Overview of article}
In section 2.1, we present a ``vanilla'' greedy algorithm GREEDY for choosing $k$ columns of a matrix $A$ that can approximate all of $A$'s other columns linearly. This provides intuition for our proposed algorithm ALG, presented in section 2.2. ALG is a much more efficient version of GREEDY because of three optimizations: (1) an efficient calculation of marginal gain; (2) a random projection to compress the ambient dimension of $A$'s columns; and (3) only looking over a small random subset of all $n$ columns in each iteration. These optimizations make analyzing ALG slightly more involved than GREEDY. To this end, we first analyze GREEDY in section 3, and then use those results to analyze ALG in section 4.
\fi
\section{Greedy Algorithm for Column Selection} \label{section-2}
Let us state our algorithm and analysis in a slightly general form. Suppose we have two matrices $A, B$ with the same number of rows and $n_A$, $n_B$ columns respectively. The $\textsf{GCSS}(A, B, k)$ problem is that of finding a subset $S$ of columns of $B$, that maximizes $f_A(S)$ subject to $|S|=k$.
Clearly, if $B = A$, we recover the CSS problem stated earlier. Also, note that scaling the columns of $B$ will not affect the solution, so let us assume that the columns of $B$ are all unit vectors. The greedy procedure iteratively picks columns of $B$ as follows:
\begin{algorithm} \label{alg:greedy}
\caption{$\textsc{Greedy}$($A \! \in \! \mathbb{R}^{m \times n_A}$, $B \! \in\! \mathbb{R}^{m \times n_B}$, $k \leq n_B$)}
\begin{algorithmic}[1]
\STATE $S \leftarrow \emptyset$
\FOR{$i = 1:k$}
\STATE Pick column $B_j$ that maximizes $f_A(S \cup B_j)$
\STATE $S \leftarrow S \cup \{B_j\}$
\ENDFOR
\STATE Return $S$
\end{algorithmic}
\end{algorithm}
Step (3) is the computationally intensive step in $\textsc{Greedy}$ -- we need to find the column that gives the most {\em marginal gain}, i.e., $f_A(S \cup B_j) - f_A(S)$.
In Section~\ref{section-5}, we describe different techniques to speed up the calculation of marginal gain, while obtaining a $1-\varepsilon$ approximation to the optimum $f(\cdot)$ value. Let us briefly mention them here.
{\em Projection to reduce the number of rows.} We can left-multiply both $A$ and $B$ with an $r \times n$ Gaussian random matrix. For $r \ge \frac{k\log n}{\varepsilon^2}$, this process is well-known to preserve $f_A(\cdot)$, for any $k$-subset of the columns of $B$ (see~\cite{Sarlos} or Appendix Section~\ref{app:random-projections} for details)
{\em Projection-cost preserving sketches.}
Using recent results from \cite{Cohen}, we can project each {\em row} of $A$ onto a random $O(\frac{k}{\varepsilon^2})$ dimensional space, and then work with the resulting matrix. Thus we may assume that the number of columns in $A$ is $O(\frac{k}{\varepsilon^2})$. This allows us to efficiently compute $f_A(\cdot)$.
\iffalse
\textbf{Random projections to reduce the number of rows.} We can project each column of $A$ and $B$ onto a random $O(\frac{k\log (\max(n', n_B))}{\varepsilon^2})$ dimensional space, and then work with the resulting matrices. Thus we may assume that the number of rows in both $A$ and $B$ is $\min\{ m, O(\frac{k\log( \max(n', n_B))}{\varepsilon^2}) \}$, which can be a big improvement. This can be obtained by a union bound on Lemma 10 from \cite{Sarlos}. (Full details in appendix.)
\fi
{\em Lazier-than-lazy greedy.}
\cite{Mirzasoleiman} recently proposed the first algorithm that achieves a constant factor approximation for maximizing submodular functions with a {\em linear} number of marginal gain evaluations. We show that a similar analysis holds for $\textsf{GCSS}$, even though the cost function is not submodular.
We also use some simple yet useful ideas from \cite{Farahat2} to compute the marginal gains (see Section~\ref{section-5}).
\subsection{Distributed Implementation}
We also study a distributed version of the greedy algorithm, shown below (Algorithm~\ref{alg:cs-greedy}). $\ell$ is the number of machines
\begin{algorithm} \label{alg:cs-greedy}
\caption{$\textsc{Distgreedy}$($A$, $B$, $k$, $\ell$)}
\begin{algorithmic}[1]
\STATE {\em Randomly} partition the columns of $B$ into $T_1, \dots, T_{\ell}$
\STATE (Parallel) compute $S_i \leftarrow \textsc{Greedy}(A, T_i, \frac{32k}{\sigma_{\min}(OPT)})$
\STATE (Single machine) aggregate the $S_i$, and compute $S \leftarrow \textsc{Greedy}(A, \cup_{i=1}^{\ell}S_i, \frac{12k}{\sigma_{\min}(OPT)})$
\STATE Return $\text{arg max}_{S' \in \{S, S_1,\dots, S_{\ell}\}} f_A(S')$
\end{algorithmic}
\end{algorithm}
As mentioned in the introduction, the key here is that the partitioning is done {\em randomly}, in contrast to most results on {\em composable summaries}.
We also note that machine $i$ only sees columns $T_i$ of $B$, but requires evaluating $f_A(\cdot)$ on the full matrix $A$ when running \textsc{Greedy}.\footnote{It is easy to construct examples in which splitting both $A$ and $B$ fails badly.} The way to implement this is again by using projection-cost preserving sketches. (In practice, keeping a small sample of the columns of $A$ works as well.) The sketch is first passed to all the machines, and they all use it to evaluate $f_A(\cdot)$.
We now turn to the analysis of the single-machine and the distributed versions of the greedy algorithm.
\section{Peformance analysis of GREEDY} \label{section-3}
The main result we prove is the following, which shows that by taking only slightly more than $k$ columns, we are within a $1 -\varepsilon$ factor of the optimal solution of size $k$.
\begin{thm} \label{thm:greedy-main}
Let $A \in \mathbb{R}^{m \times n_A}$ and $B \in \mathbb{R}^{m \times n_B}$. Let $OPT_k$ be a set of columns from $B$ that maximizes $f_A(S)$ subject to $|S| = k$. Let $\varepsilon > 0$ be any constant, and let $T_r$ be the set of columns output by $\textsc{Greedy}(A, B, r)$, for $r = \frac{16k}{\varepsilon \sigma_{\min}(OPT_k)}$. Then we hav
$$f_A(T_r) \geq (1 - \varepsilon) f_A(OPT_k).$$
\end{thm}
We show in Appendix Section \ref{app:tight-ex} that this bound is tight up to a constant factor, with respect to $\varepsilon$ and $\sigma_{\min}(OPT_k)$. Also, we note that $\textsf{GCSS}$ is a harder problem than $\textsf{MAX-COVERAGE}$, implying that if we can choose only $k$ columns, it is impossible to approximate to a ratio better than $(1-\frac{1}{e}) \approx 0.63$, unless P=NP. (In practice, $\textsc{Greedy}$ does much better, as we will see.)
The basic proof strategy for Theorem~\ref{thm:greedy-main} is similar to that of maximizing submodular functions, namely showing that in every iteration, the value of $f(\cdot)$ increases significantly. The key lemma is the following.
\begin{lemma} \label{lem:large-gain}
Let $S, T$ be two sets of columns, with $f_A(S) \ge f_A(T)$. Then there exists $v \in S$ such that
\[ f_A(T \cup v) - f_A(T) \ge \sigma_{\min}(S) \frac{\big(f_A(S) - f_A(T)\big)^2}{4|S|f_A(S)}.\]
\end{lemma}
Theorem~\ref{thm:greedy-main} follows easily from Lemma~\ref{lem:large-gain}, which we show at the end of the section. Thus let us first focus on proving the lemma. Note that for submodular $f$, the analogous lemma simply has $\frac{f(S) - f(T)}{|S|}$ on the right-hand side (RHS).
The main ingredient in the proof of Lemma~\ref{lem:large-gain} is its {\em single vector} version:
\begin{lemma}\label{lem:one-vector}
Let $S, T$ be two sets of columns, with $f_u(S) \ge f_u(T)$. Suppose $S=\{v_1, \dots, v_k\}$. Then
\[ \sum_{i=1}^k \Big( f_u(T \cup v_i) - f_u(T) \Big) \ge \sigma_{\min}(S) \frac{\big(f_u(S) - f_u(T)\big)^2}{4f_u(S)}.\]
\end{lemma}
Let us first see why Lemma~\ref{lem:one-vector} implies Lemma~\ref{lem:large-gain}. Observe that for any set of columns $T$, $f_A (T) = \sum_{j} f_{A_j} (T)$ (sum over the columns), by definition. For a column $j$, let us define $\delta_j = \min \{ 1, \frac{f_{A_j}(T)}{f_{A_j}(S)}\}$. Now, using Lemma~\ref{lem:one-vector} and plugging in the definition of $\delta_j$, we have
\begin{align}
& \frac{1}{\sigma_{\min}(S)} \sum_{i=1}^k \big( f_A(T \cup v_i) - f_A(T) \big) \label{eq:start}\\
& \quad = \frac{1}{\sigma_{\min}(S)} \sum_{j = 1}^n \sum_{i=1}^k \big( f_{A_j}(T \cup v_i) - f_{A_j}(T) \big) \notag\\
& \quad \geq \sum_{j=1}^n \frac{ (1-\delta_j)^2 f_{A_j}(S)}{4} \label{eq:temp3}\\
& \quad = \frac{f_A(S)}{4} \sum_{j=1}^n (1 - \delta_j)^2 \frac{f_{A_j}(S)}{f_A(S)} \label{eq:temp4}\\
& \quad \geq \frac{f_A(S)}{4} \left( \sum_{j=1}^n (1 - \delta_j) \frac{f_{A_j}(S)}{f_A(S)}\right)^2 \label{eq:temp5} \\
& \quad = \frac{1}{4 f_A(S)} \Big(\sum_{j=1}^n \max\{ 0, f_{A_j}(S) - f_{A_j}(T) \}\Big)^2 \label{eq:temp6} \\
& \quad \ge \frac{1}{4 f_A(S)} \Big(f_A(S) - f_A(T) \Big)^2 \label{eq:temp8}
\end{align}
To get \eqref{eq:temp5}, we used Jensen's inequality ($\mathbb{E}[X^2] \geq ( \mathbb{E}[X])^2$) treating $\frac{f_{A_j}(S)}{f_{A}(S)}$ as a probability distribution over indices $j$. Thus it follows that there exists an index $i$ for which the gain is at least a $\frac{1}{|S|}$ factor, proving Lemma~\ref{lem:large-gain}.
\begin{proof}[Proof of Lemma~\ref{lem:one-vector}]
Let us first analyze the quantity $f_u(T \cup v_i) - f_u(T)$, for some $v_i \in S$. As mentioned earlier, we may assume the $v_i$ are normalized. If $v_i \in \text{span}(T)$, this quantity is $0$. Thus we can assume that such $v_i$ have been removed from $S$. Now, adding $v_i$ to $T$ gives a gain because of the component of $v_i$ orthogonal to $T$, i.e., $v_i - \Pi_T v_i$, where $\Pi_T$ denotes the projector onto $\text{span}(T)$. Define
\[ v_i' = \frac{v_i - \Pi_T v_i}{\norm{v_i - \Pi_T v_i}}_2.\] By definition, $\text{span}(T \cup v_i) = \text{span}(T \cup v_i')$. Thus the projection of a vector $u$ onto $\text{span}(T \cup v_i')$ is $\Pi_T u + \iprod{u, v_i'} v_i'$, which is a vector whose squared length is
$\norm{\Pi_T u}^2 + \iprod{u, v_i'}^2 = f_u(T) + \iprod{u, v_i'}^2$.
This implies that
\begin{equation}\label{eq:gain-single}
f_u(T \cup v_i ) - f_u(T) = \iprod{u, v_i'}^2.
\end{equation}
Thus, to show the lemma, we need a lower bound on $\sum_i \iprod{u, v_i'}^2$. Let us start by observing
that a more explicit definition of $f_u(S)$ is the squared-length of the projection of $u$ onto $\text{span}(S)$, i.e. $f_u(S) = \max_{x \in \text{span}(S), \norm{x}_2 = 1} \iprod{u, x}^2$. Let $x = \sum_{i=1}^k \alpha_i v_i$ be a maximizer. Since $\norm{x}_2=1$, by the definition of the smallest squared singular value, we have $\sum_i \alpha_i^2 \le \frac{1}{\sigma_{\min}(S)}$. Now, decomposing $x = \Pi_T x + x'$, we have
\[ f_u(S) = \langle x, u \rangle^2
= \langle x' + \Pi_T x, \; u \rangle^2
= (\langle x', u \rangle + \langle \Pi_T x, u \rangle)^2.\]
Thus (since the worst case is when all signs align),
\begin{align}
|\iprod{ x', u}| &\ge \sqrt{f_u(S)} - |\langle \Pi_T x, u \rangle|
\ge \sqrt{f_u(S)} - \sqrt{f_u(T)} \notag \\
&= \frac{f_u(S) - f_u(T)}{\sqrt{f_u(S)}+ \sqrt{f_u(T)}} \ge \frac{f_u(S) - f_u(T)}{2\sqrt{f_u(S)}}. \label{eq:dotprod-lb2}
\end{align}
where we have used the fact that $|\iprod{\Pi_T x, u}|^2 \le f_u(T)$, which is true from the definition of $f_u(T)$ (and since $\Pi_T x$ is a vector of length $\le 1$ in $\text{span}(T)$).
Now, because $x = \sum_i \alpha_i v_i$, we have $x' = x - \Pi_T x = \sum_i \alpha_i(v_i - \Pi_T v_i) = \sum_i \alpha_i\norm{v_i - \Pi_Tv_i}_2v_i'$. Thus,
\begin{align*}
\iprod{x', u}^2 &=
\big( \sum_i \alpha_i \norm{v_i - \Pi_Tv_i}_2 \iprod{v_i' , u} \big)^2
\\ &\le \big( \sum_i \alpha_i^2 \norm{v_i - \Pi_Tv_i}_2^2 \big) \big( \sum_i \iprod{v_i', u}^2 \big)
\\ &\le \big( \sum_i \alpha_i^2 \big) \big( \sum_i \iprod{v_i', u}^2 \big).
\end{align*}
where we have used Cauchy-Schwartz, and then the fact that $\|v_i - \Pi_Tv_i\|_2 \leq 1$ (because $v_i$ are unit vectors). Finally, we know that $\sum_i \alpha_i^2 \le \frac{1}{\sigma_{\min}(S)}$, which implies
\[ \sum_i \iprod{v_i', u}^2 \ge \sigma_{\min}(S) \iprod{x', u}^2 \ge \sigma_{\min}(S) \frac{(f_u(S)- f_u(T))^2}{4f_u(S)}\]
Combined with~\eqref{eq:gain-single}, this proves the lemma.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:greedy-main}]
For notational convenience, let $\sigma = \sigma_{min}(OPT_k)$ and $F = f_A(OPT_k)$. Define $\Delta_0 = F$, $\Delta_1 = \frac{\Delta_0}{2}$, $\dots$, $\Delta_{i+1} = \frac{\Delta_i}{2}$ until $\Delta_N \leq \varepsilon F$. Note that the gap $f_A(OPT_k) - f_A(T_0) = \Delta_0$. We show that it takes at most $\frac{8kF}{\sigma \Delta_i}$ iterations (i.e. additional columns selected) to reduce the gap from $\Delta_i$ to $\frac{\Delta_i}{2} = \Delta_{i+1}$. To prove this, we invoke Lemma \ref{lem:large-gain} to see that the gap filled by $\frac{8kF}{\sigma \Delta_i}$ iterations is at least $\frac{8kF}{\sigma \Delta_i} \cdot \sigma \frac{(\frac{\Delta_i}{2})^2}{4kF}
= \frac{\Delta_i}{2} = \Delta_{i+1}$. Thus the total number of iterations $r$ required to get a gap of at most $\Delta_N \leq \varepsilon F$ is:
\vspace{-0.1cm}
\[
r
\leq \sum_{i=0}^{N-1} \frac{8kF}{\sigma \Delta_i}
= \frac{8kF}{\sigma} \sum_{i=0}^{N-1} \frac{2^{i-N+1}}{\Delta_{N-1}}
< \frac{16k}{\varepsilon \sigma}
.\]
where the last step is due to $\Delta_{N-1} > \varepsilon F$ and $\sum_{i=0}^{N-1} 2^{i-N+1} < 2$. Therefore, after $r < \frac{16k}{\varepsilon \sigma}$ iterations, we have $f_A(OPT_k) - f_A(T_r) \leq \varepsilon f_A(OPT_k)$. Rearranging proves the lemma.
\end{proof}
\section{Distributed Greedy Algorithm} \label{section-4}
We will now analyze the distributed version of the greedy algorithm that was discussed earlier. We show that in one {\em round}, we will find a set of size $O(k)$ as before, that has an objective value $\Omega(f(OPT_k)/\kappa)$, where $\kappa$ is a condition number (defined below). We also combine this with our earlier ideas to say that if we perform $O(\kappa / \varepsilon)$ {\em rounds} of \textsc{Distgreedy}, we get a $(1-\varepsilon)$ approximation (Theorem~\ref{thm:core-set-2}).
\subsection{Analyzing one round}
\iffalse
Often modern data sets are too large for us to run GREEDY or ALG on a single processor. The approach of randomized composable core-sets has received increased popularity as a provably good way to subdivide a large problem into smaller problems, each of which can be approached by a separate machine in a computer cluster. This type of approach is especially scalable because it can easily be used in tandem with mainstream distributed paradigms such as MapReduce.
\\ \\ Recently, Mirrokni and Zadimoghaddam showed that randomized composable core-sets give a constant factor approximation for maximizing submodular functions \cite{Mirrokni}. We show that here that although our cost function is not submodular, a similar result still holds (up to the condition number).
\\ \\ Formally, consider as before Column Subset Section on a matrix $A \in \mathbb{R}^{m \times n}$ with target rank $k < n$. The difference is that now we have $\ell$ machines at our disposal. The approach we take is to randomly partition the $m \times n$ columns into $\ell$ submatrices, each of dimension roughly $m \times \frac{n}{\ell}$. We give each submatrix to a different machine, and run GREEDY on each of these machines with target rank $k'$ to produce a ``core-set'' solution. Finally, we collect the $\ell$ core-sets, and run GREEDY a final time on a single machine to produce our final set of $k$ columns.
\\ \\ The parameter $k'$ must be chosen with care, since there is a natural tradeoff between precision and efficiency. Importantly, it must be sufficiently large such that we have enough options in the final stage to choose a a set of $k$ columns with a constant factor approximation. But from a practical standpoint, $k'$ must be small enough such that the final stage with approximately $\ell k'$ can be run on a single machine. If this is not possible, we can run the above procedure with multiple rounds.
\subsection{Constant factor approximation (up to condition number)}
\fi
We consider an instance of $\textsf{GCSS}(A, B, k)$, and let $OPT$ denote an optimum set of $k$ columns.
Let $\ell$ denote the number of machines available.
The columns (of $B$) are partitioned across machines, such that machine $i$ is given columns $T_i$. It runs $\textsc{Greedy}$ as explained earlier and outputs $S_i \subset T_i$ of size $k' = \frac{32k}{\sigma_{\min}(OPT)}$. Finally, all the $S_i$ are moved to one machine and we run $\textsc{Greedy}$ on their union and output a set $S$ of size $k'' = \frac{12k}{\sigma_{\min}(OPT)}$.
Let us define $\kappa(OPT) = \frac{\sigma_{\max}(OPT)}{\sigma_{\min}(OPT)}$.
\begin{thm}\label{thm:distributed-main}
Consider running $\textsc{Distgreedy}$ on an instance of $\textsf{GCSS}(A, B, k)$. We have
\[ \mathbb{E}[ \max\{ f_A(S), \max_i \{f_A(S_i)\}\}] \ge \frac{f(OPT)}{8\cdot \kappa(OPT)}.\]
\end{thm}
The key to our proof are the following definitions:
\begin{align*}
OPT_i^S &= \{x \in OPT \; : \; x \in \textsc{Greedy}(A, T_i \cup x, k') \}\\
OPT_i^{NS} &= \{x \in OPT \; : \; x \not \in \textsc{Greedy}(A, T_i \cup x, k') \}
\end{align*}
In other words, $OPT_i^S$ contains all the vectors in $OPT$ that would have been selected by machine $i$ if they had been added to the input set $T_i$. By definition, the sets $(OPT_i^S, OPT_i^{NS})$ form a partition of $OPT$ for every $i$.
\iffalse
\begin{remark} \label{rem-opt-partition}
By definition, $OPT = OPT_i^S \cup OPT_i^{NS}$ is a union of disjoint sets for all $i \in [m]$. That is, $OPT_i^S$ and $OPT_i^{NS}$ partition $OPT$.
\end{remark}
\fi
{\bf Proof outline.} Consider any partitioning $T_1, \dots, T_\ell$, and consider the sets $OPT_i^{NS}$. Suppose one of them (say the $i$th) had a large value of $f_A(OPT_i^{NS})$. Then, we claim that $f_A(S_i)$ is also large. The reason is that the greedy algorithm does {\em not} choose to pick the elements of $OPT_i^{NS}$ (by definition) -- this can only happen if it ended up picking vectors that are ``at least as good''. This is made formal in Lemma~\ref{lem:opt-ns}. Thus, we can restrict to the case when {\em none} of $f_A(OPT_i^{NS})$ is large. In this case, Lemma~\ref{lem:additivity} shows that $f_A(OPT_i^{S})$ needs to be large for each $i$. Intuitively, it means that most of the vectors in $OPT$ will, in fact, be picked by $\textsc{Greedy}$ (on the corresponding machines), and will be considered when computing $S$. The caveat is that we might be unlucky, and for every $x \in OPT$, it might have happened that it was sent to machine $j$ for which it was not part of $OPT_j^{S}$. We show that this happens with low probability, and this is where the random partitioning is crucial (Lemma~\ref{lem:opt-s}). This implies that either $S$, or one of the $S_i$ has a large value of $f_A(\cdot)$.
Let us now state two lemmas, and defer their proofs to Sections~\ref{app:opt-ns} and~\ref{app:additivity} respectively.
\begin{lemma} \label{lem:opt-ns}
For $S_i$ of size $k' = \frac{32 k}{\sigma_{\min}(OPT)}$, we have
\[ f(S_i) \geq \frac{f_A(OPT_i^{NS})}{2} ~\text{ for all $i$.}\]
\end{lemma}
\iffalse
Fix $i \in [m]$. Let $S_{i, r}$ denote the output of $GREEDY$ on $T_i$ for target rank $r$. By definition of $k'$ and $S_i$, we know $S_i = S_{i, k'}$. Denote $F = f(OPT_i^{NS})$ for shorthand. Recall from Lemma \ref{Large marginal gain lemma} that for any $r \in \mathbb{N}$ such that $f(S_{i, r}) \geq F$, then:
\begin{align}
f(S_{i, r+1}) - f(S_{i, r}) \geq \sigma_{min}(OPT_i^{NS}) \frac{\big(F - f(S_{i, r})\big)^2}{4F \; |OPT_i^{NS}|}
\end{align}
Note that since $OPT_i^{NS} \subset OPT$ by construction, so $\sigma_{min}(OPT_i^{NS}) \geq \sigma_{min}(OPT)$ and also $|OPT_i^{NS}| \leq |OPT| = k$. Thus:
\begin{align}
f(S_{i, r+1}) - f(S_{i, r}) \geq \sigma_{min}(OPT) \frac{\big(F - f(S_{i, r})\big)^2}{4kF} \label{eq-lem-coreset-2-gain}
\end{align}
Let $\Delta_0 = F$, $\Delta_1 = \frac{\Delta_0}{2}$, \dots, $\Delta_{i+1} = \frac{\Delta_i}{2}$ until $\Delta_N \leq cF$. We show that it takes at most $\frac{8kF}{\sigma_{min}(OPT) \Delta_i}$ iterations (i.e. additional columns selected) to reduce the gap from $\Delta_i$ to $\frac{\Delta_i}{2} = \Delta_{i+1}$. To see this, invoke equation \eqref{eq-lem-coreset-2-gain} above to see that:
\begin{align}
\text{Gap filled by $\frac{8kF}{\sigma_{min}(OPT) \Delta_i}$ iterations} \geq \frac{8kF}{\sigma_{min}(OPT) \Delta_i} \cdot \sigma_{min}(OPT) \frac{(\frac{\Delta_i}{2})^2}{4kF}
= \frac{\Delta_i}{2} = \Delta_{i+1}
\end{align}
So the total number of iterations $k'$ required to get a gap of at most $\Delta_N \leq cF$ is:
\begin{align}
k'
&\leq \sum_{i=0}^{N-1} \frac{8kF}{\sigma_{min}(OPT) \Delta_i}
\\ &= \frac{8kF}{\sigma_{min}(OPT)} \sum_{i=0}^{N-1} \frac{2^{i-N+1}}{\Delta_{N-1}} \label{eq-lem-coreset-2-geo-1}
\\ &< \frac{16k}{c \sigma_{min}(OPT)} \label{eq-lem-coreset-2-geo-2}
\end{align}
Equation \eqref{eq-lem-coreset-2-geo-1} is due to the fact that $\{\Delta_i\}_{i=0}^{N}$ is a geometric series by construction. Equation \ref{eq-lem-coreset-2-geo-2} is because $\Delta_{N-1} > cF$ and $\sum_{i=0}^{N-1} 2^{i-N+1} < 2$.
\\ \\ Thus, after $k' < \frac{16k}{c \sigma_{min}(OPT)}$ iterations, we have $f(S_i) - f(OPT_i^{NS}) \leq c f(OPT_i^{NS})$. Rearranging proves the lemma.
\end{proof}
\fi
\begin{lemma} \label{lem:additivity}
For any matrix $A$, and any partition $(I, J)$ of $OPT$:
\begin{equation}\label{eq:to-prove-additivity}
f_A(I) + f_A(J) \geq \frac{f_A(OPT)}{2\kappa(OPT)}.
\end{equation}
\end{lemma}
\iffalse
Let $x = \sum_{i=1}^{t} \alpha_i v_i$ be a maximizer. Define $x_A = \sum_{i=1}^s \alpha_i v_i \in \text{span}(A)$ and $x_B = \sum_{i=s+1}^t \alpha_i v_i \in \text{span}(B)$. Note that by bilinearity of the inner product:
\begin{align}
f_u(A \cup B)
= \langle u, \; x \rangle^2
= \big(\langle u, \; x_a \rangle + \langle u, \; x_b \rangle \big)^2
\end{align}
Thus by a simple averaging argument, either $\langle u, \; x_a \rangle ^2 \geq \frac{f_u(A \cup B)}{4}$ or $\langle u, \; x_b \rangle^2 \geq \frac{f_u(A \cup B)}{4}$. WLOG, let the first one be true. Let $x_a' = \frac{x_a}{\|x_a\|_2}$ denote the normalization of $x_a$. Then:
\begin{align}
\frac{f_u(A \cup B)}{4}
\leq \langle u, \; x_a \rangle ^2
= \|x_a\|_2^2 \cdot \langle u, \; x_a' \rangle ^2
\leq \|x_a\|_2^2 \cdot f_u(A) \label{eq-lem-coreset-final-1}
\end{align}
Now observe that since $\|x\|_2^2 = 1$, so:
\begin{align}
\sigma_{min}(OPT)
\leq \sigma_{min}(A \cup B)
= \inf_{(\beta_1, \dots, \beta_t)} \frac{\|\sum_{i=1}^t \beta_i v_i \|_2^2}{\sum_{i=1}^{t} \beta_i^2}
\leq \frac{\|\sum_{i=1}^t \alpha_i v_i \|_2^2}{\sum_{i=1}^{t} \alpha_i^2}
= \frac{\|x\|_2^2}{\sum_{i=1}^{t} \alpha_i^2}
= \frac{1}{\sum_{i=1}^{t} \alpha_i^2}
\end{align}
In particular, this implies that:
\begin{align}
\sigma_{min}(OPT)
\leq \frac{1}{\sum_{i=1}^{t} \alpha_i^2}
\leq \frac{1}{\sum_{i=1}^{s} \alpha_i^2}
\leq \frac{1}{\|x_a\|_2^2} \label{eq-lem-coreset-final-2}
\end{align}
Combining equations \eqref{eq-lem-coreset-final-1} and \eqref{eq-lem-coreset-final-2} finishes the proof, since $f_u(B) \geq 0$ is non-negative by Lemma \ref{f structure lemma}.
\end{proof}
Let $S$ be the set of $16k/\sigma_{min}(OPT)$ items that Greedy selects on set $\cup_{i=1}^{\ell} S_i$, i.e. $S =$ $GREEDY(\cup_{i=1}^{\ell} S_i,$ $16k/\sigma_{min}(OPT))$.
\fi
Our final lemma is relevant when none of $f_A(OPT_i^{NS})$ are large and, thus, $f_A(OPT_i^{S})$ is large for {\em all} $i$ (due to Lemma~\ref{lem:additivity}). In this case, Lemma~\ref{lem:opt-s} will imply that the expected value of $f(S)$ is large.
Note that $T_i$ is a random partition, so the $T_i$, the $OPT_i^{S}$, $OPT_i^{NS}$, $S_i$, and $S$ are all random variables. However, all of these value are fixed given a partition $\{T_i\}$. In what follows, we will write $f(\cdot)$ to mean $f_A(\cdot)$.
\begin{lemma}\label{lem:opt-s}
For a random partitioning $\{T_i\}$, and $S$ of size $k'' = \frac{12 k}{\sigma_{\min}(OPT)}$, we have
\begin{equation}\label{eq:main-lem-to-show}
\mathbb{E}[f(S)] \ge \frac{1}{2}\mathbb{E} \left[ \frac{\sum_{i=1}^{\ell} f(OPT_i^S)}{\ell}\right].
\end{equation}
\end{lemma}
\begin{proof}
At a high level, the intuition behind the analysis is that many of the vectors in $OPT$ are selected in the first phase, i.e., in $\cup_i S_i$. For an $x \in OPT$, let $I_x$ denote the indicator for $x \in \cup_i S_i$.
Suppose we have a partition $\{T_i\}$. Then if $x$ had gone to a machine $i$ for which $x \in OPT_i^{S}$, then by definition, $x$ will be in $S_i$. Now the key is to observe (see definitions) that the event $x \in OPT_i^S$ does not depend on where $x$ is in the partition! In particular, we could think of partitioning all the elements except $x$ (and at this point, we know if $x \in OPT_i^S$ for all $i$), and {\em then} randomly place $x$. Thus
\begin{equation}\label{eq:n14}
\mathbb{E}[ I_x ] = \mathbb{E} \left[ \frac{1}{\ell} \sum_{i=1}^{\ell} [[x \in OPT_i^S]] \right],
\end{equation}
where $[[~\cdot~]]$ denotes the indicator.
We now use this observation to analyze $f(S)$. Consider the execution of the greedy algorithm on $\cup_i S_i$, and suppose $V^t$ denotes the set of vectors picked at the $t$th step (so $V^t$ has $t$ vectors). The main idea is to give a lower bound o
\begin{equation}
\label{eq:diff-expectations}
\mathbb{E}[ f(V^{t+1}) - f(V^t) ],
\end{equation}
where the expectation is over the partitioning $\{T_i\}$. Let us denote by $Q$ the RHS of \eqref{eq:main-lem-to-show}, for convenience. Now, the trick is to show that for {\em any} $V^t$ such that $f(V^t) \le Q$, the expectation in~\eqref{eq:diff-expectations} is large. One lower bound on $f(V^{t+1}) - f(V^t)$ is (where $I_x$ is the indicator as above)
\[\frac{1}{k} \sum_{x\in OPT} I_x \big( f(V^t \cup x) - f(V^t) \big).\]
Now for every $V$, we can use~\eqref{eq:n14} to obtain
\begin{align}
\mathbb{E}[ &f(V^{t+1}) - f(V^t) | V^t = V] \notag \\
&\ge \frac{1}{k\ell} \!\! \sum_{x \inOPT} \!\!\!\! \mathbb{E} \left[ \sum_{i=1}^\ell [[x \in OPT_i^S]]\right] \big( f(V \cup x) \!-\! f(V) \big) \notag \\
&= \frac{1}{k\ell} \mathbb{E} \left[ \sum_{i=1}^\ell \sum_{x \in OPT_i^S} \big( f(V \cup x) - f(V) \big) \right] \,. \notag
\end{align}
Now, using~\eqref{eq:start}-\eqref{eq:temp6}, we can bound the inner sum by
\[ \sigma_{\min}(OPT_i^S) \frac{ ( \max\{ 0, f(OPT_i^S) - f(V)\} )^2}{4f(OPT_i^S)} \,.\]
Now, we use $\sigma_{\min}(OPT_i^S) \ge \sigma_{\min}(OPT)$ and the identity that for any two nonnegative reals $a, b$: $(\max\{ 0, a-b \})^2/a \ge a/2 - 2b/3$.
Together, these imply
\begin{align*}
\mathbb{E}[ & f(V^{t+1}) - f(V^t) | V^t = V ] \\
&\ge \frac{\sigma_{\min}(OPT)}{4k\ell} \mathbb{E}\left[ \sum_{i=1}^\ell \frac{f(OPT_i^S)}{2} - \frac{2 f(V)}{3} \right].
\end{align*}
and consequently:
$\mathbb{E}[ f(V^{t+1}) - f(V^t)]
\ge \alpha (Q - \frac{2}{3}\mathbb{E}[f(V^t)]$
for $\alpha = \sigma_{\min}(OPT)/4k$.
If for some $t$, we have $\mathbb{E}[f(V^t)] \geq Q$, the proof is complete because $f$ is monotone, and $V^t \subseteq S$.
Otherwise, $\mathbb{E}[f(V^{t+1}) - f(V^t)]$ is at least $\alpha Q/3$ for each of the $k'' = 12 k/\sigma_{\min}(OPT) = 3/\alpha$ values of $t$. We conclude that $\mathbb{E}[f(S)]$ should be at least $(\alpha Q/3) \times (3/\alpha) = Q$ which completes the proof.
\iffalse
For any $0 \leq t \leq \frac{16k}{\sigma_{min}(OPT)}$, we define $S^t$ to be the set of first $t$ columns that Greedy chooses from $\cup_{i=1}^{\ell} S_i$.
In particular, we have $\emptyset = S^0 \subset S^1 \subset S^2 \cdots \subset S^t = S$.
To lower bound $\mathbb{E}[f(S)]$, we apply linearity of expectation and lower bound the expected marginal value $\mathbb{E}[f(S^t) - f(S^{t-1})]$ at step $1 \leq t \leq \frac{16k}{\sigma_{min}(OPT)}$.
Since algorithm Greedy chooses the column with the maximum marginal value at each step, we know that the marginal value at step $t$ is at least $f(\{x\} \cup S^{t-1}) - f(S^{t-1})$ for any column $x \in \cup_{i=1}^{\ell} S_i$.
In particular we are interested in lower bounding the marginal values using $x \in OPT^S$ where $OPT^S$ is the set of columns in $OPT$ that are selected for the second stage, i.e. $OPT^S = OPT \cap (\cup_{i=1}^{\ell} S_i)$. For the sake of analysis, suppose $x$ is a random column in $OPT$.
This column is sent to a random machine $1 \leq i \leq \ell$, i.e. $x \in T_i$.
By definition, $x$ is selected ($x \in OPT^S$) if and only if $x \in OPT_i^{S}$, and in that case, we know the marginal value at this step is lower bounded by $f(\{x\} \cup S^{t-1}) - f(S^{t-1})$,
otherwise we just lower bound the marginal by zero.
Therefore, $\mathbb{E}[f(S^t) - f(S^{t-1})]$ is at least $\mathbb{E}_{x \in OPT, 1 \leq i \leq \ell}[(f(\{x\} \cup S^{t-1}) - f(S^{t-1})) \cdot
[x \in OPT_i^S]]$ where $[x \in OPT_i^S]$ is the indicator function, and is one when $x$ is in $OPT_i^S$, and zero otherwise. Since the choice of $x$ and $i$ are independent of each other, we can imagine that machine $i$ is chosen randomly at first, then a random column $x$ is drawn from $OPT$. We conclude that:
\begin{align}\label{eq:marginal-1}
\mathbb{E}[f(S^t) - f(S^{t-1})] \geq \frac{1}{\ell k}\mathbb{E}[\sum_{i=1}^{\ell} \sum_{x \in OPT_i^S} f(\{x\} \cup S^{t-1}) - f(S^{t-1})]
\end{align}
Using Lemma~\ref{Large marginal gain helper lemma}, we know the sum $\sum_{x \in OPT_i^S} f(\{x\} \cup S^{t-1}) - f(S^{t-1})$ is lower bounded by $\sigma_{min}(OPT) \frac{(f(OPT_i^S) - f(S^{t-1}))^2} {(4f(OPT_i^S))}$. The quadratic form of this lower bound makes it harder to use it for expectations. Therefore we use the following relaxed inequality:
\begin{align}
\sum_{x \in OPT_i^S} f(\{x\} \cup S^{t-1}) - f(S^{t-1})
& \geq \sigma_{min}(OPT) (\frac{f(OPT_i^S)}{16} - \frac{f(S^{t-1})}{8}) \label{eq:marginal-2}
\\ &\geq \sigma_{min}(OPT) (\frac{f(OPT_i^S)}{16} - \frac{f(S)}{8}) \label{eq:marginal-3}
\end{align}
where the first inequality can be seen by the identity $\frac{(a-b)^2}{a} \geq \frac{a}{4} - \frac{b}{2}$ for any $a$ and $b$, and the second inequality holds because $S^{t-1}$ is a subset of $S$, and $f$ is a non-decreasing function. Taking the expected value of both sides of equations Equations \eqref{eq:marginal-2}, and \eqref{eq:marginal-3}, and combining the result with equation \eqref{eq:marginal-1} yields the following lower bound on the expected value of marginal value at each step:
\begin{align}
\mathbb{E}[f(S^t) - f(S^{t-1})]
&\geq \frac{\sigma_{min}(OPT)}{\ell k} \mathbb{E}[\sum_{i=1}^{\ell} (\frac{f(OPT_i^S)}{16} - \frac{f(S)}{8})] \\
&= \frac{\sigma_{min}(OPT)}{k} \mathbb{E}[\frac{\sum_{i=1}^{\ell} f(OPT_i^S)}{16\ell} - \frac{f(S)}{8}]
\end{align}
We note that $\mathbb{E}[f(S)]$ is at least ${16k}{\sigma_{min}(OPT)}$ times the above lower bound since we select that many items in the second round. Therefore we have $\mathbb{E}[f(S)] \geq \mathbb{E}[\frac{\sum_{i=1}^{\ell} f(OPT_i^S)}{\ell}] - 2\mathbb{E}[f(S)]$ which completes the proof.
\fi
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:distributed-main}]
If $f_A( OPT_i^{NS}) \ge \frac{ f(OPT)}{4\kappa(OPT)}$ for some $i$, then we are done, because Lemma~\ref{lem:opt-ns} implies that $f_A(S_i)$ is large enough. Otherwise, by Lemma~\ref{lem:additivity}, $f_A(OPT_i^{S}) \ge \frac{ f(OPT)}{4\kappa(OPT)}$ for all $i$.
Now we can use Lemma~\ref{lem:opt-s} to conclude that $\mathbb{E}[ f_A(S) ] \ge \frac{ f(OPT)}{8\kappa(OPT)}$, completing the proof.
\end{proof}
\subsection{Multi-round algorithm}
We now show that repeating the above algorithm helps achieve a $(1-\varepsilon)$-factor approximation.
We propose a framework with $r$ epochs for some integer $r>0$. In each epoch $t \in [r]$, we run the $\textsc{Distgreedy}$ algorithm to select set $S^t$. The only thing that changes in different epochs is the objective function: in epoch $t$, the algorithm selects columns based on the function $f^t$ which is defined to be: $f^t(V) = f_A(V \cup S^1 \cup S^2 \cdots \cup S^{t-1})$ for any $t$. We note that function $f^1$ is indeed the same as $f_A$. The final solution is the union of solutions: $\cup_{t=1}^r S^t$.
\begin{thm}\label{thm:core-set-2}
For any $\varepsilon <1$, the expected value of the solution of the $r$-epoch $\textsc{Distgreedy}$ algorithm, for $r = O(\kappa(OPT)/\epsilon)$, is at least $(1-\varepsilon)f(OPT)$.
\end{thm}
The proof is provided in Section~\ref{app:core-set-2} of the appendix.
{\em Necessity of Random Partitioning.} We point out that the random partitioning step of our algorithm is crucial for the $\textsf{GCSS}(A, B, k)$ problem. We adapt the instance from~\cite{VahabPODS2014} and show that even if each machine can compute $f_A(\cdot)$ exactly, and is allowed to output $\text{poly}(k)$ columns, it cannot compete with the optimum. Intuitively, this is because the partition of the columns in $B$ could ensure that in each partition $i$, the best way of covering $A$ involve picking some vectors $S_i$, but the $S_i$'s for different $i$ could overlap heavily, while the global optimum should use different $i$ to capture different {\em parts} of the space to be covered. (See Theorem \ref{thm:rand-part} in Appendix~\ref{app:rand-part} for details.)
\section{Further optimizations for \textsc{Greedy}} \label{section-5}
We now elaborate on some of the techniques discussed in Section~\ref{section-2} for improving the running time of $\textsc{Greedy}$.
We first assume that we left-multiply both $A$ and $B$ by a random Gaussian matrix of dimension $r \times m$, for $r \approx k \log n/\varepsilon^2$. Working with the new instance suffices for the purposes of $(1-\varepsilon)$ approximation to CSS (for picking $O(k)$ columns). (Details in the Appendix, Section~\ref{app:random-projections})
\subsection{Projection-Cost Preserving Sketches}
Marginal gain evaluations of the form $f_A(S \cup v) - f_A(S)$ require summing the marginal gain of $v$ onto each column of $A$. When $A$ has a large number of columns, this can be very expensive. To deal with this, we use a {\em sketch} of $A$ instead of $A$ itself. This idea has been explored in several recent works; we use the following notation and result:
\begin{defin}[\cite{Cohen}] \label{defin:pcps} For a matrix $A \in \mathbb{R}^{m \times n}$,
$A' \in \mathbb{R}^{m \times n'}$ is a \emph{rank-$k$ Projection-Cost Preserving Sketch (PCPS)} with error $0 \leq \varepsilon < 1$ if for any set of $k$ vectors $S$, we have: $(1 - \varepsilon) f_A(S) \leq f_{A'}(S) + c \leq (1 + \varepsilon) f_A(S)$
where $c \geq 0$ is a constant that may depend on $A$ and $A'$ but is independent of $S$.
\end{defin}
\begin{thm}\label{thm:pcps}[Theorem 12 of \cite{Cohen}]
Let $R$ be a random matrix with $n$ rows and $n' = O(\frac{k + \log\frac{1}{\delta}}{\varepsilon^2})$ columns, where each entry is set independently and uniformly to $\raisebox{.2ex}{$\scriptstyle\pm$} \sqrt{\frac{1}{n'}}$. Then for any matrix $A \in \mathbb{R}^{m \times n}$, with probability at least $1 - O(\delta)$, $AR$ is a rank-$k$ PCPS for $A$.
\end{thm}
Thus, we can use PCPS to sketch the matrix $A$ to have roughly $k/\varepsilon^2$ columns, and use it to compute $f_A(S)$ to a $(1\pm \varepsilon)$ accuracy for any $S$ of size $\le k$. This is also used in our distributed algorithm, where we send the sketch to every machine.
\subsection{Lazier-than-lazy Greedy}
The natural implementation of $\textsc{Greedy}$ requires $O(nk)$ evaluations of $f(\cdot)$ since we compute the marginal gain of all $n$ candidate columns in each of the $k$ iterations. For submodular functions, one can do better: the recently proposed $\textsc{Lazier-than-lazy Greedy}$ algorithm obtains a $1 - \frac{1}{e} - \delta$ approximation with only a linear number $O(n \log (1/\delta))$ of marginal gain evaluations \cite{Mirzasoleiman}. We show that a similar result holds for \textsf{GCSS}, even though our cost function $f(\cdot)$ is not submodular.
The idea is as follows. Let $T$ be the current solution set. To find the next element to add to $T$, we draw a sized $\frac{n_B \log (1/\delta)}{k} $ subset uniformly at random from the columns in $B \setminus T$. We then take from this set the column with largest marginal gain, add it to $T$, and repeat. We show this gives the following guarantee (details in Appendix Section~\ref{app:thm-lazier-than-lazy}.)
\begin{thm} \label{thm-lazier-than-lazy:main}
Let $A\in \mathbb{R}^{m \times n_A}$ and $B \in \mathbb{R}^{m \times n_B}$. Let $OPT_k$ be the set of columns from $B$ that maximizes $f_A(S)$ subject to $|S| = k$. Let $\varepsilon, \delta > 0$ be any constants such that $\epsilon + \delta \leq 1$. Let $T_r$ be the set of columns output by $\textsc{Lazier-than-lazy Greedy} (A, B, r)$, for $r = \frac{16k}{\varepsilon \sigma_{\min}(OPT_k)}$. Then we have:
$$\mathbb{E}[f_A(T_r)] \geq (1 - \varepsilon - \delta) f_A(OPT_k)$$
Further, this algorithm evaluates marginal gain only a linear number $\frac{16n_B \log (1/\delta)}{\varepsilon \sigma_{\min}(OPT_k)}$ of times.
\end{thm}
Note that this guarantee is nearly identical to our analysis of $\textsc{Greedy}$ in Theorem \ref{thm:greedy-main}, except that it is in expectation.
The proof strategy is very similar to that of Theorem \ref{thm:greedy-main}, namely showing that the value of $f(\cdot)$ increases significantly in every iteration (see Appendix Section~\ref{app:thm-lazier-than-lazy}).
{\bf Calculating marginal gain faster.} We defer the discussion to Appendix Section~\ref{sec:app:marginal}.
\section{Experimental results}\label{section-6}
\begin{figure*}[t]
\centering
\includegraphics[scale=0.32]{figure.pdf}
\caption{A comparison of reconstruction accuracy, model classification
accuracy and runtime of various column selection methods (with PCA
proved as an upper bound). The runtime is shown plot shows the
relative speedup over the naive GREEDY algorithm.}
\label{fig}
\end{figure*}
In this section we present an
empirical investigation of
the GREEDY, GREEDY++ and \textsc{Distgreedy}\ algorithms.
Additionally, we will compare with several baselines:
\\
{\bf Random:}
The simplest imaginable baseline, this method selects
columns randomly. \\
{\bf 2-Phase:}
The two-phased algorithm of \cite{Boutsidis2}, which
operates by first sampling $\Theta(k \log k)$ columns based on
properties of the top-$k$ right singular space of the input matrix
(this requires computing a top-$k$ SVD), then finally selects exactly
$k$ columns via a deterministic procedure. The overall complexity is
dominated by the top-$k$ SVD, which is $O( \min\{mn^2, m^2n\})$. \\
{\bf PCA:}
The columns of the rank-$k$ PCA projection matrix will be
used to serve as an upper bound on performance, as they explicitly
minimize the Forbenius reconstruction criteria. Note this method only
serves as an upper bound and does not fall into the framework of
column subset selection.
We investigate using these algorithms using two datasets, one with a small
set of columns (mnist) that is used to compare both scalable and
non-scalable methods, as well as a sparse dataset with a large number of
columns (news20.binary) that is meant to demonstrate the scalability of the
GREEDY core-set algorithm.\footnote{Both datasets can be
found at:
www.csie.ntu.edu.tw/$\sim$cjlin/libsvmtools/datasets/multiclass.html.}
Finally, we are also interested in the effect of column selection as a
preprocessing step for supervised learning methods. To that end, we will train
a linear SVM model, using the LIBLINEAR library \citep{fan2008}, with the
subselected columns (features) and measure the effectiveness of the model on a
held out test set. For both datasets we report test error for the best
choice of regularization parameter $c \in \{10^{-3}, \ldots, 10^4\}$.
We run GREEDY++ and \textsc{Distgreedy}\ with $\frac{n}{k}\log(10)$
marginal gain evaluations per iteration and the distributed
algorithm uses $s = \sqrt{\frac{n}{k}}$ machines with each machine
recieving $\frac{n}{s}$ columns.
\subsection{Small scale dataset (mnist)}
We first consider the MNIST digit recognition task, which is a
ten-class classification problem. There are $n$ = 784 input features
(columns) that represent pixel values from the $28\times28$-pixel
images. We use $m$ = 60,000 instances to train with and 10,000 instances
for our test set.
From Figure~\ref{fig} we see that all column sampling methods, apart
from Random, select columns that approximately provide the same amount
of reconstruction and are able to reach within 1\% of the performance
of PCA after sampling 300 columns. We also see a very similar trend
with respect to classification accuracy. It is notable that, in
practice, the core-set version of GREEDY incurs almost no additional
error (apart from at the smallest values of $k$) when compared to the
standard GREEDY algorithm.
Finally, we also show the relative speed up of the competitive methods
over the standard GREEDY algorithm. In this small dataset regime, we
see that the core-set algorithm does not offer an improvement over the single
machine GREEDY++ and in fact the 2-Phase algorithm is the fastest.
This is primarily due to the overhead of the distributed core-set
algorithm and the fact that it requires two greedy selection stages
(e.g.\ map and reduce). Next, we will consider a dataset that is large enough that a distributed model is in fact necessary.
\subsection{Large scale dataset (news20.binary)}
\begin{table}
\begin{small}
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
\bf n & \bf Rand & \bf 2-Phase & \bf $\textsc{Distgreedy}$ & \bf PCA \\
\hline
\hline
500 & 54.9 & 81.8 (1.0) & 80.2 (72.3) & 85.8 (1.3) \\
\hline
1000 & 59.2 & 84.4 (1.0) & 82.9 (16.4) & 88.6 (1.4)\\
\hline
2500 & 67.6 & 87.9 (1.0) & 85.5 (2.4) & 90.6 (1.7) \\
\hline
\end{tabular}
\end{center}
\vspace{-0.1cm}
\end{small}
\caption{A comparison of the classification accuracy of selected features. Also, the relative speedup over the 2-Phase algorithm for selecting features is shown in parentheses.
\vspace{-0.4cm}
}
\label{table}
\end{table}
In this section, we show that the $\textsc{Distgreedy}$ algorithm can indeed
scale to a dataset with a large number of columns.
The news20.binary dataset is a binary class text classification
problem, where we start with $n$ = 100,000 sparse features (0.033\%
non-zero entries) that represent text trigrams, use $m$ = 14,996
examples to train with and hold-out 5,000 examples to test with.
We compare the classification accuracy and column selection runtime of
the naive random method, 2-Phase algorithm
as well as PCA (that serves as an upper bound on performance)
to the $\textsc{Distgreedy}$ algorithm.
The results are presented in Table~\ref{table}, which shows that
$\textsc{Distgreedy}$ and 2-Phase both perform significantly better than random sampling and
come relatively close to the PCA upper bound in terms of accuracy. However,
we also find that $\textsc{Distgreedy}$ can be magnitudes of order faster than the 2-Phase
algorithm. This is in a large part because the 2-Phase algorithm suffers from the bottleneck of
computing a top-$k$ SVD. We note that an approximate SVD method
could be used instead,
however, it was outside the scope of this preliminary empirical investigation.
In conclusion, we have demonstrated that \textsc{Distgreedy}\ is able to scale to
larger sized datasets while still selecting effective features.
\bibliographystyle{icml2016}
|
\section{Introduction}
\label{intro}
Non-equilibrium dynamics of closed quantum systems has been a
subject of intense theoretical and experimental research in recent
years \cite{pol1,dziar1,dutta1,marcos1}. Such systems are known to
show several interesting features which have no analogs in their
equilibrium counterparts. Some such phenomena include Kibble-Zurek
scaling of defect density upon passage through a quantum critical
point \cite{kz1,pol2,ks1,pol3,sondhi0} or a critical (gapless)
region \cite{ds1}. In addition, such drives may lead to dynamic
transitions which cannot be characterized by any local order
parameter \cite{mukh1,pol4,dytr1,dytr2,dutta2} but manifest
themselves in the vanishing of the Loschmidt overlap $F(t) = \langle
\psi_i| \exp[-i H_f t] |\psi_i\rangle$, where $|\psi_i\rangle$ is
the initial system wave function (often chosen to be the ground
state of $H_i$) and $H_f$ is the final Hamiltonian following a
quench of a Hamiltonian parameter. Such dynamical phase transitions
can be defined in terms of non-analyticities (also known as Fischer
zeroes) of the dynamical free energy of the system $f(z) = - \lim_{L
\to \infty} \ln(F(z))/L^d$, where $z$ is obtained by analytic
continuation of time $t$ in the complex plane. Finally, quantum
quenches may lead to novel properties of the work distribution of
quantum systems following the quench which are qualitatively
different from their equilibrium counterparts \cite{silva1,silva2}.
Out of the drive protocols studied theoretically and experimentally
so far, periodic drives are found to lead to a gamut of interesting
phenomena which do not have counterparts in aperiodically driven
systems. These include dynamics induced freezing where the state of
the system, after several or single cycle(s) of the drive, has close
to unity overlap with its initial state; such a phenomenon can be
related to Stuckelberg interference between quantum states of the
driven system \cite{arnab1,pekker1,uma1}. In addition, we may use
periodic drives to obtain novel steady states in many-body localized
systems where a fast drive may lead to delocalization while a slow
drive keeps the system localized \cite{arnab2}. Moreover, it was
shown that periodically driven interacting systems may lead to
stable out-of-equilibrium phases in the presence of disorder
\cite{roderich1}; such phases are, similar to their equilibrium
counterparts, amenable to symmetry based classification
\cite{sondhi1}. Furthermore, the work distribution of periodically
driven system shows an oscillatory behavior with the drive
frequency; such a behavior constitutes an example of a quantum
interference effect shaping the behavior of a thermodynamic quantity
in a closed quantum system \cite{arnab3}. Finally, periodically
driven integrable systems are known to undergo a separate class of
dynamical phase transitions; these transitions, in contrast to the
ones discussed for aperiodically driven systems, leave their mark
through a change in the convergence of local correlation functions
to their steady state values; they can be shown to be a consequence
of a change in topology of the Floquet spectrum of the driven system
as a function of the drive frequency \cite{asen1}.
Apart from the effects mentioned earlier, another widely studied
phenomenon that occurs in periodically driven clean quantum systems
involves the generation of topological phases characterized by
non-trivial edge modes even when the starting ground state of the
corresponding equilibrium Hamiltonian is topologically trivial
\cite{kita1,lind1,jiang,gu,kita2,lind2,
morell,trif,russo,basti1,liu,tong,cayssol,rudner,basti2,tomka,gomez,dora,
katan,kundu,basti3,schmidt,reynoso,wu,perez,thakurathi1,reichl,thakurathi2}.
Such systems have been treated both analytically and numerically
demonstrating the appearance of edge modes after a drive through one
or more time periods signifying that the system has entered a
topological phase. Recently, however, a more complete understanding
of generation of edge states due to periodic drives in clean systems
has been put forth in Ref.\ \onlinecite{rudner1} in terms of the
properties of the time evolution operator $U(t,0) \equiv U(t)$ given
by
\begin{eqnarray} U(t) &=& {\mathcal T}_t e^{- (i/\hbar) \int_0^t dt' H(t')},
\quad 0\le t \le T, \label{evop} \end{eqnarray}
where $H(t)$ denotes the periodically driven Hamiltonian of the system,
${\mathcal T}_t$ denotes time-ordering, we have chosen the initial time
$t_i=0$ without loss of generality, and here and in the rest of the work we
shall denote $T= 2\pi/\omega_D$ to be the drive period, where $\omega_D$ is the
drive frequency. It was pointed out in Refs.\ \onlinecite{rudner1} and
\onlinecite{rudner2} that the knowledge of $U(T)
= \exp[-i H_F T/\hbar]$, or equivalently the Floquet Hamiltonian
$H_F$, is insufficient for describing the topological properties of
the system. Such properties can be understood instead by tracking
the crossings of the phase bands $\phi_n({\vec k}, t)$ which are defined
using the expression of $U_{{\vec k}}(t)$ for $t \le T$ as
\begin{eqnarray} U_{{\vec k}}(t) = \sum_{n=1}^{n_{\rm max}} P_n({\vec k}; t) e^{i
\phi_n({\vec k}; t)}. \end{eqnarray} Here we have assumed that the
crystal momentum ${\vec k}$ is a good quantum number, $n$ is the band
index with $n_{\rm max}$ bands for each ${{\vec k}}$, $\lambda_n(\vec
k, t) = \exp[i \phi_n({\vec k},t)]$ are eigenvalues of $U_{\vec
k}(t)$, and $P_n({\vec k} t)$ projects $U_{{\vec k}}(t)$ to its $n^{\rm
th}$ eigenstate. We note that $U_{{\vec k}}(0)=1$ indicates that
$\phi_n({\vec k}, 0) = 2 \pi m$, where $ m \in Z$. Thus the phase
bands may be represented either in the repeated zone scheme or the
reduced zone scheme where the $(n_{\rm max} +1)^{\rm th}$ band is
identified with the $n=1$ band. In what follows, we shall adopt the
latter representation.
It was shown in Ref. \onlinecite{rudner1} that the topological
properties of such periodic driven systems may be understood in
terms of phase band crossings. As argued in Ref.\
\onlinecite{rudner1}, phase band crossings are topologically
significant only if they occur between the first and the top bands
of the reduced Brillouin zone; all other crossings can be gauged
away by simple deformations of the drive protocol. Each such
topologically non-trivial crossing is associated with a finite
topological charge $q_i$; the number of edge modes which result from
such a crossing can be directly related to $q_i$. For example in
$d=2$, where each phase band can be represented by a Chern number
$C_n$, the number of chiral edge states within the $m^{\rm th}$ bulk
Floquet band is given by $n_{\rm edge}(m) = \sum_{n=1, M} C_n -
\sum_i q_i$. Further, it was shown in several earlier works
\cite{jiang,rudner1,carp1} that the presence of particle-hole and
time-reversal symmetries may lead to further restrictions on such
crossings; for example, in the presence of particle-hole symmetry
and for one-dimensional (1D) driven Hamiltonians, the phase band
crossings can occur only at $k_0=0$ or $\pi/a$, where $a$ is the
lattice spacing of the model (which we will subsequently set equal
to $1$, unless mentioned otherwise). The number of edge modes in
these systems are completely determined by the parity of the number
of such crossings at $k_0=0$ and $\pi$ \cite{rudner1,comment1}.
However, the earlier works on phase band crossing did not
systematically study the role of the drive protocol; further the
conditions for such crossings has not been methodically investigated
in terms of the parameters of the driven Hamiltonian beyond a few
simple protocols and toy models \cite{kita1,lind1,jiang,zhao}. In
this work, we aim to fill up this gap.
To this end, we study a class of periodically driven integrable
models whose Hamiltonian can be represented by free fermions in
$d$-dimensions:
\begin{eqnarray}
H(t)= \sum_{{\vec k}} \psi_{{\vec k}}^{\dagger} H_{{\vec k}}(t) \psi_{\vec
k}, \label{fermham}
\end{eqnarray}
where $\psi_{{\vec k}}= (c_{{\vec k}}, c_{-{\vec k}}^{\dagger})^T$ is the
two-component fermionic field, $c_{{\vec k}}$ are the annihilation
operators for the fermions, and $H_k(t)$ is given by
\begin{eqnarray} H_{{\vec k}} = (g(t) - b_{{\vec k}}) \tau^z + (\Delta_{{\vec k}}
\tau^+ + \rm{H.c.} ), \label{fermhamden} \end{eqnarray} where $g(t)$
is a periodic function of time characterized by a frequency
$\omega_D$, and $\Delta_{{\vec k}}$ can be an arbitrary function of
momenta. We note that this kind of Hamiltonian represents a wide
class of spin and fermionic integrable models such as the Ising and
$XY$ models in $d=1$ \cite{subir1}, the Kitaev model in $d=2$
\cite{kit1,nussinov1}, triplet and singlet superconductors in $d>1$,
and Dirac fermions in graphene and atop topological insulator
surfaces \cite{graphene1,topo1}. In what follows, we shall obtain
our results by analyzing fermionic systems given by Eq.\
\eqref{fermhamden} and point out the relevance of these results in
the context of specific models in appropriate places.
The main results that we obtain from such an analysis are the
following. First, we obtain an expression for the phase bands
corresponding to Hamiltonians given by Eq.\ \eqref{fermhamden}
within the adiabatic-impulse approximation \cite{nori1,kanu1,child1}
for arbitrary drive protocols. Using these expressions and other
general arguments, we chart out the most general conditions that
need to be satisfied for these phase bands to cross. The conditions
that we obtain conform to those obtained earlier for particle-hole
symmetric Hamiltonians \cite{rudner1} and for specific drive
protocols \cite{thakurathi1,thakurathi2}; however it constitutes a
more general result which holds for arbitrary periodic drive
protocols and irrespective of the symmetry of the underlying
Hamiltonian. Second, we show that traversing a critical point during
such periodic dynamics may lead to qualitatively different band
crossing conditions, and we discuss its implications for the
properties of the driven system. Third, for $d>1$, we compute the
off-diagonal fermionic correlation function $F_{{\vec k}}(t) = \langle
c_{\vec k}^{\dagger} c_{-{\vec k}}^{\dagger} \rangle$ and show that the
Fourier transform of this correlator will exhibit maxima and minima
at specific frequencies $\omega_0$ for $\vec k= \vec k_0$ at which
the bands cross; we provide an explicit relation between $\omega_0$,
$\omega_D$ and the band crossing time $t_0$ for several drive
protocols. Thus we show that $F_{\vec k_0}(\omega_0)$ carries
information about the phase band crossing time $t_0$. Finally, we
comment on the applicability of our results to general Hamiltonians
with $N>2$ phase bands and present a discussion of the role of
symmetries of the underlying Hamiltonian in the phase band
crossings.
The plan of the rest of the paper is as follows. In Sec.\
\ref{pcross}, we derive explicit expressions for the phase bands
within adiabatic-impulse approximation, obtain the conditions for
their crossings, and point out the role of critical points for such
crossings. This is followed by Sec.\ \ref{corr1}, where we chart out
the behavior of $F_{{\vec k}_0}(\omega_0)$ and discuss the signatures of
phase band crossings which can be inferred from its behavior.
Finally, we discuss the significance of our results for more general
phase band models, point out the role of symmetries for such
crossings, and conclude in Sec.\ \ref{diss}.
\section {Phase band crossings}
\label{pcross}
In this section, we first obtain an expression for the phase bands
corresponding to the Hamiltonian in Eq.~\eqref{fermhamden} within the
adiabatic-impulse approximation for an arbitrary continuous time
protocol in Sec.\ \ref{ad-imp}. This will be followed by an analysis of
the obtained expression for $U_{{\vec k}}(t)$ leading to the phase band
crossing conditions in Sec.\ \ref{pbcr}.
\subsection{Expression for the phase bands}
\label{ad-imp}
To obtain an expression for $U_{{\vec k}}(t)$ for an arbitrary drive
protocol $g(t)$, which is characterized by a frequency $\omega_D$, we
use an adiabatic-impulse approximation which has been used
extensively for two-level systems \cite{nori1,kanu1,child1}. We
envisage a drive protocol which starts at $t_i=0$ and continues till
the end of one drive period $t_f= T$. In the rest of this section,
we shall mostly work in the adiabatic basis in which the wave
function at any time $t$ is given by
\begin{eqnarray} |\psi_{{\vec k}}\rangle (t) &=& c_{1 {\vec k}}(t) \left(
\begin{array}{c} u_{0{\vec k}}(t) \\ v_{0{\vec k}}(t) \end{array} \right)
+ c_{2 {\vec k}}(t) \left(
\begin{array}{c} -v_{0{\vec k}}(t) \\ u_{0{\vec k}}(t) \end{array} \right),
\label{adbase} \end{eqnarray}
where $|\psi_{{\vec k}}^g (t)\rangle= (u_{0{\vec k}}(t), v_{0{\vec k}}(t))^T$ and
$-E_{{\vec k}}(t)$ are the instantaneous ground state wave function and energy
which are given by
\begin{eqnarray}
u_{0{\vec k}}(t) &=& -\frac{\Delta_{{\vec k}}}{D_{{\vec k}}(t)}, \quad v_{\vec
k}(t) = \frac{E_{{\vec k}}(t)+g(t)-b_{{\vec k}}}{D_{{\vec k}}(t)}, \nonumber \\
E_{{\vec k}}(t)&=& \sqrt{(g(t)-b_{{\vec k}})^2 + |\Delta_{{\vec k}}|^2},
\label{wav1} \\
D_{{\vec k}}(t) &=& \sqrt{( E_{{\vec k}}(t)+ g(t)-b_{{\vec k}})^2 +
|\Delta_{{\vec k}}|^2}. \nonumber \end{eqnarray}
The corresponding excited state wave function and energies are given
by $|\psi_{{\vec k}}^e (t)\rangle= (-v_{0{\vec k}}(t), u_{0\vec
k}(t))^T$ and $E_{{\vec k}}(t)$. We note here that the adiabatic and
the diabatic bases are connected by the standard transformation
\begin{eqnarray} \left( \begin{array}{c} |\psi_{{\vec k}}^g(t)\rangle \\
|\psi_{{\vec k}}^e (t) \rangle \end{array} \right) &=& \left( \begin{array}{cc}
\mu_{{\vec k}}(t) & \sqrt{1-\mu_{{\vec k}}^2(t)} \\
-\sqrt{1-\mu^2_{{\vec k}}(t)} & \mu_{{\vec k}}(t) \end{array} \right)
\left( \begin{array}{c} |\psi_{{\vec k}}^g\rangle \\
|\psi_{{\vec k}}^e \rangle \end{array} \right) \nonumber \\
\mu_{{\vec k}}(t) &=& u_{0{\vec k}}(t) u_{0{\vec k}} + v_{0{\vec k}}(t) v_{0{\vec k}},
\label{addirel} \end{eqnarray} where $u_{0{\vec k}} \equiv u_{0{\vec k}
}(t=0)$, similar notations have been used for all other quantities,
and in the rest of this section we shall assume the system to be in
its initial ground state at $t=0$. The unitary evolution operator
$U_{{\vec k}}(t)$ relates, by definition, the wave function at time $t$
to the initial wave function and thus satisfies
\begin{eqnarray} |\psi_{{\vec k}} (t)\rangle &=& U_{{\vec k}} (t)
|\psi_{{\vec k}}^g\rangle. \label{evoldef} \end{eqnarray}
\begin{figure}
\includegraphics[width=\linewidth]{fig1.pdf}
\caption{Schematic representation of the time-evolution of a
periodically driven system for a drive cycle $T= 2 \pi/\omega_D$.
The system reaches the critical points at $t=t_{1 {\vec k}} \equiv
t_1$ and $t_{2 {\vec k}} \equiv t_2$. The half-period $t= T/2$ is
marked with a hash. The adiabatic regimes before and after the first
crossing of critical point are marked as regions I and II respectively,
while that after the second crossing of the critical point is marked
as region III. See text for details.} \label{fig1} \end{figure}
The time evolution of a system described by Eq.\ \eqref{fermhamden}
is sketched in Fig.\ \ref{fig1}. We divide the time evolution into
three distinct regions as shown in Fig.\ \ref{fig1}. Within the
adiabatic-impulse approximation, regions I, II and III are adiabatic
regions. In these regions, the system is sufficiently far way from
the critical points, crossed at times $t_{1 {\vec k}}$ and $t_{2 \vec
k}$, so that the instantaneous energy gap for any ${\vec k}$ satisfies
the Landau criterion: $2 E_{{\vec k}}^2 (t) \gg dE_{{\vec k}}(t)/dt$. It can
be shown that in this regime the system merely gathers kinetic phase
\cite{nori1}
\begin{eqnarray} U'_{{\vec k}} (t_f, t_i) &=& \exp[ - i \xi_{{\vec k}}(t_f,t_i)
\tau^z], \nonumber \\
\xi_{{\vec k}} (t_f,t_i) &=& \int_{t_i}^{t_f} dt E_{{\vec k}}(t) \nonumber \\
\left( \begin{array}{c} c_{1 {\vec k}}(t_f) \\ c_{2 {\vec k}}(t_f) \end{array}
\right) &=& U'_{{\vec k}}(t_f,t_i) \left( \begin{array}{c}
c_{1 {\vec k}}(t_i) \\ c_{2 {\vec k}}(t_i) \end{array} \right). \label{evo1}
\end{eqnarray}
In the impulse region, the adiabaticity condition given by the
Landau criteria breaks down. For slow enough drives, this happens
around the critical point and for momentum modes sufficiently close
to the critical mode. The key idea of the adiabatic-impulse
approximation is to approximate the impulse region to be an
infinitesimally small region around the critical point; such an
approximation holds good for large amplitude and low frequency
drives \cite{nori1,kanu1,child1}. The impulse region is typically
reached twice during a drive cycle for any momenta ${\vec k}$, at $t=t_{1
{\vec k}}$ and $t_{2 {\vec k}}= 2\pi/\omega_D -t_{1 {\vec k}}$, where
$g(t_{1(2){\vec k}})=b_{{\vec k}}$. Around $t=t_{1(2) {\vec k}}$, we can linearize
the diagonal terms of the Hamiltonian (Eq.\ \eqref{fermhamden}) and
obtain
\begin{eqnarray} H_{{\vec k}}^{\rm imp} &=& v_{{\vec k}} (t-t_{1(2) {\vec k}})
\tau^z + (\Delta_{{\vec k}} \tau^+ + {\rm H.c.}), \label{impham} \end{eqnarray}
where $v_{{\vec k}}= \partial g(t_{1(2){\vec k}})/\partial t$.
\cite{comment2} The probability of excitation between the ground
and excited states at each ${\vec k}$ can be read off from Eq.\
\eqref{impham} as \cite{nori1}
\begin{eqnarray} p_{{\vec k}} &=& \exp[-2 \pi \delta_{{\vec k}}], \quad
\delta_{{\vec k}} = |\Delta_{{\vec k}}|^2/|2v_{{\vec k}}|. \label{probex}
\end{eqnarray}
We note that $p_{{\vec k}}=1$ for an unavoided level crossing which
happens for the critical mode (where $\Delta_{{\vec k}}=0$ and $g(t) -
b_{{\vec k}}$ crosses zero), while it is zero if $\Delta_{{\vec k}}=0$ but
$g(t) \ne b_{{\vec k}}$ at any point of time during the drive. It was
shown in Ref.\ \onlinecite{kanu1} that within this approximations we
can define a transfer matrix
\begin{eqnarray} S_{{\vec k}} &=& \left( \begin {array} {cc} \sqrt{1-p_{{\vec k}}}
e^{-i \tilde \phi_{s{\vec k}}} & - \sqrt{p_{{\vec k}}} \\ \sqrt{p_{{\vec k}}} &
\sqrt{1-p_{{\vec k}}}e^{i \tilde \phi_{s{\vec k}}} \end {array} \right),
\label{evo4} \\
\phi_{s {\vec k}} &=& \frac{\pi}{4} + \delta_{{\vec k}} \left( \ln
\delta_{{\vec k}} -1 \right) + {\rm Arg} ~\Gamma (1 - i \delta_{{\vec k}} ),
\label{smatrix1} \end{eqnarray} where $\phi_{s {\vec k}}$ is the Stoke's
phase and $\tilde \phi_{s\vec k} = \phi_{s \vec k}-\pi$. The change
of wave functions across the first transition point can be obtained
using the transfer matrix $S_{{\vec k}}$ as
\begin{eqnarray} \left( \begin{array}{c} c_{1 {\vec k}}(t_{1 {\vec k}} +
\epsilon) \\
c_{2 {\vec k}}(t_{1 {\vec k}} + \epsilon) \end{array} \right) &=& S_{{\vec k}}
\left( \begin{array}{c} c_{1 {\vec k}}(t_{1 {\vec k}} -\epsilon) \\
c_{2 {\vec k}}(t_{1 {\vec k}} - \epsilon) \end{array} \right), \label{evo2}
\end{eqnarray}
where $\epsilon>0$ is infinitesimally small. An analogous condition
with $S$ replaced by $S^T$ (where the superscript $T$ denotes
transpose) holds for the second transition point.
Having obtained these relations, we now obtain explicit analytical
expressions for the evolution operator $U_{{\vec k}}(t)$ in each of the
three regions shown in Fig.\ \ref{fig1}. In region I, the system
starts from the ground state so that $c_{1{\vec k}}(0)=1$ and $c_{2
{\vec k}} = 0$, Before crossing the critical point for the first time, the
dynamics is purely adiabatic leading to (using Eqs.~ \eqref{adbase} and
\eqref{evo1})
\begin{eqnarray} c_{1{\vec k}}^I(t) &=& \exp[-i\xi_{{\vec k}} (t,0)], \quad
c_{2 {\vec k}}^I (t)=0. \label{creg1} \end{eqnarray}
Using Eqs.\ \eqref{adbase}, \eqref{addirel} and \eqref{creg1}, we thus obtain
\begin{eqnarray} |\psi_{{\vec k}}(t)\rangle &=& e^{-i\xi_{{\vec k}} (t,0)} \left(
\mu_{{\vec k}}(t) |\psi_{{\vec k}}^g\rangle + \sqrt{1-\mu_{{\vec k}}(t)^2}
|\psi^e_{{\vec k}}\rangle \right). \label{wavreg12} \nonumber \\
\end{eqnarray}
Using Eqs.~\eqref{evoldef} and \eqref{wavreg12} and the fact that
$U_{{\vec k}}(t)$ is a unitary matrix, we obtain for region I
\begin{eqnarray} U_{{\vec k}}^{I}(t) &=& \left( \begin{array}{cc} \mu_{{\vec k}}(t)
e^{-i\xi_{{\vec k}} (t,0)} & -\sqrt{1-\mu_{{\vec k}}^2(t)} e^{i\xi_{{\vec k}}
(t,0)} \\
\sqrt{1-\mu_{{\vec k}}^2(t)} e^{-i\xi_{{\vec k}} (t,0)} & \mu_{{\vec k}}(t)
e^{i\xi_{{\vec k}} (t,0)} \end{array} \right). \label{uexpreg1} \nonumber \\
\end{eqnarray}
The eigenvalues $\lambda_{\pm {\vec k}}^{I}(t)$ of $U_{{\vec k}}^{I}(t)$
provide an expression for the phase bands in region I and are given by
\begin{eqnarray} \lambda_{\pm {\vec k}}^{I}(t) &=& e^{\pm i \phi_{{\vec k}}^{I}
(t)}, \nonumber \\
\cos[\phi_{{\vec k}}^I(t)] &=& \mu_{{\vec k}}(t) \cos(\xi_{{\vec k}}(t,0)).
\label{eigenreg1} \end{eqnarray}
An exactly similar method can be used to compute the evolution
operators in regions II and III. For example, in region II, using
Eqs.\ \eqref{adbase}, \eqref{evo1}, \eqref{evo4}, and
\eqref{smatrix1}, we obtain
\begin{eqnarray} \left( \begin{array}{c} c_{1{\vec k}}^{II}(t) \\
c_{2{\vec k}}^{II} (t) \end{array} \right) &=& U'_{{\vec k}}(t,t_{1\vec
k}) S_{{\vec k}} U'_{{\vec k}}(t_{1 {\vec k}},0) \left( \begin{array}{c} 1\\
0 \end{array} \right), \label{ceqreg2a} \end{eqnarray}
which leads to
\begin{eqnarray} c_{1{\vec k}}^{II}(t) &=& \sqrt{1-p_{{\vec k}}} e^{-i
\zeta_{1 {\vec k}}^{II}(t)}, \quad c_{2{\vec k}}^{II}(t) = \sqrt{p_{{\vec k}}} e^{i
\zeta_{2 {\vec k}}^{II}(t)}, \nonumber \\
\zeta_{1 {\vec k}}^{II}(t) &=& \tilde \phi_{s{\vec k}} + \xi_{{\vec k}}(t,0), \nonumber \\
\zeta_{2 {\vec k}}^{II}(t) &=& \xi_{{\vec k}}(t_{1{\vec k}},0)-\xi_{{\vec k}}
(t,t_{1{\vec k}}). \label{ceqreg2b} \end{eqnarray}
Using Eqs.~\eqref{addirel}, \eqref{evoldef} and \eqref{ceqreg2b}, we then
obtain
\begin{eqnarray} (U^{II}_{{\vec k}}(t))_{11} &=& c_{1{\vec k}}^{II}(t)
\mu_{{\vec k}}(t) + c_{2{\vec k}}^{II}(t) \sqrt{1-\mu_{{\vec k}}^2(t)}, \nonumber \\
(U^{II}_{{\vec k}}(t))_{21} &=& c_{2{\vec k}}^{II}(t) \mu_{{\vec k}}(t) -
c_{1{\vec k}}^{II}(t) \sqrt{(1- \mu_{{\vec k}}^2(t)}, \label{uexpreg2} \\
(U^{II}_{{\vec k}}(t))_{22}^{\ast} &=& (U^{II}_{{\vec k}}(t))_{11},
\quad (U^{II}_{{\vec k}}(t))_{12}^{\ast} = -(U^{II}_{{\vec k}}(t))_{21}. \nonumber
\end{eqnarray}
The expression for the phase bands in region II may be obtained by
diagonalizing the evolution matrix $U_{{\vec k}}^{II}(t)$. A
straightforward calculation yields the expressions for the
eigenvalues of $U_{{\vec k}}^{II}(t)$ as $ \lambda_{\pm \vec
k}^{II}(t) = \exp[\pm i \phi_{{\vec k}}^{II}(t)]$, where
\begin{eqnarray} \cos[\phi_{{\vec k}}^{II}(t)] &=& \mu_{{\vec k}}(t)
\sqrt{1-p_{{\vec k}}} \cos(\zeta_{1{\vec k}}^{II}(t)) \nonumber \\
&& + \sqrt{p_{{\vec k}}(1-\mu_{{\vec k}}^2(t))} \cos(\zeta^{II}_{2{\vec k}}(t)).
\label{eigenreg2} \end{eqnarray}
A similar calculation can be carried out in region III. Here one
finds that
\begin{eqnarray} \left( \begin{array}{c} c_{1{\vec k}}^{III}(t) \\
c_{2{\vec k}}^{III} (t) \end{array} \right) &=& U'_{{\vec k}}(t,t_{2\vec
k}) S^T_{{\vec k}} U'_{{\vec k}}(t_{2 {\vec k}},0) \left( \begin{array}{c} 1\\
0 \end{array} \right), \label{ceqreg3a}
\end{eqnarray}
where $S_{{\vec k}}^T$ is the transpose of $S_{{\vec k}}$ (Eqs.\ \eqref{evo4}
and \eqref{smatrix1}). A straightforward calculation then leads to
\begin{eqnarray} c_{1{\vec k}}^{III}(t) &=& (1-p_{{\vec k}}) e^{-i \zeta_{1 \vec
k}^{III}(t)} + p_{{\vec k}} e^{-i \zeta_{2 {\vec k}}^{III}(t)}, \nonumber \\
c_{2{\vec k}}^{III}(t) &=& \sqrt{p_{{\vec k}}(1-p_{{\vec k}})} e^{i
\phi_{s {\vec k}}} \left( e^{-i \zeta_{1 {\vec k}}^{III}(t)} - e^{-i
\zeta_{2 {\vec k}}^{III}(t)}
\right), \nonumber \\
\zeta_{1 {\vec k}}^{III}(t) &=& 2 \tilde \phi_{s{\vec k}} + \xi_{{\vec k}}(t,0), \nonumber \\
\zeta_{2 {\vec k}}^{III}(t) &=& \xi_{{\vec k}}(t_{1{\vec k}},0)-\xi_{{\vec k}}(t_{2
{\vec k}},t_{1{\vec k}}) + \xi_{{\vec k}}(t,t_{2{\vec k}}). \label{ceqreg3b}
\end{eqnarray}
Using Eqs.~\eqref{addirel}, \eqref{evoldef} and \eqref{ceqreg2b}, we
obtain $U_{{\vec k}}^{III}(t)$ in an analogous manner. The
expressions for the phase bands in region III are then obtained by
diagonalizing $U_{{\vec k}}^{III}$ and yield $ \lambda_{\pm \vec
k}^{III}(t) = \exp[\pm i \phi_{{\vec k}}^{III}(t)]$, where
\begin{eqnarray} && \cos[\phi_{{\vec k}}^{III}(t)] = \mu_{{\vec k}}(t) \left(
(1-p_{{\vec k}})\cos(\zeta_{1{\vec k}}^{III}(t)) \right. \nonumber \\
&& + \left. p_{{\vec k}} \cos(\zeta_{2{\vec k}}^{III}(t))\right) +
\sqrt{p_{{\vec k}}(1-p_{{\vec k}})(1-\mu_{{\vec k}}^2(t))} \nonumber \\
&& \times \left( \cos(\zeta^{III}_{1{\vec k}}(t) - \tilde \phi_{s{\vec k}})-
\cos(\zeta^{III}_{2{\vec k}}(t) - \tilde \phi_{s{\vec k}}) \right).
\label{eigenreg3}
\end{eqnarray}
Eqs.\ \eqref{eigenreg1}, \eqref{eigenreg2}, and \eqref{eigenreg3}
constitute the central results of this section. These equations
provide us with analytic expressions for the phase bands for
integrable models studied for an arbitrary continuous time protocol.
In what follows, we shall analyze these expressions to obtain
conditions for phase band crossings for these models.
\begin{figure}
\includegraphics[width=\linewidth]{fig2.pdf}
\caption{A plot of $\cos(\phi_k(t))$ as a function of $t/T$ for (a)
$k=\pi/20$, (b) $k=\pi/10$, (c) $k= \pi/6$, and (d) $k= \pi/3$ for
the 1D Ising model in a transverse field for the drive protocol
$h(t)= h_0 + h_1 \cos(\omega_D t)$ with $h_0=1.1$, $h_1=-1$, and
$\omega_D=0.1$ with all energy scales measured in units of $J$, and
$\hbar$ is set equal to unity. The black solid line correspond to
exact numerical solution of Eqs.\ \eqref{fermsch}, while the red
dotted lines correspond to those obtained using adiabatic-impulse
approximation (Eqs.\ \eqref{eigenreg1}, \eqref{eigenreg2}, and
\eqref{eigenreg3}). The match between the two deteriorates with
increasing $\omega_D$ except at $k=0, \pi$ where the approximation
yields exact results.} \label{fig2} \end{figure}
Before ending this section, we provide a comparison between the
exact numerical values of the phase bands with those obtained by our
method for the $d=1$ Ising model in a transverse field. As is
well-known, Eq.\ \eqref{fermhamden} provides a fermionic
representation of the transverse field Ising model with $b_{k}= \cos
k$, $\Delta_{k} = \sin k$ and $g(t)=h(t)$, where we have set the
nearest-neighbor coupling $J$ between the spins and the lattice
spacing $a$ to unity \cite{subir1}. The critical point for this
system is located at $h=1$. In what follows, we choose $h(t)=h_0 +
h_1 \cos(\omega_D t)$ with $h_0=1.1$ and $h_1=-1$ so that the system
traverses the critical point twice during the dynamics. To obtain
the phase bands, we note that the Schr\"odinger equation
corresponding to the fermionic Hamiltonian in Eq.\
\eqref{fermhamden} is given by
\begin{eqnarray} i \partial_t u_k(t) &=& (h(t)-\cos k) u_k + i \sin k \, v_k,
\nonumber \\
i \partial_t v_k (t) &=& -(h(t)-\cos k) v_k - i \sin k \,u_k. \label{fermsch}
\end{eqnarray}
We solve these equations numerically with the initial condition
$(u_k(0),v_k(0))=(u_{0 k}, v_{0 k})$ and obtain the wave function
$(u_k(t),v_k(t))$ at any time $t \le T$. The unitary evolution
operator $U_k(t)$ is then obtained using Eq.\ \eqref{evoldef} from
the initial and final wave functions. Finally, we diagonalize
$U_k(t)$ to obtain its eigenvalues and hence the phase bands
$\lambda_{k \pm}(t) = \exp[\pm i \phi_k(t)]$. A plot of
$\cos(\phi_k(t))$ obtained in this manner is compared to their
adiabatic-impulse counterparts for several representative values of
$k$ and for $\omega_D=0.1$ as shown in Fig.\ \ref{fig2}; the figure
shows a near exact match between the two for all $k$. This feature
is expected since the adiabatic-impulse approximation becomes
accurate for small $\omega_D$ for any $k$; we note here that it is
exact at $k=0, \pi$ for all $\omega_D$. Thus our analytical approach
provides a decent approximation to the phase bands for all $k$ at
low drive frequencies. The structure of these phase bands as a
function of $k$ and $t/T$ is shown in Fig.\ \ref{fig3} for
representative values of $\omega_D$ and $h(t)$; we note that these
bands never reach the values 0 or $\pm \pi$ unless $k=0, \, \pi$. We
shall analyze this fact in detail in the next section.
\begin{figure}
\includegraphics[width=0.47\linewidth]{fig3a.pdf}
\includegraphics[height= 3.4 cm, width=0.47\linewidth]{fig3b.pdf}
\caption{Left Panel: Plot of $\cos(\phi_{k}(t))$ as a function of $k$ and
$t/T$ for the ID Ising model with $\omega_D=0.25$. All other parameters are
the same as in Fig.\ \ref{fig2}. Right Panel: A close-up of $\cos \phi_{k}(t)$
showing a phase-band crossing at $k=\pi$ and $t/T \simeq 0.245$. All
parameters are the same as those in the left panel.} \label{fig3} \end{figure}
\subsection{Phase band crossing conditions}
\label{pbcr}
The phase bands $\lambda_{ {\vec k} \pm}^{a}(t) = \exp[\pm i
\phi^a_{{\vec k}} (t)]$ (where $a=$ I, II or III) whose expressions
were obtained in Sec.\ \ref{ad-imp}, cross when $\phi_{\vec
k}^{a}(t) = n \pi$ for any integer $n$. We note that all such
crossings for the integrable models that we study constitute
examples of zone-edge singularities \cite{rudner1} and are therefore
topologically significant. To understand when such crossings can
happen, we first note that $\mu_{{\vec k}}(t) \le 1$ since it denotes the
overlap between ground state wave functions at different times. This
property of $\mu_{{\vec k}}(t)$ ensures that the right hand sides of Eqs.\
\eqref{eigenreg1}, \eqref{eigenreg2}, and \eqref{eigenreg3} can becomes
unity only when $\mu_{{\vec k}}(t)= 1 \, {\rm or} \, 0$. This, in turn,
can occur only for momenta ${\vec k} = {\vec k}_0$ for which $\Delta_{{\vec k}_0}
=0$; thus phase band crossings only occur at these momenta.
This condition for phase band crossings happens to be a general
result which may also be understood from simpler intuitive
arguments. We present two such equivalent arguments here. The first
of these involves a geometrical construction which invokes the
concept of the Bloch sphere. Since the generators of $U_{{\vec k}}(t)$
belong to the SU(2) algebra, the trajectory of any wave function
under its action can be considered as a trajectory on the Bloch
sphere characterized by a fixed momentum ${\vec k}$. Thus the requirement
for a phase band crossing at $U_{{\vec k}_0}(t_0)$ amounts to the
condition $|\psi_{{\vec k}_0}(0)\rangle = |\psi_{\vec k_0}(t_0)\rangle$,
{\it i.e.}, the trajectory of the wave function for ${\vec k} = {\vec k}_0$
must cross itself at $t=t_0$ under the action of $U_{{\vec k}_0}(t_0)$.
Since the eigenvalues of $U_{\vec k}(t)$ are independent of the
initial wave function $|\psi_{\vec k}(0)\rangle$, this condition
requires that the trajectory on the Bloch sphere, no matter where it
starts, must pass through itself at $t=t_0$ during its evolution.
This condition can be generically satisfied if that trajectory is
generated by rotation around a single axis {\it at all times}. Thus
such crossings can only occur if $\Delta_{{\vec k}}=0$.
The second, equivalent, argument showing that $\Delta_{{\vec k}}=0$
constitutes a necessary condition for phase band crossings is as
follows. A Trotter decomposition of Eq.\ \eqref{evop} enables us to
write $U_{{\vec k}}(t)$ at the time $t_0$ when the phase band crosses at
any given ${\vec k}$ as
\begin{eqnarray} U_{{\vec k}}(t_0,0) &=& \lim_{N \to \infty} \prod_{j=0, N-1}
U_{\vec k} (t_{j+1},t_j) \label{trot1} \\
&=& \lim_{N \to \infty} \prod_{j= \ell+1, N-1} U_{{\vec k}}
(t_{j+1},t_{j}) \prod_{j=0, \ell } U_{{\vec k}} (t_{j+1},t_j), \nonumber
\end{eqnarray} where $t_j-t_{j-1} = \Delta t= t_0/N$, $t_0=t_{N}$,
and in the second line we have organized the product into two groups
for which $t_j \le t_{\ell}$ and $t_j > t_{\ell}$. Note that the
choice of $ t_{\ell}$ is completely arbitrary. Now since the product
of these evolution matrices in each of the two groups must also be a
SU(2) rotation matrix, we can write
\begin{eqnarray} U_{{\vec k}}(t_0,0) &=& U_{2 \vec k}(t_0,t_{\ell+1})
U_{1 {\vec k}}(t_{\ell+1},0) \nonumber \\
&=& e^{-i(\vec \sigma \cdot {\vec n}_{2{\vec k}})\phi_{2 {\vec k}}} e^{-i(\vec
\sigma \cdot {\vec n}_{1 {\vec k}})\phi_{1 {\vec k}}}, \label{trot2} \end{eqnarray}
where ${\vec n}_{1(2) {\vec k}}$ are unit vectors and $\phi_{1(2) {\vec k}}$ are
rotation angles. All of these parameters depend, in general, on the
choice of $t_{\ell}$, the details of $H_{{\vec k}}(t)$, and the drive
protocol. We do not attempt to compute them here; instead, we merely
observe that in order to get $U_{{\vec k}}(t_0)=I$ {\it independent of
the choice of $t_{\ell}$}, we clearly require ${\vec n}_{2 {\vec k}} = {\vec n}_{1
{\vec k}}$. This is most easily seen by choosing $n_{1 \vec k}= \hat z$
which can be done without loss of generality by choosing suitable
axes, and then checking that the eigenvalues of $U_{\vec k}$ can
never be unity if $n_{2 \vec k} \ne \hat z$. Next we note that the
condition $n_{2 \vec k}= n_{1 \vec k}$ can only be satisfied for
arbitrary $t_{\ell}$ if $U_{{\vec k}}(t_i)$ commutes at all $t_i$, {\it
i.e.}, if the rotation axis of $U_{{\vec k}}(t_{j+1},t_j)$ is the same
for all $t_j$. In the context of the Hamiltonian $H_{{\vec k}}(t)$ given
in Eq.\ \eqref{fermhamden}, this is only possible if
$\Delta_{{\vec k}}=0$. In a more general context, the Hamiltonian of the
system can be written as
\begin{eqnarray} H_{{\vec k}}(t) &=& \sum_{i=1}^3 \tau_i f_{i {\vec k}}(t),
\label{genham} \end{eqnarray} where $f_{i \vec k}(t)$ are parameter
functions and the Pauli matrices $\tau_i$ are the generators. The
condition for the phase band crossings requires that any two of the
three parameter functions $f_{i {\vec k}}(t)$ vanishes at all times. The
third parameter function then determines $t_0$, and we discuss this
issue in detail below. Finally, we note that the above mentioned
arguments indicate that for arbitrary single-rate protocols and for
$d=1$, the phase band crossings can therefore only occur at $k=0$ or
$\pi$. In contrast, for $d>1$ systems such crossings can occur for a
wider range of momenta.
Next, we chart out the condition for phase band crossing at ${\vec k}=
{\vec k}_0$ which yields the band crossing time $t_0$ for each of the
regions shown in Fig.\ \ref{fig1}. In region I, we find from
Eqs.~\eqref{evo1} and \eqref{eigenreg1} that the phase bands will
cross at time $t_0 < t_{1{\vec k}_0}$ provided that
\begin{eqnarray} \xi_{{\vec k}_0}(t_0,0)= n \pi , \quad n \in Z. \label{croscon1}
\end{eqnarray}
To find the condition for phase band crossings in region II, we need
to find $p_{{\vec k}_0}$. Since at ${\vec k}= {\vec k}_0$, $\Delta_{\vec
k_0}=0$, $p_{{\vec k}_0}$ may assume the values $1$ or $0$. The former
occurs at the critical mode where the instantaneous energy levels of
$H$ cross while for the latter they do not cross. In what follows, we
denote the momenta of the critical modes to be ${\vec k}_{0}$ and for
that of the non-critical modes to be ${\vec k}'_0$. Within our
convention, in $d=1$, $k_0=0$ and $k'_0= \pi$. For the critical
modes, the crossing of the instantaneous energy levels ensures that
there is no overlap of the instantaneous ground state wave function
in region II with the initial ground state wave function. This
corresponds to $\mu_{{\vec k}_0}(t)=0$ and leads to $\phi_{\vec
k_0}^{II}(t) = \zeta^{II}_{2{\vec k}_0}(t)$. Thus using Eqs.\
\eqref{ceqreg2b} and \eqref{eigenreg2}, we obtain
\begin{eqnarray} \xi_{{\vec k}_0}(t_{1 {\vec k}_0},0)-\xi_{{\vec k}_0}(t_0,
t_{1 {\vec k}_0}) &=& n \pi. \label{croscon2} \end{eqnarray}
For the non-critical momenta $k'_0$, when the instantaneous energy
bands do not cross even though $\Delta_{k'_0}$ vanishes, $p_{\vec
k'_0}=0$ and $\mu_{{\vec k}'_0}(t)=1$ in region II. Using Eq.\
\eqref{eigenreg2}, we find that in this case $\phi_{{\vec k}'_0}^{II}(t) =
\zeta^{II}_{1{\vec k}'_0}(t)$, and the phase bands cross if Eq.\ \eqref{croscon1}
holds, with $t_{1 {\vec k}_0} \le t_0 \le t_{2 {\vec k}_0}$.
Finally, we obtain the conditions for such band crossings in region
III. In this case, since the critical point is crossed again,
$\mu_{{\vec k}}(t)=1$ for all ${\vec k} = {\vec k}_0, {\vec k}'_0$. For the
critical modes, where $p_{{\vec k}_0}=1$, using Eqs.~\eqref{evo1} and
\eqref{eigenreg3}, we find
\begin{eqnarray} && \xi_{{\vec k}_0}(t_{1{\vec k}_0},0)-\xi_{{\vec k}_0}(t_{2\vec
k_0},t_{1{\vec k}_0}) + \xi_{{\vec k}_0}(t_0, t_{2 {\vec k}_0}) = n \pi.
\label{croscon3} \end{eqnarray}
For the non-critical modes, where the instantaneous energy levels do
not cross at $t_{1 {\vec k}}$ or $t_{2 {\vec k}}$, $p_{{\vec k}'_0}=0$ and
we find from Eq.\ \eqref{eigenreg3} that the crossing condition is
given by Eq.\ \eqref{croscon1} with $t_0 \ge t_{2 {\vec k}}$.
Eqs.\ \eqref{croscon1}, \eqref{croscon2} and \eqref{croscon3} constitute
general conditions for phase band crossings for the integrable models
that we study. Although we have obtained them using the
adiabatic-impulse approximation, they are essentially exact since the
adiabatic-impulse approximation becomes exact for modes for which
$\Delta_{{\vec k}_0}=0$. Our results also demonstrate that the
conditions for the phase band crossings at the critical mode (Eqs.\
\eqref{croscon2} and \eqref{croscon3}), where we find an unavoided
crossing of the instantaneous energy levels, is different from those
of the non-critical modes with no instantaneous energy level
crossings. The origin of this difference can be easily traced to the
fact that at each such crossing $g(t)-b_{{\vec k}_0}$ changes sign;
thus the sign of the phase accumulated is reversed in region II
which leads to the difference between Eqs.\ \eqref{croscon2} and
\eqref{croscon3} with Eq.\ \eqref{croscon1}. We also note that the
difference in phase band crossing conditions for the critical modes
that we unravel here is absent for all protocols where the critical
point/region is traversed instantly, {\it i.e.}, at an infinite
rate. Examples of such protocols include periodic arrays of $\delta$-function
kicks and square pulses studied earlier
\cite{thakurathi1,thakurathi2,rudner1}; for these protocols $t_{2 \vec
k}=t_{1 {\vec k}}$ for all ${\vec k}$. Consequently Eqs.\ \eqref{croscon2}
and \eqref{croscon3} become identical to Eq.\ \eqref{croscon1}.
Having established the crossing conditions for a generic protocol,
we now study specific examples of such crossing for the 1D
transverse field Ising model and the 2D Kitaev model. For the transverse
field Ising model, the crossings occur at $k=0$ or $k=\pi$. For the
latter mode, there is no band crossing and the crossing condition is
given by Eq.\ \eqref{croscon1}, while for the former mode, the bands
cross when $h(t)=1$ and the crossing conditions in regions II and
III are given by Eqs.\ \eqref{croscon2} and \eqref{croscon3}. Using the
protocol $h(t)=h_0 + h_1 \cos(\omega_D t)$, we find that the band
crossing conditions at $k= \pi$ in all three regions is given by
\begin{eqnarray} h_1 \sin(x) + (h_0 +1) x &=& n \pi \omega_D,
\label{kpicondition} \end{eqnarray}
where $x= \omega_D t$ and $n$ is an integer. For the $k=0$ mode, the
conditions in region I, II and III can be obtained from Eqs.\
\eqref{croscon1}, \eqref{croscon2} and \eqref{croscon3}. Noting that
$E_0(t) = |h(t)-1|$ and $h(t)-1$ changes sign at $t_1$ and $t_2$,
we can combine the conditions in Eqs.\ \eqref{croscon1}, \eqref{croscon2},
and \eqref{croscon3} to obtain
\begin{eqnarray} h_1 \sin(x) + (h_0 -1) x &=& n \pi \omega_D.
\label{k0condition} \end{eqnarray}
We note that for $n=0$, the crossing conditions imply that
$\omega_D t_0$ is constant which indicates that such a crossing spans
over a range of frequencies. This observation is verified from the plot
of the phase bands at $k=0$ as a function of $\omega_D$ and $t/T$.
We find that the crossing time $t/T$ corresponding to
$\phi_0(t_0)=0$ which occurs in region II is independent of $
\omega_D$; thus $\cos[\phi_0(t)] =1 $ for any $\omega_D$ at
$t_0/T=0.45$ as can be seen in Fig.\ \ref{fig4}. A similar plot for
$k=\pi$ (Fig. \ref{fig5}) does not show this behavior; for $k= \pi$
and for our choice of $h_0$ and $h_1$ the crossing always occur at
$n \ne 0$ (Eq.\ \eqref{kpicondition}).
\begin{figure}
\includegraphics[width=0.47\linewidth]{fig4a.pdf}
\includegraphics[height=3.2cm, width=0.47\linewidth]{fig4b.pdf}
\caption{Left Panel: A plot of $\cos(\phi_0(t))$ as a function of
$\omega_D$ and $t/T$ for the 1D transverse field Ising model for the
drive protocol $h(t)= h_0 + h_1 \cos(\omega_D t)$, with $h_0=1.1$, and
$h_1=-1$. Note the extended region around $t/T=0.45$ where
$\phi_0(t)$ vanishes independently of the value of $\omega_D$. See text
for details. Right Panel: A closer look at a phase band crossing
corresponding to $n=1$ highlighting the local nature of the crossing
as a function of $t/T$ and $\omega_D$.} \label{fig4} \end{figure}
Next, we consider the Kitaev model in $d=2$.~\cite{kit1,nussinov1} This model
consists of spin-1/2's on a honeycomb lattice with nearest-neighbor
interactions described by the Hamiltonian
\bea H &=& \sum_{j+l={\rm even}} \left( J_1 \sigma_{j+1,l}^x \sigma_{j,l}^x ~+~ J_2
\sigma_{j-1,l}^y \sigma_{j,l}^y \right. \nonumber \\
&& ~~~~~~~~~~~~~~~~\left. + ~J_3 \sigma_{j,l+1}^z \sigma_{j,l}^z \right),
\label{kitham} \eea
where $(j,l)$ denotes coordinates of a site on the honeycomb lattice, and
$J_{1,2,3}$ are the couplings between the $x,y,z$ components of neighboring
spins. The unit cell of the system has two sites which we label as $a$ and
$b$. Denoting the location of a
unit cell by ${\vec n}$, a Jordan-Wigner transformation takes us from spin-1/2's to
two Hermitian (Majorana) fermion operators in each unit cell labeled
as $a_{\vec n}$ and $b_{\vec n}$. We can go to momentum space by defining
\bea a_{\vec n} &=& \sqrt{\frac{4}N{}} ~\sum_{\vec k} ~[a_{\vec k} ~e^{i {\vec k} \cdot {\vec n}} ~+~
a_{\vec k}^\dagger ~e^{-i {\vec k} \cdot {\vec n}}], \nonumber \\
b_{\vec n} &=& \sqrt{\frac{4}N{}} ~\sum_{\vec k} ~[b_{\vec k} ~e^{i {\vec k} \cdot {\vec n}} ~+~
b_{\vec k}^\dagger ~e^{-i {\vec k} \cdot {\vec n}}], \label{abkn} \eea
where $N$ is the number of sites (the number of unit cells is $N/2$), and
the sum over ${\vec k}$ goes over half the Brillouin zone.
A convenient choice of the Brillouin zone is given by a rhombus whose
vertices lie at $(k_x,k_y)= (\pm 2\pi, 0)$ and $(0,\pm 2\pi /\sqrt{3})$;
half the Brillouin zone is then an equilateral triangle with vertices at
$(0,\pm 2\pi /\sqrt{3})$ and $(2\pi,0)$.
Eq.~\eqref{kitham} leads to a fermionic Hamiltonian of the form
\bea H &=& \sum_{\vec k} ~\left( \begin{array}{cc}
a_{\vec k}^\dagger & b_{\vec k}^\dagger \end{array} \right) ~H_{\vec k} ~\left(
\begin{array}{c}
a_{\vec k} \\
b_{\vec k} \end{array} \right), \nonumber \\
H_{\vec k} &=& 2 [J_1 \sin({\vec k} \cdot \vec{M}_1)-J_2 \sin({\vec k} \cdot \vec{M}_2)]
\tau^x \nonumber \\
&& + 2 [J_3+J_1 \cos({\vec k} \cdot \vec{M}_1)+J_2 \cos({\vec k} \cdot \vec{M}_2)]
\tau^y, \label{hkitaev} \eea
where $\vec{M}_1 = \frac{1}{2} \hat{i} + \frac{\sqrt{3}}{2} \hat{j}$ and
$\vec{M}_2 =\frac{1}{2} \hat{i} - \frac{\sqrt{3}}{2} \hat{j}$ are spanning
vectors which join some neighboring unit cells ($\hat{i}$ and $\hat{j}$
denote unit vectors in the $x$ and $y$ directions, and we have taken the
nearest-neighbor spacing to be equal to $1/\sqrt{3}$), and $\tau^a$ are
Pauli matrices in the $a, ~b$ space.
We now drive $J_3$ is periodically in time as \beq J_3 (t) ~=~ h_0
~+~ h_1 \cos (\omega_D t). \label{j3t} \eeq To see what happens to the
phase bands, let us consider the case $J_1 = J_2$ for simplicity. We
then obtain
\bea H_{\vec k} &=& 4 J_1 \cos (k_x/2) \sin (\sqrt{3} k_y/2) \tau^x \nonumber \\
&& + [2J_3 (t) + 4 J_1 \cos (k_x/2) \cos (\sqrt{3} k_y/2)] \tau^y.
\label{hamkit} \eea The form of this is similar to that in
Eq.~\eqref{fermhamden}, except that $\tau^z$ has been replaced by
$\tau^y$; indeed these two Hamiltonians are related to each other by
a global unitary transformation \cite{ks1}. By the arguments given
earlier, we therefore see that phase band crossings will occur at a
momentum ${\vec k}_0$ and time $t_0$ given by
\bea \cos (k_{0x}/2) \sin (\sqrt{3} k_{0y}/2) &=& 0, \label{phasebandkitaev} \\
(h_0 + 2 J_1 \cos (\frac{k_{0x}}{2}) \cos (\frac{\sqrt{3}
k_{0y}}{2}))x + h_1 \sin(x) &=& \frac{n \pi \omega_D}{2}, \nonumber \eea
where $x= \omega_D t_0$. We now observe that the first equation in
Eq.~\eqref{phasebandkitaev} is satisfied for momenta lying on one of
two lines in half the Brillouin zone; the two lines are given by \\ \\
\noindent (i) $k_{0x} = \pi$ and $-\pi/\sqrt{3} \le k_{0y} \le \pi/\sqrt{3}$, \\
\noindent (ii) $k_{0y} = 0$ and $0 \le k_{0x} \le 2 \pi$. \\ \\
Correspondingly, the second equation in Eq.~\eqref{phasebandkitaev}
implies that we must have \beq h_0 x ~+~ h_1 \sin (x) ~=~ \frac{n
\pi \omega_D}{2} \label{line1} \eeq on line (i), and \beq (h_0 + 2
J_1 \cos (k_{0x}/2))x ~+~ h_1 \sin (x) ~=~ \frac{n \pi \omega_D}{2}
\label{line2} \eeq on line (ii). Interestingly,
Eqs.~(\ref{line1}-\ref{line2}) show that the phase band crossing
time $t_0$ is independent of the location of the momentum on line
(i) but depends on the location of the momentum $k_{0x}$ on line
(ii). Analogous conditions for the case $J_1 \ne J_2$ can be
obtained by a similar analysis of Eq.\ \eqref{hamkit};
however analytical expressions similar to Eq. \eqref{phasebandkitaev}
may be difficult to obtain in such cases.
To conclude, in the 2D Kitaev model, phase band crossings occur on
certain lines in momentum space. This is in contrast to 1D models
like the Ising model in a transverse field where phase band
crossings occur only at some discrete momenta, namely, 0 and $\pi$.
This difference constitute a concrete example of the symmetry based
arguments provided in Ref.\ \onlinecite{rudner1}.
\begin{figure}
\includegraphics[width=\linewidth]{fig5.pdf}
\caption{A plot of $\cos(\phi_{\pi}(t))$ as a function of $\omega_D$
and $t/T$ for the 1D transverse field Ising model for the drive
protocol $h(t)= h_0 + h_1 \cos(\omega_D t)$ with $h_0=1.1$, and
$h_1=-1$. Note that in contrast to Fig.\ \ref{fig4}, the phase band crossings
(bright yellow regions) depend on $\omega_D$.} \label{fig5} \end{figure}
\section{Fermionic correlators}
\label{corr1}
In this section, we show that the phase band crossings leave their
imprint on the Fourier transform of some of the off-diagonal
fermionic correlators. To this end, we recall that the
time-dependent Hamiltonians $H$ given by Eqs. \eqref{fermham} and
\eqref{fermhamden}. Here and in the rest of this section, we shall
assume $\Delta_{\vec k}$ to be real for simplicity; however our analysis
may be readily generalized to complex $\Delta_{{\vec k}}$. To obtain
the correlators, we first note that if $(u_{{\vec k}},v_{{\vec k}})^T$
is an eigenvector of $H_{\vec k}$, the eigenstates of $H_{{\vec k}}$ in
second quantized form is given by $(u_{{\vec k}} + v_{{\vec k}} c_{\vec
k}^\dagger c_{-{\vec k}}^\dagger ) | vac \rangle$. In this state, we find that
$\langle c_{{\vec k}}^\dagger c_{{\vec k}} \rangle = \langle c_{-{\vec k}}^\dagger c_{-{\vec k}}
\rangle = |v_{{\vec k}}|^2$ and $\langle c_{{\vec k}}^\dagger c_{-{\vec k}}^\dagger \rangle =
u_{{\vec k}} v_{{\vec k}}^*$. In what follows, we rewrite the
expressions for these correlators in a different way so as to point
out their connections to the phase bands. To do this, let us denote
the eigenvectors of $U_{{\vec k}}(t)$ corresponding to the eigenvalues
$\lambda_{\pm {\vec k}}(t) = \exp[i \phi_{\pm {\vec k}}(t)] = \exp[\pm i
\phi_{{\vec k}}(t)]$ as
\begin{eqnarray}
|\chi_{\pm {\vec k}}(t)\rangle = \left( \begin{array}{c} \mu_{\pm \vec
k}(t)
\\ \nu_{\pm {\vec k}}(t) \end{array} \right). \label{eigenvecU}
\end{eqnarray}
Note that $|\chi_{\pm {\vec k}}(t) \rangle$ forms a complete basis.
Using these eigenvectors and eigenvalues, we can now generic
expressions for both the off-diagonal and the diagonal correlators in
terms of these eigenvectors and eigenvalues as
\begin{eqnarray}
C_{{\vec k}}(t) &=& \langle \psi_{{\vec k}}(0)|U^{\dagger}_{{\vec k}}(t)
c_{{\vec k}}^{\dagger} c_{{\vec k}}U_{{\vec k}}(t) |\psi_{{\vec k}}(0) \rangle \nonumber \\
&=& \sum_{a,b= \pm} (u^0_{{\vec k}} \mu_{a {\vec k}}(t) + v_{{\vec k}}^0
\nu_{a{\vec k}}(t)) \nu^{\ast}_{a {\vec k}}(t) \nu_{b{\vec k}}(t) \nonumber \\
&& \times e^{-i(\phi_{a {\vec k}}(t)- \phi_{b {\vec k}}(t))} (u^0_{\vec
k} \mu_{b {\vec k}}(t) + v_{{\vec k}}^0 \nu_{b{\vec k}}(t)), \nonumber \\
F_{{\vec k}}(t) &=& \langle \psi_{{\vec k}}(0)|U^{\dagger}_{{\vec k}}(t)
c^{\dagger}_{-{\vec k}} c^{\dagger}_{{\vec k}} U_{{\vec k}}(t) |\psi_{\vec
k}(0) \rangle \nonumber \\
&=& \sum_{a,b= \pm} (u^0_{{\vec k}} \mu_{a {\vec k}}(t) + v_{{\vec k}}^0
\nu_{a{\vec k}}(t)) \mu^{\ast}_{a {\vec k}}(t) \nu_{b{\vec k}}(t) \nonumber \\
&& \times e^{-i(\phi_{a {\vec k}}(t)- \phi_{b {\vec k}}(t))} (u^0_{\vec
k} \mu_{b {\vec k}}(t) + v_{{\vec k}}^0 \nu_{b{\vec k}}(t)). \label{corexp0}
\end{eqnarray}
Thus we find that the phase bands contribute to the terms in the correlators
for $a \ne b$. For ${\vec k}= {\vec k}_0$, for which $|\chi_{\pm {\vec k}}(t)
\rangle$ are eigenstates of $\tau_z$ at all times, time-dependent
contributions from the phase bands only appear in $F_{{\vec k}}(t)$, since
$C_{{\vec k}}(t)$ receives contribution only from the $a=b$ terms
which are time-independent. Finally, we note that for $d=1$ where the phase
bands cross at $k=0, \pi$, where $c_{k_0}^{\dagger} c_{-k_0}^{\dagger}=0$ due
to Pauli exclusion; however these correlators are non-zero for $d>1$ where
such crossing may occur at ${\vec k}_0 \ne 0, \pi$.
For ${\vec k} = {\vec k}_0$, the phase bands are given by $\phi_{\pm {\vec k}_0}(t) = \pm
\int_0 ^t dt' (g(t') -b_{{\vec k}_0})$, and the corresponding eigenvectors are
$|\chi_{+ \vec k_0}(t)\rangle = (1,0)^T$ and $|\chi_{- {\vec k}_0}(t)\rangle=
(0,1)^T$. Substituting these in Eq.\ \eqref{corexp0}, we obtain
$C_{{\vec k}_0}(t) = |v^0_{{\vec k}_0}|^2$ (which is independent of time) and
\begin{eqnarray}
F_{{\vec k}_0}(t) = u_{{\vec k}_0}^0 v_{{\vec k}_0}^{0 \ast} \exp [-2 i \int_0^t dt' (g(t')
- b_{{\vec k}_0})]. \label{corexp1} \end{eqnarray}
We note that if the phase bands cross at $t=t_0$, we have $F_{\vec
k_0}(t_0)= F_{{\vec k}_0}(0) = u_{{\vec k}_0}^0 v_{{\vec k}_0}^0$. In what
follows, we study the Fourier transform of these correlators given by
\begin{eqnarray}
F_{{\vec k}_0}(\omega_0) = \int_0^T dt e^{i \omega_0 t} F_{{\vec k}_0} (t),
\label{correxp2} \end{eqnarray}
and show that the phase band crossings leave their imprint on the Fourier
transform.
For the sake of concreteness, we apply these ideas to the Kitaev
model with the Hamiltonian given in Eq.~\eqref{hamkit}. As remarked
earlier, the structure of this is similar to \eqref{fermham} except
that $\tau^y$ is replaced by $\tau^z$, and the Kitaev model has an
off-diagonal couplings like $a_{\vec k}^\dagger b_{\vec k}$ (which conserve
fermion number) instead of superconducting pairing terms like
$c_{\vec k}^\dagger c_{-{\vec k}}^\dagger$. We can take care of this difference by
transforming to the basis of $\tau^y$ given by
\beq c_{\vec k} ~=~ \frac{1}{\sqrt{2}} ~(a_{\vec k} + i b_{\vec k}), ~~~~d_{\vec k} ~=~
\frac{1}{\sqrt{2}} ~(a_{\vec k} - i b_{\vec k}). \eeq
At time $t=0$, we take the wave function to be $(u_{{\vec k}}^0,v_{{\vec k}}^0)^T$
which denotes the state $(u_{{\vec k}}^0 c_{\vec k}^\dagger + v_{{\vec k}}^0
d_{\vec k}^\dagger) | vac \rangle$ in second quantized notation. We can then
calculate the time-dependent fermionic correlators in the Kitaev model
in the same way as in the previous paragraph.
For the drive protocol given in Eq.~\eqref{j3t}, we have seen that
phase band crossings can occur when the term proportional to $\tau^x$ in
Eq.~\eqref{hamkit} vanishes, and these happens on one of two lines in
momentum space. For definiteness, let us consider a point ${\vec k}_0$ on
the second line, with $k_{0y} = 0$ and $0 \le k_{0x} \le 2 \pi$.
Using Eq.~\eqref{corexp0} and \eqref{corexp1}, we then find that
$\langle c_{{\vec k}_0}^\dagger c_{{\vec k}_0} \rangle_t$ and $\langle d_{{\vec k}_0}^\dagger d_{{\vec k}_0} \rangle_t$
are independent of time. In contrast the off-diagonal correlator is given by
$F_{{\vec k}}(t) = \langle c_{{\vec k}}^{\dagger} d_{{\vec k}}\rangle_t$ and reads
\bea F^{(1)}_{{\vec k}_0}(t) &=& u_0^* v_0 \exp [- i\omega_{{\vec k}_0} t + i
(4 h_1 /\omega_D) \sin (\omega_D t)], \nonumber \\
\omega_{{\vec k}_0} &=& - 4 h_0 - 8 J_1 \cos (k_{0x}/2) \label{fkt1} \eea
The Fourier transform of \eqref{fkt1} for one time period $T$ is
given by \beq F^{(1)}_{{\vec k}_0}(\omega_0) ~=~ - i u_0^* v_0
\sum_{n=-\infty}^\infty J_n \left( \frac{4h_1}{\omega_D} \right)
~\frac{e^{i (\omega - \omega_{{\vec k}_0} + n \omega_D) T} ~-~ 1}{\omega -
\omega_{{\vec k}_0} + n \omega_D}, \label{fkom1} \eeq where we have used the
identity \cite{abram} \beq e^{i z \sin \theta} ~=~
\sum_{n=-\infty}^\infty ~J_n (z) ~e^{i n \theta}. \eeq If
$4h_1/\omega_D \ll 1$, the $n=0$ term will dominate in the sum in
Eq.~\eqref{fkom1}. We then find that as a function of $\omega$, the
magnitude of $F^{(1)}_{{\vec k}_0} (\omega)$ has a maximum at $\omega_0 \simeq
\omega_{{\vec k}_0}$ where $|F^{(1)}_{{\vec k}_0} (\omega_0)| \simeq |u_0^* v_0 T J_0
(4h_1/\omega_D)|$, and minima at $\omega_{0m} \simeq \omega_{{\vec k}_0} + m \omega_D$
(with $m$ being a non-zero integer) where $F^{(1)}_{{\vec k}_0} (\omega_{0m}) = 0$.
Since the phase band crossings occur at times $t_0$ given by
$\omega_{{\vec k}_0} t_0 - (4 h_1 /\omega_D) \sin (\omega_D t_0) = 2 p \pi$ (where
$p \in Z$), we therefore obtain that $F^{(1)}_{{\vec k}_0}(\omega_0)$ will be
maximum if
\begin{eqnarray} \omega_0 t_0 &=& (4 h_1/\omega_D) \sin[\omega_D t_0] + 2 \pi p,
\label{maxcond} \end{eqnarray}
and display minima at
\begin{eqnarray} \omega_{0m} t_0 &=& (4 h_1/\omega_D) \sin[\omega_D t_0] +m
\omega_D t_0 + 2 \pi p \label{mincond} \end{eqnarray} for non-zero
integer $m$. Thus the maxima-minima pattern of $F^{(1)}_{\vec
k_0}(\omega_0)$ contains information about the phase band crossing
time. The precise relation between $t_0$, $\omega_D$ and $\omega_0$ which
leads to these maxima and minima depends on the drive protocol used.
This is demonstrated in Fig.\ \ref{fig6} where $|F_{\vec
k_0}^{(1)}(\omega)|$ is plotted as a function of the frequency
$\omega$ in the left panel; the right panel shows the position of
$t_0$ as obtained from Eqs.\ \eqref{line1} and \eqref{line2} for the
specific drive parameters used. We note that the maxima and the
minima of $|F_{\vec k_0}^{(1)}(\omega)|$ occurs at frequencies which
are given by Eq.\ \eqref{maxcond} and \eqref{mincond} with $t_0=0.4
T$.
\begin{figure}
\includegraphics[width=0.47\linewidth]{fig6a.pdf}
\includegraphics[width=0.47\linewidth]{fig6b.pdf}
\caption{Left Panel: Plot of $|F^{(1)}_{\bf k_0}(\omega_0)|$ vs
$\omega_0$ for the Kitaev model showing maxima and minima of the
off-diagonal correlation function with $(k_{x0}, k_{y0})=
(\pi/9,0)$. The protocol used is $J_3(t)= h_0 + h_1 \cos(\omega_D
t)$ with $h_0=h_1=0.5$, and $\omega_D=20$. Here we have chosen
$u_0=v_0=1/\sqrt{2}$, $J_1=J_2$, and all energies are scaled in
units of $J_1$. The right panel shows that phase bands cross at
$t_0=0.4T$; we find that the maxima and minima obtained coincides
with those predicted from Eqs.\ \eqref{maxcond} with $p=0$ (maxima)
and Eq.\ \eqref{mincond} with $p=0$ and $m=1,2$ (minima). The dotted
lines are guides to the eye.} \label{fig6} \end{figure}
To elucidate the protocol dependence stated above, we consider a
drive protocol consisting of periodic $\delta$-function
kicks~\cite{thakurathi2}, \beq J_3 (t) ~=~ h_0 ~+~ h_1
~\sum_{n=-\infty}^\infty ~\delta (t - nT), \eeq for which the phase
band crossings occur at $\omega_{{\vec k}_0} t_0 - 4 h_1 = 2 n \pi$, where
$\omega_{{\vec k}_0}$ is given in Eq.~\eqref{fkt1}. A calculation similar to
the one outlined above yields \beq F^{(2)}_{{\vec k}_0}(\omega) ~=~ - i
u_0^* v_0 ~e^{i 4 h_1} ~\frac{e^{i (\omega - \omega_{{\vec k}_0}) T} ~-~ 1}{\omega
- \omega_{{\vec k}_0}}. \label{fkom2} \eeq The magnitude of
$F^{(2)}_{{\vec k}_0}(\omega_0)$ has a maximum at $\omega_0 = \omega_{{\vec k}_0}$
where $|F^{(2)}_{vk_0}(\omega_0)| = |u_0^* v_0 T|$, and minima at
$\omega_{0m} = \omega_{{\vec k}_0} + m \omega_D$ (with $m \ne 0$) where
$F^{(2)}_{{\vec k}_0}(\omega_{0m}) = 0$. This leads to the relations
\begin{eqnarray} \omega_0 t_0 = 4 h_1 + 2 \pi n \label{maxkick}\end{eqnarray}
for maxima, and
\begin{eqnarray} \omega_{0m} t_0 = 4 h_1 + 2 \pi n + m \omega_D \label{minkick}\end{eqnarray}
for minima of $F^{(2)}_{{\vec k}_0}(\omega_0)$.
\begin{figure}
\includegraphics[width=0.47\linewidth]{fig7a.pdf}
\includegraphics[width=0.47\linewidth]{fig7b.pdf}
\caption{Left Panel: Plot of $|F^{(2)}_{\bf k_0}(\omega_0)|$ vs
$\omega_0$ for the Kitaev model showing maxima and minima of the
off-diagonal correlation function for a delta function protocol with
periodic kicks with $(k_{x0}, k_{y0})= (\pi/2,0)$, $h_0=1$ and
$h_1=0.5$. All other parameters are the same as in Fig.\ \ref{fig6}.
The right panel shows that $t_0=0.41 T$; we find that the maxima and
minima obtained coincides with those predicted from Eqs.\
\eqref{maxkick} with $n=0$ (maxima) and Eq.\ \eqref{minkick} with
$n=0$ and $m=1,2,3$ (minima). The dotted lines are guides to the
eye.} \label{fig7} \end{figure}
A plot of $|F_{\vec k_0}^{(2)}(\omega)|$ is plotted as a function of
the frequency $\omega$ in the left panel; the right panel shows the
position of $t_0$ for the specific drive parameters used. We note
that the maxima and the minima of $|F_{\vec k_0}^{(1)}(\omega)|$
occurs at frequencies which are given by Eq.\ \eqref{maxkick} and
\eqref{minkick} with $t_0=0.41 T$.
Before ending this section, we note that the fermionic correlators
in the Kitaev model can be related to the correlators of the spins
appearing in the Hamiltonian in Eq.~\eqref{kitham}.\cite{ks1} For
two neighboring sites given by $b$ located at ${\vec n}$ and $a$ located
at ${\vec n} + \vec r$ (where $\vec r$ can take three possible values given
by $(0,0)$, $-{\vec M}_1$ and ${\vec M}_2$), we can use
Eq.~\eqref{abkn} to relate the fermionic correlators in real and
momentum space, \beq \langle i b_{\vec n} a_{{\vec n} + \vec r} \rangle_t ~=~
\frac{i4}{N} ~\sum_{\vec k} ~\langle b_{\vec k}^\dagger a_{\vec k} e^{i {\vec k} \cdot \vec r} ~-~
a_{\vec k}^\dagger b_{\vec k} e^{-i {\vec k} \cdot \vec r} \rangle_t, \eeq where we have used
the relations $\langle b_{\vec k} a_{{\vec k}'} \rangle_t = 0$ for all ${\vec k}$, ${\vec k}'$,
and $\langle b_{\vec k}^\dagger a_{{\vec k}'} \rangle_t = 0$ if ${\vec k} \ne {\vec k}'$. Rewriting
$a_{\vec k}$ and $b_{\vec k}$ in terms of $c_{\vec k}$ and $d_{\vec k}$, we find that
the correlator is given by \bea \langle i b_{\vec n} a_{{\vec n} + \vec r} \rangle_t &=&
\frac{4}{N} ~\sum_{\vec k} ~ [ \cos ({\vec k} \cdot \vec r) (\langle d_{\vec k}^\dagger d_{\vec k}
\rangle_t - \langle c_{\vec k}^\dagger c_{\vec k} \rangle_t)
\nonumber \\
&& ~~~~ -~i \sin ({\vec k} \cdot \vec r) (\langle c_{\vec k}^\dagger d_{\vec k} \rangle_t - \langle
d_{\vec k}^\dagger c_{\vec k} \rangle_t) ]. \label{abn} \eea We thus see that the
terms proportional to $\sin ({\vec k} \cdot \vec r)$ are related to the
off-diagonal fermion correlators. Next, we note that for $\vec r =
(0,0)$, $-{\vec M}_1$ and ${\vec M}_2$, $i b_{\vec n} a_{{\vec n} + \vec r}$ are
given by $\sigma^z_{b,{\vec n}} \sigma^z_{a,{\vec n}}$, $\sigma^y_{b,{\vec n}} \sigma^y_{a,{\vec n}
- {\vec M}_1}$ and $\sigma^x_{b,{\vec n}} \sigma^x_{a,{\vec n} + {\vec M}_2}$
respectively~\cite{ks1}. Hence in the last two cases, where $\vec r =
-{\vec M}_1$ or ${\vec M}_2$, the nearest-neighbor spin correlators
are related to the off-diagonal fermion correlators through
Eq.~\eqref{abn}. (If $b_{\vec n}$ and $a_{{\vec n} + \vec r}$ are not on
nearest-neighbor sites, the relation between $i b_{\vec n} a_{{\vec n} +
\vec r}$ and spin correlators is more complicated. Namely, $i b_{\vec n}
a_{{\vec n} + \vec r}$ is given by a product of $\sigma^x$ or $\sigma^y$ at ${\vec n}$
and ${\vec n} + \vec r$ multiplied by a Jordan-Wigner string of $\sigma^z$'s
running between those two sites).
To summarize, at the momenta ${\vec k}_0$ where phase band crossings can
occur, we find that the Fourier transform of the off-diagonal
correlator, $F_{{\vec k}_0}(\omega_0)$ has maxima and minima at some
particular frequencies; these frequencies are related to
$\omega_{{\vec k}_0}$ (which is a function of ${\vec k}_0$) by integer multiples
of the drive frequency $\omega_D$. The maxima and minima of these correlators
provide us with a relation between $\omega_0$, the phase
band crossing time $t_0$, and the drive frequency $\omega_D$ whose
precise form depends on the drive protocol used. We note that the
standard signature of the phase band crossings shows up in the form of
localized subgap states at the ends of a finite sample
\cite{rudner1}; however, such a signature cannot identify the
crossing time $t_0$ which can be done by tracking maxima and minima
of $F_{{\vec k}_0} (\omega_0)$.
\section{Discussion}
\label{diss}
In this work, we have provided an analytic expression for the
phase bands for a class of periodically driven integrable models for
arbitrary drive protocols within the adiabatic-impulse approximation.
Using this expression, and other more generic arguments, we have
outlined the conditions for phase band crossings in these models for
arbitrary drive protocols. As we have argued in this work, although the
expressions for the phase bands are derived within the
adiabatic-impulse approximation, the crossing conditions derived are
exact for two reasons. First, such conditions can be derived from
intuitive arguments which do not depend on the approximations used
and second, the momenta at which these crossings occur are the ones
in which the adiabatic-impulse approximation becomes exact. We also
show that for a class of these crossings, the time of crossing
$t_0/T$ is independent of the drive frequency $\omega_D$, and we provide
an analytical explanation of this phenomenon. We also point out that
the crossing conditions for the critical modes, where the
instantaneous energy levels of the Hamiltonian undergo an unavoided
level crossing, are different compared to those for the non-critical
modes where no such level crossings occur. Finally, we point out
that the off-diagonal fermionic correlators carry a signature of
the phase band crossings, and we provide analytical relations between
the frequencies $\omega_0$ (at which $F_{{\vec k}_0}(\omega_0)$ either
shows a maxima or vanishes), the phase band crossing time $t_0$, and
the drive frequency $\omega_D$.
Our results regarding the phase band crossing conditions have some
implications which we briefly discuss. First, although our
results are derived for $N=2$ phase bands, they may provide an
insight into generic crossing conditions for systems with $N>2$. To
see this, let us consider a situation where there are $N$ phase
bands for any given quasi-momentum ${\vec k}$ in a system. Let us
consider a phase band crossing corresponding to a zone-edge
singularity between the top ($N^{\rm th}$) and the bottom ($1^{\rm
st}$) bands. The dynamics of these bands can always be described by
an effective $2 \times 2$ matrix Hamiltonian $H_{{\vec k}}^{\rm eff}$
which can be obtained, in principle, by integrating out all other
degrees of freedom of the system Hamiltonian. Then the effective
evolution operator describing the crossing of these two bands may
always be written, sufficiently near the band crossing point, as $
U_{{\vec k}}^{\rm eff}(t) = {\mathcal T}_t \exp[- (i/\hbar) \int_0^t dt'
H^{\rm eff}_{{\vec k}} (t')]$. The most general form of such an
effective Hamiltonian is given by
\begin{eqnarray} H_{{\vec k}}^{\rm eff} = \sum_{i=1,3} g_{i {\vec k}}(t) \tau_i,
\end{eqnarray}
where $g_{i {\vec k}}(t)$ are parameter functions which depend on ${\vec k}$
and $t$ and whose precise form depends on the details of the system
Hamiltonian and the drive protocol. In terms of these $g_{i {\vec k}}$'s,
the condition for such band crossings, as can be inferred from our
results for integrable models of a similar form of $H$, is that at
least two of the three $g_{i {\vec k}}$ (which we may choose to label as
$g_{1 \vec k}(t)$ and $g_{2 {\vec k}}(t)$ without loss of generality) are
zero for ${\vec k}= {\vec k}_0$ and for appropriate choice of Hamiltonian
parameters. The crossing time $t_0$ is then determined from the
third generator using
\begin{eqnarray} \int_0^{t_0} dt' g_{3 {\vec k}_0}(t') = 0. \end{eqnarray}
Our results for integrable models show that these conditions need
to be satisfied for any $N$ for the system to have a band crossing
corresponding to a zone-edge singularity.
The second implication of our results constitutes the relation of
the symmetry classes of the underlying Hamiltonian to the condition for
phase band crossings. Our results indicate that Hamiltonians
belonging to the same symmetry class \cite{ryu1} and driven by
identical protocols may have different behaviors of the phase bands.
This becomes evident by considering the class CI which contains
models of time-reversal and SU(2) symmetric superconductors that
include both $d$- and $s$-wave pairing symmetries \cite{ryu1}. The
Hamiltonians of such superconductors are given by Eqs.\
\eqref{fermham} and \eqref{fermhamden} where $g = -\mu_0$ and
$b_{{\vec k}} = \epsilon_{\vec k}$, and $\mu_0$ and $\epsilon_{{\vec k}}$ are
the chemical potential and energy dispersion of the fermions. For
$d$-wave superconductors $\Delta_{{\vec k}} = \Delta_0 (k_x^2
-k_y^2)/k_F^2$, while for $s-$wave $\Delta_{{\vec k}}= \Delta_0$. Now
consider periodically driving such a system by changing the chemical
potential: $\mu_0 \equiv \mu_0(t)$. The phase bands corresponding to
$U_{{\vec k}}(t)$ may cross at momenta given by four isolated points on
the Fermi surface ${\vec k}_0 = (k_{0x}, k_{0y})= k_F(\pm 1, \pm
1)/\sqrt{2}$ for $d$-wave superconductors, while they will never
cross for $s-$wave superconductors. Thus the dynamics of these two
models will be very different even if they belong to the same
symmetry class. Our results seem to indicate that for similar
dynamics a necessary condition is that the set of zeroes of two of
the parameter functions $f_{i\vec k}(t)$ or $g_{i\vec k}(t)$ defined
earlier must be identical; for example, if the parameter functions
never vanish so that the set of zeroes is a null set, there will be
no phase band crossing for any drive protocol.
In conclusion, we have presented analytic expressions for the phase bands
for a class of integrable models driven by arbitrary periodic
protocols within the adiabatic-impulse approximation. Using these
expression and other more general arguments, we have listed the
conditions for phase band crossings for such models. We have also
shown that such phase band crossings leave their mark on the Fourier
transform of the off-diagonal fermionic correlators; the positions
of the zeroes and maxima of such correlators provide information about
the band crossing time $t_0$. Finally we have discussed the relevance of
the derived band crossing conditions in the context of generic
models with $N>2$ phase bands, and we have discussed the role of
symmetry in such band crossings.
\vspace{.5cm}
\centerline{\bf Acknowledgments}
\vspace{.5cm}
D.S. thanks DST, India for Project No. SR/S2/JCB-44/2010 for financial support.
\vspace{.5cm}
|
\section{Introduction}
Stochastic searching~\cite{BLMV11} underlies many biological
processes~\cite{BWV81,V07,M08}, animal
foraging~\cite{C76,B91,OBE90,VBH99,BLMV06}, as well as operations to find
missing persons or lost items~\cite{RS71,FS01,S09}. In these settings basic
goals are to maximize the probability that the target is actually found and
to minimize the time and/or the cost required to find the target. In
response to these challenges, a wide variety of search algorithms have been
extensively investigated and rich dynamical behaviors have been
uncovered~\cite{MOS11,MMMO12}.
Typically, one or perhaps multiple searchers move in some fashion through a
search domain to locate either a single target or a series of targets. The
most naive setting is that of a single searcher has no information about the
target and moves by random-walk, or equivalently, diffusive motion. Such a
search is generally hopelessly inefficient because the target may not be
found, for spatial dimension $d\geq 3$, or the average search time is
infinite. Thus much effort has been directed to uncover more effective
search strategies
Many such possibilities have been investigated. One natural mechanism is to
allow the searcher to move according to a L\'evy flight (see, e.g.,
\cite{VBH99,SK86,VLRS11}), so that the search can quickly cover large
distances between targets. A somewhat related example is that of
intermittent search, in which the search process is partitioned into periods
of intensive search, during which the searcher moves slowly, and superficial
search, during which the searcher moves quickly~\cite{BCMSV05}. In the
context of searching for nourishment, the essential tradeoff is how long to
continue to exploit resources in a local area and when to move to a new area
as local resources become depleted~\cite{C76,BR14,CBR16}. These notions also
underlie the search for a target along a DNA by a diffusing
protein~\cite{BWV81,V07,WV81,WBV81,HM04,HGS06}, where the tradeoff is for the
search to diffuse along the DNA or unbind and reattach at some distant point
along the DNA.
Very recently, the mechanism of search that is augmented by ``resetting'' was
introduced~\cite{EM11a,EM11b,EM14}. In this model, a target is placed at the
origin (without loss of generality) and a searcher starts at some arbitrary
point. In addition, the searcher returns to a fixed ``home base'' at a given
rate during the search. If the distance between the home base and the target
is known with certainty, then a search based on a stochastically moving
searcher is not a pertinent approach. However, the natural situation is that
the target location is only partially known; for example, the target is
somewhere within a finite body of water. In this case, a relevant parameter
is the maximum possible distance between the target and the home base. As
shown in~\cite{EM14}, the basic properties of search with resetting when the
distance between the target and home base is known precisely are
qualitatively the same as the situation where only the probability
distribution of this distance is known. Thus for the purposes of
tractability we restrict ourselves to the idealized (and admittedly
unrealistic) situation where the distance between the target and home base is
known.
In general, resetting is known to have a dramatic effect on the search. A
diffusing particle requires an infinite average time to reach a target in
spatial dimensions $d=1$ and $d=2$, and the searcher may not even reach a
finite-size target for $d>2$. However, resetting ensures that: (i) the
searcher can always find the target in any dimension and (ii) the average
search time is finite. Overall, therefore, resetting gives rise to a more
efficient search. One of the basic results of recent investigations of
search with resetting~\cite{EM11a,EM11b,EM14} was to determine the conditions
that optimize the search time. This resetting mechanism has also been quite
fruitful conceptually and a variety of interesting consequences of resetting
have been
elucidated~\cite{MV13,AG13,BS14,CS15,CM15,MSS15,P15,KG15,TRU15,EM16}.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.6\textwidth]{1d-illust}}
\caption{Contrast of the trajectories in: (a) Poisson reset and (b) deterministic
reset with reset time $T$. The target is at the origin and the reset point
is $x_0$.}
\label{1d-illust}
\end{figure}
In this work, we investigate two as yet unexplored features of search with
resetting: (i) $N$ searchers, each of which is reset at the same Poisson
rate, to a ``home base'', and (ii) deterministic reset, in which the
searchers return to the home base after a fixed operation time, rather than
the searchers being reset according to a fixed-rate Poisson process
(Fig.~\ref{1d-illust}). The related situation of many searchers that are
uniformly distributed in space, each of which is reset to its own starting
position at a fixed rate, was investigated in~\cite{EM14}. However, in the
context of search for a missing person, it is natural that all the searchers
return to a single home base (or perhaps a small number of such bases).
Moreover, in such a search, activities are typically suspended at the end of
daylight or when searchers reach their physical limits. Thus it is also
realistic to investigate the situation in which all the searchers are reset
to a given location at a fixed reset time.
In the next section, we start by briefly reviewing known results about
stochastic search by a single searcher in one dimension ($d=1$), with the
additional feature that the searcher is reset to its home base at a fixed
rate. We will also present our renewal-based approach to solve this problem
that will be employed throughout this work. Next, we treat the case of $N$
searchers in one dimension, each of which is independently reset to the same
location at a fixed rate $r$. We show that for $N<N^*$ there is an optimal
non-zero reset rate $r^*_N$ that minimizes the search time as well as the
search cost, while for $N>N^*$ resetting always hinders the search. We also
determine this critical number $N^*$ analytically. In Sec.~\ref{sec:d>1}, we
turn to search with stochastic resetting in spatial dimensions $d=2$ and
$d=3$. For the case of $d=3$, we exploit a well-known construction to reduce
the diffusion equation in three dimensions to an effective diffusion equation
in one dimension. This allows us to obtain results about three-dimensional
search in terms of the corresponding one-dimensional system.
To probe the properties of the $N$-searcher system in a convincing way, we
outline, in Sec.~\ref{sec:EDS}, an efficient event-driven simulation in one
dimension that obviates the need to microscopically follow the trajectories
of each searcher between reset events. By exploiting the aforementioned
dimensional reduction of the three-dimensional diffusion equation, we can
also directly adapt our event-driven approach to three dimensions. For
$d=2$, no such dimensional reduction exists and our simulations are based on
a more direct approach. From these numerical approaches, we determine the
condition for optimal search for both a single searcher and for many
searchers in spatial dimensions $d=1$ and 3, and then $d=2$.
Finally, in Sec.~\ref{sec:DR}, we investigate the case of deterministic
resetting, for both a single and for many searchers, again for the cases of
$d=1$, $d=3$, and then $d=2$. The salient feature of deterministic resetting
is that it leads to a quicker search than stochastic resetting at their
respective optimal resetting rates (or times). Moreover, deterministic
resetting leads to a search cost that, for large $N$, is nearly independent
of the reset rate $r$ over a wide range of $r$. Concluding remarks are given
in Sec.~\ref{sec:disc}.
\section{Poisson Resetting in One Dimension}
\label{sec:SR1d}
As mentioned above, the situation where a searcher is reset at a fixed rate
$r$ has already been extensively investigated~\cite{EM11a,EM11b,EM14}. For
completeness, we quote the main results for this type of search and also
derive them by an independent method. We then investigate the case of $N$
independent searchers, each of which is reset to a common point at the same
rate $r$.
\subsection{One searcher}
Consider a target that is fixed at the origin and a diffusing searcher that
is reset to a point $x_0$ at rate $r$. For simplicity, we assume that the
searcher begins at this reset point. For this system, the first two moments
of the search time are (see Refs.~\cite{EM11a,EM11b} and also \ref{PD})
\begin{align}
\begin{split}
\label{t}
\langle t_1\rangle &=\frac{1}{r}\left(e^{x_0 \sqrt{{r}/{D}}}-1\right)\,,\\
\langle t_1^2\rangle &= \frac{1}{r^2}\left[e^{x_0 \sqrt{{r}/{D}}} \big(2\, e^{x_0
\sqrt{{r}/{D}}}-x_0 \sqrt{{r}/{D}}-2\big)\right]\,,
\end{split}
\end{align}
and higher moments can be extracted straightforwardly. The subscript 1
signifies one searcher. The basic feature of \eqref{t} is that
$\langle t_1\rangle$ is minimized at an optimal reset rate $r^*$ that is of
the order of $D/x_0^2$. The inverse of the optimal rate gives the typical
time between resets as roughly $T_D\equiv x_0^2/D$, the time for a diffusing
particle to reach a distance $x_0$. If the searcher does not find the target
within this time, then it is likely wandering in the wrong direction and will
reach the target at a time much greater than $T_D$. In this case, it is
better to reset this errant searcher back to its home base than allowing it
to continue on its current trajectory.
We now give an independent derivation for the average search time
$\langle t_1\rangle$ that relies on the renewal nature of the search process;
a similar approach was very recently developed in Ref.~\cite{R16}. Namely,
whenever a reset occurs, the process restarts at the initial condition, but
with the proviso that the time is incremented appropriately to account for
the return to the reset point. As a preliminary, we need the following:
\begin{align}
R(t)&\equiv\text{prob.\ reset time is greater than}\ t\,,\nonumber\\
S(x_0,t)&\equiv\text{prob.\ hitting time is greater than}\ t\,.\nonumber
\end{align}
For diffusive motion, $S(x_0,t)$ is also the ``survival probability'' that a
searcher initially at $x_0$ has not reached the origin by time $t$
is~\cite{R01}
\begin{subequations}
\begin{equation}
\label{S}
S(x_0,t)= \text{erf}\Big(\frac{x_0}{\sqrt{4 D t}}\Big)\,.
\end{equation}
Correspondingly, the first-passage probability that a diffusing particle
initially at $x_0$ first reaches the origin at time $t$ is
\begin{equation}
\label{FPP}
F(x_0,t)=-\frac{d S(x_0,t)}{dt}= \frac{x_0}{\sqrt{4\pi D t^3}}\,\, e^{-x_0^2/4Dt}\,\,.
\end{equation}
\end{subequations}
Since the search process is specified by whether the target is reached before
a reset occurs and vice versa, we also define two fundamental probabilities:
\begin{align}
\begin{split}
\label{q}
P&\equiv\text{prob.\ target hit before reset}
=\int_0^\infty\!\!dt
\underbrace{~\left(-\frac{dS(x_0,t)}{dt}\right)~}_{\text{prob.\ hit
in $(t,t+dt)$}}\times\
\underbrace{~~~R(t)~~~}_{\text{prob.\
reset time $>t$}}\!\!\!\!\!,\\
Q&\equiv\text{prob.\ reset before target hit}
=\int_0^\infty\!\!dt \underbrace{~~S(x_0,t)~~}_{\text{prob.\ hit time $>t$}}\,\times\,
\underbrace{~~\left(-\frac{dR(t)}{dt}\right)~~}_{\text{prob.\
reset in $(t,t\!+\!dt)$}}\!\!\!\!.
\end{split}
\end{align}
Thus the ``direct'' time for the target to be reached before a reset occurs,
which we define as $t_d$, is
\begin{subequations}
\begin{align}
\label{tdtr}
t_d =\int_0^\infty\!\!dt\, t\,
\left(-\frac{dS(x_0,t)}{dt}\right)\,\times\, R(t)\Big/P\,.
\end{align}
Similarly, the ``reset'' time $t_r$ for a reset to occur before the target is
reached is
\begin{align}
t_r =\int_0^\infty\!\!dt\, t\,
\left(-\frac{dR(t)}{dt}\right)\,\times\, S(x_0,t)\Big/Q\,.
\end{align}
\end{subequations}
Using the renewal nature of the search, $\langle t_1\rangle$ satisfies the
recursion
\begin{subequations}
\begin{equation}
\langle t_1\rangle= Pt_d+Q\big(t_r+ \langle t_1\rangle\big)\,.
\end{equation}
The first term accounts for hitting the target before a reset occurs, while
the second term accounts for the search restarting after a reset. In this
latter case, the search is delayed by $T_r$. For notational convenience, we
define $T_d=P\,t_d$ and $T_r=Q\, t_r$. Solving for $\langle t_1\rangle$
gives
\begin{equation}
\label{t-recur}
\langle t_1\rangle= \frac{T_d+T_r}{1-Q}\,\,.
\end{equation}
\end{subequations}
This result is general and can be applied to higher dimensions and to
different reset mechanisms, as will be discussed later.
Since only the sum $T_d+T_r$ appears in the expression for
$\langle t_1\rangle$, we add the two lines in \eqref{tdtr} and integrate by
parts to give
\begin{align}
\label{TdTr}
T_d+T_r= \int_0^\infty dt\, R(t)\, S(x_0,t)\,.
\end{align}
We now specialize to Poisson resetting with rate $r$, for which
$R(t)=e^{-rt}$. Using this, as well as \eqref{S} for $S(x_0,t)$, the
integrals in \eqref{q} and \eqref{TdTr} are
\begin{align}
T_d+T_r&= \frac{1}{r}\!\left(1\!-\!e^{-x_0\sqrt{r/D}}\right)\,, \nonumber\\
Q&= 1-e^{-x_0\sqrt{r/D}}\,.\nonumber
\end{align}
from which \eqref{t-recur} gives
\begin{align}
\label{t-recur-b}
\langle t_1\rangle =\frac{1}{r}\left(e^{x_0\,\sqrt{r/D}}-1\right)\,.
\end{align}
This reproduces Eq.~\eqref{t} as it must.
\subsection{Multiple searchers}
Since all searchers are independent and the target location is fixed, it is
theoretically possible to obtain the survival probability of the target in
the presence of $N$ searchers as the $N^{\rm th}$ power of the target
survival probability due to a single searcher. However, while the Laplace
transform of the target survival probability with one searcher is known
exactly, Eq.~\eqref{S1s}, it does not appear possible to Laplace invert this
expression exactly. Nevertheless, we can invert this Laplace transform in
the limit $r\to 0$ to provide information about the dependence of the search
cost as $r\to 0$; this feature will be discussed below.
Thus we resort to simulations to map out the behavior of the search time as a
function of the reset rate $r$ for multiple searchers. Because it is
inherently wasteful to simulate directly the microscopic motion of each
searcher between reset events, we developed an efficient event-driven
simulation, whose details are given in Sec.~\ref{sec:EDS}. Our focus is on
the rich features of the search time and the search cost as a function of $r$
and $N$. Under the assumption that each searcher has the same fixed cost per
unit time of operation, the search cost for $N$ searchers, $C_N$, is merely
$C_N=N\langle t_N\rangle$, where $\langle t_N\rangle$ is the average search
time for $N$ searchers. This cost has a weak dependence on $N$, so that it
is more convenient to focus on cost rather than time in the following.
\begin{figure}[ht]
\centerline{\subfigure[]{\includegraphics[width=0.5\textwidth]{Cost-1D-stoch}}
\subfigure[]{\includegraphics[width=0.48\textwidth]{minima-1D-stoch}}}
\caption{(a) Average search cost $C_N$ scaled by the diffusion time $T_D$
versus the scaled reset rate $rT_D$ in one dimension for various $N$. Each
curve represents an average over $10^9$ trajectories. The dashed line
indicates the minimum cost of 1.544 for $N=1$. (b) The minimum search cost
for each $N\leq 6$.}
\label{1Dsim}
\end{figure}
Figure~\ref{1Dsim} shows the search cost in one dimension as a function of
$r$ for $1 \leq N\leq 8$. Noteworthy features of the search cost include:
\begin{enumerate}
\item The lowest scaled search cost of $1.544$ is achieved by a single
searcher that is reset at the scaled optimal rate $r^*_1 \approx 2.540$.
At this optimum, there are typically $r^*_1\langle t_1\rangle\approx 3.657$
resets before the searcher reaches the target.
\item For $N=2,3,\ldots 7$, there is a unique, non-zero optimal reset rate
$r^*_N$ for each $N$ that minimizes the search cost and the search time.
This optimal rate is generally of the order of the inverse diffusion time
between the home base and the target, $T_D=x_0^2/D$. The optimal cost for
all $N\leq 5$ is within 2\% of the optimal cost for a single searcher.
\item For $N\!\geq\! 8$, the search time strictly increases with $r$; that
is, no resetting is optimal. This behavior arises because at least one
searcher is systematically moving toward the target, once the number of
searchers is sufficiently large, so that any resetting increases the search
time. The demonstration of the sign change in the initial slope of
$\langle t_N\rangle$ versus $r$ between $N=7$ and $8$ (see
Fig.~\ref{fig:dtdr}) is given in \ref{7-8}.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.5\textwidth]{slope-at-r=0}}
\caption{ Slope of the average search time $\langle t_N\rangle$ at $r=0$ as a
function of $N$ in one dimension when each searcher is independently reset
at rate $r$.}
\label{fig:dtdr}
\end{figure}
\item For $N=1$, the average search time diverges as $r\to 0$. This property
reflects the divergence of the first-passage time for a diffusing particle
to hit an arbitrary point in one dimension~\cite{F68,R01}. Since the
survival probability for the diffusing particle to not hit the target by
time $t$, $S(x_0,t)$ in Eq.~\eqref{S}, asymptotically decays as $t^{-1/2}$,
we estimate the hitting time for small $r$ as
$\int^{1/r} S(x_0,t)\, dt\sim r^{-1/2}$. This reproduces the behavior that
arises from a small-$r$ expansion of search time in Eq.~\eqref{t}.
\item For $N=2$, the search time again diverges as $r\to 0$. Because of the
independence of the searchers, the survival probability of the target
asymptotically decays as $\big[t^{-1/2}\big]^2=t^{-1}$. In the $r\to 0$
limit, the same argument as that given for $N=1$ leads to an average search
time that diverges as $-\ln r$ for $r\to 0$.
\item The probability that the target is not hit by any of $N$ searchers
asymptotically decays as $\big[t^{-1/2}\big]^N$. Thus when there are at
least 3 searchers, the search time is finite for $r\to 0$,
\end{enumerate}
\section{Poisson Resetting in Higher Dimensions}
\label{sec:d>1}
\subsection{Three dimensions}
There are two important physical differences between the one-dimensional and
three-dimensional system: (a) First, the target must have a non-zero size to
be detected; we take the target to be an absorbing sphere of radius $a$. (b)
Second, the existence of a non-zero target radius introduces an additional
parameter---the ratio of the target radius to the radius of the reset point
$a/r_0$.
In spite of these two complications, the above approach for one dimension can
be straightforwardly adapted to three dimensions because of the well-known
correspondence between the diffusion equation in three dimensions and in one
dimension~\cite{S43,R01}. Namely, the three-dimensional radial Laplacian
operator is related to the one-dimensional Laplacian by
$r\,\nabla^2_{3d}\, P= \nabla^2_{1d}\,(rP)$. Using this mapping, we can
write basic quantities for search with resetting in three dimensions in terms
of corresponding one-dimensional expressions.
\begin{figure}[ht]
\centerline{\subfigure[]{\includegraphics[width=0.50\textwidth]{Cost-3D-stoch-a0p1}}
\subfigure[]{\includegraphics[width=0.48\textwidth]{minima-3D-stoch-a0p1}}}
\caption{(a) Average scaled search cost $C_N/T_D$ versus scaled reset rate
$rT_D$ in three dimensions for various $N$ and for $a/r_0=10^{-1}$. The
curves for $N = 1$ and 5 are averaged over $10^8$ trajectories, while those
for $N = 10, 20, 30$ and $50$ are averaged over $10^7$ trajectories. (b)
Minimum search cost as a function of $N$.}
\label{3d-stoch}
\end{figure}
Because of computational limitations, most of our numerical results for
stochastic resetting in $d=3$ are for the case of $a/r_0=10^{-1}$
(Fig.~\ref{3d-stoch}). Simulations for different $a/r_0$ give a
qualitatively similar dependence of the search time and search cost on the
reset rate. As in the corresponding one-dimensional system, several features
of these results are worth highlighting:
\begin{enumerate}
\item For $N\lesssim 10$, the minimum cost changes so slowly with $N$ that is
is not possible to determine the value of $N$ at which the minimum cost is
achieved. For example, for $N=1$, the minimum cost is $14.760\pm 0.0015$
while for $N=10$, the minimum cost is $14.762\pm 0.0015$. By
$N\approx 15$, however, the minimum cost has a systematic increasing trend
that is larger than the error bars in the data.
\item In distinction to one dimension, the typical number of reset
events before the target is found by a single searcher is of the order of
$r_0/a$, which can be large. To understand this behavior, we start with
the hitting probability $H_{\rm 3d}(r_0,t)$ that a single searcher that is
a distance $r_0$ from the center of the target finds it within time
$t$~\cite{R01}:
\begin{equation*}
H_{\rm 3d}(r_0,t)= 1-S_{\rm 3d}(r_0,t)= \frac{a}{r_0}\erfc{\Big(\frac{r_0-a}{\sqrt{4Dt}}\Big)}\,.
\end{equation*}
Here $S_{\rm 3d}(r_0,t)$ is the probability that the searcher does not find
the target within time $t$ (see Sec.~\ref{subsec:3d}). For $r_0\gg a$ and
reset rate near the optimal value of $D/r_0^2$, the above hitting probability
reduces to $H_{\rm 3d}(r_0,t)\approx \frac{a}{r_0}\text{erfc}(\frac{1}{2})$.
Thus for $a/r_0\ll 1$, the number of reset events until target is found is of
the order of the inverse of this hitting probability, namely, of the order of
${r_0}/{a}$.
\item Because of the transience of diffusion in three dimensions, the search
time and search cost diverge as $r\to 0$ for any number of searchers. Thus
infinitesimal resetting always leads to a more efficient search than no
resetting.
\end{enumerate}
\subsection{Two dimensions}
\begin{figure}[ht]
\centerline{\subfigure[]{\includegraphics[width=0.50\textwidth]{Cost-2D-stoch-a0p1}}
\subfigure[]{\includegraphics[width=0.48\textwidth]{minima-2D-stoch-a0p1}}}
\caption{(a) Average scaled search cost $C_N/T_D$ versus scaled reset rate
$rT_D$ in two dimensions for various $N$ and for $a/r_0=10^{-1}$. The data
are averaged over $10^8$ trajectories. (b) Minimum search cost as a
function of $N$. }
\label{2d-stoch}
\end{figure}
In two dimensions, it is practically not feasible to implement an
event-driven simulation because the first-passage and hitting probabilities
that form the kernel of the event-driven algorithm are not known in closed
form as a function of time. (They are known, however, in the Laplace
domain~\cite{R01}, and were used in~\cite{EM11b,EM14} to provide an
analytical expression for the search time for the case of a single
searcher.)~ Thus we implement an alternative simulational approach, as will
be discussed in the next section.
The primary features of search with stochastic resetting in two dimensions
are (see Fig.~\ref{2d-stoch}):
\begin{enumerate}
\item Resetting again always leads to a more efficient search compared to the
case of no resetting. As in three dimensions, this feature is a
consequence of the divergent average search time at $r=0$ in two dimensions
for any number of searchers.
\item The dependence of the search cost as a function of reset rate is
qualitatively similar to that in three dimensions. The minimum cost is
nearly constant for $N\lesssim 10$, with the difference in the cost values
for adjacent $N$ values less than the simulation error bars. For example,
the minimum costs for $N=1$ and $10$ are $3.59\pm 0.01$ and $3.60\pm 0.01$
respectively, while for $N=15$, the minimum cost in $3.67\pm 0.02$.
\end{enumerate}
\section{Event-Driven Simulations}
\label{sec:EDS}
The one- and three-dimensional numerical results are based on an event-driven
algorithm that allows us to efficiently simulate $N$ independently resetting
searchers. In our approach, each searcher is propagated by a single
(typically macroscopic) time step between reset events until one of the
searchers finds the target. Thus each update is ``useful'' in that either
the target is found or a reset event occurs. No time is expended in diffusively
propagating searchers between resets.
\subsection{One dimension}
In one spatial dimension, the elemental steps of our algorithm are the
following:
\begin{enumerate}
\item Start with all the searchers at a distance $x_0$ from the target.
\item For each searcher, with the $i^{\rm th}$ one located at $x_i$, draw a
random time value from the first-passage distribution $F(x_i,t)$ given in
Eq.~\eqref{FPP}. Also choose a reset time $t_r$ from a Poisson
distribution according to the reset rate $Nr$. This gives the time for one
of the $N$ searchers to be reset.
\item If the minimum among these $N+1$ times is the reset time, choose a
random searcher and reset it to $x_0$. Each of the remaining searchers is
moved from its current position $x_i$ to a new position that is drawn from
the conditional probability
\begin{equation*}
P(x_i,t_r) = \frac{1}{\sqrt{4 \pi D t_r}}
\left[e^{-(x_i-x_0)^2/4 D t_r} - e^{-(x_i+x_0)^2/4 D t_r}\right]
\Big/S(x_0,t_r)\,,
\end{equation*}
where $S(x_0,t_r)$ is again the probability that a diffusing particle that
starts at $x_0$ does not hit the target up to time $t_r$. The distribution
$P(x_i,t_r)$ corresponds to the diffusive propagation of each searcher over
the time increment $t_r$, subject to the constraint that no searcher can
reach the target. After all the searchers are moved, increment the elapsed
time by the reset time $t_r$ and return to step (i).
\item If the minimum among the $N+1$ random times is one of the first-passage
times, then the target is found. The total search time is the current
elapsed time plus this first-passage time.
\end{enumerate}
\subsection{Three dimensions}
\label{subsec:3d}
By exploiting the dimensional reduction of the diffusion equation in three
dimensions to one dimension, the algorithm outlined above can be directly
adapted to the three-dimensional system. The algorithmic steps of our
event-driven simulation are now:
\begin{enumerate}
\item Start with all the searchers at a distance $r_0$ from the target.
\item For each searcher, with the $i^{\rm th}$ searcher at a radial distance
$r_i$, draw a random time value from the three-dimensional first-passage
distribution to a sphere of radius $a$~\cite{R01}:
\begin{subequations}
\begin{equation}
\label{F3d}
F_{\rm 3d}(r_i,t)=\frac{a}{r_i}\frac{r_i-a}{\sqrt{4\pi D t^3}}\,\, e^{-(r_i-a)^2/4Dt}\,.
\end{equation}
Also choose a reset time $t_r$ from a Poisson distribution with reset rate
$Nr$.
\item If the minimum among these $N+1$ times is the reset time, choose a
random searcher and reset it to $r_0$. Each of the remaining searchers is
moved from its current position $r_i$ to a new position that is drawn from
the conditional probability
\begin{equation}
\label{P3d}
P_{\rm 3d}(r_i,t_r) = \frac{r_i}{r_0}\frac{1}{\sqrt{4 \pi D t_r}}
\left[e^{-(r_i-r_0)^2/4 D t_r} - e^{-(r_i+r_0-2a)^2/4 D t_r}\right]
\Big/S(r_0,t_r)\,,
\end{equation}
with survival probability now equal to~\cite{R01}
\begin{equation}
\label{S3d}
S_{\rm 3d}(r_0,t_r)= 1-\frac{a}{r_0}\erfc{\bigg(\frac{r_0-a}{\sqrt{4 D t_r}}\bigg)}\,.
\end{equation}
\end{subequations}
Here $P_{\rm 3d}(r_i,t_r)$ corresponds to diffusive propagation of each
searcher over a time $t_r$, subject to the constraint that each searcher
cannot reach a spherical target of radius $a$. After all the searchers are
moved, increment the elapsed time by $t_r$ and return to step (i).
\item If the minimum among the $N+1$ random times is one of the first-passage
times, then the target is found. The total search time is the current
elapsed time plus this first-passage time.
\end{enumerate}
\subsection{Two dimensions}
As mentioned in the previously, it is impractical to implement an
event-driven simulation in two dimensions because the exact expression for
the first-passage probability to a circular target of radius $a$ as a
function of time is not known in closed form. While this first-passage
probability can be expressed as an inverse Laplace transform, the slow
convergence properties of this integral render it not useful as the kernel
for an event-driven simulation. However, we do know the first-passage
probability in the form of a well-converged series to the circumference of a
circle centered around the current position of the target. Thus our
simulation is based on propagating the searcher to the circumference of a
circle whose radius adaptively varies depending on the distance to the target
(Fig.~\ref{2d-illust}). This circle should just touch the target, so that
the radius of this circle is large when the searcher is far from the target
and small when the searcher is close to the target.
For a single searcher that is a distance $b$ from the circumference of the
target, the steps in our algorithm are the following (Fig.~\ref{2d-illust}):
\begin{enumerate}
\item Draw two random times. One is from the distribution of first-passage times
$\mathcal{F}(b,t)$
\begin{equation}
\mathcal{F}(b,t) = \sum_{n=1}^\infty \frac{2}{\mu_n
{J}_1\left(\mu_n\right)} e^{-\mu_n^2 D t/b^2}
\end{equation}
to the circumference of a circle of radius $b$ (\ref{app:F}). Here $J_1$ is
the ordinary Bessel function of index 1 and $\mu_n$ is the $n^{\rm th}$ zero
of this Bessel function. Because the jump distance is large if the searcher
is far from the target, little time is spent in simulating the motion of the
searcher when it is wandering aimlessly far from the target. The second
random time is drawn from the reset time distribution.
\item If the minimum of these two times is the reset time, reset the searcher
to $r_0$.
\item Otherwise, move the searcher to a random point on the circumference of
the circle of radius $b$. If the searcher is within a radius
$a(1+\epsilon)$ of the center of the target, then we define the target as
being found. If the target is not found after the searcher has been moved,
return to step (i).
\end{enumerate}
\begin{figure}[ht]
\centerline{\includegraphics[width=0.4\textwidth]{2d-illust}}
\caption{Illustration of a simulation event in two dimensions for a searcher
that is a distance $b$ from the circumference of a target of radius $a$.}
\label{2d-illust}
\end{figure}
We need to introduce an absorbing shell of thickness $\epsilon a$ around the
target to ensure that the searcher actually finds the target. Clearly, the
apparent search time decreases as $\epsilon$ is increased. To determine the
appropriate choice of $\epsilon$, we simulate the search with successively
smaller values of $\epsilon$ until the results do not change within the
statistical errors of the simulation and then use the largest of this set of
$\epsilon$ values for simulational efficiency. For the case of
$a/r_0=10^{-1}$, this value is $\epsilon=10^{-3}$.
By averaging over many trajectories, we construct an accurate numerical
estimate for survival probability due to a single searcher, $S_1(t)$. Due to
the independence of the searchers, we construct the survival probability for
$N$ searchers by $S_N(t) = \big[S_1(t)\big]^N$. The average search time is
then given by $\left\langle t_N \right\rangle = \int_0^\infty S_N(t)dt$.
\section{Deterministic Resetting}
\label{sec:DR}
We now investigate the situation where all searchers are reset to their
starting point after a fixed time $T$. As we shall see, this deterministic
resetting typically leads to a more efficient search compared to stochastic
reset. Moreover, because all searchers are reset simultaneously, we are able
to obtain numerically exact results for the search time for deterministic
resetting in both one and three dimensions.
\subsection{One dimension}
\subsubsection{Single searcher.}\,\, We follow the renewal approach of
Sec.~\ref{sec:SR1d} to calculate the average search time for a single
searcher that is reset to $x_0$ after a fixed time $T$. In this case $R(t)$,
the probability that reset time is greater than $t$, is just the Heaviside
step function $H(T-t)$, where $H(z)=1$ for $z>0$ and $H(z)=0$ for $z<0$.
From Eqs.~\eqref{q} and \eqref{TdTr}, we have
\begin{equation}
\begin{split}
\label{QT}
Q=-\int_0^\infty dt\, S(x_0,t)\, \frac{dR(t)}{dt} = S(x_0,T)\,,\\
T_d +T_r = \int_0^\infty dt\, R(t)\, S(x_0,t)= \int_0^T S(x_0,t)\, dt \,.
\end{split}
\end{equation}
Substituting these expressions into Eq.~\eqref{t-recur}, the average search
time for deterministic resetting of a single searcher is
(Fig.~\ref{1D-determ})
\begin{align}
\label{t1-det}
\langle t_1\rangle &= \int_0^T S(x_0,t)\, dt/\big(1-S(x_0,T)\big)\nonumber \\
&=\left[\!\sqrt{\frac{x_0^2\,T }{\pi D}}\, e^{-{x_0^2}/{4DT}}-\frac{x_0^2}{2D}\,
\text{erfc}\Big(\frac{x_0}{
\sqrt{4DT}}\Big)+ T \,\mathrm{erf}\Big(\frac{x_0}{\sqrt{4 DT}}\Big)\!\right]\!\bigg/\!
\mathrm{erfc}\Big(\frac{x_0}{\sqrt{4 DT}}\Big)\,.
\end{align}
\subsubsection{Multiple searchers.}
For multiple searchers, we merely need to use $[S(x_0,t)\big]^N$ rather than
$S(x_0,t)$ in Eq.~\eqref{QT} to account for $N$ independent searchers that
all must have their hitting time exceed a given threshold. Thus we have
\begin{equation}
\begin{split}
\label{QTN}
Q=-\int_0^\infty dt\, \big[S(x_0,t)]^N\, \frac{dR(t)}{dt} = \big[S(x_0,T)\big]^N\,,\\
T_d +T_r = \int_0^\infty dt\, R(t)\, \big[S(x_0,t)\big]N= \int_0^T \big[S(x_0,t)\big]^N\, dt \,.
\end{split}
\end{equation}
Substituting these in \eqref{t-recur}, the average search time for $N$
searchers with deterministic reset in one dimension has the simple form
\begin{equation}
\label{tN-det}
\left\langle t_N \right\rangle = \frac{\int_0^T \big[S(x_0,t)\big]^Ndt}
{1 - \big[S(x_0,T)\big]^N }\,.
\end{equation}
While we can compute the integral in \eqref{tN-det} analytically for the
cases $N=1,2$ and $N\to\infty$ (\ref{t2inf}), Mathematica can perform the
integration numerically to arbitrary precision to give $\langle t_N\rangle$
for any $N$.
\begin{figure}[ht]
\centerline{\subfigure[]{\includegraphics[width=0.48\textwidth]{Cost-1D-determ}}
\subfigure[]{\includegraphics[width=0.48\textwidth]{Cost-1D-determ-logr}}}
\caption{Average search cost scaled by the diffusion time $T_D$ versus
$T_D/T$ in one dimension for various $N$ and for deterministic resetting.
The abscissa scale is linear in (a) and logarithmic in (b).}
\label{1D-determ}
\end{figure}
Figure~\ref{1D-determ} shows the search cost as a function of $1/T$ for
representative $N$ values. We plot the search cost versus $1/T$ because
$1/T$ plays the same role as the rate $r$ in stochastic resetting. Several
new features of the search cost for deterministic reset in one dimension
are worth emphasizing:
\begin{enumerate}
\item The lowest search cost is achieved by a single searcher (as in
stochastic reset) in which the optimal scaled reset time is
$T_1^*/T_D\approx 0.458$, leading to an optimal scaled search time
$\langle t_1\rangle \approx 1.336\, T_D$, compared to the optimal cost
$1.544\, T_D$ from stochastic resetting. This optimal time corresponds to
approximately 3 reset events before the target is found.
\item For large $N$, the search cost becomes nearly independent of $1/T$ over
a wide range (Fig.~\ref{1D-determ}(b)). This behavior is characterized by
progressively more derivatives of the cost with respect to $1/T$ becoming
zero as $1/T\to 0$. In particular, the first derivative is zero for $N>4$,
the second is zero for $N>6$ and the third is zero for $N>8$. In general,
we find that the $k^{\rm th}$ derivative becomes zero for $N= 2k+3$. We
can understand this pattern of behavior by differentiating
Eq.~\eqref{tN-det} with respect to $1/T$, using Mathematica, to find the
leading behavior
\begin{align}
\frac{\partial \left\langle t_N \right\rangle}{\partial (1/T)}
&\simeq A_1 T^{3/2}\Big( \erf{\sqrt{{x_0^2 }/{4 D T}}} \Big)^{N-1}
- A_2 T^2 \Big( \erf{\sqrt{{x_0^2 }/{4 D T}}} \Big)^{N}
\end{align}
where $A_1$ and $A_2$ are $\mathcal{O}(1)$ in $T$. As $1/T\rightarrow 0$,
the error function is proportional to its argument, so that the above
leading terms scale as a negative power of $T$ for $N>4$ and as a positive
power for $N<4$. Analogous behavior arises for higher derivatives.
\begin{figure}[ht]
\centerline{\includegraphics[width=0.55\textwidth]{Asymptotic-formula}}
\caption{Comparison between the exact formula \eqref{tN-det} and the
asymptotic formula \eqref{tN-asymp} for the mean hitting time in one
dimension with deterministic reset.}
\label{1D-determ-compare}
\end{figure}
\item For $T_D/T\to\infty$, the search time asymptotically increases as
$e^{T_D/T}$, whereas for stochastic resetting the corresponding
$r\to\infty$ behavior is the search time growing as $e^{\sqrt{r}}$. The
$e^{T_D/T}$ growth has a simple origin. As $T_D/T\to\infty$, the probability
that one searcher does not reach the target within the reset time $T$ is
$S=\text{erf}(x_0/\sqrt{4DT})$. The probability that none of the searchers
reaches the target within time $T$ is $S^N$, so that the probability that
at least one of the searchers reaches with time $T$ is $1-S^N$. In the
limit $T_D/T\to\infty$, the asymptotic behavior of this probability is
\begin{align*}
1-S^N\simeq \frac{\sqrt{4DT}\, N\, e^{-x^2/4DT}}{\sqrt{\pi}\, x_0}\ll 1\,.
\end{align*}
The number of reset events until a searcher finds the target is the inverse
of this expression. Multiplying by $NT$ gives the asymptotic behavior of the
average search cost
\begin{equation}
\label{tN-asymp}
\langle C_N\rangle \simeq \frac{NT}{1-S^N}\simeq \,T\,\sqrt{\frac{\pi x_0^2}{4DT}} \,\,e^{x_0^2/4DT}\,.
\end{equation}
This asymptotics matches the numerically exact expression for the search time
when $T_D/T\gg 1$ (Fig.~\ref{1D-determ-compare})
\end{enumerate}
\subsection{Three dimensions}
\label{det-3d}
In three dimensions, we again obtain numerically exact results for
deterministic reset for all parameter values and thus can probe the role of
the reset time, as well as the parameter $a/r_0$ on the search cost and
search time. While the qualitative dependence of the search cost and time on
$1/T$ is the same for all $a/r_0$, there are quantitative anomalies that are
worth highlighting.
We again use the correspondence between one-dimensional and three-dimensional
diffusion to determine the search time in three dimensions. In the renewal
formula~\eqref{tN-det} for the search time, we now need the first-passage and
survival probabilities in three dimensions, $F_{\mathrm{3d}}(r_0,t)$ and
$ S_{\mathrm{3d}}(r_0,t)$, respectively (Eqs.~\eqref{F3d} and \eqref{S3d}).
Substituting these expressions into Eq.~\eqref{tN-det} and using Mathematica
to perform the integrals numerically, we again obtain the search time and
cost as a function of $1/T$ for any $N$ with arbitrary precision
(Fig.~\ref{3D-determ}).
\begin{figure}[ht]
\centerline{\includegraphics[width=1.\textwidth]{Cost-3D-determ}}
\caption{Average search cost scaled by $aT_D/r_0$ versus $T_D/T$ in three
dimensions for various $N$ and $a/r_0$. }
\label{3D-determ}
\end{figure}
As expected, the search cost is initially a decreasing function of $T_D/T$ for
any $N$. When $T_D/T=0$, a diffusing searcher is transient in three dimensions
and does not necessarily find the target; thus the search cost diverges in
this limiting case, even when $N$ is large. On the other hand, for
$T_D/T\to\infty$, the search again becomes inefficient because each searcher is
typically reset before it can progress towards the target.
Figure~\ref{3D-determ} also illustrates a data collapse within each panel
when $N \ll r_0/a$ and between panels in the limit of $r_0/a\gg1$. This
implies that the search cost is independent of $N$ and $a/r_0$, for small
enough target size and number of searchers, when this cost is scaled by
$a /(r_0 T_D)$. To derive this behavior, we substitute Eqs.~\eqref{F3d} and
\eqref{S3d} into \eqref{tN-det}, and approximate the argument of the error
function, $(r_0-a)/\sqrt{4 D T}$, as $\sqrt{T_D / 4 T}$. These steps lead to
\begin{equation*}
\left\langle t_N \right\rangle = \frac{\int_0^T \left[1-(a/r_0)\erfc{\Big(\sqrt{T_D/4 T}\Big)}\right]^Ndt}{1-\left[1-(a/r_0)\erfc{\Big(\sqrt{T_D/4 T}\Big)}\right]^N}~.
\end{equation*}
In the limit $r_0 / (a N) \gg 1$, the search cost
$\langle C_N\rangle = N\langle t_N\rangle$ is
\begin{equation}
\label{tN-det-approx-1}
\left\langle C_N \right\rangle = \frac{r_0 T}{a}\left[\erfc{\left(\sqrt{{T_D}/{4 T}}\right)}\right]^{-1}\,.
\end{equation}
This scaling form implies that plots of $\langle C_N\rangle$ versus $1/T$
collapse onto a single curve when $a \left\langle C_N \right\rangle/ r_0$ is
plotted against $T_D/T$. The asymptotic form~\eqref{tN-det-approx-1} may
also be derived by merely counting the number of resets until the target is
found. The probability of hitting the target within a single reset event by
a single searcher is given by,
\begin{equation*}
p \equiv 1 - S(r_0,T) = \frac{a}{r_0}\erfc{\left(\frac{r_0-a}{\sqrt{4 D
T}}\right)}\,.
\end{equation*}
The probability that any of the $N$ searchers finds the target is
$1\!-\!(1\!-\!p)^N\simeq N p \equiv P$. Since each of these hitting events is
independent, the average number of reset events before one of the searchers
reaches the target is
$\sum_{n\geq 1} n P (1-P)^n = 1/P\simeq r_0 / \big[N a \erfc(\sqrt{{T_D}/{4
T}})\big]$.
In the limit of large number of resets, the search time is just $T$ times
number of resets, as the time for the last segment of the trajectory that
actually reaches the target is negligible. This reasoning again leads to
Eq.~\eqref{tN-det-approx-1}.
\subsection{Two dimensions}
\begin{figure}[ht]
\centerline{\subfigure[]{\includegraphics[width=0.50\textwidth]{Cost-2D-determ-a0p1}}
\subfigure[]{\includegraphics[width=0.48\textwidth]{minima-2D-determ-a0p1}}}
\caption{(a) Average scaled search cost $C_N/T_D$ versus scaled inverse reset
time $T_D/T$ in two dimensions for various $N$ and for $a/r_0=10^{-1}$.
Data are averaged over $10^8$ trajectories. (b) Minimum search cost as a
function of $N$. }
\label{2d-determ}
\end{figure}
In two dimensions however, we are not able to implement the renewal process
calculation, as we do not have exact expressions for $S(r_0, t)$ or
$F(r_0, t)$. Hence, we use the same simulation as in the stochastic reset in
$d=2$. As in one and three dimensions, minimum cost is achieved for $N=1$
searcher and the cost monotonically increases with $N$
(Fig.~\ref{2d-determ}).
\section{Discussion}
\label{sec:disc}
In this work, we explored the consequences of superimposed resetting on the
performance of stochastic search processes. In this resetting, either one or
many searchers are returned to a fixed home base either at a fixed rate or at
a fixed reset time. We found a variety of intriguing and sometimes
unexpected features. When each searcher is independently reset to the home
base at a fixed rate $r$, the search cost is minimal for a single searcher
when the reset rate is of the order of the inverse diffusion time
$T_D=x_0^2/D$. We also found that resetting always hinders the search for
$N\geq 8$ searchers, while for $N\leq 7$, there is an optimal nonzero reset
rate for each $N$.
By exploiting the well-known relation between the diffusion equation in one
and in three dimensions, we obtained analogous results for search with
stochastic resetting in three dimensions. The primary new result for $d=3$
is that the cost of the search is nearly independent of the number of
searchers for $N<10$ for the case of $a/r_0 = 10^{-1}$. In two dimensions,
we developed an alternative procedure in which we simulate the target
survival probability in the presence of a single searcher and then take the
$N^{\rm th}$ power of this quantity to obtain the target survival probability
in the presence of $N$ searchers. In this case, similar to three dimensions,
the cost is nearly independent of the number of searchers up to $N=10$ and
for $N>10$, the cost increases monotonically.
We also explored a related model in which all searchers are simultaneously
reset to the home base after a fixed operation time $T$. This deterministic
resetting is theoretically and computationally simpler than stochastic
resetting, and we are able to obtain explicit formulae for the average search
time for any $N$ that can be numerically integrated to arbitrary precision.
In one dimension, deterministic resetting gives a search time that becomes
independent of the reset time $T$ when $N$ is sufficiently large, a behavior
that can be understood from the small-$r$ behavior of the search time. In
three dimensions, we showed by a simple extremal argument that the search
cost versus $1/T$ becomes independent of $N$.
There are a wide range of extensions of the basic model to practically and
theoretically interesting situations. It would be worthwhile to extend
search with resetting to the cases where either the target is diffusing
and/or the target is mortal~\cite{YAL13,AYL13,CAMYL15,MR15}. These
generalizations would naturally describe, e.g., the occupants of a lifeboat
that is adrift in the ocean. When the target is also moving, the basic
question is again whether the reset helps or hinders the search. For a mortal
target with any reasonable distribution of mortality, there will always be a
non-zero probability that the target will die before being found and the
relevant issue is to construct appropriate criteria that lead to a
well-defined optimization problem.
Financial support for this research was also provided in part by the grants
DMR-1623243 from the National Science Foundation (UB and SR), from the John
Templeton Foundation (CDB and SR), and Grant No.\ 2012145 from the United
States—Israel Binational Science Foundation (UB). We also thank B. Meerson
for helpful discussions and S. Reuveni for useful comments on the manuscript.
\emph{Added Note:} As final revisions were being made, we became aware of
related work~\cite{PKE16}, in which the authors mathematically showed that
deterministic reset leads to the smallest search time for the case of a
single searcher, as we also observed.
|
\section{Introduction}
It has long been observed, in experiments and simulations, that scale-invariant statistics of grain boundary networks appear to be universal.~\cite{1986mullins,2012lazar} That is, the statistics of their long-term scale-invariant properties are largely independent of the initial conditions. This is particularly interesting because the universal condition is both statistical and transient - a finite graph that approaches it will not stop flowing, but will rather continue evolving until it has very few edges. This observation motivates the introduction of two notions of graph limit for embedded graphs, one topological and one geometric. We use these to formally state universality conjectures for the network flow on graphs. We computationally test the local topological and local geometric convergence of simulations of the network flow in two dimensions.
In Section~\ref{sec_curvature_flow} we briefly introduce the network flow for embedded graphs in $\R^n,$ which is also known as curvature flow on embedded graphs. The planar case is the simplest model of the physical phenomenon of grain growth in a polycrystal~\cite{1956mullins}, and the evolution of graphs in higher dimensions is also of mathematical interest. Most of the concepts defined here can be extended to general $k$-dimensional regular cell complexes in $\R^n,$ and a similar universality phenomenon has been observed for the most physically important case of $2$-dimensional cell complexes in $\R^3.$ The eventual goal of the program outlined here would be to prove the existence of universal statistics in that case, and to study their properties. However, we will focus on embedded graphs in this paper for the purpose of clarity.
We define two notions of of convergence for embedded graphs - local topological, or Benjamini-Schramm, convergence in Section~\ref{sec_swatch} and local geometric convergence in Section~\ref{sec_geoConv}. The former implies convergence of all local topological statistics, and is connected to the method of swatches introduced previously in our previous paper~\cite{2016schweinhart,2012masonB}. The latter notion was developed in analogy with the former, and implies convergence of averages of local geometric properties, in a sense to be defined in this paper. It is weak convergence of probability measures on the space of embedded graphs with a smooth topology of disjoint topological types (Section~\ref{sec_topType}). This topology is not particularly nice or natural, but we give conditions in Theorem~\ref{thm_convergence} under which weak convergence of probability measures on the space of graphs with either the Hausdorff metric topology or the varifold topology implies local geometric convergence.
In Section~\ref{sec_computations}, we introduce computational methods to test for local topological and local geometric convergence, and apply them to simulations of the network flow on planar graphs. In contrast, previous papers appearing in the materials science and physics literatures have used a few ad-hoc measures to claim convergence of simulations to the conjectural universal state, before proceeding to study its properties. The notions presented here were developed in part to motivate a more systematic method to verify the convergence of such simulations.
We formalize universality conjectures for the network flow in Section~\ref{sec_conj}. The main hypothesis we propose is that the graphs are homogeneous. Homogeneous graphs are defined in~\ref{sec_hom}, and are roughly those whose local properties have well-defined spatial averages. Computational simulations indicate that these graphs either evolve to approach a universal state, or in very special cases become stationary with respect to the flow. The main conjectures are:
\begin{conj}[Universality Conjecture for Local Topological Convergence]
\label{steady_state_hypothesisb}
There exists a probability distribution $\sigma_n$ on the space of countable, connected graphs with a root vertex specified such that any network flow $G\paren{t}$ with homogeneous initial condition $G\paren{0}\in\mathcal{G}^n$ converges in the local topological sense to $\sigma_{\Omega}$ or to a stationary state as $t\rightarrow\infty.$
\end{conj}
\begin{conj}[Universality Conjecture for Local Geometric Convergence]
\label{steady_state_hypothesisa}
There exists a probability distribution $\Xi_n$ on $\mathcal{G}^n$ such that that any network flow $G\paren{t}$ with homogeneous initial condition $G\paren{0}\in\mathcal{G}^n$converges in the scale-free local geometric sense to $\Xi_{n}$ or a stationary state as $t\rightarrow\infty.$
\end{conj}
\begin{figure}
\center
\subfloat[]{%
\label{ss_2D}{%
\includegraphics[width=7cm]{%
Graphics_vNM.pdf}}}
\hspace{20pt}
\subfloat[]{%
\label{ss_3D}{%
\includegraphics[width=7cm]{%
SteadyState_200KEdges_InitialConds_graphics_v2.pdf}}}
\caption{Small regions of the conjectured universal state in (a) 2 and (b) 3 dimensions. They were produced by simulations discussed in Section~\ref{sec_computation}}
\end{figure}
A stationary state is one where the graph is at equilibrium with respect to the flow: its edges are straight and meet at trivalent vertices at angles of $2\pi/3.$
\subsection{Definition of Embedded Graphs}
Throughout, an embedded graph will be a locally finite collection of smooth curves in $\R^n,$ subject to certain niceness conditions:
\begin{Definition}
An \textbf{embedded graph} $G\subset K$ in a compact subset $K$ of $\R^n$ is a finite collection of smoothly embedded, connected curves with boundary with the following properties:
\begin{itemize}
\item The curves only meet at their boundaries, and are called the edges of the embedded graph.
\item The boundary points of the curves are called vertices, and at least three curves meet at each vertex not in $\partial K.$
\item The outward-pointing unit tangent vectors of the curves meeting at a vertex are distinct.
\end{itemize}
An embedded graph $G$ in an open subset $U$ of $\R^n$ is a collection of smoothly embedded, compact, connected curves with boundary so that $\widebar{G\cap K}$ is an embedded graph for any compact subset $K$ of $\R^n.$
\end{Definition} The second hypothesis implies that two embedded graphs in $\R^n$ are homeomorphic if and only if they are combinatorially equivalent. The intersection of an embedded graph in $\R^n$ with any open or compact subset of $\R^n$ is also an embedded graph, though new edges may be created in the process (as well as vertices at the boundary, if the set is compact) .In general, it is not always the case that $\widebar{U\cap G}=\bar{U}\cap G.$ The sets of embedded graphs in $\R^n$ and the open $n-$ball of radius $r$ are denoted $\mathcal{G}^n$ and $\mathcal{G}_r^n,$ respectively. We will define a topology for these sets in Section~\ref{sec_graphTop}.
\subsection{Summary of Concepts}
Here, we include a table summarizing the notions of graph limits defined in this paper, and a list of definitions of the terms appearing therein.
\begin{table}[ht]
\footnotesize
\centering
\begin{tabular}{|c|c|c|c|}
\hline\textbf{Convergence} & \textbf{Local Topological} & \textbf{Local Geometric} \\ \hline
\textbf{Space of} & Abstract Graphs & Embedded Graphs in $\R^n$ \\
& with Bounded Degree & with Smooth Geometric Cloth \\ \hline
\textbf{Local} & Probability Distribution & Probability Distribution\\
\textbf{Distributions} & on $\mathcal{S}_{r,k}$ & on $\mathcal{G}_{r}^n$ \\ \hline
\textbf{Universal} & Probability Distribution & Probability Distribution\\ \hline
\textbf{Distribution}& on $\mathfrak{G}^{\bullet}$ & on $\mathcal{G}^n$\\ \hline
\textbf{Symmetry} & Involution Invariance & Translation Invariance \\ \hline
\textbf{Addnl. Symmetry} & Topological Homogeneity & Homogeneity (Ergodicity of Translations) \\
\textbf{Convergence} & Convergence of discrete & Weak Convergence of\\
& Prob. Distributions & Unif. Separating Sequence \\ \hline
\textbf{Controls} & Local Topological Stats & Local Geometric Stats \\ \hline
\end{tabular}
\end{table}
\normalsize
\begin{enumerate}
\item Local topological convergence - this is the same notion as Benjamin-Schramm convergence. We use both terms interchangeably, to emphasize the analogy with local geometric convergence. Defined in Section~\ref{sec_convergence}.
\item $\mathcal{S}_{r,k}$ - the set of combinatorial isomorphism classes of rooted graphs. Defined in Section~\ref{sec_cloth}.
\item $\mathfrak{G}^{\bullet}$ - the space of abstract graphs with a root vertex specified. Defined in Section~\ref{sec_convergence}.
\item Involution invariance - roughly that a probability distribution on $\mathfrak{G}^\bullet$ gives equal weight to different root vertices of the same graph.
\item Local topological property - Definition~\ref{local_topo_prop}. Roughly, any property that can be expressed in terms of subgraph frequencies.
\item Topological homogeneity - Definition~\ref{defn_tophomo}. That local topological statistics of an embedded graph in $\R^n$ are well-defined over more general sets than balls.
\item $\mathcal{G}_{r}^n$ ($\mathcal{G}^n$) - the space of embedded graphs in the open $n$-ball of radius $r$ ($\R^n$), with the local Hausdorff metric topology.
\item Smooth geometric cloth - Definition~\ref{def_geo_cloth}. A probability distribution on $\mathcal{G}^n$ reflecting the averages of local geometric properties of a graph over balls of increasing radius.
\item Uniformly separating sequence - Definition~\ref{defn_unifSep}. A sequence of probability measures on $\mathcal{G}^n$ whose local distributions are tight with respect to a certain collection of compact subsets of $\mathcal{G}^n.$
\item Local geometric property - Definition~\ref{def_local_prop}. A bounded, continuous function on the space of embedded graphs in the ball of radius $r$ with the smooth topology of topological types.
\item Homogeneity - Definition~\ref{defn_homo}. A graph $G\in\mathcal{G}^n$ is homogeneous if the $\R^n$ translation action is ergodic with respect to its geometric cloth.
\end{enumerate}
\section{Motivation: Curvature Flow on Graphs}
\label{sec_curvature_flow}
We provide a brief introduction to the network flow, which is curvature flow on embedded graphs in $\R^n.$ A full discussion of the current state of mathematical knowledge about this flow is beyond the scope of this paper, and we direct readers to the comprehensive survey by Mantegazza, Novaga, Pluda, and Schulze~\cite{MNPS16}. Instead, we focus on a qualitative description of the properties observed in simulations, in order to motivate the graph limits defined later in the paper. We should note that the mathematical analysis of these systems is very difficult, but significant progress has recently been made for the case of graphs embedded in two dimensions~\cite{2014ilmanen}.
\begin{figure}
\center
\includegraphics[width=.9\textwidth]{grainGrowth_evolution.pdf}
\caption{\label{fig:ggEvolution} A planar graph evolving by curvature flow.}
\end{figure}
We define curvature flow on embedded graphs by locally prescribing the behavior at each point, following the approach in ~\cite{2014ilmanen}. Alternatively, it can be defined as a type of ``gradient flow'' on the sum of the lengths of the edges~\cite{2014elsey}, but we omit that description here for simplicity.
A \textbf{regular graph} is an embedded graph such that every vertex has degree three, and the angles between the outward pointing tangent vectors of any edges meeting at a vertex equal $\frac{2\pi}{3}$ at that vertex. The latter condition is called the Herring Angle Condition. For such graphs, curvature flow is defined as follows:
\begin{Definition}
A continuous function $f:\paren{T_1,T_2}\rightarrow \mathcal{G}^n$ is a \textbf{curvature flow on regular graphs} if $f\paren{t}$ is a regular graph for all $t\in\paren{T_1,T_2},$ and there there exists a collection of parametrizations of the edges of $f\paren{t},$ $\set{\gamma_i:\paren{T_1,T_2}\rightarrow\R^n}$, satisfying
\s{\frac{\partial \gamma_i\paren{t}}{\partial t}=\kappa\vec{n}\;\;\;\; \forall t\in\paren{T_1,T_2}}
where $\kappa$ is the curvature and $\vec{n}$ is the unit normal vector with orientation given by the parametrization.
\end{Definition}
The restriction to regular graphs is quite severe, but the definition of the the network flow can be extended to more general graphs as follows:
\begin{Definition}
A continuous function $f:\brac{T_1,T_2}\rightarrow \mathcal{G}^n$ is a \textbf{curvature flow on graphs} and a \textbf{network flow} if for any bounded region $B\subset\R^n,$ there is a finite collection of times $\set{T_1=s_1,\ldots,s_m=T_2}$ such that $f_B\paren{t}:=f\paren{t}\cap B$ defines a curvature flow on regular graphs on the interval $\paren{s_i,s_{i+1}}$ and
\s{\lim_{t\rightarrow s_i} f\paren{t}=f\paren{s_i}}
for each $1\leq i \leq n,$ in the varifold topology. The $\set{s_i}$ are called singular times.
\end{Definition}
\cite{2014ilmanen} discusses hypotheses under which a network flow exists starting from a non-regular initial condition in two dimensions.Note that the behavior of the flow after a singular time is not necessarily uniquely determined.
\subsection{Topological Changes}
A key feature of curvature flow on graphs is the existence of singularities in time which result in topological changes. The simplest of these changes, occurs when an edge shrinks to a point and two trivalent vertices collide, resulting in a vertex of degree four. This vertex will split into two vertices of degree three, with the adjacent edges shuffled. Another topological change occurs when a triangle shrinks to a point, resulting in the deletion of three edges and three vertices. Ilmanen, Neves and Schulze~\cite{2014ilmanen} conjecture that all singularities of the network flow in two dimensions occurs when when a connected collection of edges shrinks to a point to form a vertex of degree greater than or equal to three. In particular, the flow never leads to an edge with multiplicity greater than one.
\subsection{Qualitative Behavior}
\begin{figure}
\center
\subfloat[]{%
\label{fig:vor}{%
\includegraphics[width=7cm]{%
/Voronoi_2D_plain.pdf}}}
\hspace{20pt}
\subfloat[]{%
\label{perturbed_lattice}{%
\includegraphics[width=7cm]{%
/Perturbed_2D_plain.pdf}}}
\caption{\label{fig_initialConditions}Two initial conditions that appear to exhibit universal behavior under curvature flow: (a) a Voronoi diagram of Poisson distributed points and (b) a perturbed hexagonal lattice.}
\end{figure}
A mathematical understanding of the long-term behavior of the network flow is beyond current knowledge. However, materials scientists and physicists have put considerable effort toward simulating these systems. Here, we briefly discuss the observed behavior for graphs embedded in two dimensions.
For many random initial conditions, such as those depicted in Figure~\ref{fig_initialConditions}, the network flow appears to exhibit universal behavior in the long-term. That is, the statistics of their long-term scale-invariant properties converge to values that are independent of the initial conditions. We will show evidence for this in Section~\ref{sec_computations}. Another important behavior of these systems is that the average edge length is observed to decrease at a rate of $t^{-1/2}.$. This causes the area of components of the complement (or ``grains'') to increase, in a process called coarsening. Simulated systems of periodic graphs coarsen until until they reach a very small stationary state.
Stationary states of the network flow are embedded graphs with straight line edges meeting at trivalent vertices at angles of $2\pi/3.$ An example is the regular hexagonal lattice. There also exist non stationary flows that share the local statistics a stationary state. For example, an infinite hexagonal lattice with an edge removed will have one grain that expands forever, but the graph will always look hexagonal sufficiently far away from it.
Finally, one can construct examples that neither become stationary nor appear to exhibit universal behavior. Two such initial conditions are shown in Figure~\ref{fig_halfHex2}. We introduce the hypothesis of homogeneity in Section~\ref{sec_hom} to rule out this behavior.
\begin{figure}
\subfloat[]{%
\label{fig_halfHex2}{%
\includegraphics[width=7cm]{%
/InhomogeneousGraph.pdf}}}
\hspace{20pt}
\subfloat[]{%
\label{fig_1Din2D}{%
\includegraphics[width=7cm]{%
/1Din2D.pdf}}}
\caption{\label{fig_inhomogenous} Finite regions of two graphs that do not appear to exhibit universal behavior. }
\end{figure}
\section{Local Topological Convergence}
\label{chp_Swatches}
\label{sec_swatch}
Local topological convergence, or Benjamini-Schramm convergence, is a notion of weak local limit for abstract graphs. It associates to a graph a family of probability distribution of local topological configurations of radius $r$ for each $r\in\N.$ Local topological convergence is the convergence of these discrete probability distributions. It was first noted that curvature flow on graphs could correspond to a Benjamini-Schramm graph limit in our paper ``Topology of Random Cell Complexes, and Applications''~\cite{2016schweinhart}, building on the work in~\cite{2012masonB}. In that paper, we defined local topological convergence for general regular cell complexes, but here we briefly review the concept for graphs.
\subsection{Swatches and Topological Cloths}
\label{sec_cloth}
In the following, the graph distance between two cells (edges or vertices) of a graph $G$ is the minimum number of edges and vertices in a path connecting them. The graph distance from a vertex to a neighboring edge is one, between vertices sharing an edge is two, and between cells in different connected components is $\infty.$ . A swatch is a ball in the graph distance:
\begin{Definition}
Let $v$ be a vertex of a graph $G.$ The \textbf{swatch} of radius $r$ at $v$ is the ball of radius $r$ centered at $v$ in the graph distance.
\end{Definition}
Note that the swatch is not a graph on vertices contained in $G,$ but rather a bipartite graph on the set of edges and vertices of $G.$ That is, an edge may be contained in a swatch that does not contain both of its adjacent vertices. Allowing such intermediate neighborhoods does not matter for the theory of topological convergence, but it gives more flexibility for computation applications.
We will be counting graph isomorphism classes of swatches. Two swatches are said to have the same \textbf{swatch type} if they are isomorphic as bipartite, rooted graphs. A \textbf{subswatch} of a swatch is a swatch of smaller radius at the same root vertex.
\begin{figure}
\center
\subfloat[]{%
\label{free_swatch}{%
\includegraphics[height=5cm]{%
FreeSwatch_color.pdf}}}
\hspace{20pt}
\subfloat[]{%
\label{not_free_swatch}{%
\includegraphics[height=4.75cm]{%
NonFreeSwatch_color.pdf}}}
\caption{\label{fig_swatchTypes}Swatch types of radius six in an embedded graph. The vertex color indicates distance from the root, with the root colored dark blue. (a) A free swatch where there is a single path from the root to any given vertex. (b) A swatch that contains a cycle of length four and a cycle of length two.}
\end{figure}
It will be useful later to have a distance on swatches. Let the largest common subswatch of two swatches be the swatch of largest radius that is a subswatch of both. The distance between two swatches is defined as the reciprocal of the number of vertices and edges in the largest common subswatch, or zero if the swatches are the same. For example, the largest radius for which the swatch in Figure~\ref{not_free_swatch} is free is $r = 3$, and the distance to the free swatch in Figure~\ref{free_swatch} is $1/13$.
\begin{Definition}
Let $G$ be a finite, connected graph. The \textbf{topological cloth at radius $r$}, denoted $\mathcal{S}_r\paren{G}$ is the probability measure on the set of swatch types of radius $r$ induced by the counting measure on the vertices. That is, for a swatch type $S,$ $\mathcal{S}_r\paren{G}\paren{S}$ is the proportion of vertices of $G$ whose swatch of radius $r$ is combinatorially equivalent to $S.$
\end{Definition}Letting $r$ vary, we get a family of probability distributions called the topological cloth of $G.$ The topological cloth characterizes the local topology of the graph in the sense of determining the probability at which any local configuration appears in the graph, as well as of prescribing all of its local topological properties (in a sense defined below).
\subsection{A Topological Distance on Graphs}
\label{sec_metric}
We use the earth mover's distance on the topological cloth at radius $r$ to define a family of distances on graphs. Suppose $\paren{X,\rho}$ is a finite metric space, and $P_1$ and $P_2$ are two probability distributions on it. A matching $M$ of $P_1$ and $P_2$ is a probability distribution on $X\times X$ with marginals are $P_1$ and $P_2.$ The earth mover's distance $d_{\text{em}}$, or first Wasserstein metric, is the minimum cost of a matching between $P_1$ and $P_2$ \cite{1781Monge,1998Rubner}:
\s{d_\text{em}\paren{P_1,P_2}=\min_{M}\sum_{x_1,x_2\in X}\rho\paren{x_1,x_2}M\paren{x_1,x_2}}
Given two graphs $G_1$ and $G_2,$ define the distance at radius $r$ to be the earth mover's distance between their topological cloths of radius $r$:
\s{d_r\paren{G_1,G_2}=d_{\text{em}}\paren{\mathcal{S}_r\paren{G_1},\mathcal{S}_r\paren{G_2}}}
$d_r$ is uniformly bounded and non-decreasing in $r$, and it stabilizes for some finite $r$ if the graphs are finite. The limit distance on graphs is defined as the limit of $d_r$ with increasing $r$, or
\begin{equation}
\label{eq_clothDist}
d_\infty\paren{G_1,G_2}=\lim_{r\rightarrow\infty}d_r\paren{G_1,G_2}
\end{equation}Note that a a sequence $\set{G_i}$ is Cauchy in $d_\infty$ if and only if all swatch frequencies converge.
\subsection{Local Topological Convergence}
Let $\mathfrak{G}^\bullet$ be the space of countable, connected graphs with a root vertex specified. The theory of Benjamini-Schramm graph limits associates to a Cauchy sequence in $d_\infty$ a probability distribution on $\mathfrak{G}^\bullet:$
\label{sec_convergence}
\begin{theorem}[Benjamini-Schramm Convergence~\cite{2012lovasz,2001benjamini}]
\label{thm_BS}
Let $G_1,G_2,\ldots=\left\{G_i\right\}$ be a Cauchy sequence of finite graphs in $d_\infty.$ There exists a unique probability distribution $\sigma$ on $\mathfrak{G}^\bullet$ such that $\left\{G_i\right\}$ converges to $\sigma$ in $d.$ The value of $\sigma$ on a basis set $E_S$ is given by $\sigma\paren{E_S}=\sigma_r\paren{S}$ where $r$ is the radius of the swatch $S$ and $\sigma_r$ is the limiting distribution on the set of swatches of radius $r.$ $\sigma$ is called the \textbf{Benjamini-Schramm limit} or \textbf{local topological limit} of the sequence of graphs.
\end{theorem}
\subsection{Consequences of Topological Convergence}
If $G$ is a finite graph, and $H$ is a finite, two-colored graph graph, let $\text{hom}\paren{H,G}$ be the number of graph homomorphisms $H$ into the adjacency graph on the union of the vertex and edge sets of $G,$ and let $\text{inj}\paren{H,G}$ be the number that are injective. Similarly, $H'$ and $G'$ are (possibly infinite) rooted graphs, let $\text{hom}\paren{H',G'}$ and $\text{inj}\paren{H',G'}$ be the homomorphisms that map root to root.
\begin{Definition}
If $H$ is a finite, two-colored graph and $G$ is a finite graph, the \textbf{injective homomorphism frequency} of $H$ in $G$ is
\s{rho\paren{H,G}=\frac{\text{inj}\paren{H,G}}{\abs{G}}=\sum_{v\in G}\frac{\text{inj}\paren{H',G^v}}{\abs{G}}}
where $G^v$ is $G$ with root vertex $v$ specified.If $\sigma$ is a probability distribution on $\mathcal{G}^\bullet,$ and $H'$ is a rooted graph, the injective homomorphism frequency of $H'$ in $\sigma$ is
\s{\rho\paren{H',\sigma}=\mathbb{E}_{G'\in G}\text{inj}\paren{H',G'}}
where the expectation is taken with respect to $\sigma.$
\end{Definition}The probability distributions occurring as Benjamini-Schramm graph limits have a property called involution invariance, which implies that $\rho\paren{H',\sigma}$ does not depend on the choice of root of $H.$
\begin{Definition}
\label{local_topo_prop}
A \textbf{local topological property} of graphs is any property that can be expressed in terms of a finite combination of injective homomorphism frequencies.
\end{Definition}For example, any quantity of the form $\frac{\text{hom}\paren{H,G}}{\abs{G}}$ is a local topological property~\cite{2012lovasz}.
\begin{theorem}
A sequence of graphs $\set{G_i}$ converges in the local topological sense to a limit distribution $\sigma$ if and only if all local topological properties to the corresponding quantities defined for the Benjamini-Schramm limit.~\cite{2012lovasz}
\end{theorem}
\subsection{Cloths and Limits of Embedded Cell Complexes}
\label{sec_infiniteCloth}
Here, we extend the notions of topological cloth and local topological convergence from finite graphs to to countable, locally finite graphs embedded in $\R^n.$ Let $G$ be an embedded graph in $\R^n.$ We associate to $G$ to sequences of finite graphs: $I_r\paren{G}$ is the union the vertices and edges of $G$ completely contained in the ball of radius $r$ centered at the origin, and $O_r\paren{G}$ is the union of the vertices and edges of $G$ intersecting that ball.
\begin{Definition}If $\lim_{r\rightarrow\infty}O_r\paren{G}$ is locally topologically convergent, we say that $G$ has an \textbf{outer-regular topological cloth}. If $I_r\paren{G}$ does, $G$ has an \textbf{inner-regular topological cloth}. If they are equal, $G$ has a \textbf{well-defined topological cloth}, which we call $\mathcal{S}\paren{G}.$
\end{Definition}
Note that $\mathcal{S}\paren{G}$ does not depend on the choice of origin in $\R^n.$ If $\set{G_i}$ is a sequence of embedded graphs with well-defined topological cloths, we say that $\set{G_i}$ is locally topologically convergent if the probability distributions $\mathcal{S}\paren{G_i}$ converge strongly to a limit distribution.
\section{Topologies on the Space of Embedded Graphs}
\label{sec_graphTop}
In this section, we will study the properties of the several topologies on the space of embedded graphs, as preparation for the definition of local geometric convergence. The varifold topology is the most natural topology in the context of curvature flow on embedded graphs in $\R^n,$ but the topology induced by the local Hausdorff metric is perhaps more natural for computational applications. In this section, we study the properties of these two topologies, and relate them to a third topology called the smooth topology of topological types.
Many of the proofs of the propositions in this section are quite technical, and are deferred to an appendix at the end of the paper.
\subsection{The Varifold Topology}
The varifold topology is the vague (or weak*) topology on certain Radon measures associated to embedded graphs. If $X$ is a locally compact Hausdorff space, let $\mathcal{M}^+\paren{X}$ be the space of positive, locally finite Radon measures on $X$. Give $\mathcal{M}^+\paren{X}$ the initial topology with respect to the integrals of compactly, supported, continuous functions. That is, a sequence of Radon measures $\mu_i$ converges to $\mu$ if and only if for every continuous, compactly supported function $f:X\rightarrow\R$
\s{\lim_{i\rightarrow\infty}\int_{X}f\;d\mu_i\rightarrow \int_{X}f\;d\mu}
$\mathcal{M}^+\paren{X}$ is metrizable, and if $Y\subset X$ the restriction map $\mathcal{M}^+\paren{X}\rightarrow\mathcal{M}^+\paren{Y}$ is continuous. If $X$ is compact, the vague topology~\cite{2002kallenberg,1976dieudonne}.
If $e$ is an edge of a graph $G,$ we say that $e\in G$ and denote the tangent bundle of $e$ as $T_x\paren{e}.$ The tangent bundle $T\paren{G}\subset \R^n\times\mathbb{RP}^{n-1}$ of an embedded graph $G$ is
\s{T\paren{G}=\set{\paren{x,v:x\in e \in G, v\in T_x\paren{e}}}}
That is, the set of pairs $\paren{x,v}$ where $x$ is a point of $G$ and $v$ is a unit tangent vector of an edge containing $x.$ The projection map $T\paren{G}\rightarrow G$ is injective away from vertices.
An embedded graph with multiplicity $m$ is a graph $G$ together with a function $m:G\rightarrow \N$ that is constant on each edge. To such a graph, we associate a positive Radon measure $\mu_G$ on $\R^n\times \mathbb{RP}^{n-1}$ that assigns to a Borel subset of $Y$ of $\R^n$
\s{\mu_G\paren{Y}=\int_{T\paren{G\cap X}}m\;d\mathcal{H}^1}
where $\mathcal{H}^1$ is the first-dimensional Hausdorff measure. The \textbf{varifold topology} on the space of embedded graphs is the vague topology on the measures $\mu_G.$ associated measures. In other words a sequence of graphs $\set{G_i}$ converges to $G$ if and only if for any continuous, compactly supported function $f:\R^n\times\mathbb{RP}^{n-1}\rightarrow\R,$
\s{\lim_{i\rightarrow\infty} \int_{\R^n}f\mu_{G_i} = \lim_{i\rightarrow\infty} \int_{T\paren{G_i}}m_i f\;d\mathcal{H}^1\rightarrow \int_{T\paren{G}}m f\;d\mathcal{H}^1= \int_{\R^n}f\mu_{G}}
\begin{Definition}
Let $\check{\mathcal{G}}^n_r$ be the space of embedded graphs with multiplicity in the open $n$-ball of radius $r$ with the varifold topology. Let $\mathcal{G}^n$ be the space of embedded graphs with multiplicity in $\R^n$ with the same topology.
\end{Definition}For more information on the properties of the varifold topology, we suggest~\cite{1983simon,deLellis,1972allard}.
\subsection{The Hausdorff Metric Topology}
If $X$ and $Y$ are subsets of $\R^n$ the Hausdorff distance between them is
\s{d_H\paren{X,Y}=\max\Big(\sup_{x\in X}\inf_{y\in Y} d\paren{x,y},\sup_{y\in Y}\inf_{x\in X} d\paren{x,y}\Big).}That is, the smallest $\epsilon$ so that $Y$ is contained in the closed $\epsilon$ neighborhood of $X$ and visa versa. $d_H$ makes the set of compact subsets of of any bounded subset of $\R^n$ into a complete metric space.
Let $B_r$ be the open ball of radius $r$ centered at the origin in $\R^n,$ and $K\paren{B_r}$ be the space of relatively compact subsets of $B_r$ with the Hausdorff metric topology. If $r<s,$ the map from $K\paren{B_s}$ to $K\paren{B_s}$ given by intersecting a subset of $B_s$ with $B_r$ is not continuous. We introduce a different metric inducing at topology for which this map is continuous:
\begin{Definition}
If $X_1$ and $X_2$ are subsets of the open ball of radius $r,$ $B_r,$ the \textbf{local Hausdorff distance} between them is
\s{d_L\paren{X_1,X_2}=d_H\paren{X_1\cup \partial B_r,X_2\cup \partial B_r}}
\end{Definition}
\begin{Definition}
$\mathcal{G}^n_r$ is the space of embedded graphs in the open ball of radius $r$ in $\R^n$ with the topology induced by the local Hausdorff distance. Let $p_{s,r}:\mathcal{G}_s^n\rightarrow \mathcal{G}_r^n$ for $s\geq r$ be the continuous map given by intersecting an embedded graph in $B_s$ with $B_r.$ Also, define $\mathcal{G}^n$ to be the inverse limit of the spaces $\mathcal{G}_r^n$ with respect to this family of maps.
\end{Definition}
$\mathcal{G}_r^n$ is separable, with a dense subset given by graphs whose vertices have rational coordinates, and whose edges are given by polynomial maps from $\brac{0,1}$ to $B_r.$ $\mathcal{G}^n$ can be equivalently defined as the space of embedded graphs in $\R^n$ with the initial topology with respect to the intersection maps $p_r:\mathcal{G}^n\rightarrow \mathcal{G}_r^n$ for $n\in \N.$
Let $r:\check{\mathcal{G}}_r^n\rightarrow\mathcal{G}_r^n$ be the map induced by forgetting the multiplicity.
\begin{restatable}{proposition}{propOne}
\label{prop_HausdorffVarifold}
Let $\set{\check{G}_i}$ be a sequence of graphs in $\check{\mathcal{G}}^n_r$ that converges to a limit $\check{G}\in\check{\mathcal{G}}^n_r$. Then $\set{r\paren{G_i}}$ converges to $r\paren{\check{G}}.$ That is, $r$ is continuous.
\end{restatable} See Appendix~\ref{appendix} for the proof. A partial converse to this result follows from the Allard compactness theorem~\cite{1972allard}. If $G_i\rightarrow G$ in the local Hausdorff metric, and the $G_i$ have uniformly bounded curvatures and masses, that result implies that there is a convergent subsequence in the varifold topology, and that limit must equal $G$ modulo multiplicity.
\subsection{The Smooth Topology of Topological Types}
\label{sec_topType}
A \textbf{topological type} $S$ of $\mathcal{G}_{r}^{n}$ is, as a set, equal to the collection of graphs in $\mathcal{G}_{r}^{n}$ sharing a single combinatorial isomorphism type. We will define a topology of smooth convergence on $S$ in which a sequence of graphs converges if and only if their edges converge smoothly to edges.
The non-linear Grassmanian $\text{Gr}\paren{\brac{0,1},\R^n}$ of smoothly embedded curves with boundary is the space
\s{\text{Gr}\paren{\brac{0,1},\R^n}=\frac{\text{Emb}\paren{\brac{0,1},\R^n}}{\text{Diff}\paren{\brac{0,1},\brac{0,1}}}}
where $\text{Emb}{\brac{0,1},\R^n}\subset C^{\infty}\paren{\brac{0,1},\R^n}$ is is the space of embeddings of $\brac{0,1}$ in $\R^n$ and $\text{Diff}\paren{\brac{0,1},\brac{0,1}}$ is the space of diffeomorphisms of the interval. $\text{Gr}\paren{\brac{0,1},n}$ is a smooth Frechet manifold~\cite{2014balmaz}, and can be metrized as follows. $C^\infty\paren{\brac{0,1},\R^n}$ is a Frechet space, there is a metric $\hat{d}$ inducing its topology. For a curve $x\in \text{Gr}\paren{\brac{0,1},\R^n},$ let $\phi_x^1:\brac{0,1}\rightarrow\R^n$ and $\phi_x^2$ the two unit-speed parametrizations. Then
\s{d\paren{x,y}=\min\paren{\hat{d}\paren{\phi_{x}^1,\phi_y^1},\hat{d}\paren{\phi_{x}^2,\phi_y^1}}}
induces the topology on $\text{Gr}\paren{\brac{0,1},n}.$\footnote{We will use metrizability of this space to apply the Portmanteau theorem. A difficulty in generalizing the concept of local geometric convergence to higher dimensional regular cell complexes is showing that the corresponding non-linear Grassmanians are completely regular.}
If $S$ is a topological type with $l$ edges, there is an injective function $\psi:S\rightarrow\text{Sym}^{l}\paren{\text{Gr}^{D}\paren{k,n}}$ given by sending an embedded graph in $S$ to the tuple of the closures of its edges inside the closed ball of radius $r.$ Topologize $S$ as a subspace of $\text{Sym}^{l}\paren{\text{Gr}^{D}\paren{k,n}},$ so a sequence of graphs converges if and only if each of its edges converges in the smooth topology. $S$ is metrizable and second-countable.
To be precise, let $\set{S^i_{n,r}}$ be the set of topological types of $\mathcal{G}_r^n,$ one for each graph isomorphism class, and let $s_i:S^i_{n,r}\rightarrow \mathcal{G}^{n}_r$ be the map induced by inclusion.
\begin{Definition}
The space of embedded graphs in the ball of radius $r$ in $\R^n$ with the \textbf{smooth topology of topological types}, $\hat{\mathcal{G}}^{n}_r$ is the set $\mathcal{G}^n_r$ with the final topology with respect to the maps $s_i.$
\end{Definition}
\begin{Definition}
\label{def_local_prop}
A \textbf{local geometric property} of an embedded graph in $\R^n$ is a function of the form $f=\hat{f} \circ s^{-1}\circ p_r$ where $\hat{f}:\hat{\mathcal{G}}_r^n\rightarrow \R$ is a bounded, continuous function on $\hat{\mathcal{G}}_r^n.$
\end{Definition}
A \textbf{topological type with multiplicity} of $\check{\mathcal{G}}_r^n$ is as a set, equal to the collection of graphs in $\check{\mathcal{G}}_{r}^{n}$ sharing the same combinatorial isomorphism type as colored graphs, with the multiplicity giving the coloring. The construction of a smooth topology of topological types proceeds identically to that for graphs without multiplicity. In the next section, we introduce a concept of $\delta$-thickenability that allows us to mediate between these smooth topologies and the Hausdorff metric and varifold topologies.
\subsection{Thickenability of Embedded Graphs}
A $\delta$-thickenable graph in $\mathcal{G}_r^n$ is, roughly speaking, one whose combinatorial properties are unchanged if each vertex is replaced by a ball of radius $\epsilon<\delta/2,$ each edge is replaced by its $\epsilon^2/2$-tube, and it is intersected with the ball of radius $r-\epsilon/2.$ It is related to the notion of control data for a stratified space. First, we require two concepts:
\begin{itemize}
\item The (open) $\epsilon$-neighborhood of a a subset $X$ of $\R^n$ is the set of all points within distance $\epsilon$ of it. It is denoted $X_\epsilon.$
\item A smooth curve $e$ in $\R^n$ has an (open) tubular neighborhood of radius $\epsilon$ if every point in $e_\epsilon$ is closest to a unique point of $e.$ This neighborhood is denoted $T_\epsilon\paren{e}.$ If $e$ has a tubular neighborhood of radius $\epsilon,$ then the curvature of $e$ is above by $1/\epsilon,$ and that the intersection of $e$ with any ball of radius less than $\epsilon/2$ is connected.
\end{itemize}
\begin{figure}
\center
\subfloat[]{%
\label{fig_beforeThicken}{%
\includegraphics[width=.3\textwidth]{%
/beforeThicken.pdf}}}
\hspace{20pt}
\subfloat[]{%
\label{fig_afterThicken}{%
\includegraphics[width=.3\textwidth]{%
/afterThicken.pdf}}}
\caption{\label{fig_thicken} (a) An embedded graph in $B_r$ that is $\delta$-thickenable for $\delta\leq \Delta=.133r.$ The limiting factor for thickenability is a pair of vertices of distance $\Delta$ apart. (b) The graph obtained by replacing each vertex with a ball of radius $\Delta/2$ and each edge with a tube of radius $\Delta^2/4.$}
\end{figure}
\begin{Definition}
$G\in\mathcal{G}^n_r$ is \textbf{$\delta$-thickenable} if it satisfies the following for all $\epsilon<\delta$:
\begin{enumerate}
\item If $v$ and $w$ are vertices of $G,$ then $d\paren{v,w}\geq\delta.$ Also, the only edges of $G$ that intersect $B_\delta\paren{v}$ are those adjacent to $v.$
\item Each edge has a tubular neighborhood of radius $2\epsilon.$
\item If $e$ is an edge of $G$ adjacent to vertices $v_1$ and $v_2$ then
\s{T_{\epsilon^2}\paren{e}\cap \paren{G-e}\subset B_{\epsilon}\paren{v_1}\cup B_{\epsilon}\paren{v_2}\cup \paren{B_r-B_{r-\epsilon}}}
\item Every vertex and edge of $G$ intersects $B_{r-\delta}.$
\item If $e$ is an an edge and $\paren{B_r-B_{r-\epsilon}}\cap e\neq \emptyset,$ then each connected connected component of $\paren{B_r-B_{r-\epsilon}}\cap e$ intersects $\partial B_{r-\epsilon}$ in exactly one point.
\end{enumerate}Note that if $G$ is $\delta$-thickenable, then it is $\epsilon$-thickenable for all $\epsilon\leq \delta.$ An embedded graph $G\in\mathcal{G}^n$ graph is \textbf{locally thickenable} if its intersection with any open ball is $\delta$-thickenable for some $\delta>0.$
\end{Definition} An example of a graph in $B_r$ for which vertex separation is the limiting factor for thickenability is shown in Figure~\ref{fig_beforeThicken}.
\begin{example}
Let $G$ be an embedded graph in $\R^2$ whose vertices are at least distance $D$ apart, and whose edges are straight line segments meeting at angles of at least $\theta.$ Then $G$ is $\min\paren{D,2\sin\paren{\theta/2}}$-thickenable in $\R^2.$
\end{example}
The proof of the following lemma and proposition are included in Appendix~\ref{appendix}.
\begin{restatable}{lemma}{lemmaTwo}
All embedded graphs $G\in\mathcal{G}^n$ are locally thickenable.
\end{restatable}
\begin{restatable}{proposition}{propTwo}
\label{prop2}
Let $X_{\delta,D}\subset \mathcal{G}_r^n$ be the set of $\delta$-thickenable embedded graphs such that the $n-1$ Frenet-Serret curvatures of each edge are bounded above by $D.$
\begin{enumerate}
\item $X_{\delta,D}$ is compact, and equals the disconnected union of sets of constant topological type.
\item If $\hat{X}_{\delta,D}\subset \hat{\mathcal{G}}_r^n$ is the same set with the smooth topology, then $\hat{X}_{\delta,D}$ is homeomorphic to $X_{\delta,D}.$
\end{enumerate}
The corresponding statements are also true for embedded graphs with multiplicity in $\check{\mathcal{G}}_r^n.$
\end{restatable}
\section{Local Geometric Convergence}
\label{sec_geoConv}
In this section, we propose a notion of convergence for embedded graphs that implies the convergence of averages of a large class of local geometric properties. This is called \textbf{local geometric convergence}, and was developed in analogy with Benjamini-Schramm convergence. In short, we associate to an embedded graph $G\in\mathcal{G}^n$ an empirical probability distribution $\mathcal{P}\paren{G}$ on $\mathcal{G}^n$, called the the \textbf{geometric cloth}. Local geometric convergence is related to weak convergence of geometric cloths.
The definitions of local geometric convergence for embedded graphs in $\mathcal{G}^n$ and embedded graphs with multiplicity in $\check{\mathcal{G}_r^n}$ are nearly identical, as are the proofs of their properties. For brevity, we will work with $\mathcal{G}^n$ here.
\subsection{Preliminaries: Weak Convergence and the Ergodicity}
\label{sec_prelim}
We will require the following definitions and results.
\subsubsection{Weak Convergence}
\begin{definition}
A sequence of probability measures $\set{\mu_i}$ on a completely regular space $X$ \textbf{converges weakly} to a probability measure $\mu$ if
\s{\lim_{i\rightarrow\infty}\int_{X}f\;d\mu_i=\int_{X}f\;d\mu}
for all bounded continuous functions $f:X\rightarrow \R.$
\end{definition}
\begin{theorem}[Portmanteau theorem~\cite{1970topsoe}]
Let $X$ be a completely regular topological space, $\mu$ a probability measure, and $\mu_i$ a sequence of probability measures on $Y.$ The following are equivalent:
\begin{enumerate}
\item $\mu_i\rightarrow \mu$ weakly.
\item $\limsup\mu_i\paren{C}\leq \mu\paren{C}$ for all closed subsets $C\subset X.$
\item $\liminf\mu_i\paren{U}\geq\mu\paren{U}$ for all open subsets $U\paren X.$
\item $\mu_i\paren{Z}\rightarrow \mu_i\paren{Z}$ for all $Z\subset X$ with $\mu\paren{\partial Z}=0.$
\end{enumerate}
\end{theorem}
Note that all metric spaces are completely regular, and that the hypotheses do not require $\mu$ to be a Radon measure.
\begin{Definition}
A collection of probability measures $M$ on a space $X$ is \textbf{tight} if for all $\epsilon>0$ there is a compact set $K_{\epsilon}$ so that for all $\mu\in M$
\s{\mu\paren{K_{\epsilon}}>1-\epsilon}
\end{Definition}
\begin{theorem}(Prokhorov's Theorem)
The closure of a tight collection of probability measures on a separable metric space is compact in the weak topology.
\end{theorem}
\subsection{Uniformly Separating Sequences}
\label{sec_USS}
Let $s:\hat{\mathcal{G}}^{r}_n\rightarrow\mathcal{G}^{r}_n$ be the identity map. It is bijective and continuous, and is a homeomorphism when restricted to the set $\hat{X}_m$ of $1/m$-thickenable graphs whose Frenet-Serret curvatures bounded above in magnitude by $m.$ Let $X_m=s\paren{\hat{X}_m}.$
\begin{lemma}
$s^{-1}$ is a Borel function.
\end{lemma}
\begin{proof}
If $B$ be a Borel set of $S$ then
\s{s\paren{B}=s\paren{\bigcup_{m\in\N}{\hat{X}_m\cap B}}=\bigcup_{m\in\N}s\paren{\hat{X}_m\cap B}=\bigcup_{m\in\N}X_m\cap s\paren{\hat{X_m}\cap B}}
which is a Borel set, because the countable union and homeomorphic image of Borel sets are Borel.
\end{proof}The previous lemma implies that Borel measures on $\mathcal{G}^{r}_n$ pull back to Borel measures on $\hat{\mathcal{G}}^{r}_n,$ so we can make the following definition:
\begin{Definition}
Let $\mu$ be a probability measure on $\mathcal{G}^n_r.$ The \textbf{induced smooth measure} $s^*\mu$ on $\hat{\mathcal{G}}^{r}_n$ is the pushforward of $\mu$ by $s^{-1}.$
\end{Definition}
\begin{Definition}
\label{defn_unifSep}
A sequence of probability measures $\set{\mu_i}$ on $\mathcal{G}_{r}^n$ is \textbf{uniformly separating} for all $\epsilon>0$ there is an $M$ so that each $\mu_i$ satisfies
\s{\mu_i\paren{X_M}>1-\epsilon}
Similarly, a sequence of probability measures $\set{\nu_i}$ on $\mathcal{G}^n$ is uniformly separating if $\set{{p_r}_*\nu_i}$ is for each $r\in\N.$
\end{Definition} $X_M$ is compact, so a uniformly separating sequence on $\mathcal{G}_{r}^n$ is tight.
\begin{proposition}
\label{unifProp}
Let $\set{\mu_i}$ be a uniformly separating sequence of probability measures on $\mathcal{G}_{r}^n.$ $\mu_i$ converges weakly to $\mu$ if and only if $s^*\mu_i$ converge weakly to $s^*\mu.$
\end{proposition}
\begin{proof}
The smooth topology is finer than that of $\mathcal{G}_{r}^n,$ so weak convergence of $s^*\mu_i$ to $s^*\mu$ implies weak convergence of $\mu_i$ to $\mu,$ even without the requirement of uniform separation.
For the other direction, assume that $\mu_i\rightarrow \mu$ weakly and let $U$ be an open set of $\hat{\mathcal{G}}^{r}_n.$ We would like to show that
\s{\liminf_i \mu_i\paren{s\paren{U}}\geq \mu\paren{s\paren{U}}} Let $\epsilon>0.$ $\set{\mu_i}$ is uniformly separating, so we can find an $M$ so that $\mu_i\paren{X_M}>1-\epsilon$ for all $i.$ $X_M$ is closed so $\mu\paren{X_M}>1-\epsilon,$ as well. $s$ is a homeomorphism when restricted to $\hat{X}_{M},$ so $s\paren{U}\cap X_M=s\paren{U\cap\hat{X}_{M}}$ is open in $X_{M}.$ It follows that there is an open set $V$ of $\mathcal{G}_{r}^n$ with $X_M\cap s\paren{U}=X_{M}\cap V.$ Choose $I$ so that $\mu_i\paren{V}>\mu\paren{V}-\epsilon$ for all $i>I.$ Then
\begin{align*}
\mu_i\paren{s\paren{U}}&\\
&\geq \mu_i\paren{X_{M}\cap s\paren{U}}-\epsilon\\
&= \mu_i\paren{X_M\cap V}-\epsilon\\
&\geq \mu_i\paren{V}-2\epsilon\\
&\geq\mu\paren{V}-3\epsilon\\
& \geq \mu\paren{X_{M}\cap s\paren{U}}-4\epsilon\\
& \geq \mu\paren{s\paren{U}}-5\epsilon
\end{align*}
So
\s{\liminf_i \mu_i\paren{s\paren{U}}\geq \mu\paren{s\paren{U}}}
and $s^*\mu_i\rightarrow s^*\mu$ weakly, as desired.
\end{proof}
\subsection{The Geometric Cloth}
\label{sec_geoCloth}
$\R^n$ acts on $\mathcal{G}^{n}$ by translation. Let $G\in\mathcal{G}^n$ be an embedded graph and let $\phi^G:\R^n\rightarrow\mathcal{G}^{n}$ be the map sending $x$ to the graph $xG.$ Also, let $\phi_r^G:\R^n\rightarrow\mathcal{G}_r^n$ be the composition of $p_r\circ\phi^G,$ as shown in Figure~\ref{fig:geometric_cloth}.
\begin{figure}
\center
\includegraphics[width=.6\textwidth]{/GeometricCloth.pdf}
\caption{\label{fig:geometric_cloth} An embedded graph $G\in\mathcal{G}^2$ induces a ``window map'' $\phi_r^G:\R^2\rightarrow\mathcal{G}_r^2.$}
\end{figure}
\begin{Definition}
\label{def_geo_cloth}
Let $G\in\mathcal{G}^{n}.$ If $K$ is a bounded subset of $\R^n,$ define the \textbf{geometric cloth of $G$ in K}, let $P\paren{G,K}$ be the pushforward of the normalized Lebesgue measure on $K$ via $\psi_G.$ If $G$ is infinite, and the sequence $\set{P\paren{G,B_i}}_{i\in\N}$ converges weakly to a non-zero limit distribution $\mathcal{P}\paren{G},$ we call $\mathcal{P}\paren{G}$ the \textbf{geometric cloth} of $G.$ Otherwise, if $G$ is finite, we refer to the geometric cloth of $G$ in its convex hull its geometric cloth $\mathcal{P}\paren{G}.$
\end{Definition} That is, $P\paren{G,K}$ is the probability distribution of graphs embedded in balls of radius $r$ centered at points of $K.$ Note that the definition of geometric cloth, and none of the following definitions in this section, depend on the choice of origin in $\R^n$ because the percentage overlap between the balls of radius $r$ centered at any two points in $\R^n$ goes to $1$ as $r\rightarrow\infty.$
We are interested in infinite graphs because the conjectures about curvature flow on graphs in Section~\ref{sec_conj} are most natural in that context. An infinite graph has smooth geometric cloth if its local geometric properties have well-defined averages over balls:
\begin{Definition}
If $G$ is infinite and the induced smooth measures $\hat{P}_r\paren{G,B_i}=s^*{p_r}_*P\paren{G,B_i}$ on $\hat{\mathcal{G}}_r^n$ also converge weakly to limit distributions, we say that $G$ has a \textbf{smooth geometric cloth} (for example, if $P\paren{G,B_i}$ is a uniformly separating sequence). Finite graphs always have smooth geometric cloth.
\end{Definition}
An equivalent and perhaps more intuitive way to define the geometric cloth is via local probability distributions:
\begin{proposition}
\label{localProp}
An infinite embedded graph $G$ has a geometric cloth if and only if the probability distributions $P_r\paren{G,B_i}$ converge weakly to non-zero limit distributions $\mathcal{P}_r\paren{G}$ as $i\rightarrow\infty$ for all $r\in\N.$
\end{proposition}
\begin{proof}
Recall that the topology on $\mathcal{G}^{n}$ is the initial topology with respect to the maps $p_r:\mathcal{G}^{n}\rightarrow \mathcal{G}_r^{n}$ for $r\in\N.$ For $r\leq s,$ let $p_{s,r}:\mathcal{G}_s^{n}\rightarrow \mathcal{G}_r^{n}$ be the natural map given by intersecting a graph in $B_s$ with $B_r.$ $\mathcal{G}^{n}$ is the inverse limit with respect to the maps $p_{s,r}.$
Weak convergence of $\mathcal{P}\paren{G,B_i}$ to a limit clearly implies weak convergence of the local distributions $\mathcal{P}_r\paren{G,B_i}.$
Conversely, suppose that $P_r\paren{G,B_i}$ converges weakly to $\mu^r$ for each $r\in\N.$ Note that $\mu_r={p_{s,r}}_*\paren{\mu_s}$ for any $s>r.$ Let $\mathcal{B}\paren{\mathcal{G}_r^n}$ be the Borel $\sigma$-algebra of $\mathcal{G}_r^n.$ The collection of sets
\s{\mathcal{A}=\bigcup_{r\in\N}\set{p_r^{-1}\paren{B}: B\in \mathcal{B}\paren{\mathcal{G}_r^{n}}}}
is a semi-ring of Borel sets of $\mathcal{G}^{n}$ generating its topology. Define a pre-measure $\nu_0$ on $\mathcal{A}$ by $\nu_0\paren{B}=\mu^r\paren{p_r\paren{B}}$ if $B=p_r^{-1}\paren{B_0}$ for some $B_0\in\mathcal{B}.$ The consistency of the measures $\mu^r$ and $\mu^s$ implies that this is well-defined. By the Caratheodory extension theorem, $\nu_0$ extends to a unique probability measure $\mu$ on $\mathcal{G}^{n}.$ Weak convergence in a $\pi$-system of sets implies weak convergence with respect to the $\sigma$-algebra it generates~\cite{1999billingsley}, so $\set{P\paren{G,B_i}}$ converges weakly to $\mu.$
\end{proof}
The importance of the geometric cloth is that its local geometric properties have well-defined averages over balls in the following sense:
\begin{Definition}
Let $G\in\mathcal{G}^n$ be an embedded graph, $f$ be a local geometric property, and let $\set{F_i}$ be an increasing sequence of compact sets whose union is $R^n.$ $f$ has a \textbf{well-defined average} $\bar{f}$ over $\set{F_i}$ if
\s{\bar{f}=\lim_{i\rightarrow\infty}\frac{1}{\abs{F_i}}\int_{F_i}f\circ \psi_G\;d\mu}
where $\mu$ is the standard Lebesgue measure on $\R^n.$ If the average of $f$ over the sequence of balls centered at a point in $\R^n$ exists, is simply referred to as the \textbf{well-defined average} of $f$ for $G,$ $\widebar{f_G}.$
\end{Definition} If $G$ has a smooth geometric cloth, and $f$ is a local geometric property, then $f$ has a well-defined $\widebar{f_G}$ and
\s{\widebar{f_G}=\int_{\R^n}f\;d\mathcal{P}\paren{G}}. The converse is true for graphs with smooth geometric cloth for which $\set{P_r\paren{G,B_i}}$ is uniformly separating:
\begin{proposition}
An embedded graph $G\in\mathcal{G}^n$ for which $P\paren{G,B_i}$ is a uniformly separating sequence has a smooth geometric cloth if and only if the averages of local geometric properties over the sequence $\set{B_r}$ converge as $r\rightarrow\infty.$
\end{proposition}
\begin{proof}
The uniform separation property implies that the sequence $\set{P\paren{G,B_r}}$ is tight, and has a convergent subsequence $\mu.$ The convergence of averages of all local geometric properties implies that the limit of any convergent subsequence must be $\mu,$ so $P\paren{G,B_r}\rightarrow \mu$ and $\mu=\mathcal{P}\paren{G}.$
\end{proof} We will examine hypotheses under which the local properties of graphs with smooth geometric cloth have averages over more general sequences in~\ref{sec_hom}.
Local geometric properties include a wide range of important properties, from subgraph densities to the percentage of edges with curvature above a certain value. In contrast, the set of continuous functions on $\mathcal{G}^r_n$ or $\check{\mathcal{G}}_r^n$ is relatively depaupaerate and, for example, does not include most topological properties. That is why weak convergence of probabilities measures on those spaces alone is not sufficient for our purposes.
However, the restriction to bounded functions does exclude many important properties. For example, the number of vertices contained in a ball of radius $r$ in $\R^n$ is continuous with respect to $\hat{\mathcal{G}}_r^n,$ but is unbounded. To obtain more local geometric properties, one can restrict consideration to a smaller class of embedded graphs, such as graphs whose edge length is uniformly bounded.
In summary, embedded graphs with smooth geometric cloth have local geometric properties which have well-defined averages over balls, and those averages are captured by the cloth.
\subsection{Examples}
\label{sec_geoClothExamples}
\subsubsection{Periodic Graphs}
A periodic embedded graph $G$ is one that is invariant under the translation action of a sublattice $\Z^n\subset\R^n.$ Such a graph has a unit cell $T\subset\R^n$ of least volume so that
\s{G=\bigcup_{t\in\Z^n}t\paren{G\cap T}}
The the geometric cloth of $G$ equals empirical probability distribution $P\paren{G, T}.$
\subsubsection{Poisson Point Process}
Let $\mathcal{C}^{0,n}$ and $\mathcal{C}_r^{0,n}$ and be the spaces of locally finite point collections in $\R^n$ and $B_r,$ respectively. A topological type $S_i$ of $\mathcal{C}_r^{0,n}$ is the subset consisting of all point collections with $i$ points. There is a natural map $\phi_i:S_i\rightarrow \text{Sym}^{i}B_r,$ and the Lebesgue measure on $B_r$ induces a probability measure $\nu_i$ on $S_i$ that selects $i$ points in the ball uniformly and independently.
The Poisson Point Process $\mu_\lambda$ is defined via marginal distributions $\mu_{\lambda,r}$ on $\mathcal{C}_r^{0,n}.$ $\mu_{\lambda,r}$ assigns each set $S_i$ probability $\frac{\lambda^n}{n!},$ and its condition distribution on each $S_i$ is $\nu_i.$
\subsubsection{Voronoi Graphs}
\label{ex_Voronoi}
Let $:V^n:\mathcal{C}^{0,n}\rightarrow\mathcal{G}^{n}$ be the function sending a locally finite point collection to the one-skeleton of the associated Voronoi tessellation. The probability distribution of Poisson-Voronoi graphs in $\R^n$ with intensity $\lambda$ is
\s{\mathcal{V}_\lambda^n=V^n_{*}\paren{\mu_\lambda}} The translation action of $\R^n$ is ergodic with respect to $\mathcal{V}_\lambda^n,$ so for $\mathcal{V}_\lambda^n$-almost all $G,$
\s{\mathcal{P}\paren{G}=\mathcal{V}_\lambda^n} $\mathcal{P}\paren{G}$ assigns probability zero to the orbit of $G$ under the $\R^n$ action. An example of a Voronoi graph in the plane is shown in Figure~\ref{fig:vor}.
\subsubsection{A Merged Configuration}
Let $\nu$ be the probability distribution on $\mathcal{C}^{0,2}$ given by sampling points from a Poisson point process with density one in the right half of the plane, and placing points in a fixed triangular lattice in the left half of the plane. Let $\hat{\nu}={V^2}_*\nu.$ $\hat{\nu}$ samples embedded graphs that look like Voronoi graph in the right half of the plane, and a hexagonal lattice in the left half. A finite region of such a graph is shown in~\ref{fig_halfHex2}.
Graphs sampled from $\hat{\nu}$ have a smooth geometric cloth with probability one, but that cloth does not equal $\hat{\nu}.$ Let $\mu_1$ be the probability distribution of Voronoi graphs in the planar with intensity 1, and let $\mu_2$ be the cloth of the hexagonal lattice in the left half of the plane. Then, with probability one the cloth of graph $G$ sampled from $\hat{\nu}$ is given by
\s{\mathcal{P}\paren{G}\paren{B}=\frac{\mu_1+\mu_2}{2}}
\subsubsection{Penrose tilings}
As is the case for Benjamini-Schramm graph limits~\cite{2012lovasz}, Penrose tilings provide an example of embedded graphs with interesting geometric cloths. A full analysis of their properties is beyond the scope of this paper but can be derived based on the work of de Bruijin~\cite{1981bruijin,1981bruijin2}.
\subsection{Local Geometric Convergence}
\label{sec_GeoConv}
We are now ready to define local geometric convergence.
\begin{Definition}
\label{defn_geoConv}
A sequence $\set{G_i}$ of embedded graphs in $\mathcal{G}^n$ is \textbf{locally geometrically convergent} if each $G_i$ has smooth geometric cloth, and $s^*\mathcal{P}_r\paren{G_i}$ converges weakly to a limit $\Xi_r$ for all $r\in \N.$
\end{Definition}
For example, if $G$ is an infinite graph with smooth geometric cloth, the sequence $G\cap B_i$ is locally geometrically convergent. It follows immediately from the definition that if $\set{G_i}\rightarrow \Xi$ in the local geometric sense and $f$ is a local geometric property then
\s{\lim_{i\rightarrow\infty}\widebar{f_{G_i}}=\int_{\mathcal{G}^n}f\;d\Xi}
Furthermore, if the $\R^n$ translation action is ergodic with respect to a geometric graph limit $\Xi,$ then $\Xi$-almost all $G$ have local geometric properties whose averages equal the limiting values of the sequence averages. We will consider the implications of this hypothesis in more detail in Section~\ref{sec_hom}
The smooth topology of topological types is not particularly nice nor natural. However, weak convergence of a uniformly separating sequence on $\mathcal{G}^n$ implies local geometric convergence:
\begin{theorem}
\label{thm_convergence}
Let $\set{G_i}$ be a uniformly separating sequence of graphs in $\mathcal{G}^n.$ The following are equivalent:
\begin{enumerate}
\item $\set{G_i}$ is locally geometrically convergent
\item $\set{s^*\mathcal{P}_r\paren{G_i}}$ is weakly convergent for all $r\in\N.$
\item $\mathcal{P}_r\paren{G_i}$ is weakly convergent for all $r\in \N.$
\item $\mathcal{P}\paren{G_i}$ is weakly convergent
\item The average values of all local geometric properties $f,$ $\widebar{f_{G_i}}$ converge.
\end{enumerate}
\end{theorem}
\begin{proof}
The second condition is the definition of local geometric convergence. The following three are equivalent because of Propositions~\ref{localProp} and~\ref{unifProp}. The last condition is equivalent to the first because a uniformly separating sequence is tight.
\end{proof}
The following question can be viewed as an analogue to the Aldous-Lyons conjecture~\cite{2012lovasz} in the local topological context:
\begin{question}
Is any translation-invariant probability distribution on $\mathcal{G}^n$ the geometric cloth of an embedded graph $G$? More specifically, does this hold if the distribution $\mu$ is the the local geometric limit of a sequence of embedded graphs $\set{G_i}$?
\end{question}
\subsection{Scale-free Convergence}
\label{sec_scaleFree}
The concept of local geometric convergence is too strong for curvature flow on graphs - computational simulations indicate that scale-dependent properties of change as graphs evolve. To define a notion of scale-free convergence, we need a definition of a characteristic length scale:
\begin{Definition}
Let $\rho_{x,\alpha}:\mathcal{G}^n\rightarrow \mathcal{G}^n$ be the map induced the the dilation of $\R^n$ centered at $x$ by a factor $\alpha.$ A \textbf{characteristic length scale} of graphs is a continuous function $g:Y\rightarrow\R^+$ defined on a dilation-invariant subset $Y$ of $\mathcal{G}^n$ that scales with dilations:
\s{g\circ\rho_{x,\alpha}=\alpha g \;\;\;\forall \alpha\in\R^+,x\in\R^n}
An embedded graph $G$ has a \textbf{well-defined characteristic length scale} $\widebar{g}_G$ if $g$ is a characteristic length scale and
\s{0<\widebar{g}=\lim_{r\rightarrow\infty} \int_{\mathcal{G}^{n}}g\;P\paren{G,B_r}=\int_{\mathcal{G}^{n,k}}g\;\mathcal{P}\paren{G}<\infty}
\end{Definition}
A characteristic length scale is necessarily unbounded, so we will need to impose some hypotheses on an infinite embedded graph to ensure it has a well-defined characteristic lengthscale. For example, the the function that assigns to a graph the infimal $r$ so that the probability that a ball of radius $r$ fully contains an edge of $G$ is a well-defined characteristic lengthscale for graphs with smooth geometric cloth whose maximum edge length is bounded above.
\begin{Definition}
Suppose $\set{G_i}$ is a sequence of embedded graphs in $\mathcal{G}^n$ which share a characteristic length-scale $g.$ Then $\set{G_i}$ converges in the \textbf{scale-free local geometric sense} if the sequence of rescaled graphs
\s{\set{\rho_{\frac{1}{\widebar{g}_{G_i}} G_i}}}
is locally geometrically convergent, where $\rho_\alpha$ is the dilation of $\R^n$ by $\alpha$ centered at the origin. A \textbf{scale-free local geometric property} of cell complexes is a local geometric property that is dilation-invariant.
\end{Definition}
\subsection{Relation to Local Topological Convergence}
\label{sec_geoAndBS}
Local topological properties are not in general local geometric properties, for two reasons: small topological neighborhoods may be arbitrarily large geometric ones, and local topological properties are normalized by topological quantities rather than by volume. However, one may place hypotheses on the tail of the edge length distribution and the variation of vertex density under which local geometric convergence implies local topological convergence.
There are several ways to define a topological cloth of an infinite embedded graph, given certain consistency assumptions. These notions coincide for finite graphs. We consider one such definition here, but very similar statements can be made about others. Let $G\in\mathcal{G}^n$ be an embedded graph, and let $I_r\paren{G}$ be the graph consisting of all edges and vertices of $G$ fully contained in $B_r.$ Recall that $G$ is said to have an inner regular topological cloth if $\set{I_r\paren{G}}$ is locally topologically convergent.
Let $e\paren{G,s,M}$ be the percentage of edges in $I_s\paren{G}$ of length greater than $M.$ Also, define the vertex density function $\rho\paren{G,s,r,M}$ to be the percentage of vertices $v\in I_s\paren{G}$ so that $B_r\paren{v}$ contains more than $M$ vertices of $G.$
\begin{proposition}
Let $\set{G_j}$ be a locally geometrically convergent sequence of graphs in $\mathcal{G}^n,$ whose vertex degrees are bounded above by $D$ (with probability one with respect to local distributions $\mathcal{P}_r\paren{G_j}$). Also, let $V_j\paren{B_r}$ be the vertex set of $I_r\paren{G_j}.$
Assume there is a constant $C_1>0$ and bounded, real-valued functions $M\paren{\epsilon}$ and $N\paren{\epsilon}$ so that for all $\epsilon>0$ and all sufficiently large $s$ and $j$
\begin{enumerate}
\item $\abs{\frac{\abs{V_j\paren{B_s}}}{\vol{B_s}}-C_1}<\epsilon$
\item $e\paren{G_j,s,M\paren{\epsilon}}<\epsilon$
\item $\rho\paren{G_j,s,r,N\paren{\epsilon}}<\epsilon$
\end{enumerate}then each $G_j$ has an inner-regular topological cloth and
\s{\lim_{j\rightarrow\infty} \lim_{r\rightarrow\infty}I\paren{G_j,B_r}}
converges in the Benjamini-Schramm sense.
\end{proposition}The proof is long and technical, but straightforward, and we omit it here.
\section{Homogeneous Graphs}
\label{sec_hom}
A homogeneous graph is an embedded graph with smooth geometric cloth for which the translation action on $\R^n$ is ergodic. This will be the key hypothesis for the universality conjectures we propose for curvature flow on graphs in Section~\ref{sec_conj}. Here, we study the properties of homogeneous graphs. In particular, we show that the local geometric properties of a homogeneous graph have well-defined averages over a large class of sequences of increasing subsets of $\R^n.$ Note that a homogenous graph is necessarily infinite.
\subsection{Preliminaries: Ergodic Theory}
\begin{Definition}
Let $\paren{X,\textit{F},\nu}$ be a measure space, and suppose a group $H$ acts on $X.$ The action of $H$ is \textbf{measure-preserving} if $\nu\paren{hB}=\nu\paren{B}$ for all $B\in\textit{F}$ and all $h\in H.$ In this case $\nu$ is said to be \textbf{$H$-invariant}. The action is \textbf{ergodic} if every subset of $X$ such that $hX=X\;\forall h\in H$ has measure 0 or 1.
\end{Definition}We require a generalized version of the Pointwise Ergodic Theorem due to Lindenstrauss~\cite{2001lindenstrauss}.
\begin{theorem}
(Lindenstrauss)
Let $T$ be a locally compact, amenable group acting on a measure space $\paren{X,\textit{F},\nu}$ with left-invariant Haar measure $\mu,$ and suppose $\set{F_n}$ is a sequence of compact subsets of $T$ that is
\begin{enumerate}
\item F{\o}lner: for all $t\in T$
\s{\lim_{i\rightarrow \infty} \frac{\mu\paren{tF_i\triangle F_i}}{\mu\paren{F_i}}=0}
where $\triangle$ denotes the symmetric difference.
\item Tempered: There exists a $C>0$ such that for all $n,$
\s{\mu\paren{\cup_{k\leq n} F_k^{-1}F_{n+1}}\leq C\mu\paren{F_{n+1}}}
\item \s{\abs{F_n}\geq n}
\end{enumerate}
If the action of $H$ is measure-preserving, then for any $f\in L^1\paren{\nu}$ there is a well-defined, $H$-invariant average $\widebar{f}\in L^1\paren{\nu}$ so that for $\nu$-almost all $x$
\s{\lim_{i\rightarrow\infty}\frac{1}{\mu\paren{F_i}}\int_{F_i}f\paren{gx}\;d\mu\paren{g}=\widebar{f}\paren{x}}
In addition, if the action is ergodic
\s{\widebar{f}\paren{x}=\int_{X}f\;d\nu\paren{x}}
\end{theorem}
\subsection{Homogeneous Graphs}
\begin{definition}
\label{defn_homo}
An embedded graph $G\in\mathcal{G}^n$ is \textbf{homogeneous} if it has a smooth geometric cloth, and the translation action of $\R^n$ is ergodic with respect to $\mathcal{P}\paren{G}.$
\end{definition} The following is an immediate consequence of the definition of ergodicity:
\begin{proposition}
If $G$ is homogeneous and $f$ is a local geometric property then $\mathcal{P}\paren{G}$-almost all $H$ have $\widebar{f_G}=\widebar{f_H}.$
\end{proposition} That is, the geometric cloth of a homogeneous graph $G$ is a probability distribution sampling graphs with the same local geometric properties as $G.$ Note the resemblance to the topological cloth.
One might hope that if $f$ is a local geometric property and $G$ is a homogeneous graph, then $f$ would have a well-defined average over any tempered F{\o}lner sequence. A counterexample is presented in Section~\ref{sec_counterexample}. However, the local properties of homogeneous graphs have well-defined averages over a smaller set of averaging sequences:
\begin{Definition}
An increasing sequence $\set{F_i}_{i\in\N}$ of compact, convex sets in $\R^n$ is an \textbf{admissible averaging sequence} if
\begin{enumerate}
\item$\lim_{i\rightarrow\infty}\vol{F_i}=\infty$
\item $\limsup_{i\rightarrow\infty}\frac{\vol{B\paren{F_i}}}{\vol{F_i}}<\infty$
where where $B\paren{F_i}$ is the smallest ball centered at the origin containing $F_i.$
\end{enumerate}
\end{Definition}
\begin{proposition}
Let $X$ be a completely regular space with $\R^n$ action. For $x\in X$ and any bounded subset $K$ of $\R^n,$ define the empirical measure $P\paren{x,K}$ to the pushforward of the normalized Lebesgue measure on $K$ via the map $\phi_x:\R^n\rightarrow X$ given by $\phi_x\paren{a}=ax.$ If $\lim_{r\rightarrow\infty}P\paren{X,B_r}$ converges weakly to a probability distribution $\sigma$ for which the $\R^n$ action is ergodic, then $\lim_{i\rightarrow\infty}P\paren{X,F_i}$ converges weakly to the same limit for any admissible averaging sequence $\set{F_i}.$
\end{proposition}
\begin{proof}
Let $\set{F_i}$ be an admissible averaging sequence, and $\mu$ be the standard Lebesgue measure on $\R^n.$ Also, let $f:X\rightarrow\R$ be a bounded continuous function, and let $\bar{f}$ be the expected value of $f$ for $x\in X,$ taken over a sequence of balls centered at the origin. We will show that $f$ has the same expected value for $f$ for any admissible averaging sequence. To do so, we will replace $f$ with an averaging function $f_r$:
\s{f_r\paren{x}=\frac{1}{\vol{B_r}}\int_{B_r}f\paren{ax}\;d\mu\paren{a}.}
Also, let
\s{\hat{F}_i^r=\set{a\in F_i : B_r\paren{a}\subset F_i}}
Then
\begin{align*}
\int_{F_i}f_r\paren{ax}\;d\mu\paren{a}&\\
&=\frac{1}{\vol{B_r}}\int_{F_i}\int_{B_r}f\paren{abx}\;d\mu\paren{b}\;d\mu\paren{a}\\
&=\frac{1}{\vol{B_r}}\int_{B_r}\int_{\hat{F}_i^r}f\paren{abx}\;d\mu\paren{a}d\mu\paren{b}+\int_{F_i-\hat{F}_i^r}f_r\paren{ax}\;d\mu\paren{a}\\
&=\int_{\hat{F}_i^r}f\paren{ax}\;d\mu\paren{a}+O\paren{\text{vol}_{n-1}\paren{\partial F_i}}\\
&=\int_{F_i}f\paren{ax}\;d\mu\paren{a}+O\paren{\text{vol}_{n-1}\paren{\partial F_i}}
\end{align*}
because $f$ is bounded and the volume of $F_i-\hat{F}_i^r$ is $O\paren{\text{vol}_{n-1}\paren{\partial F_i}}.$ Therefore if the error of approximating $f$ by $f_r$ is denoted
\s{\text{err}\paren{F_i,r}=\frac{1}{\vol{F_i}}\abs{\int_{F_i}f_r\paren{ay}\;d\mu\paren{a}-\int_{F_i}f\paren{y}},} then $text{err}\paren{F_i,r}$ goes to zero as $i\rightarrow\infty.$
For any $\epsilon>0,$ define a sequence of open subsets of $X:$
\s{A_{s,\epsilon}=\set{x\in X:\abs{\frac{1}{\vol{B_s}}\int_{B_s}f\paren{ax}\;d\mu\paren{a}-\widebar{f}}<\epsilon}}
Ergodicity implies that for any $\delta>0$ and sufficiently large $s,$
\s{\sigma\paren{A_{s,\epsilon}\paren{f}}>1-\delta} Let $\epsilon>0$ and choose $s$ such that $\sigma\paren{A_{s,\epsilon}}>1-\epsilon/2.$ Because $\mu_i$ converges weakly to $\mu$ as $r\rightarrow\infty$, the Portmanteau lemma implies that for all sufficiently large $r,$
\s{P\paren{x,B_r}\paren{A_{s,\epsilon}}>1-\epsilon}
The second hypothesis in the definition of an admissible averaging sequence implies that
\s{\limsup_{i\rightarrow\infty}{\frac{\text{vol}_{n-1}\paren{\partial F_i}}{\text{vol}_n\paren{F_i}}}<\infty}
via the Blaschke selection theorem~\cite{2007gruber}. Thus, for sufficiently large $i,$
\s{P\paren{x,F_i}\paren{A_{s,\epsilon}}>1-\epsilon}It follows that there is an $I_1$ so that for all $i>I_1$ we can write $F_i=E_i\cup D_i$ where $E_{i}\subset A_{s,\epsilon}$ and $\sup\abs{f}\mu\paren{D_i}<\epsilon\mu\paren{F_i}$ This also implies $\frac{\mu\paren{E_i}}{\mu\paren{F_i}}>1-\frac{\epsilon}{\sup\abs{f}}$.
Let $M=\sup\abs{f}$, and choose $I_2$ so that $\text{err}\paren{F_i,s}<\epsilon$ for all $i>I_2.$ Then for all $i>\max\paren{I_1,I_2},$
\begin{align*}
\abs{\frac{1}{\vol{F_i}}\int_{F_i}f\paren{ax}\;d\mu\paren{a}-\widebar{f}}\\
&=\abs{\frac{1}{\vol{F_i}}\int_{F_i}f_s\paren{ay}\;d\mu\paren{a}+\text{err}\paren{F_i,s}-\widebar{f}}\\
&\leq\abs{\frac{1}{\vol{F_i}}\big(\int_{E_i}f_s\paren{ay}\;d\mu\paren{a}+\int_{D_i}f_s\paren{ay}\;d\mu\paren{a}\big)-\widebar{f}}+\epsilon\\
&\leq \abs{\frac{\vol{E_i}}{\vol{F_i}}\frac{1}{\vol{E_i}}\int_{E_i}f_s\paren{ay}\;d\mu\paren{a}-\widebar{f}}+M\frac{\mu\paren{D_i}}{\mu\paren{F_i}}+\epsilon\\
&\leq\abs{\frac{1}{\vol{E_i}}\int_{E_i}f_s\paren{ay}\;d\mu\paren{a}-\widebar{f}}+\frac{\epsilon}{M\vol{E_i}}\abs{\int_{E_i}f\paren{ay}\;d\mu\paren{a}}+2\epsilon\\
&\leq4\epsilon
\end{align*} So $P\paren{x,F_i}\rightarrow\sigma,$ weakly, as desired.
\end{proof} The following is an immediate consequence:
\begin{theorem}
If $G$ is a homogeneous graph then then $s^*\mathcal{P}\paren{G,F_i}$ converges weakly to $\mathcal{P}\paren{G}$ for any admissible averaging sequenced. Furthermore, $\mathcal{P}\paren{G}$-almost all graphs have local geometric properties whose averages are equal to those of $G,$ and are well-defined over any tempered F{\o}lner sequence.
\end{theorem}
\subsection{A Counterexample}
\label{sec_counterexample}
\begin{figure}[ht]
\center
\includegraphics[width=.4\textwidth]{/CounterexamplePoisson.pdf}
\caption{\label{fig_counterex} A Poisson distributed point collection, with points from the region $M$ removed.}
\end{figure}
We will provide an example of a homogeneous locally finite point collection whose local properties do not have well-defined averages over all tempered F{\o}lner sequences. The Voronoi diagram of such a point collection will also share that property. Note that the concepts of geometric cloth and homogeneity extend automatically to locally finite point collections.
Let $\mu$ be the Poisson Point Process with intensity one on $\R^2.$ Let $M\subset\R^2$ be the region bounded by the inequality
\s{\text{log}\paren{\abs{x}+1}\leq\abs{y}}
Also for any $w\in Y,$ let $\hat{w}$ be the locally finite point collection obtained from $w$ by deleting all points contained in $M.$ An example is shown in Figure~\ref{fig_counterex}. The volume fraction
\s{\frac{\vol{M\cap B_r}}{\vol{B_r}}\leq\frac{4r\text{log}\paren{r+1}}{\pi r^2}}
goes to zero as $r\rightarrow\infty.$ It follows from the pointwise ergodic theorem that for $\mu$-almost all $w,$ $\hat{w}$ has a well-defined geometric cloth.
Let $E_r$ be the rectangle centered at the origin bounded by the lines $x=\pm r$ and $y=\pm\text{log}\paren{r+1}.$ $E_r$ is a nested sequence of convex sets whose inradius goes to $\infty,$ so it is a tempered F{\o}lner sequence. However, the volume ratio
\s{\frac{\vol{M\cap E_r}}{\vol{E_r}}=\frac{4\int_{0}^{r}\text{log}\paren{s+1}\;d\mu\paren{s}}{4r\text{log}\paren{r}}=\frac{4\paren{\paren{r+1}\text{log}\paren{r+1}-r}}{4r\text{log}\paren{r+1}}}
goes to one as $r\rightarrow\infty,$ so $P\paren{\hat{w},E_r}$ cannot converge weakly to $\mu.$
\section{The Universality Conjectures}
\label{sec_conj}
We use the concepts of Local Geometric Convergence and Local Topological Convergence to make several onjectures about the long-term behavior of graphs evolving by the network flow of curvature flow on embedded graphs in $\R^n.$
\begin{conj}[Universality Conjecture for Local Topological Convergence]
\label{steady_state_hypothesisb1}
There exists a probability distribution $\sigma_n$ on the space of countable, connected graphs with a root vertex specified such that any network flow $G\paren{t}$ with homogeneous initial condition $G\paren{0}\in\mathcal{G}^n$ converges in the local topological sense to $\sigma_{\Omega}$ or to a stationary state as $t\rightarrow\infty.$
\end{conj} Note that $G\paren{t}$ might converge topologically to a stationary state in a local sense without ever being stationary itself. For example, the flow starting with an infinite regular hexagonal lattice with a single edge removed is never stationary, but always has the the local statistics of a stationary state. Assuming the previous conjecture, let $\sigma_{n}$ be the universal probability distribution on graphs in $\R^n.$ We conjecture all that all embedded graphs with homogeneous topological cloth, as defined below, converge to $\sigma_{n}$ for some $n.$ This allows, for example, an embedded graph in $\R^3$ to converge to the universal distribution for planar embedded graphs.
\begin{Definition}
\label{defn_tophomo}
$G\in\mathcal{G}^n$ has a \textbf{homogeneous topological cloth} if there is a probability distribution $\mathcal{S}\paren{G}$ on $\mathcal{G}^\bullet$ such that the sequences $I\paren{C,G_i}$ and $O\paren{C,G_i}$ converge in the Benjamini-Schramm sense to $\mathcal{S}\paren{G}$ for every admissible averaging sequence $F_i.$
\end{Definition}
\begin{conj}[Classification of Orbits]
\label{steady_state_hypothesis10}
Suppose $G\in\mathcal{G}^n$ is an embedded graph with a homogeneous topological cloth and a well-defined time evolution under curvature flow $G\paren{t}.$ Then either $G\paren{t}$ converges topologically to $\sigma_{m}$ for some $m\leq n,$ or it converges to a stationary state.
\end{conj} We would not be surprised if $\sigma_{n}=\sigma_{m}$ for all $m,n\geq 3.$
\begin{conj}[Universality Conjecture for Local Geometric Convergence]
\label{steady_state_hypothesisa1}
There exists a probability distribution $\Xi_n$ on $\mathcal{G}^n$ such that that any network flow $G\paren{t}$ with homogeneous initial condition $G\paren{0}\in\mathcal{G}^n$converges in the scale-free local geometric sense to $\Xi_{n}$ or a stationary state as $t\rightarrow\infty.$
\end{conj} It is conjectured that curvature flow on graphs never results in an edge with multiplicity greater than one~\cite{2014ilmanen}. Under this assumption, the corresponding universality conjecture in terms of local geometric convergence of graphs with multiplicity in $\check{\mathcal{G}}^n$ would be equivalent to the above.
\subsection{Statement in Terms of Invariant Measures}
We defined local geometric convergence in terms of individual graphs in order to find an analogue of Benjamini-Schramm graph limits for embedded graphs, and to develop a language that would be amenable to computation. However, a perhaps more traditional way to state a universality conjecture about the behavior of graphs evolving under curvature flow is in terms of invariant measures:
\begin{conj}
\label{steady_state_hypothesis_ergodic}
Let $\hat{\Omega}$ be curvature-flow on embedded graphs in $\R^n,$ rescaled to keep a characteristic length scale equal to one. There exists a unique $\hat{\Omega}$-invariant, homogeneous probability measure $\nu$ on $\mathcal{G}^n$ such that $\nu$-almost all graphs do not evolve to a stationary state under $\hat{\Omega}.$ Furthermore, $\hat{\Omega}$ is ergodic with respect to $\nu.$
Such a measure $\nu$ can be obtained by choosing any homogeneous initial condition $G\in\mathcal{G}^n$ that does not converge to a stationary state, and taking the limit of $\mathcal{P}\paren{G\paren{t}}$ as $t\rightarrow\infty.$
\end{conj}
\section{Computational Evidence}
\label{sec_data}
\label{sec_comp}
\label{sec_computation}
We test the universality conjectures stated in the previous section for embedded graphs in two dimensions We simulated curvature flow of graphs embedded in $\mathbb{T}^2$ using Jeremy Mason's implementation of the algorithm developed by Lazar et al in~\cite{2010lazar}. It is based on the von Neumann - Mullins Relation, which gives the rate of change of the area of a component of the complement in terms of the number of edges on its boundary.
We use two kinds of random initial conditions for our simulations. The first is the Voronoi diagram of Poisson distributed points, as depicted in Figure~\ref{fig:vor}. The other, shown in \ref{perturbed_lattice} is a perturbed hexagonal lattice. It is generated by perturbing the vertices of the dual triangular lattice, then computing a Voronoi diagram.
\subsection{Testing for Local Topological Convergence}
\label{sec_swatchComp}
\label{sec_computations}
\begin{figure}
\center
\includegraphics[height=7cm]{/distanceData2D_bars.pdf}
\caption{\label{fig_2D1}Distances of a simulations starting with both initial conditions to the candidate universal state. Error bars show the standard deviation of the distance of a representative subsample of the candidate state to the candidate state.}
\end{figure}
We test the for local topological convergence by tracking the distance $d_r$ of a simulated graph throughout its evolution to a candidate universal state. The candidate universal state is taken from a large simulation for which all measured properties have converged. We sample swatches of a given radius at each vertex, and use the program \textit{Nauty} to classify them up to graph isomorphisms~\cite{nauty}.
Separate implementations of these algorithms were programmed by both the author and Jeremy Mason. The two programs work for general regular cell complexes, not just those of dimension one. More details on the implementation is included in~\cite{2016schweinhart}.
Figure~\ref{fig_2D1} shows the distance from from two simulations - one starting from a Voronoi graph and the other a perturbed hexagonal lattice - to a large candidate universal state. The candidate universal state was taken from a different simulation starting with a Voronoi graph with $6.0 \times 10^7$ edges. The evolution is parametrized by the number of edges in the graph, a decreasing function of time. Thus, these figures are read from right to left. All simulations start at initial conditions far from the candidate state, before decreasing toward the universal state. Finite size effects cause the distances to increase toward the end of the simulation.
The distance from both simulations to the candidate universal state decreases very rapidly. However, in none of these systems does the distance to the candidate universal state go to zero. To quantify the finite size effects, we constructed representative subsamples of the candidate universal state. The error bars in Figure~\ref{fig_2D1} show one standard deviation above and below the mean of the distance of a representative subsample of a certain size to the candidate universal state. The two simulations are within the bars from between 100,000 and 1000 edges, indicating that the deviation of the distance from zero is explainable due to finite size effects.
\cite{2016schweinhart} contains a more detailed analysis of these results, as well as similar ones for the network flow in three dimensions.
\subsection{Testing for Local Geometric Convergence}
\begin{figure}
\center
\subfloat[]{%
\label{HEMD_1}{%
\includegraphics[width=.45\textwidth]{%
HausdorffEMD_Voronoi_1.jpg}}}
\hspace{20pt}
\subfloat[]{%
\label{HEMD_2}{%
\includegraphics[width=.45\textwidth]{%
HausdorffEMD_Voronoi_2.jpg}}}
\caption{\label{Fig_HEMD}Plots showing the local Hausdorff EMD from a simulation to a candidate universal state with (a) $r=1$ and (b) $r=2.$ Data from perturbed lattice initial conditions will also be included in a final draft.}
\end{figure}
To test for local geometric convergence, we rescale a sequence of graphs $\set{G_i}$ so that the average edge length of each equals one, and test the weak convergence and uniform separability of $\mathcal{P}_r\paren{G_i}$ for several values of $r.$
To check weak convergence, we estimate the earth mover's distance (first Wasserstein distance) induced by the local Hausdorff distance. We sample $10^4$ random $r$-balls in each $G_i$ and a candidate universal state $G,$ compute a distance matrix in $d_L$ between these samples, and use the transport package in $R$~\cite{transport} to determine the earth mover's distance. Plots of this distance for $r=1$ and $r=2$ are shown in Figure~\ref{Fig_HEMD}. As before, the number of edges decreases with time, so the plots are meant to be read from left to right. In each, the distance to the candidate state decreases rapidly. Note that this convergence is faster than in Figure~\ref{fig_2D1}, indicating that the local properties at small radius converge much faster than those at larger radii.
For $G\in\mathcal{G}_r^2,$ let $\Delta\paren{G}$ be the supremal $\delta$ so that $G$ is $\delta$-thickenable. Note that the curvature of $G$ is bounded by $\Delta\paren{G}^{-1},$ by the definition of $\delta$-thickenability. Figure~\ref{fig_thicken} depicts a graph $H\in\mathcal{G}_2^2$ sampled from the candidate universal state for which $\Delta\paren{H}=.266.$
To check uniform separation, we computed the probability distribution $\Delta$ for $10^5$ random $r-$balls in $G_i.$ Note that the conditions in the definition of $\delta$-thickenability have direct analogues for discretized graphs. For example, we use the discrete global radius of curvature~\cite{1999gonzales} as a stand-in for the maximum $\epsilon$ so that an edge has an $\epsilon$-tube. To study the tail of $\mu_i,$ we plotted percentile curves of these distributions as a function of the number of edges. This is shown in Figure~\ref{Fig_thickenability_check} for $r=1.$ Each plot appear to approach to limiting values in the mid-range of the evolution, and are bounded uniformly away from zero. They are certainly uniformly bounded away from $0.$ This supports the hypothesis the sequence is uniformly separated. Interestingly, the percentage of samples of the candidate universal state where each of the following was the limiting factor for $\Delta$ is approximately 89.4\% for the minimum distance of an edge or vertex to the boundary, 5.2\% for the minimal separation between vertices, 2.2\% for the global radius of curvature, and .6\% for the distance between edges outside of a neighborhood of the vertices. The remaining 2.6\% of the samples were empty. This is unsurprising because edges meet at vertices at angles very close to $2\pi/3,$ and the grains in the candidate universal state appear to be quite ``round'' (as can be seen in Figure~\ref{ss_2D}).
\begin{figure}
\center
\includegraphics[height=7cm]{/ThickenablePercentile.pdf}
\caption{\label{Fig_thickenability_check}Percentile values of $\Delta\paren{G}$ for a graph evolving by curvature flow.}
\end{figure}
We used the the spatial searching module of CGAL to improve the speed of both the computation of the local Hausdorff distance and of $\Delta\paren{G}$~\cite{CGAL,CGALspatial}. In the future, we hope to further optimize computation of the Hausdorff earth mover's distance by taking $d_L$ mod rotations.
\section{Conclusion}
Curvature flow on graphs and cell complexes is a process of great mathematical interest, with important physical applications in materials science. We defined notions of local topological convergence and local geometric convergence for embedded graphs in $\R^n,$ and studied their properties. Using these concepts, we stated several universality conjectures for curvature flow on graphs, formalizing observations from experiments and computer simulations of these systems.
It is our hope that this paper will stimulate interest in these problems, eventually leading to a mathematically rigorous understanding of this beautiful phenomenon.
\section*{Acknowledgments}
The author would like to thank Robert MacPherson and Jeremy Mason and in particular for discussions that led to the ideas in this paper. The method of swatches was originally developed in collaboration with them in~\cite{2016schweinhart}. Professor MacPherson was the author's PhD advisor, and suggested curvature flow on graphs as a system of interest. The author would also like to thank Tom Ilmanen, Ryan Peckner, and Philippe Sosoe for interesting and useful discussions.
This research was supported by an NSF Mathematical Sciences Postdoctoral Research Fellowship and, before that, the Center of Mathematical Sciences and Applications at Harvard University.
|
\section{Preliminaries}
In our work, we mainly focus on the learning problem in two stages, i.e., C-stage and E-stage with feature and instance evolution simultaneously. In C-stage, data are collected in mini-batch style. Assume that there are totally $T_1$ batches. In each batch, the features of each instance can be divided into two parts. The first part contains the features which will vanish in E-stage and the other part consists of the features which survive for both stages. They are named as \emph{vanished feature} and \emph{survived feature}. In E-stage, we assume that there are two batches of data. One batch is employed for training and the other one is used for testing. The features of each instance in this stage can also be divided into two parts. The first part contains the features that are survived in both stages. The second part is augmented features, which is referred as \emph{augmented feature} in the following.
\begin{figure}[!t]
\centering
\label{fig:1}
\includegraphics[width=5.in]{notation.pdf}
\caption{Notations.}
\vskip -0.2in
\end{figure}
Formally, as pictured in Fig. 1, in the $i$-batch of C-stage, data points can be represented by two matrices, i.e., $\mathbf{X}_i^{\textrm{(v)}} \in \mathbb{R}^{n_i\times d^{\textrm{(v)}}}$ and $\mathbf{X}_i^{\textrm{(s)}} \in \mathbb{R}^{n_i\times d^{\textrm{(s)}}}$, where $n_i$ is the number of points in this batch, $d^{\textrm{(v)}}$ and $d^{\textrm{(s)}}$ are the numbers of vanished features and survived features respectively. Here, the superscripts "$\textrm{(v)}$" and "$\textrm{(s)}$" correspond to vanished features and survived features. The $j$-th row of $\mathbf{X}_i^{\textrm{(v)}}$ is an instance with only vanished features. Correspondingly, the $j$-th row of $\mathbf{X}_i^{\textrm{(s)}}$ is an instance with only survived features. The label matrix of instances in the $i$-batch is denoted by $\mathbf{Y}_i \in \mathbb{R}^{n_i \times c}$ with $c$ as the number of class. Its $(k,l)$-element $Y_i(k,l)=1$ if and only if the $k$-th instance in the $i$-batch belongs to the $l$-th category and $Y_i(k,l)=0$ otherwise.
In E-stage, it is often that we want to make prediction when we only get one batch training data. Thus, there are only two batches in this stage. The first batch contains training data, represented by $\mathbf{X}_{T_1+1}^{\textrm{(s)}} \in \mathbb{R}^{n_{T_1+1} \times d^{\textrm{(s)}}}$ and $\mathbf{X}_{T_1+1}^{\textrm{(a)}} \in \mathbb{R}^{n_{T_1+1} \times d^{\textrm{(a)}}}$, where $n_{T_1+1}$ is the number of training points in this stage, $d^{\textrm{(a)}}$ is the numbers of augmented features and the superscript "$\textrm{(a)}$" corresponds to augmented features. Similarly, the $j$-th row of $\mathbf{X}_{T_1+1}^{\textrm{(s)}}$ consists of survived features and the $j$-th row of $\mathbf{X}_{T_1+1}^{\textrm{(a)}}$ contains augmented features. The label matrix is denoted by $\mathbf{Y}_{T_1+1}$. Similarly, the second batch contains testing points with the same structure as training.
According to above notations, \textbf{our main task} is to classify data points represented by $\bar{\mathbf{X}}_{T_1+2} \triangleq [\mathbf{X}_{T_1+2}^{\textrm{(s)}}, \mathbf{X}_{T_1+2}^{\textrm{(a)}}]$, based on $\{ \bar{\mathbf{X}}_{T_1+1} \triangleq [\mathbf{X}_{T_1+1}^{\textrm{(s)}}, \mathbf{X}_{T_1+1}^{\textrm{(a)}}],~\mathbf{Y}_{T_1+1}\}$. Besides, we can also use models learned on $\{\tilde{\mathbf{X}}_i \triangleq [\mathbf{X}_i^{\textrm{(v)}}, \mathbf{X}_i^{\textrm{(s)}}],~\mathbf{Y}_{i}\}_{i=1}^{T_1}$, without saving all the data. In the $T$-th step, we can only access the data $\{\tilde{\mathbf{X}}_T, ~\mathbf{Y}_T\}$, without saving the past data $\{\tilde{\mathbf{X}}_i, ~\mathbf{Y}_{i}\}_{i=1}^{T-1}$. Here, \emph{a tilde above the symbol represents variable in C-stage and a bar represents variable in E-stage}.
It is noteworthy to mention that we can also use $\{\bar{\mathbf{X}}_{T_1+1}, \mathbf{Y}_{T_1+1}\}$ merely to train a classifier in E-stage. Nevertheless, in many real applications, since $n_{T_1+1}$ is often comparable to that of $n_{i}$ for $i=1,2,\cdots,n_{T_1}$ and they are often small due to the limitation of storage, training with only the instance in E-stage trends to be over-fitting. It is better to learn a classifier with the assistance of model trained in C-stage.
\section{The OPID Approach}
We investigate a real application problem with complicated settings, and it is difficult to use traditional approaches to solve this problem directly. There are two stages and both the instances and features are changed. In our paper, we tackle this problem in following way.
~~1. In C-stage, we learn a classifier based on $\{\tilde{\mathbf{X}}_i, \mathbf{Y}_{i}\}_{i=1}^{T_1}$ in one-pass way. That is, we only access training examples once. Besides, the learned classifier should also provide useful classification information to help the training in E-stage.
~~2. In E-stage, we learn a classifier based on $\{\bar{\mathbf{X}}_{T_1+1}, \mathbf{Y}_{T_1+1}\}$, under the supervision of classifier learned in C-stage.
\textbf{C-stage}: In this stage, there are two different kinds of features, i.e., vanished feature and survived feature. If we combine them to learn a unified classifier, it will be difficult to use it in E-stage, since the features are different in two stages. Notice that, there are features surviving in both stages. It is better to compress important information of vanished features into functions of survival features. In other words, we want to use the model trained in survived features to represent important information of both vanished features and survived features.
Let $\tilde{\mathcal{H}}$ and $\mathcal{H}^{\textrm{(s)}}$ be the function space for all features (vanished features and survived features) and the survived features respectively. In C-stage, we try to learn two classifiers $ \tilde{h} \in \tilde{\mathcal{H}}$ and $ \tilde{h}^{\textrm{(s)}} \in \mathcal{H}^{\textrm{(s)}}$ with some consistency constraints between them. Denote the loss function on all features and survived features as $\tilde{\ell}$ and $\tilde{\ell}^{\textrm{(s)}}$, the expected classifiers in C-stage can be learned by optimizing
{
\begin{equation}
\label{eq1}
\begin{split}
\min_{\tilde{h} \in \mathcal{H}, \tilde{h}^{\textrm{(s)}} \in \mathcal{H}^{\textrm{(s)}} }~~& \sum_{i=1}^{T_1}
\tilde{\ell} \left( \tilde{h} (\tilde{\mathbf{X}}_i), \mathbf{Y}_i \right)+\tilde{\ell}^{\textrm{(s)}}\left( \tilde{h}^{\textrm{(s)}}(\mathbf{X}_i^{\textrm{(s)}}), \mathbf{Y}_i \right) \\
\textrm{s.t.}~~ & \mathcal{D}\left( \tilde{h}(\tilde{\mathbf{X}}_i) , \tilde{h}^{\textrm{(s)}}(\mathbf{X}_i^{\textrm{(s)}}) \right) \leq \epsilon. \textrm{~~for~~} i\in[T_1]~.\\
\end{split}
\end{equation}}
Here $\tilde{\mathbf{X}}_i = [\mathbf{X}_i^{\textrm{(v)}}, \mathbf{X}_i^{\textrm{(s)}}]$. $\mathcal{D}$ is employed to measure the consistency between two classifiers on every batch. We denote $[T_1]=\{1,2,\cdots,T_1\}$ for the convenience of presentation.
For example, we assume $\tilde{h}$ and $\tilde{h}^{\textrm{(s)}}$ are two linear classifiers with square loss. Besides, the Frobenius norm (denoted by $\|\cdot\|$) is employed as the measurement of consistency. The optimization problem in Eq. (\ref{eq1}) becomes
{
\begin{equation}
\label{eq2}
\begin{split}
\min_{\tilde{\mathbf{W}}, \mathbf{W}^{\textrm{(s)}}}~~& \sum_{i=1}^{T_1} \| \langle \tilde{\mathbf{W}}, \tilde{\mathbf{X}}_{i} \rangle - \mathbf{Y}_{i} \|^2 + \sum_{i=1}^{T_1} \| \langle \mathbf{W}^{\textrm{(s)}}, \mathbf{X}_{i}^{\textrm{(s)}} \rangle- \mathbf{Y}_{i} \|^2 \\
&+\lambda \sum_{i=1}^{T_1} \| \langle \tilde{\mathbf{W}}, \tilde{\mathbf{X}}_{i} \rangle - \langle \mathbf{W}^{\textrm{(s)}}, \mathbf{X}_{i}^{\textrm{(s)}} \rangle \|^2+
\rho(\| \tilde{\mathbf{W}} \|^2 + \| \mathbf{W}^{\textrm{(s)}}\|^2),
\end{split}
\end{equation}}
where $\tilde{\mathbf{W}}$ and $\mathbf{W}^{\textrm{(s)}}$ are classifier coefficients defined on all features and survived features in C-stage respectively, $\lambda>0$ is the parameter to tune the importance of consistency constraint and $\rho>0$ is the parameter for regularization. It is an extension of traditional regularized least square classifier by adding the consistency constraint. We will solve it by two types of one-pass learning methods.
\textbf{E-stage}: We expand the classifier trained on survived features, i.e., $\tilde{h}^{\textrm{(s)}}$, in C-stage to accommodate augmented features. It can take $\mathbf{X}_{T_1+1}^{\textrm{(s)}}$ as the input directly. Denote $\tilde{h}_{*}^{\textrm{(s)}}$ as the optimal classifier and $\mathbf{W}_{*}^{\textrm{(s)}}$ as its coefficient in C-stage, then $\mathbf{Z}_{T_1+1}^{\textrm{(s)}} \triangleq \tilde{h}_{*}^{\textrm{(s)}}(\mathbf{X}_{T_1+1}^{\textrm{(s)}}) = \mathbf{X}_{T_1+1}^{\textrm{(s)}} \mathbf{W}_{*}^{\textrm{(s)}}$ is the prediction by employing the classifier training in C-stage. We take this prediction as new representations of $\mathbf{X}_{T_1+1}^{\textrm{(s)}}$ as in stacking \cite{stacking, Zhou:2012:EMF}. After that, we train a classifier $\bar{h}^{\textrm{(s)}}$ on $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ and simultaneously, another classifier $\bar{h}$ is trained on $ \bar{\mathbf{Z}}_{T_1+1} \triangleq [\mathbf{Z}_{T_1+1}^{\textrm{(s)}}, \mathbf{X}_{T_1+1}^{\textrm{(a)}}]$ for accommodation. It can be regarded as expansion of the optimal classifier in C-stage to include augmented features.
To inherit advantages of $\tilde{h}_{*}^{\textrm{(s)}}$ trained in C-stage, we combine two classifiers like ensemble methods \cite{Zhou:2012:EMF}. At first, we employ this strategy to unify two classifiers by optimizing the following problem.
{
\begin{equation}
\label{eq11}
\begin{split}
\min_{\bar{h}^{\textrm{(s)}}, \bar{h}} w_1~\bar{\ell}^{\textrm{(s)}} \left( \bar{h}^{\textrm{(s)}}(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}),\mathbf{Y}_{T_1+1} \right) + w_2~\bar{\ell} \left( \bar{h} (\bar{\mathbf{Z}}_{T_1+1}),\mathbf{Y}_{T_1+1} \right)~,
\end{split}
\end{equation}}
where $w_1 \geq 0, w_2\geq 0, w_1+w_2=1$, are the weights to balance two classifiers. $\bar{\ell}^{\textrm{(s)}}$ and $\bar{h}$ are two surrogate loss function defined on $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ and $\bar{\mathbf{Z}}_{T_1+1}$ respectively.
For simplicity, we use the L$_2$-regularized Logistic Regression model for each classifier. Take the binary classification problem as an example, the objective functions are
{
\begin{equation}
\label{eq19}
\begin{split}
&\bar{\ell}^{\textrm{(s)}} \left( \bar{h}^{\textrm{(s)}}(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}),\mathbf{y}_{T_1+1} \right) = \frac{1}{2} \mathbf{v}^{\textrm{(s)}}(\mathbf{v}^{\textrm{(s)}})^\top+ \alpha_1 \sum_{j=1}^{n_{T_1+1}} \textrm{log}(1+\exp (- y_{T_1+1,j}\mathbf{z}_{T_1+1,j}^{\textrm{(s)}}\mathbf{v}^{\textrm{(s)}} )) \\
&\bar{\ell} \left( \bar{h} (\bar{\mathbf{Z}}_{T_1+1}),\mathbf{y}_{T_1+1} \right) = \frac{1}{2} \bar{\mathbf{v}} \bar{\mathbf{v}}^\top+ \alpha_2 \sum_{j=1}^{n_{T_1+1}} \textrm{log}(1+\exp (- y_{T_1+1,j} \bar{\mathbf{z}}_{T_1+1,j} \bar{\mathbf{v}} ))
\end{split}
\end{equation}}
where $\mathbf{v}^{\textrm{(s)}}$ and $\bar{\mathbf{v}}$ are coefficients. $\mathbf{z}_{T_1+1,j}^{\textrm{(s)}}$ and $\bar{\mathbf{z}}_{T_1+1,j}$ are the $j$-th row (the $j$-th instances) of $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ and $\bar{\mathbf{Z}}_{T_1+1}$. $\alpha_1$ and $\alpha_2$ are balance parameters. $y_{T_1+1,j}$ is 1 or -1 for binary classification.
One point should be mentioned here. We take $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ as the new representation, although it can be regarded as the classification results directly. The reasons are: (1) The prediction $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ is computed by the classifier trained on C-stage, we can use training data in E-stage to improve $\tilde{h}_{*}^{\textrm{(s)}}$. This composite works since $\tilde{h}_{*}^{\textrm{(s)}}$ and $\bar{h}^{\textrm{(s)}}$ are trained on different data sets. (2) As $\sum_{i=1}^{T_1} n_{i}$ is often much larger than $n_{T_1+1}$, by contrast with the classifier trained on $\mathbf{X}_{T_1+1}^{\textrm{(s)}}$ merely, $\tilde{h}_{*}^{\textrm{(s)}}$ is a better classifier and it could extract more discriminative information. In other words, compared with $\mathbf{X}_{T_1+1}^{\textrm{(s)}}$, $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ is a more compact and accurate representation. (3) If $\tilde{h}_{*}^{\textrm{(s)}}$ is good enough, the composite of $\bar{h}^{\textrm{(s)}}$ will not degrade the performance by our following strategy.
\section{Optimization and Extension}
\subsection{Optimization}
\textbf{C-stage:} The optimization problem in Eq.(\ref{eq1}) can be divided into $T_1$ subproblems, thus, it is direct to use the online learning method, such as online ADMM \cite{icml/WangB12}, to solve it by scanning the data only once. Nevertheless, we aim to train $\tilde{h}^{\textrm{(s)}}$ to assist the classification in E-stage, and the most direct way is to employ a linear classifier. In this case, we will provide more effective one-pass learning methods than using ADMM.
\begin{lemma}
\label{lemma1}
The optimal solution to Eq. (\ref{eq2}) can be obtained by solving
{
\begin{equation}
\label{eq4}
\begin{split}
\mathbf{A}_{[T_1]} \left[ {\begin{array}{*{20}{c}}
\tilde{\mathbf{W}} \\
\mathbf{W}^{\textrm{(s)}}
\end{array}} \right]=\mathbf{B}_{[T_1]},
\textrm{~~with~~}
\mathbf{B}_{[T]} \triangleq \sum_{i=1}^T
\left[ {\begin{array}{*{20}{c}}
\tilde{\mathbf{X}}_i^\top \\
(\mathbf{X}_i^{\textrm{(s)}})^\top \\
\end{array}} \right]\mathbf{Y}_i~,
\end{split}
\end{equation}}
{
\begin{equation}
\label{eq5}
\begin{split}
\mathbf{A}_{[T]} \triangleq
\left[ {\begin{array}{*{20}{c}}
(1+\lambda) \sum_{i=1}^{T} \tilde{\mathbf{X}}_{i}^\top \tilde{\mathbf{X}}_{i} +\rho \mathbf{I} & -\lambda \sum_{i=1}^{T} \tilde{\mathbf{X}}_{i}^\top \mathbf{X}_{i}^{\textrm{(s)}} \\
-\lambda \sum_{i=1}^{T} (\mathbf{X}_{i}^{\textrm{(s)}})^\top \tilde{\mathbf{X}}_{i} & (1+\lambda) \sum_{i=1}^{T} (\mathbf{X}_{i}^{\textrm{(s)}})^\top \mathbf{X}_{i}^{\textrm{(s)}} +\rho \mathbf{I}
\end{array}}
\right],
\end{split}
\end{equation}}
and $\mathbf{I}$ is an identity matrix.
\end{lemma}
\begin{proof}
Take the derivative of the objective function in Eq. (\ref{eq2}) with respect to $\tilde{\mathbf{W}}$ and $\mathbf{W}^{\textrm{(s)}}$ and set them to zeros, we have the following equations.
{
\begin{equation}
\label{eq3-s}
\begin{split}
&\sum_{i=1}^{T_1} \tilde{\mathbf{X}}_{i}^\top (\tilde{\mathbf{X}}_{i} \tilde{\mathbf{W}} - \mathbf{Y}_{i}) + \lambda \sum_{i=1}^{T_1} \tilde{\mathbf{X}}_{i}^\top (\tilde{\mathbf{X}}_{i} \tilde{\mathbf{W}} - \mathbf{X}_{i}^{\textrm{(s)}}\mathbf{W}^{\textrm{(s)}}) +\rho \tilde{\mathbf{W}} = \mathbf{0},\\
&\sum_{i=1}^{T_1} (\mathbf{X}_{i}^{\textrm{(s)}})^\top (\mathbf{X}_{i}^{\textrm{(s)}}\mathbf{W}^{\textrm{(s)}}- \mathbf{Y}_{i})+\lambda \sum_{i=1}^{T_1} (\mathbf{X}_{i}^{\textrm{(s)}})^\top ( \mathbf{X}_{i}^{\textrm{(s)}} \mathbf{W}^{\textrm{(s)}}- \tilde{\mathbf{X}}_{i} \tilde{\mathbf{W}}) + \rho \mathbf{W}^{\textrm{(s)}}=\mathbf{0}.\\
\end{split}
\end{equation}
}
Denote $\mathbf{A}_{[T]}$ and $\mathbf{B}_{[T]}$ as shown in Eq. (\ref{eq5}) and Eq. (\ref{eq4}). The optimization problem problem in Eq. (\ref{eq3-s}) becomes
{
\begin{equation}
\label{eq6-s}
\begin{split}
\mathbf{A}_{[T_1]} \left[ {\begin{array}{*{20}{c}}
\tilde{\mathbf{W}} \\
\mathbf{W}^{\textrm{(s)}}
\end{array}} \right]=\mathbf{B}_{[T_1]}.
\end{split}
\end{equation}
}
It is just the results shown in Proposition 1.
\end{proof}
Based on the above deduction, we turn to solve the problem in Eq. (\ref{eq2}) in one-pass way quickly. In the $T$-th time, we only access the instance $\{\tilde{\mathbf{X}}_i, ~\mathbf{Y}_{i}\}_{i=1}^{T}$. The counterpart optimization problem is the same as Eq. (\ref{eq2}), except that the sum of subscript $i$ is from 1 to $T$.
Notice that the solution to problem in Eq. (\ref{eq2}) is determined by $\mathbf{A}_{[T]}$ and $\mathbf{B}_{[T]}$ defined in Eq. (\ref{eq5}) and Eq. (\ref{eq4}). This evokes us to get the following updating rule.
{
\begin{equation}
\label{eq7}
\begin{split}
\mathbf{A}_{[T+1]} =& \mathbf{A}_{[T]}+
\left[\begin{array}{*{20}{c}}
(1+\lambda) \tilde{\mathbf{X}}_{T+1}^\top \tilde{\mathbf{X}}_{T+1} & -\lambda \tilde{\mathbf{X}}_{T+1}^\top \mathbf{X}_{T+1}^{\textrm{(s)}} \\
-\lambda (\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \tilde{\mathbf{X}}_{T+1} & (1+\lambda)(\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \mathbf{X}_{T+1}^{\textrm{(s)}}
\end{array}
\right],\\
\mathbf{B}_{[T+1]} =& \mathbf{B}_{[T]}+
\left[ {\begin{array}{*{20}{c}}
\tilde{\mathbf{X}}_{T+1}^\top \\
(\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \\
\end{array}} \right]\mathbf{Y}_{T+1},
\end{split}
\end{equation}}
{
\begin{equation}
\label{eq8}
\begin{split}
\mathbf{A}_{1} =
\left[\begin{array}{*{20}{c}}
(1+\lambda) \tilde{\mathbf{X}}_{1}^\top \tilde{\mathbf{X}}_{1}+\rho \mathbf{I} & -\lambda \tilde{\mathbf{X}}_{1}^\top \mathbf{X}_{1}^{\textrm{(s)}} \\
-\lambda (\mathbf{X}_{1}^{\textrm{(s)}})^\top \tilde{\mathbf{X}}_{1} & (1+\lambda)(\mathbf{X}_{1}^{\textrm{(s)}})^\top \mathbf{X}_{1}^{\textrm{(s)}}+\rho \mathbf{I}
\end{array}
\right],
\mathbf{B}_{1} =
\left[ {\begin{array}{*{20}{c}}
\tilde{\mathbf{X}}_{1}^\top \\
(\mathbf{X}_{1}^{\textrm{(s)}})^\top \\
\end{array}} \right]\mathbf{Y}_{1}.
\end{split}
\end{equation}}
According to this result, in time $T+1$, we only need to update $\mathbf{A}_{[T+1]}$ and $\mathbf{B}_{[T+1]}$ by adding the matrices calculated based on the data in batch $T+1$. In other words, we just need to store the matrices $\mathbf{A}_{[T]}$ and $\mathbf{B}_{[T]}$ and update them based on Eq.(\ref{eq7}) in each iteration.
This approach has the following advantages: (1) It just needs to store two matrices with size $(d^{\textrm{(v)}}+2d^{\textrm{(s)}})\times (d^{\textrm{(v)}}+2d^{\textrm{(s)}})$ and $(d^{\textrm{(v)}}+2d^{\textrm{(s)}})\times c$. When $d^{\textrm{(v)}}+2d^{\textrm{(s)}} < \sum_{i=1}^{T_1} n_{i}$, it needs less space than storing the whole data. Through this way, we can get the optimal solution to Eq. (\ref{eq2}) by scanning the total data only once. (2) We only make matrix multiplication in updating $\mathbf{A}_{[T]}$ and $\mathbf{B}_{[T]}$, the computational cost is small. In solving Eq. (\ref{eq4}), the most time-consuming step is computing the inverse of a matrix with size $(d^{\textrm{(v)}}+2d^{\textrm{(s)}})\times (d^{\textrm{(v)}}+2d^{\textrm{(s)}})$. Thus, this method is very efficient with large data number and small data dimensionality.
Compared with the data size, when the number of features, i.e., $d^{\textrm{(v)}}+d^{\textrm{(s)}}$, is rather large, it is unwise to compute the inverse of $\mathbf{A}_{[T_1]}$ directly. In this case, we propose another one-pass learning approach in solving the optimization problem in Eq. (\ref{eq4}).
Define $\mathbf{A}_{[0]}=\rho \mathbf{I}$ and $\mathbf{B}_{[0]}= \mathbf{0}$, the updating rule shown in Eq. (\ref{eq7}) can be initialized from $T=0$. Note that $\mathbf{A}_{[0]}^{-1} = (1/\rho) \mathbf{I}$. If we can replace the updating rule of $\mathbf{A}_{[T+1]}$ in Eq. (\ref{eq7}) by the updating rule of $\mathbf{A}_{[T+1]}^{-1}$ with less computational cost in each iteration, the computational burden in calculating $\mathbf{A}_{[T_1]}^{-1}$ will release. Notice that, the added matrix in updating $\mathbf{A}_{[T+1]}$ is not a full rank matrix if $n_{T+1}$ is smaller than $\min\{d^{\textrm{(v)}},d^{\textrm{(s)}}\}$. We will use this property to compute the inverse with low cost.
\begin{lemma}
\label{lemma2}
The updating rule of $\mathbf{A}_{[T+1]}^{-1}$ is
{
\begin{equation}
\label{eq10}
\begin{split}
\mathbf{A}_{[T+1]}^{-1} =
\mathbf{A}_{[T]}^{-1}-\mathbf{A}_{[T]}^{-1}\mathbf{U}_{T+1}\left(\mathbf{I}+\mathbf{U}_{T+1}^\top\mathbf{A}_{[T]}^{-1}\mathbf{U}_{T+1}\right)^{-1} \mathbf{U}_{T+1}^\top \mathbf{A}_{[T]}^{-1}
\end{split}
\end{equation}}
where $\mathbf{U}_{T+1} = [\mathbf{U}_{T+1,1}, \mathbf{U}_{T+1,2}, \mathbf{U}_{T+1,3}]$ and
{
\begin{equation*}
\label{eq9}
\begin{split}
\mathbf{U}_{T+1,1}=\left[ {\begin{array}{*{20}{c}}
\tilde{\mathbf{X}}_{T+1}^\top \\
\mathbf{0}
\end{array}} \right],
\mathbf{U}_{T+1,2}=\left[ {\begin{array}{*{20}{c}}
\mathbf{0}\\
(\mathbf{X}_{T+1}^{\textrm{(s)}})^\top
\end{array}} \right],
\mathbf{U}_{T+1,3}=\left[ {\begin{array}{*{20}{c}}
\sqrt{\lambda} \tilde{\mathbf{X}}_{T+1}^\top\\
-\sqrt{\lambda}(\mathbf{X}_{T+1}^{\textrm{(s)}})^\top
\end{array}} \right].
\end{split}
\end{equation*}}
\end{lemma}
\begin{proof}
Note that, the updating rule of $\mathbf{A}_{[T+1]}$ is shown in Eq. (\ref{eq7}). We now decompose the adding part as
{
\begin{equation}
\label{eq7-ss}
\begin{split}
& \left[\begin{array}{*{20}{c}}
(1+\lambda) \tilde{\mathbf{X}}_{T+1}^\top \tilde{\mathbf{X}}_{T+1} & -\lambda \tilde{\mathbf{X}}_{T+1}^\top \mathbf{X}_{T+1}^{\textrm{(s)}} \\
-\lambda (\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \tilde{\mathbf{X}}_{T+1} & (1+\lambda)(\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \mathbf{X}_{T+1}^{\textrm{(s)}}
\end{array}
\right] =
\left[\begin{array}{*{20}{c}}
\tilde{\mathbf{X}}_{T+1}^\top \tilde{\mathbf{X}}_{T+1} & \mathbf{0} \\
\mathbf{0} & \mathbf{0}
\end{array}
\right] \\
+& \left[\begin{array}{*{20}{c}}
\mathbf{0} & \mathbf{0} \\
\mathbf{0} & (\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \mathbf{X}_{T+1}^{\textrm{(s)}}
\end{array}
\right]+
\left[\begin{array}{*{20}{c}}
\lambda \tilde{\mathbf{X}}_{T+1}^\top \tilde{\mathbf{X}}_{T+1} & -\lambda \tilde{\mathbf{X}}_{T+1}^\top \mathbf{X}_{T+1}^{\textrm{(s)}} \\
-\lambda (\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \tilde{\mathbf{X}}_{T+1} & \lambda (\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \mathbf{X}_{T+1}^{\textrm{(s)}}
\end{array}
\right]
\end{split}
\end{equation}
}
Denote $\mathbf{U}_{T+1,1}$, $\mathbf{U}_{T+1,2}$ and $\mathbf{U}_{T+1,3}$ as shown in Proposition 2, we have
{
\begin{equation}
\label{eq9-s}
\begin{split}
& \left[\begin{array}{*{20}{c}}
(1+\lambda) \tilde{\mathbf{X}}_{T+1}^\top \tilde{\mathbf{X}}_{T+1} & -\lambda \tilde{\mathbf{X}}_{T+1}^\top \mathbf{X}_{T+1}^{\textrm{(s)}} \\
-\lambda (\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \tilde{\mathbf{X}}_{T+1} & (1+\lambda)(\mathbf{X}_{T+1}^{\textrm{(s)}})^\top \mathbf{X}_{T+1}^{\textrm{(s)}}
\end{array}
\right] \\
= &\mathbf{U}_{T+1,1} \mathbf{U}_{T+1,1}^\top + \mathbf{U}_{T+1,2} \mathbf{U}_{T+1,2}^\top+\mathbf{U}_{T+1,3} \mathbf{U}_{T+1,3}^\top \\
= & \mathbf{U}_{T+1} \mathbf{U}_{T+1}^\top,
\end{split}
\end{equation}
}
with $\mathbf{U}_{T+1} = [\mathbf{U}_{T+1,1}, \mathbf{U}_{T+1,2}, \mathbf{U}_{T+1,3}]$.
Using the Woodbury equation \cite{Higham:2002}, we have the results as follows.
{
\begin{equation}
\label{eq10-s}
\begin{split}
\mathbf{A}_{[T+1]}^{-1} = (\mathbf{A}_{[T]}+ \mathbf{U}_{T+1} \mathbf{U}_{T+1}^\top )^{-1} =
\mathbf{A}_{[T]}^{-1}-\mathbf{A}_{[T]}^{-1}\mathbf{U}_{T+1}\left(\mathbf{I}+\mathbf{U}_{T+1}^\top \mathbf{A}_{[T]}^{-1}\mathbf{U}_{T+1}\right)^{-1} \mathbf{U}_{T+1}^\top \mathbf{A}_{[T]}^{-1}
\end{split}
\end{equation}}
It is the results shown in Proposition 2.
\end{proof}
Similarly, it is also the one-pass way in updating $\mathbf{A}_{[T+1]}^{-1}$ and we need to access the whole data only once. In each iteration shown in Eq. (\ref{eq10}), the most computational step is calculating the inverse of a matrix with size $3n_{T+1}\times 3n_{T+1}$. If the batch size $n_{T+1}$ is small, its computational cost is limited and we can compute $\mathbf{A}_{[T_1]}^{-1}$ in a quick way. Especially, if $n_i=1$ for $i\in [T_1]$ as in traditional online learning, we only need to compute the inverse of a $3\times 3$ matrix.
Besides, a byproduct of this kind of iteration is that we can get the optimal solution at any time $T$, since we have derived $\mathbf{A}_{[T]}^{-1}$ directly. If we use the iteration method shown in Eq. (\ref{eq7}), we need to calculate $\mathbf{A}_{[T]}^{-1}$ at each time $T$. When this requirement is frequent, the computational cost will increase since we need to compute the matrix inverse for each requirement.
\textbf{E-stage:} The optimization problem in Eq. (\ref{eq11}) with concrete forms defined in Eq. (\ref{eq19}) has been widely investigated in previous works and the details are omitted. We use the implementation of LibLinear \cite{REF08a} to solve them and the parameters $w_1$ and $w_2$ are turned by cross validation.
\subsection{Extension}
Note that in Eq. (\ref{eq11}), the interaction between two classifiers is a balance of classification results by turning the weights. We think the more direct way is combining all the features as in stacking \cite{stacking} and training a unified classifier on the stacked representations. This evokes the following formulation.
{
\begin{equation}
\label{eq12}
\begin{split}
\min_{\bar{h}^{\textrm{(s)}}, \bar{h}, w_1, w_2} \ell \left( \sqrt{w_1} \bar{h}^{\textrm{(s)}}(\mathbf{Z}_{T_1+1}^{\textrm{(s)}})+
\sqrt{w_2} \bar{h}(\bar{\mathbf{Z}}_{T_1+1}), \mathbf{Y}_{T_1+1} \right)~,
\end{split}
\end{equation}}
where $\ell$ is a general loss on the joint representations. Here, we use the square root of balance parameters to guarantee their convexity and avoid the trivial solution. They can be learned automatically without extra hyper-parameters.
Taking regression with kernels as an example, we have the following concrete formulation.
{
\begin{equation}
\label{eq13}
\begin{split}
\min_{\mathbf{V}^{\textrm{(s)}},\bar{\mathbf{V}},w_1,w_2 }
& \left \| \sqrt{w_1} \langle {\mathbf{V}}^{\textrm{(s)}}, \Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}) \rangle +
\sqrt{w_2} \langle \bar{\mathbf{V}}, \Phi( \bar{\mathbf{Z}}_{T_1+1}) \rangle - \mathbf{Y}_{T_1+1} \right\|^2\\
+& \gamma (\frac{1}{c}\| \mathbf{V}^{\textrm{(s)}}\|^2+ \frac{1}{c+d^{\textrm{(a)}}}\| \bar{\mathbf{V}} \|^2),~\textrm{with}~w_1+w_2=1,~w_1\geq 0,~w_2\geq 0~,
\end{split}
\end{equation}}
where $\Phi(\cdot)$ is a mapping function and $\gamma$ is the parameter for regularization. The feature numbers of $\mathbf{Z}_{T_1+1}^{\textrm{(s)}}$ and $\bar{\mathbf{Z}}_{T_1+1}$ are $c$, $c+d^{\textrm{(a)}}$, and $c$ is often much smaller than $c+d^{\textrm{(a)}}$. To alleviate the influence caused by the unbalance, each regularizer is divided by the corresponding feature number.
It is not easy to solve the problem in Eq. (\ref{eq13}) directly. After some deductions, it is equal to
{
\begin{equation}
\label{eq14}
\begin{split}
\min_{\mathbf{V}^{\textrm{(s)}},\bar{\mathbf{V}}, w_1, w_2 }
& \left\| \langle {\mathbf{V}}^{\textrm{(s)}}, \Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}) \rangle +
\langle \bar{\mathbf{V}}, \Phi( \bar{\mathbf{Z}}_{T_1+1}) \rangle - \mathbf{Y}_{T_1+1} \right\|^2\\
+\gamma (\frac{1}{c \times w_1} &\| \mathbf{V}^{\textrm{(s)}}\|^2+ \frac{1}{(d^{\textrm{(a)}}+c)\times w_2}\| \bar{\mathbf{V}} \|^2),~\textrm{with}~w_1+w_2=1,~w_1\geq 0,~w_2\geq 0.
\end{split}
\end{equation}}
The optimization problem in Eq. (\ref{eq14}) seems very similar to traditional regularized square loss regression. Nevertheless, they are different since the regularization parameter is fixed in traditional method, while it is also an optimization parameter in our setting.
There are two groups of optimization variables, i.e., the coefficients for classification and the balance parameters. It is difficult to optimize them together and we optimize them alternatively.
When $\mathbf{V}^{\textrm{(s)}}$ and $\bar{\mathbf{V}}$ are fixed, the optimal balance parameters $w_1$ and $w_2$ can be computed by the following proposition directly.
\begin{lemma}
\label{lemma3}
When $w_1+w_2=1,~w_1\geq 0,~w_2\geq 0$,
{
\begin{equation}
\label{eq15}
\begin{split}
\min_{w_1,w_2}~(\frac{1}{ w_1 \times c} \| \mathbf{V}^{\textrm{(s)}}\|^2+ \frac{1}{w_2 \times (d^{\textrm{(a)}}+c)} \|\bar{\mathbf{V}}\|^2) =
(\frac{1}{\sqrt{c}} \| \mathbf{V}^{\textrm{(s)}}\|+ \frac{1}{\sqrt{d^{\textrm{(a)}}+c}} \| \bar{\mathbf{V}}\|)^2.
\end{split}
\end{equation}}
The optimal solution is
{
\begin{equation}
\label{eq16}
\begin{split}
w_1^*=\frac{\|\mathbf{V}^{\textrm{(s)}}\|/ \sqrt{c}} {\|\mathbf{V}^{\textrm{(s)}}\|/ \sqrt{c}+ \|\bar{\mathbf{V}}\| / \sqrt{d^{\textrm{(a)}}+c} },~
w_2^*=\frac{\|\bar{\mathbf{V}} \|/ \sqrt{d^{\textrm{(a)}}+c}} {\|\mathbf{V}^{\textrm{(s)}}\|/ \sqrt{c}+\|\bar{\mathbf{V}}\| / \sqrt{d^{\textrm{(a)}}+c} }.
\end{split}
\end{equation}}
\end{lemma}
\begin{proof}
The optimization problem is
\begin{equation}
\label{eq11-s}
\begin{split}
\min_{w_1,w_2}~&(\frac{1}{ w_1 \times c} \| \mathbf{V}^{\textrm{(s)}}\|^2+ \frac{1}{w_2\times (d^{\textrm{(a)}}+c)} \|\bar{\mathbf{V}} \|^2)~, \\
\textrm{s.t.}~&w_1+w_2=1,~w_1\geq 0,~w_2\geq 0.
\end{split}
\end{equation}
Replace $w_2$ with $w_2=1-w_1$ in Eq. (\ref{eq11-s}), take derivative of this objective function with respect to $w_1$ and set it to zero, we have
{
\begin{equation}
\label{eq12-s}
\begin{split}
-\frac{1}{ w_1^2 \times c} \| \mathbf{V}^{\textrm{(s)}}\|^2 + \frac{1}{(1-w_1)^2 \times (d^{\textrm{(a)}}+c)} \| \bar{\mathbf{V}}\|^2)= 0. \\
\end{split}
\end{equation}
}
By solving this problem and using $w_2=1-w_1$, we have
{
\begin{equation}
\label{eq13-s}
\begin{split}
w_1=\frac{\|\mathbf{V}^{\textrm{(s)}}\|/ \sqrt{c}} {\|\mathbf{V}^{\textrm{(s)}}\|/ \sqrt{c}+ \|\bar{\mathbf{V}}\| / \sqrt{d^{\textrm{(a)}}+c} },~
w_2=\frac{\|\bar{\mathbf{V}} \|/ \sqrt{d^{\textrm{(a)}}+c}} {\|\mathbf{V}^{\textrm{(s)}}\|/ \sqrt{c}+\|\bar{\mathbf{V}}\| / \sqrt{d^{\textrm{(a)}}+c} }.
\end{split}
\end{equation}
}
Note that the solution in Eq. (\ref{eq13-s}) satisfies the constraint $w_1\geq 0,~w_2\geq 0$ automatically. Thus, it is the optimal solution and the results in Proposition 3 hold.
\end{proof}
When $w_1$ and $w_2$ are fixed, the optimal solution to the problem in Eq. (\ref{eq14}) can be obtained in a close form as shown in the following proposition.
\begin{lemma}
\label{lemma4}
When $w_1$ and $w_2$ are fixed, the optimal solution to the problem in Eq. (\ref{eq14}) can be derived by solving
{
\begin{equation}
\label{eq17}
\begin{split}
\left[\begin{array}{*{20}{c}}
(\Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}))^\top \Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}})+ \frac{\gamma} {c \times w_1}\mathbf{I}
& (\Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}))^\top \Phi( \bar{\mathbf{Z}}_{T_1+1}) \\
(\Phi(\bar{\mathbf{Z}}_{T_1+1}))^\top \Phi( \mathbf{Z}_{T_1+1}^{\textrm{(s)}})
&(\Phi( \bar{\mathbf{Z}}_{T_1+1} ))^\top \Phi(\bar{\mathbf{Z}}_{T_1+1})+ \frac{\gamma} {(d^{\textrm{(a)}}+c) \times w_2}\mathbf{I}
\end{array}
\right] \mathbf{V} = \mathbf{D},
\end{split}
\end{equation}}
{
\begin{equation}
\label{eq18}
\begin{split}
\mathbf{V}=
\left[ {\begin{array}{*{20}{c}}
\mathbf{V}^{\textrm{(s)}} \\
\bar{\mathbf{V}} \\
\end{array}} \right],~~
\mathbf{D} = \left[ {\begin{array}{*{20}{c}}
(\Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}))^\top \\
(\Phi(\bar{\mathbf{Z}}_{T_1+1}))^\top \\
\end{array}} \right] \mathbf{Y}_{T_1+1}.
\end{split}
\end{equation}}
\end{lemma}
\begin{proof}
When $w_1$ and $w_2$ are fixed, the optimization problem is shown in Eq. (\ref{eq14}). Take the derivative of it with respect respect to $\mathbf{V}^{\textrm{(s)}}$ and $\bar{\mathbf{V}}$, set them to zeros, we have
{
\begin{equation}
\label{eq17-s}
\begin{split}
&\left( (\Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}))^\top \Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}})+ \frac{\gamma} {c \times w_1}\mathbf{I}\right) \mathbf{V}^{\textrm{(s)}}+
(\Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}))^\top \Phi(\bar{\mathbf{Z}}_{T_1+1}) \bar{\mathbf{V}} = (\Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}))^\top \mathbf{Y}_{T_1+1}, \\
&\left( (\Phi(\bar{\mathbf{Z}}_{T_1+1}))^\top \Phi(\bar{\mathbf{Z}}_{T_1+1})+ \frac{\gamma} {(d^{\textrm{(a)}}+c) \times w_2}\mathbf{I}\right) \bar{\mathbf{V}} +
(\Phi( \bar{\mathbf{Z}}_{T_1+1}))^\top \Phi(\mathbf{Z}_{T_1+1}^{\textrm{(s)}}) \mathbf{V}^{\textrm{(s)}}
= (\Phi(\bar{\mathbf{Z}}_{T_1+1}))^\top \mathbf{Y}_{T_1+1}.
\end{split}
\end{equation}
}
After making some notations shown in Proposition 4, we can get the results in Proposition 4.
\end{proof}
There is a mapping function $\Phi$ in this formulation. If we do not know its concrete form, the kernel trick \cite{Scholkopf:1999:AKM:299094} can be employed to make prediction for testing data.
When we obtain the final classifier, it can be used on the testing data $\bar{\mathbf{X}}_{T_1+2} = [\mathbf{X}_{T_1+2}^{\textrm{(s)}}, \mathbf{X}_{T_1+2}^{\textrm{(a)}}]$ by first computing the new representations of $\mathbf{X}_{T_1+2}^{\textrm{(s)}}$ as ($\tilde{h}_{*}^{\textrm{(s)}}(\mathbf{X}_{T_1+2}^{\textrm{(s)}})$) and then employing this classifier. If we take the updating rule in Eq. (\ref{eq7}) and train a classifier in E-stage by optimizing Eq. (\ref{eq13}), the procedure of OPID is listed in Algorithm \ref{alg1:opid}.
\begin{algorithm}[!t]
\caption{OPID}
\label{alg1:opid}
{
\begin{algorithmic}
\STATE \textbf{Input:} Training data and label $\{\tilde{\mathbf{X}}_i, \mathbf{Y}_{i}\}_{i=1}^{T_1}$, $\{\bar{\mathbf{X}}_{T_1+1}, \mathbf{Y}_{T_1+1}\}$, testing data $\bar{\mathbf{X}}_{T_1+2}$, parameters $\lambda$, $\rho$, $\gamma$.
\STATE \textbf{Output:} The optimal $w_1$, $w_2$, $\mathbf{V}^{\textrm{(s)}}$ and $\bar{\mathbf{V}}$. \\
\STATE \textbf{Training:} \\
\STATE 1: Update $\mathbf{A}_{[T]}$ and $\mathbf{B}_{[T]}$ using Eq. (\ref{eq7}) in the one-pass way. \\
\STATE 2: Compute the optimal $\mathbf{W}_{*}^{\textrm{(s)}}$ by solving Eq. (\ref{eq4}).\\
\STATE 3: Compute the new representation of $\mathbf{X}_{T_1+1}^{\textrm{(s)}}$ by $\mathbf{Z}_{T_1+1}^{\textrm{(s)}} = \mathbf{X}_{T_1+1}^{\textrm{(s)}} \mathbf{W}_{*}^{\textrm{(s)}}$.\\
\STATE 4: Initialize $w_1$=$w_2$=1/2\\
\STATE \textbf{Repeat} \\
\STATE 5: Update $\mathbf{V}^{\textrm{(s)}}$ and $\bar{\mathbf{V}}$ by solving the problem in Eq. (\ref{eq17}). \\
\STATE 6: Update $w_1$ and $w_2$ by Eq. (\ref{eq16}). \\
\STATE \textbf{Until converges}
\STATE \textbf{Testing}
\STATE 7: Compute the new representation of $\mathbf{X}_{T_1+2}^{\textrm{(s)}}$ by $\mathbf{X}_{T_1+2}^{\textrm{(s)}} \mathbf{W}_{*}^{\textrm{(s)}}$. \\
\STATE 8: Compute the predicted label matrix of $\mathbf{X}_{T_1+2}^{\textrm{(s)}}$ by using the optimal $w_1$, $w_2$, $\mathbf{V}^{\textrm{(s)}}$, $\bar{\mathbf{V}}$ and kernel trick.
\end{algorithmic}}
\end{algorithm}
\section{Experimental Results}
There are two implementations (Eq. (\ref{eq11}) and Eq. (\ref{eq13})) of our algorithm. We name the OPID method with ensemble (Eq. (\ref{eq11})) as OPIDe and the implementation in Eq. (\ref{eq13}) is still named as OPID. For simplicity, in Eq. (\ref{eq13}), we take $\Phi(\mathbf{x})=\mathbf{x}$ and thus the classifiers are linear. For fairness, the classifiers in Eq. (\ref{eq11}) are also linear. We implement it using the LibLinear toolbox \cite{REF08a} with L$_2$-regularized logistic regression and the parameters are tuned by five-fold cross validation. The multi-class problem is tackled by one-vs-rest strategy.
We will compare OPID with other related methods. To show whether the classifier trained in C-stage is helpful, we also train a linear classifier with L$_2$-regularized logistic regression on the whole training data in E-stage, i.e., $\{ \bar{\mathbf{X}}_{T_1+1} \triangleq [\mathbf{X}_{T_1+1}^{\textrm{(s)}}, \mathbf{X}_{T_1+1}^{\textrm{(a)}}],~\mathbf{Y}_{T_1+1}\}$. For simplicity, this method is notated by SVM. Besides, to show the effectiveness of expanding in E-stage, we also train the same kind of linear classifiers on data with survived feature $\{ \mathbf{X}_{T_1+1}^{\textrm{(s)}}, \mathbf{Y}_{T_1+1}\}$ and augmented feature $\{\mathbf{X}_{T_1+1}^{\textrm{(a)}}, \mathbf{Y}_{T_1+1}\}$ respectively. They are named as SVM$^{\textrm{(s)}}$ and SVM$^{\textrm{(a)}}$.
We have evaluated our approach on eight different kinds of data sets. They are three digit data sets: Mnist\footnote{http://yann.lecun.com/exdb/mnist/}, Gisette\footnote{http://clopinet.com/isabelle/Projects/NIPS2003/\#challenge} and
USPS\footnote{http://yann.lecun.com/exdb/mnist/},
three DNA data sets:
DNA\footnote{http://archive.ics.uci.edu/ml/machine-learning-databases/statlog/}, Splice\footnote{http://www.cs.toronto.edu/~delve/data/splice/desc.html} and
Protein\footnote{http://archive.ics.uci.edu/ml/machine-learning-databases/statlog/},
the Vehicle data: SensIT Vehicle\footnote{ http://www.ecs.umass.edu/~mduarte/Software.html}, and
the image data set: Satimage\footnote{http://www.cs.toronto.edu/~delve/data/splice/desc.html}.
The digit data sets are employed as toy examples and the rest data sets are all collected in a sequential way. For example, the SensIT Vehicle data are collected from a sensor networks. Due to the life variance of different kinds of sensors, it is a typical feature and instance evolution system.
For simplicity, we assume that (1) $n_1=n_2=,\cdots,n_{T_1}=n_{T_1+1}=n_{T_1+2}$. In C-stage, the number of training points in each category is equal, whereas in E-stage, we randomly split $n_{T_1+1}+n_{T_1+2}$ examples into two equal parts and assign them as training and testing samples respectively. (2) In C-stage, we fix the total number of examples and vary the points number in each batch. Thus, companied with this, the number of training and testing examples also changes in E-stage. (3) We select the first $(\sum_{i=1}^{T_1} n_i)/c$ samples in each category as the training data in C-stage. (4) We assign the first $d^{\textrm{(v)}}$ features as vanished features, the next $d^{\textrm{(s)}}$ features as survived features and the rest as augmented features. Without specification, in our experiments, the first quarter and the last quarter are vanished features and augmented features, except for DNA and Splice, since their original dimensionality is low. Experimental setting details are shown in Table \ref{table1}.
\begin{table}[ht]
\caption{{ The details about experimental setting.}}
\label{table1}
\vskip 0.1in
\centering
{
\begin{tabular}{cccccccc}
\hline
Data & c & $\sum_{i=1}^{T_1} n_i$ & $n_i$ & $d^{\textrm{(v)}}$ & $d^{\textrm{(s)}}$ & $d^{\textrm{(a)}}$ & $n_{T_1+1}=n_{T_1+2}$\\
\hline
Mnist0vs5 & 2 & 3200 & 40, 80, 160, 320 & 114 & 228 & 113 & 40, 80, 160, 320 \\
Mnist0vs3vs5 & 3 & 4800 & 60, 120, 240, 480 & 123 & 245 & 121 & 60, 120, 240, 480 \\
DNA & 3 & 1200 & 60, 120, 240, 300 & 50 & 80 & 50 & 60, 120, 240, 300 \\
Splice & 2 & 2240 & 40, 80, 160, 320 & 10 & 40 & 10 & 40, 80, 160, 320 \\
SensIT Vehicle & 3 & 48000& 60, 120, 240, 480 & 25 & 50 & 25 & 60, 120, 240, 480 \\
Gisette & 2 & 6000 & 40, 100, 200, 300 & 1239 & 2478 & 1238 & 40, 100, 200, 300\\
USPS0vs5 & 2 & 960 & 20, 40, 60, 80 & 64 & 128 & 64 & 20, 40, 60, 80\\
USPS0vs3vs5 & 3 & 1440 & 30, 60, 90, 120 & 64 & 128 & 64 & 30, 60, 90, 120 \\
Protein & 3 & 4500 & 60, 150, 300, 450 & 70 & 200 & 86 & 60, 150, 300, 450 \\
Satimage & 3 & 1080 & 30, 60, 90, 120 & 10 & 18 & 8 & 30, 60, 90, 120\\
\hline
\end{tabular}}
\vskip -0.1in
\end{table}
There are totally two groups of experiments. In the first group, we would like to report the classification accuracy comparison on the testing data in E-stage. In the second group, we will focus on the performance variation caused by the number of survived features.
\subsection{Classification Accuracy Comparison}
To show the results both intuitively and qualitatively, we report results on the first six data in the form of table and the rest with figures. The results of different methods on different data sets are presented in Table \ref{table_accuracy} and Fig. \ref{acc_comparison-s} respectively.
\begin{table*}[!t]
\renewcommand{\arraystretch}{1.1}
\caption{ The testing accuracies (mean$\pm$std.) of the compared methods on 6 data sets with different number of training and testing examples. '$\bullet/${\tiny $\odot$}$/\circ$' denote respectively that OPID(or OPIDe) is significantly better/tied/worse than the compared method by the $t$-test\cite{t-test} with confidence level 0.05. '-' means that the result is unavailable. From the third column to the fifth column, two symbols are the comparisons to IPODe and IPOD respectively. In the sixth column, the symbols are the comparisons between OPIDe and OPID. The highest mean accuracy is also boldfaced.}
\vskip 0.1in
\label{table_accuracy}
\centering
{\scriptsize
\begin{tabular}{c|| c| c c c |c c }
\hline
Data set &$n_i$ & $\textrm{SVM}$ & $\textrm{SVM}^{\textrm{(s)}} $ & $\textrm{SVM}^{\textrm{(a)}}$ & OPIDe & OPID \\
\hline
Mnist& 40&.9485(.0309)$\bullet\bullet$& .9455(.0334)$\bullet\bullet$& .7105(.0569)$\bullet\bullet$& .9700(.0286)$\bullet$& \textbf{.9840}(.0158) \\
0vs5 & 80&.9630(.0196)$\bullet\bullet$& .9633(.0196)$\bullet\bullet$& .6403(.0434)$\bullet\bullet$& .9868(.0088){\tiny $\odot$}& \textbf{.9888}(.0099) \\
&160&.9571(.0137)$\bullet\bullet$& .9528(.0136)$\bullet\bullet$& .6691(.0271)$\bullet\bullet$& .9794(.0097){\tiny $\odot$}& \textbf{.9875}(.0090) \\
&320&.9604(.0104)$\bullet\bullet$& .9585(.0102)$\bullet\bullet$& .6508(.0220)$\bullet\bullet$& \textbf{.9738}(.0058){\tiny $\odot$}& .9721(.0066) \\
\hline
Mnist &60& .9043(.0424)$\bullet\bullet$& .9093(.0399)$\bullet\bullet$& .5137(.0534)$\bullet\bullet$& .9280(.0339)$\bullet$& \textbf{.9403}(.0265) \\
0vs3vs5&120&.9233(.0218)$\bullet\bullet$& .9240(.0243)$\bullet\bullet$& .4852(.0381)$\bullet\bullet$& .9458(.0178){\tiny $\odot$}& \textbf{.9497}(.0130) \\
&240&.9125(.0165)$\bullet\bullet$& .9133(.0134)$\bullet\bullet$& .4653(.0266)$\bullet\bullet$& .9345(.0141){\tiny $\odot$}& \textbf{.9348}(.0135) \\
&480&.9265(.0110)$\bullet\bullet$& .9232(.0101)$\bullet\bullet$& .4843(.0182)$\bullet\bullet$& .9330(.0086){\tiny $\odot$}& \textbf{.9337}(.0079) \\
\hline
DNA&60& .7693(.0495)$\bullet\bullet$& .8017(.0511)$\bullet\bullet$& .3063(.0499)$\bullet\bullet$& .9183(.0246){\tiny $\odot$}& \textbf{.9253}(.0248) \\
&120&.8418(.0327)$\bullet\bullet$& .8630(.0308)$\bullet\bullet$& .3855(.0428)$\bullet\bullet$& .9315(.0191){\tiny $\odot$}& \textbf{.9322}(.0172) \\
&240&.8732(.0182)$\bullet\bullet$& .8890(.0203)$\bullet\bullet$& .3778(.0256)$\bullet\bullet$& \textbf{.9385}(.0111){\tiny $\odot$}& .9343(.0132) \\
&300&.8857(.0174)$\bullet\bullet$& .8969(.0188)$\bullet\bullet$& .3957(.0242)$\bullet\bullet$& \textbf{.9405}(.0120)$\circ$&
.9348(.0118) \\
\hline
Splice&40&.6625(.0621)$\bullet\bullet$& .6755(.0618)$\bullet\bullet$& .4640(.0513)$\bullet\bullet$& \textbf{.8150}(.0468){\tiny $\odot$}& .8025(.0455) \\
&80&.7512(.0403)$\bullet\bullet$& .7597(.0446)$\bullet\bullet$& .4680(.0486)$\bullet\bullet$& \textbf{.8122}(.0373){\tiny $\odot$}& .8050(.0353) \\
&160&.8084(.0258)$\bullet\bullet$& .8116(.0244)$\bullet\bullet$& .4829(.0354)$\bullet\bullet$& \textbf{.8400}(.0203){\tiny $\odot$}& .8391(.0205) \\
&320&.8408(.0174)$\bullet\bullet$& .8387(.0171)$\bullet\bullet$& .4961(.0232)$\bullet\bullet$& .8555(.0132){\tiny $\odot$}& \textbf{.8594}(.0138) \\
\hline
SensIT & 60&.6533(.0536)$\bullet\bullet$& .6513(.0561)$\bullet\bullet$& .5533(.0727)$\bullet\bullet$& .7113(.0443){\tiny $\odot$}& \textbf{.7147}(.0482) \\
Vehicle& 120&.6418(.0482)$\bullet\bullet$& .6550(.0439)$\bullet\bullet$& .5777(.0512)$\bullet\bullet$& \textbf{.7253}(.0410){\tiny $\odot$}& .7218(.0365) \\
& 240&.6816(.0287)$\bullet\bullet$& .6805(.0293)$\bullet\bullet$& .5942(.0352)$\bullet\bullet$& .7237(.0187){\tiny $\odot$}& \textbf{.7242}(.0183) \\
& 480&.7088(.0177)$\bullet\bullet$& .7076(.0164)$\bullet\bullet$& .6065(.0214)$\bullet\bullet$& .7206(.0132){\tiny $\odot$}& \textbf{.7223}(.0139) \\
\hline
Gisette
& 40&.8795(.0564)$\bullet\bullet$& .8710(.0566)$\bullet\bullet$& .8550(.0627)$\bullet\bullet$& .9600(.0331)$\bullet$& \textbf{.9710}(.0185) \\
& 100&.9172(.0276)$\bullet\bullet$& .9028(.0286)$\bullet\bullet$& .9116(.0264)$\bullet\bullet$& .9714(.0128){\tiny $\odot$}& \textbf{.9756}(.0126) \\
& 200&.9133(.0155)$\bullet\bullet$& .9039(.0181)$\bullet\bullet$& .8998(.0201)$\bullet\bullet$& \textbf{.9559}(.0095){\tiny $\odot$}& .9539(.0106) \\
& 300&.9413(.0106)$\bullet\bullet$& .9308(.0109)$\bullet\bullet$& .9231(.0126)$\bullet\bullet$& \textbf{.9636}(.0069)$\circ$& .9533(.0093) \\
\hline
\multicolumn{2}{c}{OPIDe: win/tie/loss}& 24/0/0 & 24/0/0 & 24/0/0 & - & 2/19/3 \\
\multicolumn{2}{c}{OPID: win/tie/loss}& 24/0/0 & 24/0/0 & 24/0/0 & 3/19/2 & - \\
\hline
\end{tabular}}
\end{table*}
\begin{figure*}[!ht]
\vskip 0in
\centering
\begin{minipage}[h]{7.5 cm}
\centering
\includegraphics[height= 2.5 in]{usps0vs5_acc_total.pdf}\\
\mbox{\footnotesize~~~~(a) USPS0vs5}
\end{minipage}
\begin{minipage}[h]{7.5 cm}
\centering
\includegraphics[height= 2.5 in]{usps0vs3vs5_acc_total.pdf}\\
\mbox{\footnotesize~~~~(b) USPS0vs3vs5}
\end{minipage}
\begin{minipage}[h]{7.5 cm}
\centering
\includegraphics[height= 2.5 in]{protein_acc_total.pdf}\\
\mbox{\footnotesize~~~~(c) Protein}
\end{minipage}
\begin{minipage}[h]{7.5 cm}
\centering
\includegraphics[height= 2.5 in]{satimage_acc_total.pdf}\\
\mbox{\footnotesize~~~~(d) Satimage}
\end{minipage}
\caption{Testing accuracy of different methods on different data sets with different numbers of training and testing instances.}
\label{acc_comparison-s}
\vskip -0.0in
\end{figure*}
There are several observations from these results.
(1) When we use the classifier trained in C-stage to assist learning in E-stage, the testing performance will increase significantly, especially when the training points in E-stage are rare. This is consistent with intuition since the assistance from C-stage will be weaker with the increase of training points.
(2) Compared with the accuracy of SVM$^{\textrm{(a)}}$, our results have a remarkable improvement. This validates that our methods could, to some extent, inherit the metric from C-stage.
(3) It seems that the improvement of our method with respect to other approaches is much larger in multi-class scenario. The reason may be that the binary classification accuracy is high enough and it is hard to make further improvement.
(4) Compared OPID with OPIDe, it seems that OPID performs slightly better than OPIDe in most data sets. The $t$-test results show that their performances tie in most cases. Nevertheless, their performances are also data dependent.
(5) It seems that our method achieves more significant improvement on biological data sets (DNA, Splice, Protein) than image data sets (Mnist, Gisette and Satimage). It may be caused by the fact that the biological data is more time dependent than the image data and it is more consistent with our settings.
\begin{figure*}[!ht]
\vskip 0in
\centering
\begin{minipage}[h]{8 cm}
\centering
\includegraphics[height= 2.5 in]{SensIT_Dif_Par_total.pdf}\\
\mbox{\footnotesize (a) SensIT Vehicle}
\end{minipage}
\begin{minipage}[h]{8 cm}
\centering
\includegraphics[height= 2.5 in]{Gisette_Dif_Pa_total.pdf}\\
\mbox{\footnotesize (b) Gisette}
\end{minipage}
\caption{Accuracy comparison of different methods with different percents of vanished features, survived features and augmented features.}
\label{dif_par-s}
\end{figure*}
\subsection{The Influence of the Number of Survived Features}
Different from traditional problems, there are three kinds of features in our settings. To illustrate the effectiveness of our methods, we vary the percentages of survived features and compare our methods with other related works. As in previous subsection, we also conduct experiments on SensIT Vehicle and Gisette. Similar to the way of assigning three kinds of features, we select different percentages of features in the middle as the survived feature. The rest are assigned as vanished feature and augmented feature with equal feature number. Comparison results are shown in Fig. (\ref{dif_par-s}).
There are at least two observations from Fig. (\ref{dif_par-s}).
(1) Our proposed methods outperform traditional methods, no matter what the percentage of survived features is. It validates the effectiveness of our method in dealing the problem in this setting.
(2) With the increase number of survived features, OPID and OPIDe achieve higher accuracies on SensIT Vehicle data. Nevertheless, their performances have a little turbulence on Gisette data. The reason may be that compared with SensIT Vehicle data whose total feature number is 100, the feature number of Gisette is 4955. It trends to be suffered from the curse of dimensionality \cite{Donoho00}.
\section{Conclusion}
In this paper, we study the problem of learning with incremental and decremental features, and propose the one-pass learning approach that does not need to keep the whole data for optimization. Our approach is particularly useful when both the data and features are evolving, while robust learning performance are needed, and is well scalable because it only needs to scan each instance once. In this paper we focus on one-shot feature change where the survived and augmented features do not vanish. It will be interesting to extend to multi-shot feature change where the survived and augmented features can vanish later.
|
\section{Introduction}
Let $G$ be a totally disconnected locally compact group and $H$ a closed subgroup. A smooth, complex valued representation $(\pi,V)$ of $G$ is called $H$-distinguished if there exists a non-zero linear form $\ell$ on $V$ such that $\ell(\pi(h)v)=\ell(v)$ for all $h\in H$ and $v\in V$.
If $\pi$ is irreducible, then such a linear form realizes $\pi$ in a space of functions on $G$, to wit,
\[
\pi\simeq\{g\mapsto \ell(\pi(g^{-1})v): v\in \pi\}\subseteq C^\infty(G/H).
\]
The class of $H$-distinguished representations play an important role in the harmonic analysis of the homogeneous space $G/H$ (see \cite{MR1075727} for instance).
Furthermore, distinguished representations are crucial for the global theory of period integrals of automorphic forms, have applications to the study of special values of $L$-functions and to the description of the image of functorial lifts in the Langlands program.
This paper continues the study of \cite{MR2417789} of distinguished representations of $\operatorname{GL}_n$ over a non-archimedean local field $F$ with respect to Klyachko subgroups.
A Klyachko model is an induced representation $\operatorname{Ind}_{H_{2k,r}}^G(\psi)$ of $G=\operatorname{GL}_n(F)$. Here, $n=2k+r$,
$H_{2k,r}$ is the subgroup of $G$ consisting of matrices of the form $\sm{h}{X}{0}{u}$ where $h\in \operatorname{Sp}_{2k}(F)$, $X\in M_{2k\times r}(F)$, $u$ is an upper-triangular unipotent matrix, and $\psi$ is the character of $H_{2k,r}$ trivially extending a non-degenerate character on the upper-triangular unipotent matrices in $\operatorname{GL}_r(F)$. For a fixed $n$, as $k$ varies, these models `interpolate' between the well known Whittaker model (the case $k=0$) and, if $n$ is even, the sympectic model (the case $k=n/2$).
When $F$ is a finite field, they were introduced by Klyachko in \cite{MR691984}. Together they form a complete model in that case (\cite{MR1129515}), that is, they satisfy
\[
\oplus_{k=0}^{\lfloor n/2 \rfloor} \operatorname{Ind}_{H_{2k,r}}^G(\psi)=\oplus_{\pi\in \hat{G}}\pi.
\]
Klyachko models over non-archimedean local fields were first studied in \cite{MR1078382} where it was observed, among other things, that some irreducible representations do not imbed into a Klyachko model. In \cite{MR2414223} it is proved that the sum $\oplus_{k=0}^{\lfloor n/2 \rfloor} \operatorname{Ind}_{H_{2k,r}}^G(\psi)$ is multiplicity free.
We shall say that an irreducible representation $\pi$ admits a Klyachko model if it can be embedded into $\oplus_{k=0}^{\lfloor n/2 \rfloor} \operatorname{Ind}_{H_{2k,r}}^G(\psi).$
The main result in \cite{MR2417789} prescribes a Klyachko model for any irreducible representation in the unitary dual of $\operatorname{GL}_n(F)$ based on Tadi\'c classification.
This was achieved by first prescribing a Klyachko model to any Speh representation. For a general unitary representation, a hereditary property of Klyachko models for representations parabolically induced from the Speh representations is applied.
Recently, Lapid and M\'inguez introduced a class of irreducible representations of $\operatorname{GL}_n$ over a non-archimedean local field \cite{MR3163355}. Inspired by their presentation in the Zelevinsky classification scheme, they called them {\it ladder} representations.
(see \S \ref{sss: ladder} for the definition). The class of ladder representations contains the Speh representations, the building blocks of the unitary dual (see Tadi\'c classification of the unitary dual of $\operatorname{GL}_n(F)$ \cite{MR870688}). Thus any irreducible unitarizable representation of $\operatorname{GL}_{n}(F)$ is a product of ladder representations.
In \cite{MR2515933} we found a connection between the Klyachko model and a partition naturally obtained from the Langlands parameter of a representation. Inspired by this relation, we were led to extend our study of Klyachko models to the entire admissible dual.
The present paper provides a collection of results regarding distinction of representations of finite length with respect to the Klyachko groups.
In particular, in Theorem \ref{thmC:main} below we classify the distinguished ladder representations in the context of Klyachko models over a non-archimedean local field $F$ of characteristic different then $2$. The special case, when $G=\operatorname{GL}_{2n}(F)$ and $H=\operatorname{Sp}_{2n}(F)$ is described in Theorem \ref{thmB:main} below. These results, together with the Hereditary property established in Thereom \ref{thmA:main} below, recovers, using only local methods, our recipe for the Klyachko model of any representation in the unitary dual of a general linear group.
To help understand the motivation for our results and techniques, we mention two general strategies that one could employ to approach the problem of classifying distinguished admissible representations in the context of a reductive $p$-adic group. The first strategy is based on Langlands classification, the second based on the notion of imprimitive representations.
We start with the strategy based on the Langlands classification and the notion of standard modules. The smooth dual of $G$ was classified by Langlands in terms of tempered representations of Levi subgroups: Every irreducible smooth representation of $G$ is the unique irreducible quotient of a unique standard module.
Clearly, every non-zero $H$-invariant linear form on an irreducible representation produces such a linear form on its standard module.
A possible strategy for classifying $H$-distinguished representations is based on the Langlands classification:
\begin{enumerate}
\item Classify all $H$-distinguished standard modules;
\item Determine if an $H$-invariant linear form on the standard module
descends to one on the irreducible quotient.
\end{enumerate}
To implement the first part of this strategy one can use the geometric lemma of Bernstein and Zelevinsky to analyze distinction of induced representations.
We refer to \cite{MR2930996} and \cite{MR3421655} for cases where a complete classification of distinguished standard modules was achieved.
Implementing the second step turns out to be subtler. An $H$-invariant linear form on a standard module will induce such a linear form on some irreducible component. To determine whether the irreducible quotient admits
an $H$-invariant linear form is equivalent to determining whether one of the $H$-invariant linear forms on the standard module descends to the irreducible quotient. This problem can be approached by studying the distinction properties of the maximal proper submodule. However, in general, not enough is known about its structure.
The second strategy is based on the concept of imprimitive representation. An irreducible representation of $G$ is called {\it imprimitive} if it is not parabolically induced from any proper parabolic subgroup.
Any irreducible representation is induced from an imprimitive one.
Thus an approach to the classification problem of $H$-distinction on the smooth dual of $G$ could be:
\begin{enumerate}
\item Classify all $H$-distinguished imprimitive representations
\item Determine the relation between $H$-distinction and parabolic induction.
\end{enumerate}
We now focus on the case where $G=\operatorname{GL}_n(F)$ or a product of general linear groups. In this case, the second step might be more accessible.
We further propose to carry this step in two stages:
\begin{itemize}
\item {\bf Hereditary Property}: Showing that $H$-distinction is compatible with parabolic induction
\item {\bf Purity Lemma}: Showing that an $H$-distinguished representation that is induced from a parabolic subgroup, must be induced from a distinguished representation of the Levi subgroup of that parabolic.
\end{itemize}
The second strategy is problematic even for $\operatorname{GL}_n$ as the classification of imprimitive representations is an open problem. Nevertheless, within the class of ladder representations, the imprimitive representations are easy to describe and one of the contributions of this work is to pursue this second approach for the problem of distinction with respect to Klyachko models.
Moreover, since the maximal proper subrepresentation of the standard module associated to a ladder representation has a particularly simple description we can implement the steps in the first strategy for these representations.
This makes distinction problems for the class of irreducible representations parabolically induced from the ladder representations more accessible.
\subsection{Main Results}
To simplify our exposition we omit from the introduction, the results whose formulation will require heavy notation. The interested reader should look in the body of the paper for further results of interest.
We state our main results in the form of Theorem \ref{thmA:main}, Theorem \ref{thmB:main} and Theorem \ref{thmC:main}. Additionally we will formulate a conditional Theorem \ref{thmD:main}.
Theorem \ref{thmA:main} concerns the distinction of representations of finite length with respect to Klyachko subgroups while Theorem \ref{thmB:main} (resp.~Theorem \ref{thmC:main}) provides a classification of the distinguished ladder representations with respect to the symplectic (resp.~general Klyachko) subgroup. Theorem \ref{thmD:main}, conditional on a certain combinatorial assumption (Hypothesis \ref{hyp*}), provides a complete classification of representations induced from ladder representations that are distinguished with respect to the symplectic group. Furthermore, assuming Hypothesis \ref{hyp*}, we provide a necessary consition for a standard module to be distinguished by the symplectic group.
For the sake of notational simplification let us say that a smooth finite length representation $\pi$ of $\operatorname{GL}_{n}(F)$ is $\operatorname{Sp}$-distinguished if $n$ is even and $\pi$ is $\operatorname{Sp}_n(F)$-distinguished.
We also require some of the notation and beautiful results of Zelevinsky \cite{MR584084}.
We recall that for an irreducible cuspidal representation $\rho$ of $\operatorname{GL}_{n}(F)$ and integers $a\le b$ one considers the segment
\[
\Delta=\Delta_{\rho}=[\nu^a\rho,\nu^b\rho]=\{\nu^i\rho:i=a,\dots,b\},
\]
where $\nu(g)=|\det(g)|$ for $g\in\operatorname{GL}_n(F)$. We set $b(\Delta)=a$ for the begining, $e(\Delta)=b$ for the end and $\ell(\Delta)=b-a+1$ for the length of $\Delta$. To $\Delta$ one associates a representation $Z(\Delta)$ and a representation $L(\Delta)$ as follows: $Z(\Delta)$ is the unique irreducible subrepresentation while $L(\Delta)$ is the unique irreducible quotient of the Bernstein-Zelevinsky product $\nu^a\rho \times \dots \times \nu^b\rho$.
To a multi-set $\mathfrak{m}=\{\Delta_{1},\dots,\Delta_{t}\}$ (a set with possible repetitions) of segments of irreducible cuspidal representations one associates an irreducible representation $Z(\mathfrak{m})$ and an irreducible representation $L(\mathfrak{m})$ as follows (See \ref{info: segments}): $Z(\mathfrak{m})$ is the unique irreducible submodule of the product $Z(\Delta_1) \times Z(\Delta_2) \times \dots \times Z(\Delta_{t})$ where we have arranged the segments $\Delta \in \mathfrak{m}$ in a standard form (See \ref{sec: standard}).
Analogously, the representation $L(\mathfrak{m})$ is the unique irreducible quotient of the standard module $\lambda(\mathfrak{m})=L(\Delta_1) \times L(\Delta_2) \times \dots \times L(\Delta_{t}).$
The Zelevinsky classification implies that the map $\mathfrak{m} \mapsto Z(\mathfrak{m})$ is a bijection between the set of such muti-sets of segments and the disjoint union of admissible duals of $\operatorname{GL}_n(F)$ for all $n$, while the Langlands classification implies that the map $\mathfrak{m} \mapsto L(\mathfrak{m})$ is a bijection between these sets.
\begin{theorem}\label{thmA:main}
\begin{enumerate}
\item \label{part nec}{\bf A necessary condition for $\operatorname{Sp}$-distinction} (See Proposition \ref{prop: Z dist}):
If $Z(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\ell(\Delta)$ is even for all $\Delta\in \mathfrak{m}$.
\item \label{part hered}{\bf Hereditary property for Klyachko models} (See Proposition \ref{prop: herad}):
Let $\pi_i$ be representations of finite length and $n_i=2k_i+r_i$ be such that $\pi_i$ is $(H_{2k_i,r_i},\psi)$-distinguished for $i=1,\dots,t$. Then $\pi=\pi_1\times\cdots\times\pi_t$ is $(H_{2k,r},\psi)$-distinguished where $k=k_1+\cdots+ k_t$ and $r=r_1+\cdots+r_t$.
\item \label{part cusp line}{\bf Reduction to cuspidal lines} \label{part3}(See Proposition \ref{lem: Kly cusp}):
Let $\pi_i$ be representations of finite length, $i=1,\dots,t$, such that their cuspidal supports, ${\rm Supp}(\pi_i)$ and ${\rm Supp}(\pi_j)$ are totally disjoint for all $i\ne j$ (See \ref{def: cusp supp}). Then $\pi=\pi_1\times\cdots\times\pi_t$ admits a Klyachko model if and only if $\pi_i$ admits a Klyachko model for all $i=1,\dots,t$.
\end{enumerate}
\end{theorem}
The Zelevinsky classification implies that any irreducible representation $\pi$ of $\operatorname{GL}_n(F)$ can be written as a product $\pi=\pi_{1} \times \dots \times \pi_{t}$ where the cuspidal support ${\rm Supp}(\pi_{i})$ is contained in a cuspidal line
and ${\rm Supp}(\pi_{i})$ , ${\rm Supp}(\pi_{j})$ are disjoint for $i \ne j.$ This is sometimes called a decomposition of $\pi$ into a product of irreducible {\it rigid} representations. Now, using Theorem \ref{thmA:main} \eqref{part3}, the study of distinction with respect to the Klyachko groups is reduced to the study of distinction within the class of rigid irreducible representations.
We say that a rigid irreducible representation $\pi=L(\mathfrak{m})$ is a ladder representation if the multi-set of segments $\mathfrak{m}=\{\Delta_{1},\dots, \Delta_{t}\}$ satisfies the conditions $b(\Delta_{1})> \dots > b(\Delta_{t})$ and $e(\Delta_{1})> \dots > e(\Delta_{t}).$
The next theorems provide the classification of distinguished ladder representations in terms of Langlands classification. We begin with $\operatorname{Sp}$-distinguished representations since the result is easier to formulate.
\begin{theorem}\label{thmB:main}
{\bf $\operatorname{Sp}$-distinguished Ladder representations} (See Theorem \ref{thm: dist lad}):
Let $L(\mathfrak{m})$ be a ladder representation with $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}$. Then $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished if and only if, $t$ is even and $\Delta_{2i-1}=\nu\Delta_{2i}$ for all $i=1,\dots,t/2$.
\end{theorem}
The condition on $\mathfrak{m}$ in Theorem \ref{thmB:main} is equivalent to the existence of a multi-set of segments $\mathfrak{n}$ such that $\mathfrak{m}=\mathfrak{n}+\nu \mathfrak{n}.$
We call such $\mathfrak{m}$ a multi-set of {\it Speh} type (See \ref{def: Sp type}).
For the next theorem we need the notion of right-alignment (See \ref{def: ra seg}). For segments $\Delta=[\nu^{a}\rho, \nu^{b}\rho]$ and $\Delta'=[\nu^{a'}\rho, \nu^{b'}\rho]$ we say that $\Delta'$ is {\it right-aligned} with $\Delta$ and write $\Delta' \vdash \Delta$ if
$a\ge a'+1$ and $ b= b'+1$.
When $\rho$ is a representation of $\operatorname{GL}_{d}(F)$ we label this relation by the integer $r=d(a-a'-1)$ and write $\Delta'\vdash_r \Delta$.
Our description of ladder representations distinguished with respect to Klyachko groups will be given in two steps.
We will say that a ladder representation $\pi=L(\mathfrak{m})$ is a {\it proper ladder} if for all $i=1,\dots, t-1$ we have $e(\Delta_{i+1}) \geq b(\Delta_{i})-1.$ The proper ladder representations are imprimitive and every ladder representation is a product of proper ladders in an essentially unique way. This decomposition into proper ladders is explicit in terms of the underlying multi-set of segments associated to the ladder representation.
\begin{theorem}\label{thmC:main}
{\bf Ladder representations distinguished with respect to Klyachko groups:}
\begin{enumerate}
\item (See Proposition \ref{prop: prop lad})
Let $L(\mathfrak{m})$ be a proper ladder representation of $\operatorname{GL}_{n}(F)$ with $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}$ and let $n=2k+r$.
\begin{itemize}
\item If $t$ is even then the representation $L(\mathfrak{m})$ is $(H_{2k,r},\psi)$-distinguished if and only if $\Delta_{t-2i}\vdash_{r_i}\Delta_{t-2i-1}$
for some $r_i$ $(i=0,\dots,t/2-1)$ and $r=r_0+\cdots +r_{t/2-1}$.
\item If $t$ is odd, let $s$ be such that $L(\Delta_1)$ is an irreducible representation of $\operatorname{GL}_{s}(F).$
The representation $L(\mathfrak{m})$ is $(H_{2k,r},\psi)$-distinguished if and only if $\Delta_{t-2i}\vdash_{r_i}\Delta_{t-2i-1}$
for some $r_i$ $(i=0,\dots,(t-3)/2)$ and $r=r_0+\cdots +r_{(t-3)/2}+s$.
\end{itemize}
\item (See Theorem \ref{thm: kly lad})
Let $\pi=\pi_{1} \times \dots \times\pi_{t}$ be the decomposition of the ladder representation $\pi$ into proper ladder representations $\pi_{i}$, $i=1, \dots, t.$ Then $\pi$ admits a Klyachko model if and only if the proper ladder representations $\pi_{i}$ admit Klyachko models for all $i=1,\dots,t$.
\end{enumerate}
\end{theorem}
Our last main result contains in its formulation a certain combinatorial property of multi-sets of segments that we call
Hypothesis \ref{hyp*} (See section \S 8). Roughly speaking, it says that the restrictions imposed by the geometric lemma on $\operatorname{Sp}$-distinction of a standard module $\lambda(\mathfrak{m})$ imply that $\mathfrak{m}$ is of Speh type.
For more details see Section \ref{into: pfs}.
\begin{theorem}\label{thmD:main}
\begin{enumerate}
\item {\bf On $\operatorname{Sp}$-Distinguished Standard modules} (See Proposition \ref{prop: dist hyp}):
Suppose $\lambda(\mathfrak{m})$ is rigid and $\operatorname{Sp}$-distinguished. Assume further that $\mathfrak{m}$ satisfies Hypothesis \ref{hyp*}. Then $\mathfrak{m}$ is of Speh type.
In particular, if $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished and $\mathfrak{m}$ satisfies Hypothesis \ref{hyp*} then $\mathfrak{m}$ is of Speh type.
\item \label{part prod lad}{\bf Distinction for irreducible products of ladder representations} (See Proposition \ref{prop: li}):
Assume Hypothesis \ref{hyp**} holds true for all multisegement.
Let $\pi_1,\dots,\pi_k$ be ladder representations such that $\pi=\pi_1\times\cdots\times\pi_k$ is an irreducible representation.
If $\pi$ is $\operatorname{Sp}$-distinguished then $\pi_i$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,k$.
\item \label{part: pure}{\bf Purity of symplectic distinction within ladder class} (See Corollary \ref{cor: prd two}):
Let $\pi_1$ and $\pi_2$ be ladder representations such that $\pi=\pi_1\times\pi_2$ is irreducible.
If $\pi$ is $\operatorname{Sp}$-distinguished then $\pi_1$ and $\pi_2$ are $\operatorname{Sp}$-distinguished.
\end{enumerate}
\end{theorem}
We emphasize that Theorem \ref{thmD:main} \eqref{part: pure} is unconditional. We further show in Proposition \ref{prop: hyp set} that Hypothesis \ref{hyp*} is satisfied by multi-sets that are in fact sets.
This yields the following unconditional result.
\begin{theorem}(See Corollary \ref{cor: set case})
Let $\lambda(\mathfrak{m})=L(\Delta_1)\times\cdots\times L(\Delta_t)$ be a standard module such that $\Delta_i\ne \Delta_j$ for all $i\ne j$. If $\lambda(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type.
\end{theorem}
\subsection{Proofs and Methods}\label{into: pfs}
Let us now elaborate a bit on the techniques used in the proofs. The filtration of the geometric lemma allows us to study $\operatorname{Sp}$-distinction of induced representations from the parabolic subgroup $P$ in terms of the geometry of $P$-orbits on the symmetric space $\operatorname{GL}_{2n}(F)/\operatorname{Sp}_{2n}(F).$ In particular we show that an induced $\operatorname{Sp}$-distinguished representation admits a $P$-orbit which is {\it relevant}. Analyzing the relevant orbits together with the Jacquet module calculations of segment representations allows us to prove Theorem \ref{thmA:main} \eqref{part nec}. For Theorem \ref{thmA:main} \eqref{part hered}, we combine the theory of derivatives with a meromorphic continuation technique of Blanc and Delorme. The first is used to reduce the problem to the case of $\operatorname{Sp}$-distinction and the second to construct $\operatorname{Sp}$-invariant linear forms on families of induced representations.
A key ingredient for the proof of Theorems \ref{thmB:main} and \ref{thmD:main} is the necessary condition for a standard module to be $\operatorname{Sp}$-distinguished provided by the geometric lemma.
This allows us to reduce the problem to a purely combinatorial one on multi-sets of segments. We address it under a technical hypothesis that we can prove only for certain multi-sets (in particular, whenever they are sets). The hypothesis can be interpreted as a statement that certain orbits, of the natural action of a parabolic subgroup $P$ of $\operatorname{GL}_{2n}(F)$ on $\operatorname{GL}_{2n}(F)/\operatorname{Sp}_{2n}(F)$, do not contribute a non-trivial $\operatorname{Sp}_{2n}(F)$-invariant linear form. Explicitly, assuming the hypothesis we show that if a representation, irreducibly induced from ladder representations, is $\operatorname{Sp}$-distinguished then each ladder representation in the inducing data also admits a symplectic model.
This partial result along with the description of the maximal proper subrepresentation of the standard module associated to a ladder representation allows us to finish the
proof of Theorem \ref{thmB:main}.
The proof of Theorem \ref{thmC:main} is obtained by using the explicit knowledge of the structure of a Jacquet module of a ladder representation, and the classification of $\operatorname{Sp}$-distinguished ladders given by Theorem \ref{thmB:main}.
To prove Theorem \ref{thmD:main} \eqref{part prod lad} we use a recent irreducibility result of \cite{1411.6310}, as well as the invariance of the class of ladder representations with respect to the Zelevinsky involution, to construct an inductive set-up in the context of representations irreducibly induced from ladder representations. We hope that the techniques developed in doing so will be useful more generally. Most notably, when we attempt to study distinction by other closed subgroups of $\operatorname{GL}_{n}(F)$ for representations that are irreducibly induced from ladders.
\subsection{Related Works}
We mention here a few works where the results or the tools used have some intersection with the present work.
The present work began as an attempt to extend the distinction results of \cite{MR2332593} and \cite{MR2417789} from the unitary dual of $\operatorname{GL}_{n}(F)$ to the admissible dual of $\operatorname{GL}_{n}(F).$
We emphasize that in \cite{MR2332593} and \cite{MR2417789} obtaining a model for a Speh representation, in particular the classification of Speh representations admitting a symplectic model, was based on the global theory of period integrals of Eisenstein series and their residues obtained in \cite{MR2254544} and \cite{MR2248833}. A novel aspect of this work is that our method of proof is purely local, and therefore, independently provides a local proof for the results of the aforementioned works.
The methods employed here are very different from those works and are, in fact, closer in spirit to the techniques of \cite{MR1078382}, or to that of \cite{MR3227442} which studies admissible representations distinguished with respect to a symplectic group in small rank cases.
The focus on distinction problems within the class of ladder representations
was made in \cite{MR2930996} and later in \cite{MR3421655} and Theorem \ref{thmB:main} of the introduction could be considered as an analogue to their results.
A study parallel to our study of the distinction problem for standard modules can be found in these two references.
In \cite{MR2889169} the existence of Klyachko models is proved for unitary representations of $\operatorname{GL}_{n}(\mathbb{R})$ and $\operatorname{GL}_{n}(\mathbb{C}).$ The methods there are parallel to those in the works of the second and third author in the non-archimedean case. In particular, the proof of existence of those models for Speh representations was based on the theory of periods of automorphic forms. Recently, in \cite{MR3416438} a local construction of these invariant functionals is provided, based on tools from the theory of distributions and $D$-modules. Some of the results of the present work can be considered as a non-archimedean analogue of the main result of \cite{MR3416438}.
\subsection{Structure of the Paper}
Let us now delineate the contents of this paper. A large part of it (\S \ref{notn}-\ref{s: dist ladder}) concerns symplectic models.
After setting up the general notation for this work in \S \ref{notn}, we recall some well known results concerning $\operatorname{GL}_{2n}(F)/\operatorname{Sp}_{2n}(F)$ in \S \ref{s: sym sp}, especially the structure of orbits of the natural action of a parabolic subgroup of $\operatorname{GL}_{2n}(F)$. Our main tool for studying $\operatorname{Sp}$-distinction of induced representations is an application of the geometric lemma of Bernstein and Zelevinsky. This is recalled in \S \ref{sec: gl}.
In \S \ref{ss: oc} we obtain some immediate consequences for $\operatorname{Sp}$-distinction of certain induced representations. They come from contributions to the open and to the closed orbits of the aforementioned action. In \S \ref{ss: cusp lines} we reduce the classification of $\operatorname{Sp}$-distinguished irreducible representations to those supported in a single cuspidal line viz. the rigid representations.
In order to study distinction for rigid representations, we recall in \S \ref{s: LZ cl} the segment notation of Zelevinsky and the classification of the admissible dual. In \S \ref{s: nec Z} we provide a necessary condition for an irreducible representation to be $\operatorname{Sp}$-distinguished in terms of the Zelevinsky classification (Proposition \ref{prop: Z dist}).
We then turn to the study of distinction of standard modules. A necessary condition for a standard module (and for an irreducible representation) to be $\operatorname{Sp}$-distinguished is reduced in \S \ref{s: std mod sp} to a combinatorial problem. This problem is formulated in \S \ref{s: multisets} as Hypothesis \ref{hyp*}.
Section \ref{s: multisets} is written in a way completely independent from the rest of the paper, is accessible to any mathematician, and presents a problem with applications to the study of $\operatorname{Sp}$-distinction.
Our partial results suffice in order to obtain a complete classification of $\operatorname{Sp}$-distinguished ladder representations. This is Theorem \ref{thm: dist lad}.
In \S \ref{s: dist ladder} we obtain our results on $\operatorname{Sp}$-distinction for the class of representations irreducibly induced from ladder. These are conditional on Hypothesis \ref{hyp*}. Again these results require some purely combinatorial lemmas that we prove in \S \ref{s: ladder sums}.
In \S \ref{s: Kl} we turn to the study of Klyachko models in general. We prove the hereditary property with respect to parabolic induction (Proposition \ref{prop: herad}) and reduce the problem to rigid representations Proposition \ref{lem: Kly cusp}). Finally the classification of ladder representation with a given Klyachko model is obtained in \S \ref{s: kl ladder}.
\subsection{Acknowledgments}
The authors are grateful to Erez Lapid for sharing with them his insights on ladder representations and irreducibility results.
\section{Notation and preliminaries}\label{notn}
We set the general notation in this section. More particular notation is defined in the section where it first occurs.
\subsection{Generalities}
Let $G$ be a totally disconnected, locally compact group.
\subsubsection{}
Let $\delta_G$ be the modulus function of $G$ with the convention that $\delta_G(g)dg$ is a right-invariant Haar measure if $dg$ is a left-invariant Haar measure on $G$.
\subsubsection{}
Let $H$ be a closed subgroup of $G$ and $\sigma$ a smooth, complex-valued representation of $H$.
We denote by $\operatorname{Ind}_H^G(\sigma)$ the normalized induced representation. It is the representation of $G$ by right translations on the space of functions $f$ from $G$ to the space of $\sigma$ satisfying
\[
f(hg)=(\delta_H^{1/2}\delta_G^{-1/2})(h)\sigma(h)f(g), \ \ \ h\in H,\,g\in G
\]
and $f$ is right invariant by some open subgroup of $G$.
The representation of $G$ on the subspace of functions with compact support modulo $H$ is denoted by ${\rm ind}_H^G(\sigma)$.
\subsubsection{}
This paper is concerned with distinguished representations in the following sense.
\begin{definition}
Let $\pi$ be a smooth, complex-valued representation of $G$ and $H$ a closed subgroup of $G$.
\begin{itemize}
\item We say that $\pi$ is \emph{$H$-distinguished} if there exists a non-zero $H$-invariant linear form $\ell$ on the space of $\pi$, i.e., $\ell(\pi(h)v)=\ell(v)$ for all $h\in H$ and $v$ in the space of $\pi$. We denote by $\operatorname{Hom}_H(\pi,1)$ the space of $H$-invariant linear forms on $\pi$.
\item More generally, for a character $\chi$ of $H$ we say that $\pi$ is $(H,\chi)$-distinguished if the space $\operatorname{Hom}_H(\pi,\chi)$ of $H$-equivariant linear forms on $\pi$ is non-zero.
\end{itemize}
\end{definition}
By Frobenius reciprocity we have a natural linear isomorphism
\begin{equation}\label{eq: frob rec}
\operatorname{Hom}_H(\pi,\chi \delta_H^{1/2}\delta_G^{-1/2})\simeq\operatorname{Hom}_G(\pi,\operatorname{Ind}_H^G(\chi)).
\end{equation}
\subsubsection{} We state the following simple observation.
\begin{lemma}\label{drmk: ist quot}
Let $\pi$ and $\sigma$ be smooth, complex-valued representations of $G$ so that $\sigma$ is a quotient of $\pi$, $H$ is a closed subgroup of $G$ and $\chi$ is a character of $H$. If $\sigma$ is $(H,\chi)$-distinguished then $\pi$ is $(H,\chi)$-distinguished.
\end{lemma}
\begin{proof}
Note that composition with the projection $\pi\rightarrow \sigma$ defines an imbedding $\operatorname{Hom}_H(\sigma, \chi)\hookrightarrow \operatorname{Hom}_H(\pi,\chi)$. The lemma follows.
\end{proof}
The lemma allows us to reduce some distinction questions to induced representations (e.g. using the Langlands classification). Its converse need not be true.
\subsubsection{}
We record here another simple observation related to the converse problem, distinction of subquotients of a distinguished representation.
\begin{lemma}\label{lem: dist comp}
Let $\pi$ be a smooth complex-valued representation of $G$, $H$ a closed subgroup of $G$ and $\chi$ a character of $H$.
Let $0=\pi_0\subseteq \pi_1\subseteq\cdots\subseteq \pi_k=\pi$ be a filtration of $\pi$ by sub-representations.
If $\pi$ is $(H,\chi)$-distinguished, then there exists $i\in\{1,\dots,k\}$ such that $\pi_i/\pi_{i-1}$ is $(H,\chi)$-distinguished.
In particular, if $\pi$ is of finite length and $(H,\chi)$-distinguished then there exists an irreducible subquotient $\sigma$ of $\pi$ that is $(H,\chi)$-distinguished.
\end{lemma}
\begin{proof}
If $0\ne \ell\in\operatorname{Hom}_H(\pi,\chi)$ then there exists $i\in\{1,\dots,k\}$ minimal such that $\ell|_{\pi_i}\ne 0$. Thus, $\ell$ defines a non-zero element of $\operatorname{Hom}_H(\pi_i/\pi_{i+1},\chi)$. Since a finite length representation has such a finite filtration with irreducible quotients the rest of the lemma follows.
\end{proof}
\subsubsection{}
Let $\Pi(G)$ be the category of complex valued, smooth, admissible representations of $G$ of finite length and ${\rm Irr}(G)$ the class of irreducible representations in $\Pi(G)$.
Let $\pi^\vee$ denote the contragredient of a representation $\pi\in \Pi(G)$. Then $(\pi^\vee)^\vee\simeq\pi$ and $\pi\in {\rm Irr}(G)$ if and only if $\pi^\vee\in {\rm Irr}(G)$.
\subsection{Notation for $\operatorname{GL}_n(F)$}
Let $F$ be a non-archimedean local field of characteristic different than two. For $n\in \mathbb{N}$, let $G_n=\operatorname{GL}_n(F)$. By convention, let $G_0$ be the trivial group.
\subsubsection{}
Fix $n$ and let $G=G_n$.
Let $B=T \ltimes N$ be the standard Borel subgroup of $G$ consisting of uppertriangular matrices with its standard Levi decomposition. Here $T$ is the subgroup of diagonal matrices and $N=N_n$ is the unipotent radical of $B$.
\subsubsection{}
A parabolic subgroup of $G$ that contains $B$ is called standard. Standard parabolic subgroups of $G$ are in bijection with decompositions of $n$.
For a decomposition $\alpha=(n_1,\dots,n_k)$ of $n$ let $P_\alpha=M_\alpha\ltimes U_\alpha$ be the standard parabolic subgroup of $G$ consisting of block uppertriangular matrices with standard Levi subgroup
\[
M_\alpha=\{\operatorname{diag}(g_1,\dots,g_k):g_i\in G_{n_i},\,i=1,\dots,k\}\simeq G_{n_1}\times \cdots\times G_{n_k}
\]
and unipotent radical $U_\alpha$.
\subsubsection{}
The Weyl group $N_G(T)/T$ of $G$ is isomorphic to the permutation group $S_n$ of $n$ elements. We identify it with the subgroup $W=W_G$ of permutation matrices in $G$. Let $w_n=(\delta_{i,n+1-j})\in W$ be the longest Weyl element.
By the Bruhat decompositon $W$ is a complete set of representatives for the double coset space $B\backslash G/B$.
\subsubsection{}
More generally, for a decomposition $\alpha=(n_1,\dots,n_k)$ of $n$, $W_{M_\alpha}=W\cap M_\alpha\simeq S_{n_1}\times\cdots\times S_{n_k}$ is the Weyl group of $M_\alpha$.
If $P=M\ltimes U$ and $Q=L\ltimes V$ are standard parabolic subgroups of $G$ with their standard Levi decompositions then $w\mapsto PwQ$ defines a bijection
\[
W_M\backslash W/W_L\simeq P\backslash G/Q.
\]
Further more, every double coset in $W_M\backslash W/W_L$ contains a unique element of minimal length. Denote by ${}_MW_L$ the set of elements $w\in W$ that are of minimal length in $W_M wW_L$. Then ${}_MW_L$ is a complete set of representatives for $P\backslash G/Q$.
For every $w\in {}_MW_L$ the group
\[
P(w)=M(w)\ltimes U(w)=M\cap wQw^{-1}
\]
is a standard parabolic subgroup of $M$ with its standard Levi decomposition, where
\[
M(w)=M\cap wLw^{-1}\ \ \ \text{and}\ \ \ U(w)=M\cap wVw^{-1}.
\]
\subsection{Representations of $\operatorname{GL}_n(F)$}
We recall some well known facts and set the notation for representations of $G_n$.
Let $\Pi$ be the disjoint union of $\Pi(G_n)$ for all $n\in \mathbb{Z}_{\ge 0}$. Let ${\rm Irr}$ be the subset of irreducible representations in $\Pi$ and ${\rm Cusp}$ be the subset of cuspidal representations in ${\rm Irr}$.
\subsubsection{Parabolic Induction}
Set $G=G_n$. Let $P=M\ltimes U$ and $Q=L\ltimes V$ be standard parabolic subgroups of $G$ with their standard Levi decompositions.
Assume further that $Q$ is a subgroup of $P$. The functor ${\bf i}_{M,L}:\Pi(L)\rightarrow\Pi(M)$ of normalized parabolic induction is defined as follows. As noted above $M\cap Q=L\ltimes (M\cap V)$ is a standard parabolic subgroup of $M$. For $\rho\in \Pi(L)$ we consider $\rho$ as a representation of $M\cap Q$ trivial on its unipotent radical $M\cap V$ and set
\[
{\bf i}_{M,L}(\rho)={\rm ind}_{M\cap Q}^M(\rho).
\]
The functor ${\bf i}_{M,L}$ is exact and we have
\[
{\bf i}_{M,L}(\rho)^\vee\simeq {\bf i}_{M,L}(\rho^\vee).
\]
Let $\alpha=(n_{1},\dots,n_{k})$ be a decomposition of $n$. Assume that $M=M_\alpha$ and let $\rho_i\in\Pi(G_{n_i})$, $i=1,\dots,k$. Then $\rho=\rho_1\otimes \cdots \otimes \rho_k\in \Pi(M)$.
Set
\[
\rho_1\times\cdots \times \rho_k={\bf i}_{G,M}(\rho).
\]
\subsubsection{Jacquet module}
The functor ${\bf i}_{M,L}$ admits a left adjoint, namely, the normalized Jacquet functor ${\bf r}_{L,M}:\Pi(M)\rightarrow\Pi(L)$. For $\sigma\in \Pi(M)$, ${\bf r}_{L,M}(\sigma)$ is the representation of $L$ on the space of $V\cap M$-coinvariants of $\sigma$ induced by the action $\delta_{Q\cap M}^{-1/2}\sigma$. It is also an exact functor and for $\sigma\in \Pi(M)$ and $\rho\in \Pi(L)$ we have the natural linear isomorphism (Frobenius reciprocity):
\begin{equation}\label{eq: 1st adj}
\operatorname{Hom}_M(\sigma,{\bf i}_{M,L}(\rho))\simeq \operatorname{Hom}_L({\bf r}_{L,M}(\sigma),\rho).
\end{equation}
Let $\beta_i$ be the decomposition of $n_i$, $i=1,\dots,k$ so that $L=M_{(\beta_1,\dots,\beta_k)}$. For representations $\pi_i\in \Pi(G_{n_i})$, $i=1,\dots,k$ we have
\begin{equation}\label{eq: trans jm}
{\bf r}_{L,M}(\pi_1\otimes\cdots\otimes\pi_k)={\bf r}_{M_{\beta_1}, G_{n_1}}(\pi_1)\otimes\cdots\otimes {\bf r}_{M_{\beta_k}, G_{n_k}}(\pi_k).
\end{equation}
\subsubsection{The cuspidal support}\label{def: cusp supp}
For every $\pi\in {\rm Irr}$ there exist $\rho_1,\dots,\rho_k\in {\rm Cusp}$, unique up to rearrangement, so that $\pi$ is isomorphic to a subrepresentation of $\rho_1\times \cdots \times \rho_k$. Let ${\rm Supp}(\pi)=\{\rho_i:i=1,\dots,k\}$ be the support of $\pi$.\footnote{The support is often considered as a multi-set. Only the underlying set is relevant to us.}
For $\sigma_1,\dots,\sigma_k\in {\rm Irr}$ let ${\rm Supp}(\sigma_1\otimes\cdots\otimes\sigma_k)=\cup_{i=1}^k{\rm Supp}(\sigma_i)$.
For any standard Levi subgroup $M$ of $G$ and $\pi\in \Pi(M)$ let $\{\pi_1,\dots,\pi_t\}$ be the set of irreducible components (subquotients) of $\pi$ and set
\[
{\rm Supp}(\pi)=\cup_{i=1}^t {\rm Supp}(\pi_i).
\]
As a simple consequence of the geometric lemma of Bernstein and Zelevinsky \cite[\S 2.12]{MR0579172} and exactness we have
\begin{equation}\label{eq: supp jm}
{\rm Supp}({\bf r}_{M,G}(\pi))\subseteq{\rm Supp}(\pi), \ \pi\in \Pi(G).
\end{equation}
\subsubsection{Generic representations}\label{sss: gen}
Let $\psi$ be a non-trivial character of $F$.
We further denote by $\psi=\psi_n$ the character of $N_n$ defined by
\[
\psi(u)=\psi(\sum_{i=1}^{n-1} u_{i,i+1}), \ u=(u_{i,j})\in N_n.
\]
\begin{definition}
A representation $\pi\in\Pi(G_n)$ is called \emph{generic} if it is $(N_n,\psi)$-distinguished.
\end{definition}
\section{Non-degenerate skew-symmetric matrices and parabolic orbits}\label{s: sym sp}
We recall here the analysis of double cosets and related data that are relevant to the study of induced representations of $G_{2n}$ that are distinguished by the symplectic group.
\subsection{The symmetric space}
Fix $n\in N$ and let $G=G_{2n}$.
\subsubsection{}\label{sss: sp not}
Let
\[
H=H_n=\operatorname{Sp}_{2n}(F)=\{g\in G:{}^t g J g=J\}
\]
where
\[
J=J_n=\begin{pmatrix} & w_n \\ -w_n & \end{pmatrix}.
\]
Note that $H=G^\theta$ is the group of fixed points in $G$ of the involution $\theta$ defined by
\[
\theta(g)=J\,{}^t g^{-1} J^{-1}.
\]
\subsubsection{}
For $\pi\in\Pi(G_n)$, since $n$ will not always be specified, we adopt throughout the following convention. We say that $\pi$ is $\operatorname{Sp}$-distinguished if $n$ is even and $\pi$ is $\operatorname{Sp}_n(F)$-distinguished. If in addition $\pi\in {\rm Irr}$ we say that $\pi$ admits a symplectic model (see \eqref{eq: frob rec}).
Recall the following simple observation.
By a result of Gelfand and Kazhdan, \cite{MR0404534}, we have $\pi^\theta\simeq \pi^\vee$ for every $\pi\in {\rm Irr}$. We therefore have
\begin{lemma}\label{lem: cont dist}
A representation $\pi\in{\rm Irr}$ is $\operatorname{Sp}$-distinguished if and only if $\pi^\vee$ is $\operatorname{Sp}$-distinguished. \qed
\end{lemma}
\subsubsection{}
Consider the symmetric space
\[
X=\{x\in G: x\theta(x)=I_{2n}\}
\]
with the $G$-action
\[
g\cdot x=gx\theta(g)^{-1}.
\]
Note that $XJ$ is the space of skew-symmetric matrices in $G$ and
\[
(g\cdot x)J=g(xJ){}^tg.
\]
Therefore $X$ is a homogeneous $G$-space.
The map $(g\mapsto g\cdot I_{2n}): G\rightarrow X$ defines an isomorphism $G/H\simeq X$ of $G$-spaces.
\subsubsection{}
For a subgroup $Q$ of $G$ and a $Q$-invariant subspace $Y$ of $X$ we denote by $Q\backslash Y$ the set of $Q$-orbits in $Y$. For $x\in X$ let $Q_x=\{g\in Q:g\cdot x=x\}$ be the stabilizer in $Q$ of $x$.
Applications of the geometric lemma to the study of $H$-distinguished induced representations of $G$ require the study of orbits in $X$ by the group from which we induce. Since parabolic induction is central to the classification of ${\rm Irr}$ we recall next the study of orbits in $X$ under a standard parabolic subgroup $P=M\ltimes U$.
Of particular interest for these applications are choices of orbit representatives $x$ for which we can provide explicit description of the stabilizer $M_x$ and the restriction to $M_x$ of $\delta_P^{1/2}\delta_{P_x}^{-1}$.
We refer to \cite[\S3]{MR2254544} and \cite[\S3.1]{MR2248833} for proofs of the results presented in Sections \ref{X/B} and \ref{X/P}.
\subsection{Borel orbits in $X$}\label{X/B}
We begin with the Borel orbits.
\subsubsection{}
Note that both $B$ and $T$ are $\theta$ stable.
In particular $\theta$ defines an involution on $W$ that we continue to denote by $\theta$. We have
\[
\theta(w)=w_{2n}ww_{2n}^{-1},\,w\in W.
\]
It also follows that the map $( B\cdot x\mapsto BxB ): B\backslash X \rightarrow B\backslash G/B$ is well defined. By the Bruhat decomposition this defines a map from $B\backslash X$ to $W$. Since $\theta(BxB)=(BxB)^{-1}$ it follows that every $w$ in the image of this map satisfies $w\theta(w)=e$ (the identity element of $W$). We refer to such permutations as twisted involutions.
\subsubsection{}
Let
\[
[w_{2n}]=\{ww_{2n}w^{-1}:w\in W\}
\]
be the $W$-conjugacy class of the longest Weyl element. It is the set of involutions without fixed points in $W$.
Note that the set of twisted involutions in $W$ is precisely $[w_{2n}]w_{2n}$. In fact we have
\begin{lemma}
The map $( B\cdot x\mapsto B\cdot x\cap T ):B\backslash X\rightarrow T\backslash (X\cap N_G(T))$ and the natural map from $T\backslash (X\cap N_G(T))$ to $[w_{2n}]w_{2n}$ are both bijective. \qed
\end{lemma}
\subsubsection{}
Fix a Borel orbit $B\cdot x\in B\backslash X$. We may and do assume that $x\in X\cap N_G(T)$ and let $w\in [w_{2n}]$ be such that $x\in T ww_{2n}$.
Below is a description of $T_x$ and of the restriction to $T_x$ of $\delta_B^{1/2}\delta_{B_x}^{-1}$.
Note first that
\[
T_{I_{2n}}=T\cap H=\{\operatorname{diag}(a_1,\dots,a_n,a_n^{-1},\dots,a_1^{-1}): a_i\in F^*,\,i=1,\dots,n\}
\]
and
\[
(\delta_B^{-1/2}\delta_{B_{I_{2n}}})(\operatorname{diag}(a_1,\dots,a_n,a_n^{-1},\dots,a_1^{-1}))=\abs{a_1}_F\cdots\abs{a_n}_F.
\]
We summarize the relevant results.
\begin{lemma}
For every orbit in $B\backslash X$ there exists a unique $\tau\in [w_{2n}]$ and a representative $x\in N_G(T)$ such that
\begin{enumerate}
\item $T_x=\{\operatorname{diag}(a_1,\dots,a_{2n})\in T: a_{\tau(i)}=a_i^{-1},\,i=1,\dots,n\}$;
\item $(\delta_B^{-1/2}\delta_{B_x})(t)=\prod_{i<\tau(i)}\abs{a_i}_F$ for every $t=\operatorname{diag}(a_1,\dots,a_{2n})\in T_x$. \qed
\end{enumerate}
\end{lemma}
\subsection{$P$-orbits in $X$}\label{X/P}
Let $\alpha=(n_1,\dots,n_k)$ be a decomposition of $2n$ and let $P=P_\alpha=M\ltimes U$. Note that $\theta(P)=P_{(n_k,\dots,n_1)}$ and $\theta(M)=w_{2n}Mw_{2n}^{-1}=M_{(n_k,\dots,n_1)}$.
\subsubsection{} The $P$-orbits in $X$ are in bijection with certain twisted involutions.
\begin{lemma}\label{lem: P-bij}
The map $(P\cdot x\mapsto Px\theta(P) ) :P\backslash X\rightarrow P\backslash G/\theta(P)\simeq {}_MW_{\theta(M)}$ defines a bijection
\[
P\backslash X\simeq {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}.
\] \qed
\end{lemma}
\subsubsection{}
Let
\[
\imath_M: {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}\rightarrow P\backslash X
\]
denote the bijection of Lemma \ref{lem: P-bij}.
Recall that for $w\in {}_MW_{\theta(M)}$
\[
M(w)=M\cap w\theta(M)w^{-1}
\]
is a standard parabolic subgroup of $M$.
\begin{lemma}\label{lem good reps}
For every $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ we have that $\imath_M(w)\cap M(w)w$ is a single $M(w)$-orbit. In particular, it is not empty. \qed
\end{lemma}
\subsubsection{Admissible orbits}
\begin{definition}
We say that $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ (or the corresponding $P$-orbit $\imath_M(w)$) is $M$-admissible if $M(w)=M$, i.e., if $ww_{2n}\in N_G(M)$.
\end{definition}
Thus, $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ is $M$-admissible if and only if the intersection $\imath_M(w)\cap N_G(M)w_{2n}$ is not empty.
In particular $\imath_M$ restricts to a bijection
\[
({}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}\cap N_G(M)w_{2n})\ \ \simeq \ \ (M-\text{admissible orbits in}\ X).
\]
The $P$-orbits in $X$ are studied in terms of certain $L$-admissible orbits for Levi subgroups $L$ of $M$. More precisely, to the $P$-orbit $\imath_M(w)$ we associate a certain $M(w)$-admissible orbit.
We therefore begin by describing the relevant data for $M$-admissible orbits.
Let
\[
S_2[\alpha]=\{\tau\in S_k: \tau^2=e,\,n_{\tau(i)}=n_i,\,i=1,\dots,k\ \text{and}\ n_i\ \text{is even if} \ \tau(i)=i\}.
\]
The admissible $M$-orbits are in bijection with $S_2[\alpha]$. Before we state the general results we provide examples of prototypes of admissible orbits.
\subsubsection{}
Assume that $k=s+2t$, $n_i=n_{k+1-i}$, $i=1,\dots,t$ and $n_i$ is even for $i=t+1,\dots,t+s$, i.e., $\alpha$ is of the form
\[
\alpha=(n_1,\dots,n_t, 2m_1,\dots,2m_s,n_t,\dots,n_1).
\]
Let
\[
x=\operatorname{diag}(I_N,J_{(m_1,\dots,m_s)}J_m^{-1},I_N)=\begin{pmatrix} & & w_N \\ & J_{(m_1,\dots,m_s)} & \\ -w_N & & \end{pmatrix}J_n^{-1}\in X
\]
where
\[
J_{(m_1,\dots,m_s)}=\operatorname{diag}(J_{m_1},\dots,J_{m_s}),\ \ \ N=n_1+\cdots+n_t \ \ \ \text{and}\ \ \ m=m_1+\cdots+m_s.
\]
Note that $xJ_n$ is a skew-symmetric matrix in $N_G(M)$ and therefore $P\cdot x$ is $M$-admissible.
For every $d\in \mathbb{N}$ consider the involution $g\mapsto g^*$ on $G_d$ defined by $g^*=w_d\,{}^tg^{-1}w_d^{-1}$. We have
\[
M_x=\{\operatorname{diag}(g_1,\dots,g_t,h_1,\dots,h_s,g_t^*,\dots,g_1^*):g_1,\dots,g_t\in G_{n_i},\,h_1,\dots,h_s\in H_{m_j}\}
\]
and
\[
(\delta_P^{-1/2}\delta_{P_x})(\operatorname{diag}(g_1,\dots,g_t,h_1,\dots,h_s,g_t^*,\dots,g_1^*))=\prod_{i=1}^t \abs{\det g_i}_F.
\]
\subsubsection{}
We now return to the general setting where $\alpha$ is any decomposition of $2n$.
\begin{lemma}\label{admP-orb}
There is a bijection between the $M$-admissible $P$-orbits in $X$ and $S_2[\alpha]$ that satisfies the following properties. Let $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}\cap N_G(M)w_{2n}$ and let $\tau\in S_2[\alpha]$ correspond to $\imath_M(w)$. Then, there exists $x=x_{M,w}\in X\cap Mw$ such that:
\begin{enumerate}
\item $M_x=\{\operatorname{diag}(g_1,\dots,g_k):g_{\tau(i)}=g_i^* \ \text{if}\ \tau(i)\ne i\ \text{and} \ g_i\in H_{n_i/2}\ \text{if}\ \tau(i)=i\}$
\item $(\delta_P^{-1/2}\delta_{P_x})(\operatorname{diag}(g_1,\dots,g_k))=\prod_{i<\tau(i)} \abs{\det g_i}_F$. \qed
\end{enumerate}
\end{lemma}
\subsubsection{}
Every $\tau\in S_k$ defines a unique $w_\tau\in W$ such that for every $g=\operatorname{diag}(g_1,\dots,g_k)\in M$ we have
\[
w_\tau gw_\tau^{-1}=\operatorname{diag}(g_{\tau^{-1}(1)},\dots,g_{\tau^{-1}(k)}).
\]
\begin{remark}
In fact, the relation between $w$ and $\tau$ in the above lemma is characterized by $\operatorname{diag}(w_{n_1},\dots,w_{n_k}) w_\tau w_{2n}= w$.
\end{remark}
\subsubsection{General orbits}\label{General orbits}
Fix $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ and let $L=M(w)$ be the standard Levi subgroup of $M$ we associated with $w$.
Let $\beta=(\beta_1,\dots,\beta_k)$ be such that $L=M_\beta$ where $\beta_i=(m_{i,1},\dots, m_{i,k_i})$ is a decomposition of $n_i$. On the set of indices
\[
\mathfrak{I}=\{(i,j):i=1,\dots,k,\ j=1,\dots,k_i\}
\]
we consider the lexicographic order $(i,j) \prec (i',j')$ if either $i<i'$ or $i=i'$ and $j<j'$. We further consider the partial order $(i,j)\ll(i',j')$ if $i<i'$.
Recall that by Lemmas \ref{lem: P-bij} and \ref{lem good reps}, $X\cap Lw$ is an $L$-orbit. Note that $w$ is $L$-admissible. Furthermore, for $x\in X\cap Lw$ we have $M_x=L_x$ and $P_x=Q_x$ where $Q=P_\beta$ is the standard parabolic subgroup of $G$ with Levi subgroup $L$. We may therefore apply Lemma \ref{admP-orb} with $M$ replaced by $L$.
We consider $S_2[\beta]$ as a set of involutions on $\mathfrak{I}$, by identifying $(\mathfrak{I},\prec)$ with the linearly ordered set $\{1,2,\dots,\abs{\mathfrak{I}}\}$. Let $\tau\in S_2[\beta]$ be the involution associated with $\imath_L(w)$.
Since $w\in {}_MW_{\theta(M)}$ there are more restrictions on $\tau$, it must satisfy
\begin{equation}\label{eq: stronger cond}
\tau(i,j+1)\ll \tau(i,j),\ i=1,\dots,k,\,\ j=1,\dots,k_i-1.
\end{equation}
This implies in particular that for every $i$ there is at most one $j$ such that $\tau(i,j)=(i,j)$.
\section{The geometric lemma}\label{sec: gl}
We recall here a special case of the geometric lemma \cite[Theorem 5.2]{MR0579172} (see also \cite[Proposition 1.17]{MR2401221}).
As in the previous section, fix $n\in \mathbb{N}$ and let $G=G_{2n}$ and $H=H_n$.
Let $\alpha=(n_1,\dots,n_k)$ be a decomposition of $2n$ and let $P=P_\alpha=M\ltimes U$.
Consider the functor $\operatorname{Res}_H\circ {\bf i}_{G,M}$ from $\Pi(M)$ to the category of smooth representations of $H$ where $\operatorname{Res}_H$ stands for restriction to $H$.
\subsection{The $H$-filtration}
For every $\sigma\in \Pi(M)$ we recall here the existence of an $H$-filtration on $\operatorname{Res}_H\circ {\bf i}_{G,M}(\sigma)$ parameterized by $P\backslash X$ and explicate the factors of the filtration.
\subsubsection{} By \cite[\S1.5]{MR0425030} (see also \cite[Lemma 3.1]{MR2401221}) there is a linear ordering ${}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}=\{w_1,\dots,w_m\}$ so that
\[
X_i=\cup_{j=1}^i \imath_M(w_i)
\]
is open in $X$ for all $i=1,\dots,m$.
\subsubsection{}\label{sss: clop}
The orbit of the identity $\iota_M(I_{2n})=P\cdot I_{2n}$ is closed in $X$ and we may assume that $w_m=I_{2n}$. Furthermore, if $n_i$ is even for all $i=1,\dots,k$ then the orbit $P\cdot x_M$ where
\[
x_M=J_{(n_1/2,\dots,n_k/2)}J_n^{-1}
\]
is open in $X$ and we may assume that $\imath_M(w_1)=P\cdot x_M$. Furthermore, in this case, $w_1$ is $M$-admissible.
\subsubsection{}
Let $\sigma\in \Pi(M)$ and let $\mathcal{V}$ be the representation space of ${\bf i}_{G,M}(\sigma)$. Set
\[
\mathcal{V}_i=\{\varphi\in\mathcal{V}:{\rm Supp}(\varphi)\subseteq X_i\},\ i=1,\dots,m
\]
then $\mathcal{V}_0:=0\subseteq \mathcal{V}_1 \subseteq \cdots\subseteq\mathcal{V}_m=\mathcal{V}$ is a filtration of $\operatorname{Res}_H({\bf i}_{G,M}(\sigma))$.
For every $i$ choose (by Lemma \ref{lem good reps}) $x_i\in \imath_M(w_i)\cap M(w_i)w_i$ and $\eta_i\in G$ such that $\eta_i\cdot I_{2n}=x_i$.
For a subgroup $A$ of a group $B$, $b\in B$ and a representation $\rho$ of $A$ we denote by $\rho^b$ the representation of $b^{-1}Ab$ on the space of $\rho$ defined by $\rho^b(b^{-1}ab)=\rho(a)$. By \cite[Proposition 3]{MR2248833} we have
\begin{lemma}\label{lem comp iso}
For every $i=1,\dots,m$ we have the isomorphism of representations of $H$
\[
\mathcal{V}_i/\mathcal{V}_{i-1}\simeq {\rm ind}_{H\cap \eta_i^{-1}P_{x_i}\eta_i}^H(\delta_{P_{x_i}}^{-1/2}(\operatorname{Res}_{P_{x_i}}(\delta_P^{1/2}\sigma))^{\eta_i}).
\]
\qed
\end{lemma}
\subsubsection{Relevant orbits}
\begin{definition}
We say that $w_i\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ (or $\imath_M(w_i)$) is relevant for $\sigma$ if
\[
\operatorname{Hom}_H(\mathcal{V}_i/\mathcal{V}_{i-1},1)\ne 0.
\]
\end{definition}
\subsubsection{} The following makes this property more explicit. By \cite[Corollary 1]{MR2248833} we have
\begin{lemma}\label{lem comp frob}
Fix $i$ and let $w=w_i$, $x=x_i$, $\eta=\eta_i$, $L=M(w_i)$ and $Q$ the standard parabolic subgroup of $G$ with Levi subgroup $L$. Then
\[
\operatorname{Hom}_H({\rm ind}_{H\cap \eta^{-1}P_x\eta}^H(\delta_{P_x}^{-1/2}(\operatorname{Res}_{P_x}(\delta_P^{1/2}\sigma))^{\eta}),1)\simeq \operatorname{Hom}_{L_x}({\bf r}_{L,M}(\sigma),\delta_Q^{-1/2}\delta_{Q_x}).
\]
\qed
\end{lemma}
\subsubsection{} Combining Lemmas \ref{lem comp iso} and \ref{lem comp frob} we have
\begin{corollary}\label{cor: rel}
Let $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$. With the above notation $w$ is relevant for $\sigma$ if and only if
\[
\operatorname{Hom}_{L_x}({\bf r}_{L,M}(\sigma),\delta_Q^{-1/2}\delta_{Q_x})\ne 0.
\]
\qed
\end{corollary}
\subsubsection{} Finally, by choosing the orbit representative $x$ as in Lemma \ref{admP-orb} (where $M$ is replaced by $L$) we explicate the condition $\operatorname{Hom}_{L_x}(\rho,\delta_Q^{-1/2}\delta_{Q_x})\ne 0$ for certain pure tensor representations $\rho\in \Pi(L)$.
For a representation $\pi\in \Pi(G_r)$, let $\pi^*\in \Pi(G_r)$ be the representation on the space of $\pi$ defined by $\pi^*(g)=\pi(g^*)$. By a result of Gelfand and Kazhdan (\cite{MR0404534}) we have $\pi^*\simeq\pi^\vee$ for all $\pi\in {\rm Irr}$. In the notation of Section \ref{General orbits} let
\[
\rho=\otimes_{\imath\in (\mathfrak{I},\prec)}\rho_\imath
\]
and let $\tau\in S_2[\beta]$ be the involution on $\mathfrak{I}$ associated to $w$ by Lemma \ref{admP-orb} applied with $L$ replacing $M$.
Assume that $\rho_\imath\in {\rm Irr}(G_{n_\imath})$ whenever $\tau(\imath)\ne \imath$.
Then by Lemma \ref{admP-orb} we have
\begin{multline}\label{dist cond}
\operatorname{Hom}_{L_x}(\rho,\delta_Q^{-1/2}\delta_{Q_x})\ne 0 \ \ \ \text{if and only if for all}\ \imath\in\mathfrak{I} \ \text{we have} \\ \rho_\imath \simeq \nu\rho_{\tau(\imath)} \ \text{whenever} \ \imath \prec \tau(\imath) \ \text{and}\ \rho_\imath\ \text{is} \ H_{n_\imath}-\text{distinguished if} \ \tau(\imath)=\imath.
\end{multline}
\subsubsection{}
Combining Corollary \ref{cor: rel}, \eqref{dist cond} and Lemma \ref{lem: dist comp} we have
\begin{corollary}\label{cor: fine rel}
Let $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$. With the above notation, if $w$ is relevant for $\sigma$ then
there exists an irreducible component $\rho=\otimes_{\imath\in (\mathfrak{I},\prec)}\rho_\imath\in {\rm Irr}(L)$ of ${\bf r}_{L,M}(\sigma)$ such that
\[
\rho_\imath \simeq \nu\rho_{\tau(\imath)} \ \text{whenever} \ \imath\prec \tau(\imath) \ \text{and}\ \rho_\imath\ \text{is} \ H_{n_\imath}-\text{distinguished if} \ \tau(\imath)=\imath.
\] \qed
\end{corollary}
\section{First applications of the geometric lemma to $\operatorname{Sp}$-distinction}
In this section we apply the contribution of the open and closed orbits of the filtration defined in \S \ref{sec: gl} in order to show that certain induced representations are $\operatorname{Sp}$-distinguished. We further apply \S \ref{sec: gl} to reduce the study of $\operatorname{Sp}$-distinction on ${\rm Irr}$ to representations supported on a single cuspidal line.
\subsection{Distinction and relevant orbits} \label{ss: oc}
\subsubsection{}
The variant of the geometric lemma discussed in \S \ref{sec: gl} is often applied to show that certain induced representations are not distinguished.
This is based on the following simple observation, which is an immediate consequence of Lemma \ref{lem: dist comp} (applied with $G=H$).
\begin{lemma}\label{lem: ex orb}
Let $M$ be a standard Levi subgroup of $G$. If $\sigma\in\Pi(M)$ is such that ${\bf i}_{G,M}(\sigma)$ is $H$-distinguished then there exists a $P$-orbit in $X$ that is relevant to $\sigma$. \qed
\end{lemma}
The reverse implication need not be true. However, there are two cases in which the geometric lemma indicates distinction.
\subsubsection{} Assume that $n_i$ is even for all $i$. Then the open $P$-orbit in $X$ is $\imath_M(w_1)=P\cdot x_M$ (see \S \ref{sss: clop}) and
\[
M_{x_{M}}=\{\operatorname{diag}(h_1,\dots,h_k): h_i\in H_{n_i/2},\,i=1,\dots,k\}.
\]
Let $\sigma_i\in \Pi(G_{n_i})$ be $H_{n_i/2}$-distinguished and $0\ne \ell_i\in \operatorname{Hom}_{H_{n_i/2}}(\sigma_i,1)$ for all $i=1,\dots,k$. Let $\sigma=\sigma_1\otimes\cdots\otimes\sigma_k$ and $\ell=\ell_1\otimes\cdots\otimes\ell_k\in \operatorname{Hom}_{M_{x_M}}(\sigma,1)$. The integral
\begin{equation}\label{open int}
\tilde\ell(\varphi)=\int_{(H\cap \eta_M^{-1}M_{x_M}\eta_M)\backslash H}\varphi(\eta_Mh)\ dh
\end{equation}
where $\eta_M\in G$ is such that $\eta_M\cdot I_{2n}=x_M$, defines a non-zero linear form $\tilde\ell\in \operatorname{Hom}_H(\mathcal{V}_1,1)$. It does not necessarily extend to an $H$-invariant linear form on ${\bf i}_{G,M}(\sigma)$, but it lies in a holomorphic family of linear forms that do extend meromorphically.
Note that $G_{x_M}=G^{\theta_{x_M}}$ is the fixed point group of the involution $\theta_{x_M}(g)=x_M\theta(g)x_M^{-1}$ and that $M=P\cap \theta_{x_M}(P)$.
For $\lambda=(\lambda_1,\dots,\lambda_k)\in \mathbb{C}^k$ let $\sigma[\lambda]$ be the representation on the space of $\sigma$ defined by
\[
\sigma[\lambda](\operatorname{diag}(g_1,\dots,g_k))=\abs{\det g_1}_F^{\lambda_1}\cdots\abs{\det g_k}_F^{\lambda_k}\sigma(\operatorname{diag}(g_1,\dots,g_k)).
\]
The representations ${\bf i}_{G,M}(\sigma[\lambda])$ can all be realized in the same space $\mathcal{V}$ and then the $H$-filtration $\{\mathcal{V}_i\}_{i=1}^n$ is independent of $\lambda$. The following follows from \cite[Theorem 2.8]{MR2401221}.
\begin{lemma}\label{BD her}
With the above notation and assumptions, there is a non-zero meromorphic function $(\lambda\mapsto \ell_\lambda):\mathbb{C}^k\rightarrow \mathcal{V}^*$ that satisfies $\ell_\lambda\in \operatorname{Hom}_H({\bf i}_{G,M}(\sigma[\lambda]),1)$ whenever holomorphic at $\lambda$.
\end{lemma}
\subsubsection{Hereditary property of $\operatorname{Sp}$-distinction}
This implies the hereditary property of $\operatorname{Sp}$-distinction.
\begin{corollary}\label{open dist}
Assume that $n_i$ is even for all $i=1,\dots,k$.
Let $\sigma_i\in \Pi(G_{n_i})$ be $H_{n_i/2}$-distinguished for all $i=1,\dots,k$ Then $\sigma_1\times\cdots\times\sigma_k$ is $H$-distinguished.
\end{corollary}
\begin{proof}
This is immediate from Lemma \ref{BD her} by taking a leading term at $\lambda=0$ of $\ell_\lambda$ at a complex line through zero in a generic direction.
\end{proof}
\subsubsection{Distinction by the closed orbit}
When a closed orbit is relevant, the geometric lemma directly implies distinction.
\begin{lemma}\label{lem: closed}
Let $\sigma_1,\dots,\sigma_t \in {\rm Irr}$, $\rho_1,\dots,\rho_s\in \Pi$ and assume that $\rho_i$ is $\operatorname{Sp}$-distinguished for $i=1,\dots,s$ (allow the case $s=0$). Then
\[
\nu\sigma_1\times\cdots\times \nu\sigma_k\times \rho_1\times \cdots\times \rho_s\times \sigma_k\times\cdots\times\sigma_1
\]
is $\operatorname{Sp}$-distinguished.
\end{lemma}
\begin{proof}
It follows from Corollary \ref{open dist} that $\rho= \rho_1\times \cdots\times \rho_s$ is $\operatorname{Sp}$-distinguished. Let $G=G_{2n}$ and $M$ its standard Levi subgroup so that
\[
\sigma=\nu\sigma_1\otimes\cdots\otimes \nu\sigma_k\otimes \rho\otimes \sigma_k\times\cdots \otimes\sigma_1
\]
is a representation of $M$. Then ${\bf i}_{G,M}(\sigma)\in \Pi(G)$. Let $P$ be the standard parabolic subgroup with Levi subgroup $M$. Then the closed orbit $P\cdot I_{2n}$ is relevant to $\sigma$ by \eqref{dist cond}.
By \S \ref{sss: clop}, $\operatorname{Hom}_H(\mathcal{V}/\mathcal{V}_{m-1},1)\ne 0$ and therefore by Lemma \ref{drmk: ist quot} (applied with $G=H$), ${\bf i}_{G,M}(\sigma)$ is $\operatorname{Sp}$-distinguished.
\end{proof}
\subsection{Reduction to cuspidal $\mathbb{Z}$-lines} \label{ss: cusp lines}
Let $\nu=\abs{\det}_F$ on $G_n$ for any $n\in \mathbb{N}$.
\subsubsection{}For $\rho\in {\operatorname{cusp}}$ let $\rho^\mathbb{Z}=\{\nu^n\rho:n\in\mathbb{Z}\}$ be the $\mathbb{Z}$-line through $\rho$. Denote by $<$ the order on $\rho^\mathbb{Z}$ induced by the standard order on $\mathbb{Z}$ (so that $\rho< \nu\rho$).
\begin{definition}
A representation $\pi\in \Pi$ is called \emph{rigid} if ${\rm Supp}(\pi)\subseteq \rho^\mathbb{Z}$ for some $\rho\in{\rm Cusp}$.
\end{definition}
\subsubsection{}
Every element of ${\rm Irr}$ has a unique decomposition as a product of rigid representations supported on disjoint cuspidal lines.
Indeed, by \cite[Proposition 8.6]{MR584084} we have
\begin{lemma}\label{lem; disj is irr}
For every $\pi\in {\rm Irr}$ there exist $\rho_1,\dots,\rho_k\in {\rm Cusp}$, so that $\rho_i^\mathbb{Z}\cap \rho_j^\mathbb{Z}=\emptyset$ for all $i\ne j$, and $\pi_1,\dots,\pi_k\in{\rm Irr}$ so that ${\rm Supp}(\pi_i)\subseteq \rho_i^\mathbb{Z}$ and $\pi=\pi_1\times\cdots\times\pi_k$.
\qed
\end{lemma}
\subsubsection{}
Another application of the geometric lemma will allow us to reduce the question of $\operatorname{Sp}$-distinction of irreducible representations to those supported on a single cuspidal line. Indeed, if $\pi=\pi_1\times\cdots\times\pi_k$ is a decomposition as in Lemma \ref{lem; disj is irr} then
\begin{equation}\label{eq: tensor H}
\operatorname{Hom}_{\operatorname{Sp}}(\pi,1)\simeq\operatorname{Hom}_{\operatorname{Sp}}(\pi_1,1)\otimes \cdots\otimes \operatorname{Hom}_{\operatorname{Sp}}(\pi_k,1).
\end{equation}
Here, we write $\operatorname{Hom}_{\operatorname{Sp}}(\pi,1)=\operatorname{Hom}_{\operatorname{Sp}_{2n}(F)}(\pi,1)$ for any $\pi\in \Pi(G_{2n})$.
In fact, we prove \eqref{eq: tensor H} for a slightly more general setting for which we need to introduce some more terminology.
\subsubsection{}
Consider the graph $\mathcal{E}$ with ${\rm Cusp}$ as the set of vertices and an edge between $\rho$ and $\nu\rho$ for every $\rho\in {\rm Cusp}$.
For every finite subset $V\subseteq {\rm Cusp}$ let $\mathcal{E}_V$ be the induced graph on the set of vertices $V$ and $\mathfrak{m}_V$ the set of connected components of $\mathcal{E}_V$. Every connected component $\Delta\in\mathfrak{m}_V$ is of the form $\Delta=\{\nu^i\rho:i=a,\dots,b\}$ for some $\rho\in {\rm Cusp}$ and integers $a\le b$.
\begin{definition}\label{def: tot disj}
We say that finite subsets $V,\,V'\subseteq{\rm Cusp}$ are \emph{totally disjoint} if either $V$ and $V'$ are contained in disjoint cuspidal $\mathbb{Z}$-lines or they satisfy the following property. For every $\Delta\in \mathfrak{m}_V$ and $\Delta'\in \mathfrak{m}_{V'}$ we have that either $\nu\rho< \rho'$ for all $\rho\in\Delta$ and $\rho'\in \Delta'$ or $\nu\rho'< \rho$ for all $\rho\in\Delta$ and $\rho'\in \Delta'$. (Equivalently, $\Delta\cup \Delta'$ is not connected in $\mathcal{E}_{V\cup V'}$.)
\end{definition}
As a consequence of \cite[Proposition 8.5]{MR584084} we have
\begin{lemma}\label{lem; tot disj is irr}
If $\pi_1,\dots,\pi_k\in {\rm Irr}$ are such that ${\rm Supp}(\pi_i)$ and ${\rm Supp}(\pi_j)$ are totally disjoint for all $i\ne j$ then $\pi_1\times \cdots\times \pi_k\in {\rm Irr}$.
\qed
\end{lemma}
\subsubsection{} We now show that \eqref{eq: tensor H} holds for totally disjoint decompositions as in Lemma \ref{lem; tot disj is irr}.
The following is a small generalization of \cite[Lemma 3.4]{MR3227442}.
\begin{lemma}\label{lem: useful cusp}
Let $\pi_1,\dots,\pi_k\in \Pi$ be such that ${\rm Supp}(\pi_i)$ and ${\rm Supp}(\pi_j)$ are totally disjoint for all $i\ne j$. Then $\pi=\pi_1\times\cdots\times\pi_k$ is $\operatorname{Sp}$-distinguished if and only if $\pi_i$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,k$. In particular, if $\pi\in {\rm Irr}$ then \eqref{eq: tensor H} holds.
\end{lemma}
\begin{proof}
The `only if' part follows from Corollary \ref{open dist}. We prove the `if' part.
Let $\sigma=\pi_1\otimes\cdots\otimes \pi_k$, $n$ be such that $\pi_1\times\cdots\times\pi_k\in\Pi(G_n)$ and $\alpha$ the decomposition of $n$ such that $\sigma\in \Pi(M_\alpha)$. Set $G=G_n$ and $M=M_\alpha$. Assume that $\pi_1\times\cdots\times\pi_k$ is $\operatorname{Sp}$-distinguished.
By Lemma \ref{lem: ex orb},
$\sigma$ admits a relevant orbit, let $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ be relevant to $\sigma$.
Apply the notation of \S \ref{General orbits}.
By Corollary \ref{cor: fine rel}, there exists an irreducible component $\rho$ of ${\bf r}_{L,M}(\sigma)$ that satisfies \eqref{dist cond}.
Then $\rho=\otimes_{\imath\in (\mathfrak{I},\prec)} \rho_\imath$ where $\rho_{i,1}\otimes\cdots\otimes\rho_{i,k_i}$ is an irreducible component of ${\bf r}_{M_{\beta_i},G_{n_i}}(\pi_i)$ for all $i=1,\dots,k$ (see \eqref{eq: trans jm}). In particular, ${\rm Supp}(\rho_{i,j})\subseteq {\rm Supp}({\bf r}_{M_{\beta_i},G_{n_i}}(\pi_i))\subseteq {\rm Supp}(\pi_i)$ for all $i$ (see \eqref{eq: supp jm}).
Assume that there exists $\imath\in\mathfrak{I}$ such that $\imath\prec\tau(\imath) $ and let $\imath=(i,j)$ and $\tau(\imath)=(i',j')$. Then $\rho_\imath\simeq \nu\rho_{\tau(\imath)}$ and by \eqref{eq: stronger cond}, $i\ne i'$. In particular, there exists $\rho'\in {\rm Supp}(\rho_{\tau(\imath)})\subseteq {\rm Supp}(\pi_{i'})$ such that $\nu\rho'\in{\rm Supp}(\rho_\imath)\subseteq {\rm Supp}(\pi_i)$. This contradicts the total disjointness of ${\rm Supp}(\pi_i)$ and ${\rm Supp}(\pi_{i'})$. Therefore, $\tau$ is the trivial involution. Now \eqref{eq: stronger cond} implies that $w$ is $M$-admissible and $\pi_i$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,k$ as required.
The isomorphism \eqref{eq: tensor H} now follows from \cite[Theorem 2.4.2]{MR1078382} (local multiplicity one for symplectic models).
\end{proof}
\section{Representations of $\operatorname{GL}_n(F)$}\label{s: LZ cl}
Before we continue with further applications of the geometric lemma to $\operatorname{Sp}$-distinction, we need to introduce the segment notation of Zelevinsky, and the Langlands and Zelevinsky classifications of ${\rm Irr}$. We refer to \cite{MR584084} for the results stated in this section.
\subsection{Segment Representations}\label{info: segments}
By a segment of cuspidal representations we mean a set
\[
[a,b]_{(\rho)}=\{\nu^i\rho:i=a,\,a+1,\dots,b\}
\]
where $\rho\in {\rm Cusp}$ and $a \le b$ are integers. By convention, the empty set is also considered a segment.
\subsubsection{} For a segment $\Delta=[a,b]_{(\rho)}$ as above, the representation $\nu^a\rho\times \nu^{a+1}\rho \times\cdots\times\nu^b\rho$ has a unique irreducible subrepresentation that we denote by $Z(\Delta)$ and a unique irreducible quotient that we denote by $L(\Delta)$.
\subsubsection{} We remark that $\Delta\mapsto L(\Delta)$ is a bijection between the set of segments of cuspidal representations and the subset of essentially square-integrable representations in ${\rm Irr}$.
\subsubsection{} Also, $Z(\Delta)$ is the unique irreducible quotient and $L(\Delta)$ is the unique irreducible subrepresentation of $\nu^b\rho\times\cdots\times\nu^{a+1}\rho\times\nu^a\rho$. By convention, if the segment $\Delta$ is empty, then both $L(\Delta)$ and $Z(\Delta)$ are taken to be the trivial representation of the trivial group.
\subsubsection{} We denote by $b(\Delta)=\nu^a\rho$ the beginning, $e(\Delta)=\nu^b\rho$ the end and $\ell(\Delta)=\abs{\Delta}=b-a+1$ the length of $\Delta$.
Let $\nu\Delta=[a+1,b+1]_{(\rho)}$.
\subsubsection{}
Let $\Delta$ and $\Delta'$ be segments of cuspidal representations.
We say that $\Delta$ \emph{precedes} $\Delta'$ and write $\Delta\prec\Delta'$ if both $\Delta$ and $\Delta'$ are contained in some $\mathbb{Z}$-line $\rho^\mathbb{Z}\subseteq {\rm Cusp}$, $b(\Delta)<b(\Delta')$, $e(\Delta)<e(\Delta')$ and $e(\Delta) \ge b(\Delta')-1$.
The segments $\Delta$ and $\Delta'$ are called \emph{linked} if either $\Delta\prec\Delta'$ or $\Delta'\prec\Delta$. Equivalently, $\Delta$ and $\Delta'$ are linked if and only if neither of them is contained in the other and their union is a segment.
\subsubsection{}\label{sec: standard} For our conventions regarding multi-sets see \S \ref{sss: mult}.
Let $\OO$ be the set of multi-sets of segments of cuspidal representatons.
An order $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}\in\OO$ on a multi-set $\mathfrak{m}$ is of \emph{standard form} if $\Delta_i\not \prec\Delta_j$ for all $i<j$.
Every $\mathfrak{m}\in \OO$ admits at least one standard order. Indeed, if for example $e(\Delta_1)\ge \cdots\ge e(\Delta_t)$ then $\{\Delta_1,\dots,\Delta_t\}$ is in standard form.
\subsubsection{The Zelevinsky classification}\label{sss: Z cl} Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}\in\OO$ be ordered in standard form. The representation
\[
\zeta(\mathfrak{m})= Z(\Delta_1) \times\cdots \times Z(\Delta_t)
\]
is independent of the choice of order of standard form.
It has a unique irreducible submodule that we denote by $Z(\mathfrak{m})$.
The Zelevinsky classification says that the map $(\mathfrak{m}\mapsto Z(\mathfrak{m})):\OO\rightarrow {\rm Irr}$ is a bijection.
\subsubsection{}
The representation
\[
\tilde\zeta(\mathfrak{m})=Z(\Delta_t) \times\cdots \times Z(\Delta_1).
\]
is also independent of the choice of standard order on $\mathfrak{m}$ and $Z(\mathfrak{m})$ is the unique irreducible quotient of $\tilde\zeta(\mathfrak{m})$.
\subsubsection{The Langlands classification}\label{sss: L cl} Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}\in\OO$ be ordered in standard form. The representation
\[
\lambda(\mathfrak{m})=L(\Delta_1)\times\cdots\times L(\Delta_t)
\]
is independent of the choice of order of standard form.
It has a unique irreducible quotient that we denote by $L(\mathfrak{m})$.
The Langlands classification says that the map $( \mathfrak{m}\mapsto L(\mathfrak{m})) :\OO\rightarrow {\rm Irr}$ is a bijection.
\subsubsection{The Zelevinsky involution}
It follows from \S \ref{sss: Z cl} and \S \ref{sss: L cl} that for any $\mathfrak{m}\in \OO$ there exists a unique $\mathfrak{m}^t\in \OO$ such that $Z(\mathfrak{m})=L(\mathfrak{m}^t)$.
The function $\mathfrak{m}\mapsto \mathfrak{m}^t$ is an involution on $\OO$.
For $\pi=Z(\mathfrak{m})\in {\rm Irr}$ let $\pi^t=L(\mathfrak{m})$. Then $\pi\mapsto \pi^t$ is the corresponding involution on ${\rm Irr}$.
\subsubsection{} For $\mathfrak{m}\in\OO$ let ${\rm Supp}(\mathfrak{m})=\{\rho\in {\rm Cusp}: \rho\in\Delta\text{ for some }\Delta\in \mathfrak{m}\}$ be the support of $\mathfrak{m}$.
(Note that ${\rm Supp}(\mathfrak{m})={\rm Supp}(Z(\mathfrak{m}))={\rm Supp}(L(\mathfrak{m}))$.)
A multi-set $\mathfrak{m}\in \OO$ is called rigid if ${\rm Supp}(\mathfrak{m})\subseteq\rho^\mathbb{Z}$ for some $\rho\in {\rm Cusp}$.
Let
\[
\OO_\rho=\{\mathfrak{m}\in\OO:{\rm Supp}(\mathfrak{m})\subseteq \rho^\mathbb{Z}\}
\]
be the set of rigid multi-sets supported on $\rho^\mathbb{Z}$.
\subsubsection{}
Lemma \ref{lem: useful cusp} reduces the study of $\operatorname{Sp}$-distinguished representations in ${\rm Irr}$ to those supported on a cuspidal $\mathbb{Z}$-line.
From now on fix $\rho\in {\rm Cusp}$ once and for all. We will study $\operatorname{Sp}$-distinction of certain rigid representations supported on $\rho^\mathbb{Z}$.
\section{A necessary condition for $\operatorname{Sp}$-distinction of $Z(\mathfrak{m})$}\label{s: nec Z}
In this section we show that if $\mathfrak{m}\in \OO_\rho$ is such that $Z(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then all segments in $\mathfrak{m}$ are of even length.
The main tool is the geometric lemma of \S \ref{sec: gl}. We also apply a result of Heumos and Rallis that we first recall.
\subsection{Results of Heumos and Rallis}
\subsubsection{}The following disjointness of models is \cite[Theorem 3.2.2]{MR1078382}.
\begin{lemma}\label{gen not sp}
If $\pi\in {\rm Irr}$ is generic then it is not $\operatorname{Sp}$-distinguished.
\qed
\end{lemma}
\subsubsection{} We also recall \cite[Theorem 11.1]{MR1078382} that provides first examples of irreducible, $\operatorname{Sp}$-distinguished representations that are not necessarily one dimensional.
\begin{lemma}\label{2speh sp}
Let $\Delta$ be a segment in ${\rm Cusp}$. Then $L(\{\Delta,\nu\Delta\})$ is $\operatorname{Sp}$-distinguished.
\end{lemma}
For the convenience of the reader, and since the proof bellow will be generalized in the sequel, we recall here the argument given by Heumos and Rallis.
Recall that $L(\{\Delta,\nu\Delta\})$ lies in an exact sequence
\[
0\rightarrow\pi \rightarrow \lambda(\{\Delta,\nu\Delta\})=\nu L(\Delta)\times L(\Delta)\rightarrow L(\{\Delta,\nu\Delta\})\rightarrow 0
\]
where $\pi\in {\rm Irr}$ is generic. The representation $\nu L(\Delta)\times L(\Delta)$ is $\operatorname{Sp}$-distinguished by Lemma \ref{lem: closed}, whereas $\pi$ is not $\operatorname{Sp}$-distinguished by Lemma \ref{gen not sp}. Therefore, $L(\{\Delta,\nu\Delta\})$ is $\operatorname{Sp}$-distinguished.
\begin{remark}\label{rmk: Speh dist}
Representations of the form $L(\{\Delta,\nu\Delta,\dots,\nu^{n-1}\Delta\})$ are often referred to as Speh representations.
In \cite{MR2332593}, Lemma \ref{2speh sp} is generalized, showing that the Speh representation $L(\{\Delta,\nu\Delta,\dots,\nu^{n-1}\Delta\})$ is $\operatorname{Sp}$-distinguished if and only if $n$ is even. This characterization of $\operatorname{Sp}$-distinguished Speh representations was based on a global argument involving the period integrals of certain Eisenstein series. In this paper we provide a local proof of a generalization.
Since Lemmas \ref{gen not sp} and \ref{2speh sp} will be applied in the sequel, we emphasize that their proofs in \cite{MR1078382} are purely local.
\end{remark}
\subsection{On $\operatorname{Sp}$-distinction of $Z(\mathfrak{m})$}
\subsubsection{} Let $\Delta=[a,b]_{(\rho)}$ be a segment in $\rho^\mathbb{Z}$. A special case of Remark \ref{rmk: Speh dist} says that $Z(\Delta)=L(\{\nu^a\rho\},\{\nu^{a+1}\rho\},\dots,\{\nu^b\rho\})$ is $\operatorname{Sp}$-distinguished if and only if $\ell(\Delta)$ is even. For the sake of completeness of our local proof, we now provide a proof of the easy part of this equivalence. A local proof of the other implication will be a part of Corollary \ref{cor: dist lad Z}.
\begin{lemma}\label{lem: Z dist}
If the representation $Z(\Delta)$ is $\operatorname{Sp}$-distinguished then $\ell(\Delta)$ is even.
\end{lemma}
\begin{proof}
Assume that $Z(\Delta)$ is $\operatorname{Sp}$-distinguished. By Lemma \ref{drmk: ist quot}, $\nu^b\rho\times\cdots\times \nu^a\rho$ is also $\operatorname{Sp}$-distinguished.
Let $d$ be such that $\rho\in\Pi(G_d)$, $G=G_{d(b-a+1)}$ and $M=M_{(d,\dots,d)}$ the standard Levi subgroup of $G$ so that $\sigma=\nu^b\rho\otimes\cdots\otimes\nu^a\rho\in \Pi(M)$. By Lemma \ref{lem: ex orb}, there exists an orbit, relevant to $\sigma$.
Since $\rho$ is cuspidal, it follows from \eqref{eq: trans jm} that ${\bf r}_{L,M}(\sigma)=0$ for every proper Levi subgroup $L$ of $M$. It therefore follows from Corollary \ref{cor: rel} that an orbit relevant to $\sigma$ is $M$-admissible.
Since all elements of ${\rm Cusp}$ are generic, it now follows from Lemma \ref{gen not sp} and \eqref{dist cond} (for $L=M$) that there exists, in particular, an involution on $\Delta$ without fixed points. Therefore $\ell(\Delta)$ is even.
\end{proof}
\subsubsection{}\label{jm Z} This can be generalized to products of $Z(\Delta)$'s, but we first need to explicate their Jacquet modules. We start with the Jacquet module of $Z(\Delta)$ itself.
Let $\Delta=[a,b]_{(\rho)}$ be a segment in ${\rm Cusp}$.
Recall the description of the Jacquet module of $Z(\Delta)$ following \cite[\S3.4]{MR584084}.
Suppose that $\rho\in\Pi(G_d)$ and let $n=(b-a+1)d$ so that $Z(\Delta)\in \Pi(G_n)$.
Let $M=M_\alpha$ for a decomposition $\alpha=(n_1,\dots,n_k)$ of $n$. Then ${\bf r}_{M,G_n}(Z(\Delta))=0$ unless $d|n_i$, $i=1,\dots,k$ in which case
\[
{\bf r}_{M,G_n}(Z(\Delta))=Z(\Delta_1) \otimes\cdots\otimes Z(\Delta_k)
\]
where $\Delta_i=[a_i,b_i]_{(\rho)}$, $a_1=a$, $a_{i+1}=b_i+1$, $i=1,\dots,k-1$ and $d(b_i-a_i+1)=n_i$, $i=1,\dots,k$.
\subsubsection{}\label{Z jm on M}
Suppose that $\beta$ is a refinement of a decomposition $\alpha$ of $n$ and let $M=M_\alpha$ and $L=M_\beta$. For the parts of the decompositions and the ordered index set $\mathfrak{I}$ we apply the notation of \S \ref{General orbits}.
Let $\Delta_1,\dots,\Delta_k$ be segments of cuspidal representations so that $\sigma=Z(\Delta_1)\otimes\cdots\otimes Z(\Delta_k)$ is an irreducible representation of $M$. It follows from \eqref{eq: trans jm} and \S \ref{jm Z} that whenever non-zero
\[
{\bf r}_{L,M}(\sigma)=\tensor\limits_{\imath\in(\mathfrak{I},\prec)} Z(\Delta_{\imath})
\]
where ${\bf r}_{M_{\beta_i}, G_{n_i}}(Z(\Delta_i))=Z(\Delta_{i,1})\otimes\cdots\otimes Z(\Delta_{i,k_i})$ is prescribed by \S \ref{jm Z}.
\subsubsection{}
\begin{proposition}\label{prop: Z dist}
Let
$\mathfrak{m}\in \OO_\rho$. If $\tilde\zeta(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $Z(\Delta)$ is $\operatorname{Sp}$-distinguished and, in particular, $\ell(\Delta)$ is even for all $\Delta\in \mathfrak{m}$.
In particular, if $Z(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\ell(\Delta)$ is even for all $\Delta\in \mathfrak{m}$.
\end{proposition}
\begin{proof}
Fix an order $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}$ so that $b(\Delta_1)\le \cdots \le b(\Delta_k)$ and note that
$\{\Delta_k,\dots,\Delta_1\}$ is a standard order on $\mathfrak{m}$, i.e., that
\[
\tilde\zeta(\mathfrak{m})\simeq Z(\Delta_1)\times\cdots \times Z(\Delta_k).
\]
Let $\sigma=Z(\Delta_1)\otimes\cdots \otimes Z(\Delta_k)$, $G=G_n$ and $M=M_\alpha$ a standard Levi subgroup of $G$ such that $Z(\Delta_1)\times\cdots \times Z(\Delta_k)={\bf i}_{G,M}(\sigma)$.
Assume that $\tilde\zeta(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished. By Lemma \ref{lem: ex orb} there exists $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$ that is relevant to $\sigma$.
We show that $Z(\Delta_i)$ is $\operatorname{Sp}$-distinguished for all $i$.
Assume, by contradiction, the contrary and use the notation of \S \ref{General orbits}.
It follows from Corollary \ref{cor: fine rel} that $\tau$ is not the trivial involution. Let $\imath\in\mathfrak{I}$ be minimal such that $\tau(\imath)\ne \imath$.
Then $\imath\prec \tau(\imath)$ and it follows from \eqref{eq: stronger cond} that $\imath=(i,1)$ for some $1\le i\le k$. But then the condition $\Delta_{\imath}= \nu\Delta_{\tau(\imath)}$ (from Corollary \ref{cor: fine rel}) contradicts our choice of order on $\mathfrak{m}$ and \S \ref{Z jm on M}.
It therefore follows that $Z(\Delta_i)$ is $\operatorname{Sp}$-distinguished and from Lemma \ref{lem: Z dist} that $\ell(\Delta_i)$ is even for $i=1,\dots,k$. The last part of the proposition follows from Lemma \ref{drmk: ist quot}.
\end{proof}
Applications of the geometric lemma to the study of $\operatorname{Sp}$-distinction of representations of the form $\lambda(\mathfrak{m})$, $\mathfrak{m}\in\OO$ (standard modules)
lead to certain combinatorial problems that we are only able to solve in special cases. In the next section we formulate these problems and present the proofs for our partial results.
The section is written in such a way that it is independent of the rest of the paper and elementary. The problem we raise is accessible to every mathematician.
\section{Multi-sets of segments}\label{s: multisets}
We formulate an elementary problem on multi-sets of segments of integers that has applications to representation theory. We are only able to provide a partial solution.
\subsubsection{}\label{sss: mult}
By a multi-set $f$ of elements in a set $X$ we mean a function $f:X\rightarrow \mathbb{Z}_{\ge 0}$ of finite support. The support of $f$ is also referred to as the set underlying $f$. If $f$ takes value in $\{0,1\}$ then we identify $f$ with its support and say that it is a set.
For example, for $x\in X$ the set of one element $\{x\}$ is the characteristic function of $x$.
Denote by $\abs{f}=\sum_{x\in X} f(x)$ the size of the multi-set $f$. By abuse of notation, we sometimes write $f=\{x_1,\dots,x_t\}$ where $t=\abs{f}$ and $x\in X$ equals $x_i$ for exactly $f(x)$ indices $i$. We refer to the presentation $\{x_1,\dots,x_t\}$ as an order on $f$.
We write $x\in f$ if $x$ is in the support of $f$.
\subsubsection{}
By a segment of integers we mean a set $[a,b]=\{a,a+1,\dots,b\}$ where $a\le b$ are integers. By convention, the empty set is a segment. Let $\operatorname{Sgm}$ denote the set of all segments of integers.
Consider the operation
\[
\nu[a,b]=[a+1,b+1],\ \ \ [a,b]\in \operatorname{Sgm}
\]
and define the following relation on $\operatorname{Sgm}$. For $[a,b],\,[a',b']\in \operatorname{Sgm}$ we say that $[a,b]$ \emph{precedes} $[a',b']$ and write $[a,b]\prec[a',b']$ if $a<a'$, $b<b'$ and $b \ge a'-1$.
By a \emph{decomposition} of $[a,b]\in \operatorname{Sgm}$ we mean a $k$-tuple of segments $([a_1,b_1],\dots,[a_k,b_k])\in \operatorname{Sgm}^k$, $k\in \mathbb{N}$, such that $b_1=b$, $a_k=a$ and $b_{i+1}=a_i-1$, $i=1,\dots,k-1$.
The decomposition is called trivial if $k=1$.
\subsubsection{}
Let $\OO_\mathbb{Z}$ be the set of multi-sets of segments of integers.
We say that a multi-set $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}$ is ordered in standard form if $\Delta_i\not\prec\Delta_j$ for all $i<j$.
Given an ordered multi-set $\mathfrak{m}=\{\Delta_1\dots,\Delta_k\}\in\OO_\mathbb{Z}$, by a decomposition of $\mathfrak{m}$ we mean a decomposition of $\Delta_i$ for all $i=1,\dots,k$.
The decomposition is called trivial if the decomposition of each $\Delta_i$ is trivial.
It will be convenient to index a decomposition of an ordered multi-set as follows.
If $(\Delta_{i,1},\dots,\Delta_{i,k_i})$ is the decomposition of $\Delta_i$, $i=1,\dots,k$ let $(\mathfrak{I},\prec)$ be the linearly ordered set
\[
\mathfrak{I}=\{(i,j):i=1,\dots,k,\ j=1,\dots,k_i\}
\]
with the lexicographic order $(i,j) \prec (i',j')$ if either $i<i'$ or $i=i'$ and $j<j'$. We further consider the partial order $(i,j)\ll(i',j')$ if $i<i'$.
Thus, a decomposition of an ordered multi-set $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in \OO_\mathbb{Z}$ produces a new ordered multi-set $\{\Delta_\imath:\imath\in(\mathfrak{I},\prec)\}$.
\subsubsection{Relevant decompositions}
\begin{definition}\label{rel dec}
Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in\OO_\mathbb{Z}$ be an ordered multi-set.
We say that an ordered decomposition $\{\Delta_\imath:\imath\in (\mathfrak{I},\prec)\}$ (of the order $\{\Delta_1,\dots,\Delta_k\}$ of $\mathfrak{m}$) is \emph{relevant} to $\{\Delta_1,\dots,\Delta_k\}$ if there exists an involution $\tau$ on $\mathfrak{I}$ satisfying the following properties:
\begin{enumerate}
\item\label{order flip} $\tau(i,j+1)\ll\tau(i,j),\ i=1,\dots,k,\,j=1,\dots,k_i-1$;
\item\label{no fix} $\tau(\imath)\ne \imath, \ \imath\in\mathfrak{I}$;
\item\label{segment match} $\Delta_\imath=\nu\Delta_{\tau(\imath)}$ whenever $\imath\prec\tau(\imath)$.
\end{enumerate}
\end{definition}
\subsubsection{} The involutions $\tau$ in the above definition must satisfy the following property.
\begin{lemma}\label{lem: 1 to end}
Let $\tau$ be an involution on $\mathfrak{I}$ satisfying conditions \eqref{order flip} and \eqref{no fix} of Definition \ref{rel dec}. Then, there exist $i_1>\cdots>i_{k_1}>1$ such that $\tau(1,j)=(i_j,k_{i_j})$, $j=1,\dots,k_1$.
\end{lemma}
\begin{proof}
Let $\tau(1,j)=(i_j,r_j)$. The inequalities $i_1>\cdots>i_{k_1}>1$ are immediate from conditions \eqref{order flip} and \eqref{no fix} of Definition \ref{rel dec}.
If $r_j<k_{i_j}$ then, again by the same condition, $\tau(i_j,r_j+1)\ll (1,j)$ which is impossible. Therefore $r_j=k_{i_j}$.
\end{proof}
\subsubsection{Distinguished multi-set}
\begin{definition}\label{def: dist mult}
A multi-set $\mathfrak{m}\in\OO_\mathbb{Z}$ is called distinguished if every standard order of $\mathfrak{m}$ admits a relevant decomposition.
\end{definition}
\subsubsection{Speh type}\label{def: Sp type}
For $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in\OO_\mathbb{Z}$ and $n\in\mathbb{Z}$ let $\nu^n\mathfrak{m}=\{\nu^n\Delta_1,\dots,\nu^n\Delta_k\}$.
It is easy to see that the following conditions are equivalent for $\mathfrak{m}\in\OO_\mathbb{Z}$:
\begin{itemize}
\item $\mathfrak{m}$ is of the form $\mathfrak{n}+\nu\mathfrak{n}$ for some $\mathfrak{n}\in\OO_\mathbb{Z}$;
\item the trivial decomposition of $\mathfrak{m}$ is relevant to some standard order of $\mathfrak{m}$;
\item the trivial decomposition of $\mathfrak{m}$ is relevant to any standard order of $\mathfrak{m}$.
\end{itemize}
\begin{definition}
We say that $\mathfrak{m}\in\OO_\mathbb{Z}$ is of Speh type if it satisfies the above equivalent conditions.
\end{definition}
\subsubsection{The main Hypothesis}
Consider the following property of a multi-set $\mathfrak{m}\in\OO_\mathbb{Z}$.
\begin{hypothesis}\label{hyp*}
If $\mathfrak{m}$ is distinguished then $\mathfrak{m}$ is of Speh type.
\end{hypothesis}
Unfortunately, we do not have enough information to determine whether Hypothesis \ref{hyp*} is satisfied for all $\mathfrak{m}\in\OO_\mathbb{Z}$. We will prove it in some special cases and in particular when $\mathfrak{m}$ is a set.
\subsubsection{} In fact, for some of the representation theoretic applications we have in mind it suffices to consider a slightly weaker property.
For $[a,b]\in\operatorname{Sgm}$ let $[a,b]^\vee=[-b,-a]$ and for $\mathfrak{m}\in\OO_\mathbb{Z}$ let $\mathfrak{m}^\vee\in\OO_\mathbb{Z}$ be defined by $\mathfrak{m}^\vee(\Delta)=\mathfrak{m}(\Delta^\vee)$, $\Delta\in\operatorname{Sgm}$.
Consider the following property of a multi-set $\mathfrak{m}\in\OO_\mathbb{Z}$.
\begin{hypothesis}\label{hyp**}
If $\mathfrak{m}$ and $\mathfrak{m}^\vee$ are both distinguished then $\mathfrak{m}$ is of Speh type.
\end{hypothesis}
\subsubsection{} In order to prove that a given multi-set $\mathfrak{m}$ satisfies the Hypothesis \ref{hyp*} it is enough to show that if $\mathfrak{m}$ is distinguished then the trivial decomposition is relevant to some standard order. In particular, it is enough to show that for some standard order, no non-trivial decomposition is relevant. We can only prove Hypothesis \ref{hyp*} for some special cases by proving this stronger version. For this purpose we define a certain standard order on multi-sets.
\subsubsection{}\label{orders}
For $\Delta=[a,b]$ let $b(\Delta)=a$ be the beginning and $e(\Delta)=b$ the end of $\Delta$.
For $\mathfrak{m}\in \OO_\mathbb{Z}$ let $c_1>\cdots>c_s$ be such that we have the identity of sets $\{c_1,\dots,c_s\}=\{e(\Delta):\Delta\in \mathfrak{m}\}$. Let $\mathfrak{m}[i]\in\OO_\mathbb{Z}$ be defined by
\[
\mathfrak{m}[i](\Delta)=\begin{cases} \mathfrak{m}(\Delta) & e(\Delta)=c_i \\ 0 & \text{otherwise} \end{cases}
\]
for $i=1,\dots,s$ so that
\[
\mathfrak{m}=\mathfrak{m}[1]+\cdots+\mathfrak{m}[s]
\]
and all segments in the support of $\mathfrak{m}[i]$ end at $c_i$.
Note that any linear order on $\mathfrak{m}$ that extends the relation, $\Delta<\Delta'$ whenever $\Delta\in \mathfrak{m}[i]$ and $\Delta'\in \mathfrak{m}[j]$ for all $1\le i<j\le s$, is in standard form.
\subsubsection{}\label{oord}
In order to fix such an order, we need to linearly order each $\mathfrak{m}[i]$. We make a particular such choice recursively.
Let $\mathfrak{m}[1]$ be ordered by $\mathfrak{m}[1]=\{\Delta_1,\dots,\Delta_k\}$ where $b(\Delta_1)\le \cdots\le b(\Delta_k)$ and set $\mathfrak{m}[1]'=\mathfrak{m}[1]$. Suppose that $\mathfrak{m}[1],\dots,\mathfrak{m}[i]$ are ordered and that $\mathfrak{m}[1]',\dots,\mathfrak{m}[i]'$ are defined for some $i<s$. We order $\mathfrak{m}[i+1]$ and define $\mathfrak{m}[i+1]'$ as follows. Set $\mathfrak{m}[i+1]=\{\Delta_1,\dots,\Delta_k,\Delta_1',\dots,\Delta_m'\}$ where $b(\Delta_1)\le \cdots\le b(\Delta_k)$, $b(\Delta_1')\ge \cdots \ge b(\Delta_m')$ and $\min(\mathfrak{m}[i+1],\nu^{-1}\mathfrak{m}[i]')=\{\Delta_1',\dots,\Delta_m'\}$ and let $\mathfrak{m}[i+1]'=\{\Delta_1,\dots,\Delta_k\}$.
\subsubsection{} We now prove that Hypothesis \ref{hyp*} holds for sets.
\begin{proposition}\label{prop: hyp set}
Let $\mathfrak{m}\in \OO_\mathbb{Z}$ be a set. Then, no non-trivial decomposition is relevant to the order on $\mathfrak{m}$ defined in \S \ref{oord}. In particular, Hypothesis \ref{hyp*} holds for $\mathfrak{m}$.
\end{proposition}
\begin{proof}
We prove the statement by induction on $\abs{\mathfrak{m}}$.
If $\abs{\mathfrak{m}}\le 2$ then it follows from Lemma \ref{lem: 1 to end} that conditions \eqref{order flip} and \eqref{no fix} of Definition \ref{rel dec} cannot be satisfied by any non-trivial decomposition of $\mathfrak{m}$ and the proposition is therefore true.
Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}$ and assume by contradiction that there exists a non-trivial decomposition $\{\Delta_\imath:\imath\in \mathfrak{I}\}$ that is relevant to $\{\Delta_1,\dots,\Delta_k\}$. Let $\tau$ be the involution on $\mathfrak{I}$ satisfying
properties \eqref{order flip}, \eqref{no fix}, \eqref{segment match} of Definition \ref{rel dec} and let $i_1>\cdots>i_{k_1}>1$ be given by Lemma \ref{lem: 1 to end} so that $\tau(1,j)=(i_j,k_{i_j})$, $j=1,\dots,k_1$.
By the property of the order chosen $e(\Delta_1)\ge e(\Delta_{i_{k_1}})\ge \cdots \ge e(\Delta_{i_1})$.
The condition \eqref{segment match} of Definition \ref{rel dec} implies that $e(\Delta_{i_1})\ge e(\Delta_1)-1$ and $b(\Delta_{i_1})>\cdots>b(\Delta_{i_{k_1}})=b(\Delta_1)-1$.
Since the equality $e(\Delta_{i_{k_1}})=e(\Delta_1)$ contradicts the order chosen on $\mathfrak{m}[1]$ we must have $e(\Delta_{i_{k_1}})=e(\Delta_1)-1$, i.e., $\Delta_{i_{k_1}}=\nu^{-1}\Delta_1$ and in the notation of \S \ref{orders}, $c_2=c_1-1$ and $\Delta_{i_j}\in \mathfrak{m}[2]$ for all $j=1,\dots,k_1$.
Since $\mathfrak{m}$ is a set, we get that $\Delta_{i_{k_1}}\not\in \mathfrak{m}[2]'$.
Taking the order chosen on $\mathfrak{m}[2]$ into consideration, now implies that $k_1=1=k_{i_1}$.
Again since $\mathfrak{m}$ is a set, it is easy to see that the order defined on $\mathfrak{m}-\{\Delta_1\}-\{\Delta_{i_1}\}$ by \S \ref{oord} is that inherited from $\mathfrak{m}$, i.e., the order $\{\Delta_2,\dots,\Delta_{i_1-1},\Delta_{i_1+1},\dots,\Delta_k\}$. Furthermore, $\tau|_{\mathfrak{I}\setminus \{(1,1),(i_1,1)\}}$ shows that $\{\Delta_\imath:\imath\in \mathfrak{I}\setminus \{(1,1),(i_1,1)\}\}$ is a non-trivial decomposition relevant to $\{\Delta_2,\dots,\Delta_{i_1-1},\Delta_{i_1+1},\dots,\Delta_k\}$. This contradicts the induction hypothesis.
\end{proof}
\subsubsection{} The same proof gives another family of multi-sets satisfying Hypothesis \ref{hyp*}.
\begin{proposition} \label{prop: hyp 2}
Let $\mathfrak{m}\in \OO_\mathbb{Z}$ and set $\mathfrak{m}=\mathfrak{m}[1]+\cdots+\mathfrak{m}[s]$ as in \S \ref{orders}. Assume that $\abs{\mathfrak{m}[i]}\le 2$ for all $i=1,\dots,s$. Then, no non-trivial decomposition is relevant to the order on $\mathfrak{m}$ defined in \S \ref{oord}. In particular, Hypothesis \ref{hyp*} holds for $\mathfrak{m}$.
\end{proposition}
\begin{proof}
The first three paragraphs of the proof of Proposition \ref{prop: hyp set} hold for any multi-set and we apply the conclusions and the notation in this case.
Since $\abs{\mathfrak{m}[2]}\le 2$ we conclude that $k_1\le 2$. Furthermore, if $k_1=2$ then $\Delta_{i_2}=\nu^{-1}\Delta_1$, $\Delta_{i_1}=\nu^{-1}\Delta_{1,1}$ and $\mathfrak{m}[2]=\{\Delta_{i_2},\Delta_{i_1}\}$. But this shows that $\Delta_{i_2}\not\in\mathfrak{m}[2]'$ and contradicts the order on $\mathfrak{m}[2]$.
We therefore have $k_1=1=k_{i_1}$.
Let $\mathfrak{n}=\mathfrak{m}-\{\Delta_1\}-\{\Delta_{i_1}\}$. Again, it is easily observed that the order on $\mathfrak{n}$ defined in \S \ref{oord} is the one inherited by $\mathfrak{m}$ and the proposition follows by induction as in the last paragraphs of the proof of Proposition \ref{prop: hyp set}.
\end{proof}
\section{On $\operatorname{Sp}$-distinction of standard modules}\label{s: std mod sp}
The bijection $[a,b]\mapsto [a,b]_{(\rho)}$ from segments of integers to segments in $\rho^\mathbb{Z}$ induces a bijection on multi-sets from $\OO_\mathbb{Z}$ to $\OO_\rho$. We refer to this bijection as $\rho$-labeling and to its inverse as unlabeling.
In this section we show that if $\mathfrak{m}\in \OO_\rho$ is such that $\lambda(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then the unlabeling of $\mathfrak{m}$ is distinguished in the sense of Definition \ref{def: dist mult}. The results and Hypothesis of \S \ref{s: multisets} therefore become relevant to the study of $\operatorname{Sp}$-distinction of standard modules.
From now on we adopt the following convention. We say that a multi-set $\mathfrak{m}\in \OO_\rho$ satisfies a property defined on multi-sets in $\OO_\mathbb{Z}$ if its unlabeling satisfies this property.
\subsubsection{}\label{jm sqr}
Let $\Delta=[a,b]_{(\rho)}$ be a segment in $\rho^\mathbb{Z}$ and $\delta=L(\Delta)$.
We first recall the description of the Jacquet module of $\delta$ following \cite[\S9.5]{MR584084}.
Suppose that $\rho$ is a representation of $G_d$ and let $n=(b-a+1)d$ so that $\delta\in \Pi(G_n)$.
Let $M=M_\alpha$ for a decomposition $\alpha=(n_1,\dots,n_k)$ of $n$. Then ${\bf r}_{M,G_n}(\delta)=0$ unless $d|n_i$, $i=1,\dots,k$ in which case
\[
{\bf r}_{M,G_n}(\delta)=\delta_1 \otimes\cdots\otimes \delta_k
\]
where $\delta_i=L(\Delta_i)$, $\Delta_i=[a_i,b_i]_{(\rho)}$, $b_1=b$, $b_{i+1}=a_i-1$, $i=1,\dots,k-1$ and $d(b_i-a_i+1)=n_i$, $i=1,\dots,k$.
\subsubsection{}\label{jm on M} Suppose that $\beta$ is a refinement of a decomposition $\alpha$ of $n$ and let $M=M_\alpha$ and $L=M_\beta$. For the parts of the decompositions and the ordered index set $\mathfrak{I}$ we apply the notation of \S \ref{General orbits}.
Let $\delta=\delta_1\otimes\cdots\otimes\delta_k$ be an irreducible, essentially square-integrable representation of $M$ (i.e., $\delta_i=L(\Delta_i)$ for some segment of cuspidal representations, $i=1,\dots,k$). It follows from \eqref{eq: trans jm} that whenever non-zero
\[
{\bf r}_{L,M}(\delta)=\tensor\limits_{\imath\in(\mathfrak{I},\prec)} \delta_{\imath}
\]
where ${\bf r}_{M_{\beta_i}, G_{n_i}}(\delta_i)=\delta_{i,1}\otimes\cdots\otimes\delta_{i,k_i}$ is prescribed by \S \ref{jm sqr}.
\subsubsection{}\label{sss: for dist} Let $M$ and $\delta$ be as above and $w\in {}_MW_{\theta(M)}\cap [w_{2n}]w_{2n}$. Recall that $\delta_i$ is generic $i=1,\dots,k$. In the notation of \S \ref{General orbits} it follows from \eqref{dist cond}, Lemma \ref{gen not sp} and \S \ref{jm on M} that
\begin{equation}\label{eq: rel sqe}
w \ \text{is relevant for}\ \delta \ \text{if and only if}\ \tau(\imath)\ne\imath\ \text{for all}\ \imath\in\mathfrak{I} \ \text{and}\ \delta_\imath\simeq\nu\delta_{\tau(\imath)}\ \text{whenever}\ \imath\prec\tau(\imath).
\end{equation}
\subsubsection{} The following is an immediate consequence.
\begin{proposition}\label{prop: dist hyp}
Let $\mathfrak{m}\in\OO_\rho$ satisfy Hypothesis \ref{hyp*}. If $\lambda(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type.
In particular, if $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type.
\end{proposition}
\begin{proof}
If $\lambda(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then for any standard order on $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}$, the induced representation $L(\Delta_1)\times\cdots\times L(\Delta_k)$ is $\operatorname{Sp}$-distinguished. Therefore, by Lemma \ref{lem: ex orb}, $L(\Delta_1)\otimes\cdots\otimes L(\Delta_k)$ admits a relevant orbit. Combining the condition \ref{eq: stronger cond} with \eqref{eq: rel sqe}, \S \ref{sss: for dist} says that $\mathfrak{m}$ is distinguished in the sense of Definition \ref{def: dist mult}. Hypothesis \ref{hyp*} therefore implies that $\mathfrak{m}$ is of Speh type.
The last part of the proposition follows from Lemma \ref{drmk: ist quot}.
\end{proof}
\subsubsection{} Combining Proposition \ref{prop: dist hyp} with Propositions \ref{prop: hyp set} and \ref{prop: hyp 2} we obtain
\begin{corollary}\label{cor: set case}
Let $\mathfrak{m}\in\OO_\rho$ either be a set or, in the notation of \S \ref{orders}, satisfy $\abs{\mathfrak{m}[i]}\le 2$ for all $i$. If $\lambda(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type. In particular, if $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type. \qed
\end{corollary}
\subsubsection{} We also point out the weaker consequence of Hypothesis \ref{hyp**}.
\begin{proposition}\label{prop: dist hyp**}
Let $\mathfrak{m}\in\OO_\rho$ satisfy Hypothesis \ref{hyp**}. If $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type.
\end{proposition}
\begin{proof}
If $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $L(\mathfrak{m}^\vee)$ is $\operatorname{Sp}$-distinguished by Lemma \ref{lem: cont dist}. Therefore, by Lemma \ref{drmk: ist quot}, both $\lambda(\mathfrak{m})$ and $\lambda(\mathfrak{m}^\vee)$ are $\operatorname{Sp}$-distinguished. It now follows, as in the proof of Proposition \ref{prop: dist hyp}, that both $\mathfrak{m}$ and $\mathfrak{m}^\vee$ are distinguished in the sense of Definition \ref{def: dist mult}. Hypothesis \ref{hyp**} therefore implies that $\mathfrak{m}$ is of Speh type.
\end{proof}
\subsubsection{}
We end this section by providing an example which demonstrates that the necessary condition for distinction obtained by combining Propositions \ref{prop: Z dist} and \ref{prop: dist hyp**} is not sufficient.
\begin{example}
Let $\pi=L(\mathfrak{m})$ where
\[
\mathfrak{m}=\{[\nu^{4},\nu^{4}],[\nu^{3},\nu^{3}],[\nu^{3},\nu^{3}],[\nu^{2},\nu^{2}],[\nu,\nu^{2}],[1,\nu]\}.
\]
By Proposition \ref{prop: hyp 2}, the multi-set $\mathfrak{m}$ satisfies Hypothesis \ref{hyp*}. Using the combinatorial algorithm of M\oe{}glin and Waldspurger (\cite{MR863522}), it is easy to see that
\[
\mathfrak{m}^{t}=\{[\nu^{2},\nu^{3}],[\nu,\nu^{4}],[1,\nu]\}.
\]
We will now show that $\pi$ is not $\operatorname{Sp}$-distinguished. Assume the contrary, if possible. Let $\pi_{1}=Z([\nu^{2},\nu^{3}],[1,\nu])$ and $\pi_{2}=Z([\nu,\nu^{4}])$. The representation $\pi=Z(\mathfrak{m}^t)$ is the unique irreducible quotient of $\tilde\zeta(\mathfrak{m}^{t})$ and so it is also the unique irreducible quotient of $\pi_{1} \times \pi_{2}$. Thus, by Lemma \ref{drmk: ist quot}, $\pi_{1} \times \pi_{2}$ is $\operatorname{Sp}$-distinguished. Apply the notation of \S \ref{General orbits} with $k=2$ for an orbit that is relevant to $\pi_{1}\otimes \pi_{2}$ (by Lemma \ref{lem: ex orb}). Since $k=2$, note that $k_{2}\leq 2$.
From Corollary \ref{cor: fine rel} and \eqref{eq: trans jm} it follows that there exist irreducible components $\sigma_1$ of ${\bf r}_{M_{\beta_1},G_{n_1}}(\pi_{1})$ and $\sigma_2$ of ${\bf r}_{M_{\beta_2},G_{n_2}}(\pi_{2})$ such that writing
\[
\sigma_i=\sigma_{i,1}\otimes\cdots\otimes\sigma_{i,k_i},\ i=1,2,\ \ \ \sigma_{i,j}\in{\rm Irr},\ j=1,\dots,k_i
\]
we have $\sigma_\imath=\nu\sigma_{\tau(\imath)}$ whenever $\imath\prec \tau(\imath)$ and $\sigma_\imath$ is $\operatorname{Sp}$-distinguished if $\tau(\imath)=\imath$. Also it follows from \S \ref{jm Z} that $\nu^{4}\in {\rm Supp}(\sigma_{2,k_{2}})$. Since $\nu^{5}\notin {\rm Supp}(\pi_{1})$, we deduce that $\tau(2,k_{2})=(2,k_{2})$. By \eqref{eq: stronger cond} it follows that $k_{2}=k_{1}=1$, and so $\tau(1,1)=(1,1)$. In other words, $Z([\nu^{2},\nu^{3}],[1,\nu])$ is $\operatorname{Sp}$-distinguished. Using the algorithm of M\oe{}glin and Waldspurger again we get that
\[
Z([\nu^{2},\nu^{3}],[1,\nu])\cong L([\nu^{3},\nu^{3}],[\nu,\nu^{2}],[1,1]).
\]
By Corollary \ref{cor: set case} we obtain a contradiction.
\end{example}
\section{The $\operatorname{Sp}$-distinguished ladder representations}
We classify $\operatorname{Sp}$-distinguished representations in the class of ladder representations introduced by Lapid and M\'inguez in \cite{MR3163355}.
\subsection{Distinction of ladder representations-the L aspect}
We classify $\operatorname{Sp}$-distinction in the class of ladder representations defined below.
\subsubsection{Ladder representations}\label{sss: ladder}
\begin{definition} Let $\rho\in {\rm Cusp}$. The set $\{\Delta_1,\dots,\Delta_k\}\in\OO_\rho$ is called a \emph{ladder} if
\[
b(\Delta_1)>\cdots>b(\Delta_k)\ \ \ \text{and} \ \ \ e(\Delta_1)>\cdots>e(\Delta_k).
\]
A representation $\pi\in {\rm Irr}$ is called a ladder
representation if $\pi=L(\mathfrak{m})$ where $\mathfrak{m}\in \OO_\rho$ is a ladder
.
\end{definition}
Whenever we say that $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in\OO_\rho$ is a ladder, we implicitly assume that $\mathfrak{m}$ is already ordered as in the definition above.
\subsubsection{}\label{sss: lm}
The following property allows us to show that certain ladder representations are $\operatorname{Sp}$-distinguished.
By convention, let $L([a,a-1]_{(\rho)})$ be the trivial representation of the trivial group and let $L([a,b]_{(\rho)})=0$ if $b<a-1$. Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in \OO_\rho$ be a ladder, with $\Delta_i=[a_i,b_i]_{(\rho)}$ and for every $i=1,\dots,k-1$ let
\[
\mathcal{K}_i=L(\Delta_1)\times \cdots\times L(\Delta_{i-1})\times L([a_{i+1},b_i]_{(\rho)})\times L([a_i,b_{i+1}]_{(\rho)})\times L(\Delta_{i+2})\times\cdots\times L(\Delta_k).
\]
(Thus, $\mathcal{K}_i=0$ if $a_i>b_{i+1}+1$.) By \cite[Theorem 1]{MR3163355} we have
\begin{theorem}\label{thm: main lad}
With the above notation let $\mathcal{K}$ be the kernel of the projection $\lambda(\mathfrak{m})\rightarrow L(\mathfrak{m})$. Then $\mathcal{K}=\sum_{i=1}^{k-1} \mathcal{K}_i$. \qed
\end{theorem}
\subsubsection{} The following is the characterization of $\operatorname{Sp}$-distinguished ladder representations.
\begin{theorem}\label{thm: dist lad}
Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in \OO_\rho$ be a ladder. Then the following conditions are equivalent
\begin{enumerate}
\item $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished;
\item $k$ is even and $\Delta_{2i-1}=\nu\Delta_{2i}$ for all $i=1,\dots,k/2$;
\item $\mathfrak{m}$ is of Speh type.
\end{enumerate}
\end{theorem}
\begin{proof}
The equivalence of the last two conditions is obvious. If $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished then $\mathfrak{m}$ is of Speh type by Corollary \ref{cor: set case}.
Assume now that $k=2m$ is even and $\Delta_{2i-1}=\nu\Delta_{2i}$ for all $i=1,\dots,m$. Let $\pi_i=L(\Delta_{2i-1},\Delta_{2i})$, $i=1,\dots,m$ and $\pi=\pi_1\times\cdots\times \pi_m$. Note that $\pi$ is a quotient of $\lambda(\mathfrak{m})$. It follows from Lemma \ref{2speh sp} that $\pi_i$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,m$. Therefore, by Corollary \ref{open dist}, $\pi$ is $\operatorname{Sp}$-distinguished and by Lemma \ref{drmk: ist quot}, $\lambda(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished.
Apply the notation of \S \ref{sss: lm}. In order to show that $L(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished it is enough to show that $\mathcal{K}$ is not $\operatorname{Sp}$-distinguished (by Lemma \ref{lem: dist comp}). By Theorem \ref{thm: main lad} it is enough to show that $\mathcal{K}_i$ is not $\operatorname{Sp}$-distinguished for all $i=1,\dots,2m-1$.
Note that
\[
\mathfrak{m}_i=\{\Delta_1,\dots,\Delta_{i-1},[a_{i+1},b_i]_{(\rho)},[a_i,b_{i+1}]_{(\rho)},\Delta_{i+2},\dots,\Delta_k\}\in \OO_\rho
\]
is ordered by strictly decreasing end points and is therefore, in particular, a set and in standard form. Thus $\mathcal{K}_i\simeq\lambda(\mathfrak{m}_i)$.
By Corollary \ref{cor: set case}, it is enough to show that $\mathfrak{m}_i$ is not of Speh type for all $i=1,\dots,2m-1$.
Assume by contradiction that $\mathfrak{m}_i$ is of Speh type. Let $\Delta_j'=\Delta_j$ for $j\ne i,\,i+1$, $\Delta_i'=[a_{i+1},b_i]_{(\rho)}$ and $\Delta_{i+1}'=[a_i,b_{i+1}]_{(\rho)}$. Since $\mathfrak{m}_i=\{\Delta_1',\dots,\Delta_{2m}'\}$
is ordered by strictly decreasing end points we clearly must have $\Delta_{2j-1}'=\nu\Delta_{2j}'$ for all $j=1,\dots,m$. If $i$ is odd this implies that $a_{i+1}=a_i+1$ contradicting the inequality $a_i>a_{i+1}$. If $i$ is even this implies that $a_{i-1}=a_{i+1}+1$ contradicting the fact that $a_{i-1}>a_i>a_{i+1}$ are integers. The theorem follows.
\end{proof}
\subsection{Distinction of ladder representations-the Z aspect}
The class of ladder representations is closed under Zelevinsky involution. We now reinterpret the classification above in order to characterize the ladders $\mathfrak{m}\in\OO_\rho$ so that $Z(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished.
\subsubsection{} Recall that in \cite{MR863522}, M\oe{}glin and Waldspurger describe a combinatorial algorithm to compute $\mathfrak{m}^t$ for $\mathfrak{m}\in \OO_\rho$. This algorithm takes a particularly simple form if $\mathfrak{m}$ is a ladder, as described in \cite[\S 3.2]{MR3163355}.
In particular, Lapid and M\'inguez observe that $\mathfrak{m}\in \OO_\rho$ is a ladder if and only if $\mathfrak{m}^t$ is a ladder.
Thus, $Z(\mathfrak{m})$ is a ladder representation for a ladder $\mathfrak{m}\in\OO_\rho$ and every ladder representation is of this form for some ladder $\mathfrak{m}$.
In \S \ref{sss: MW alg} we give a recursive characterization of the M\oe{}glin and Waldspurger algorithm for ladders based on \cite[\S 3.2]{MR3163355}.
\subsubsection{}Note that for a ladder $\{\Delta_1,\dots,\Delta_k\}$, $\Delta_i\cap \Delta_{i+1}$ is either empty or a segment and therefore the length $\ell(\Delta_i\cap \Delta_{i+1})$ makes sense for all $i=1,\dots,k-1$. The following is another elementary observation that follows from \S \ref{sss: MW alg}. We omit the simple proof.
\begin{lemma}\label{lem: lad Z}
Let $\mathfrak{m}\in\OO_\rho$ be a ladder and let $\mathfrak{m}^t=\{\Delta_1,\dots,\Delta_k\}$ be ordered as a ladder. Then $\mathfrak{m}$ is of Speh type if and only if $\ell(\Delta_i)$ is even for all $i=1,\dots,k$ and $\ell(\Delta_i\cap \Delta_{i+1})$ is odd for all $1\le i\le k-1$ such that $\Delta_i\cup\Delta_{i+1}$ is a segment. \qed
\end{lemma}
\subsubsection{} Combining Lemma \ref{lem: lad Z} and Theorem \ref{thm: dist lad} we obtain another classification of $\operatorname{Sp}$-distinguished ladder representations.
\begin{corollary}\label{cor: dist lad Z}
Let $\rho\in {\rm Cusp}$ and $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in\OO_\rho$ be a ladder.
Then $Z(\mathfrak{m})$ is $\operatorname{Sp}$-distinguished if and only if we have
\begin{enumerate}
\item $\ell(\Delta_i)$ is even for all $i=1,\dots,k$ and
\item $\ell(\Delta_i\cap \Delta_{i+1})$ is odd for all $1\le i\le k-1$ such that $\Delta_{i}\cup\Delta_{i+1}$ is a segment.
\end{enumerate}
\qed
\end{corollary}
\begin{remark}
In \cite {MR2414223}, we obtained a classification of the $\operatorname{Sp}$-distinguished unitary dual in terms of Tadic's classification (\cite{MR870688}). Recall (see Remark \ref{rmk: Speh dist}) that a Speh representation $L(\{\Delta,\nu\Delta,\dots,\nu^{n-1}\Delta\})$ is $\operatorname{Sp}$-distinguished if and only if it is even (i.e. $n$ is even). Any unitary representation is a product of Speh representations and it was already proved in \cite{MR2332593} that a product of even Speh representations is $\operatorname{Sp}$-distinguished. The disjointness of Klyachko models obtained in \cite {MR2414223}, together with a prescribed model for any irreducible unitary representation \cite[Theorem 3.7]{MR2417789} imply that if an irreducible product of Speh representations is $\operatorname{Sp}$-distinguished then all Speh representations in the product are even. Based on our current results, we can reprove this implication without reference to Klyachko models as follows. If $\Delta\subseteq \rho^\mathbb{Z}$ for $\rho\in {\rm Cusp}$ and $\ell=\ell(\Delta)$ then it is well known that $L(\{\Delta,\nu\Delta,\dots,\nu^{n-1}\Delta\})=Z(\{\Delta',\nu\Delta,\dots,\nu^{\ell-1}\Delta'\})$ for some segment $\Delta' \subseteq \rho^\mathbb{Z}$ with $\ell(\Delta')=n$. If $Z(\mathfrak{m})=\pi_1\times\cdots\times \pi_k\in {\rm Irr}$ is $\operatorname{Sp}$-distinguished and $\pi_i=Z(\mathfrak{m}_i)$ is a Speh representation for all $i=1,\dots,k$, then $\mathfrak{m}=\mathfrak{m}_1+\cdots+\mathfrak{m}_k$ and it follows from Proposition \ref{prop: Z dist} and the results of \S \ref{ss: cusp lines} that all segments in $\mathfrak{m}$ are of even length, i.e., that all $\pi_i$'s are even Speh representations.
\end{remark}
ֿ
\subsubsection{} Our next goal is to study the $\operatorname{Sp}$-distinguished representations in the class of representations in ${\rm Irr}$ that are induced from ladder representations. We only obtain a classification of this class conditional to Hypothesis \ref{hyp*}. Our proof is based on certain combinatorial statements concerning the multi-sets in $\OO_\rho$ that are obtained as sums of ladders in the above context. It is more convenient, to formulate these technical results by unlabeling. We therefore, now use the convention that $\mathfrak{m}\in \OO_\mathbb{Z}$ satisfies a property on $\OO_\rho$ if its $\rho$-labeling satisfies this property.
The next section collects the required results on multi-sets in $\OO_\mathbb{Z}$.
\section{On sums of ladders of Speh type}\label{s: ladder sums}
\subsubsection{}
For segments $\Delta,\,\Delta'\in \operatorname{Sgm}$, write $\Delta\le_b \Delta'$ if either $b(\Delta)<b(\Delta')$ or $b(\Delta)=b(\Delta')$ and $e(\Delta)\le e(\Delta')$.
Thus, $\le_b$ is a linear order on $\operatorname{Sgm}$.
\subsubsection{}
Let $\ell\in \mathbb{N}$ and $\mathfrak{m}\in \OO_\mathbb{Z}$ be such that $\ell(\Delta)=\ell$ for all $\Delta\in \mathfrak{m}$. If $\Delta\in \mathfrak{m}$ is minimal with respect to $\le_b$ then we can express $\mathfrak{m}$ as a linear combination
\[
\mathfrak{m}=\sum_{n=0}^N a_n\{\nu^n\Delta\}
\]
with $a_0\in \mathbb{N}$ and $a_n\in\mathbb{Z}_{\ge 0}$ for all $n=1,\dots,N$ for some large enough $N\in \mathbb{N}$.
\begin{lemma}\label{lem: linear cond}
With the above assumptions and notation, if $\mathfrak{m}$ is of Speh type then
\begin{equation}\label{eq: lin cond}
\sum_{n=0}^N (-1)^{N-n}a_n=0.
\end{equation}
\end{lemma}
\begin{proof}
Note that
\[
\mathfrak{m}=\sum_{n=0}^{N-1} b_n(\{\nu^n\Delta\}+\{\nu^{n+1}\Delta\})+b_N \{\nu^N\Delta\}
\]
where
$b_0=a_0$ and $b_n=\sum_{i=0}^n(-1)^{n-i}a_i$. If $\mathfrak{m}$ is of Speh type, then it follows that $b_n\ge 0$ for all $n$ and that, as required, $b_N=0$.
\end{proof}
\subsubsection{}\label{sss: MW alg} As observed by M{\oe}glin and Waldspurger the Zelevinsky involution $\pi^t=L(\mathfrak{m}^t)$ of a rigid representation $\pi=L(\mathfrak{m})\in {\rm Irr}$ where $\mathfrak{m}\in \OO_\rho$ is `blind' to the cuspidal line $\rho^\mathbb{Z}$ where it is supported. Indeed, the Moeglin-Waldspurger algorithm defines an involution on $\OO_\mathbb{Z}$ that gives $\mathfrak{m}\mapsto \mathfrak{m}^t$ via $\rho$-labeling. We now explicate a recursive characterization of the Moeglin-Waldspurger algorithm of the Zelevinsky involution for ladders based on \cite[\S 3.2]{MR3163355}.
For $\Delta=[a,b]\in \operatorname{Sgm}$ we have $\{\Delta\}^t=\sum_{c=a}^b \{c\}=\{\{b\},\dots,\{a\}\}$.
Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k,\Delta_{k+1}\}$ be a ladder and let $\mathfrak{m}'=\{\Delta_1,\dots,\Delta_k\}$. Write $\Delta_i=[a_i,b_i]$.
Let $(\mathfrak{m}')^t=\{\Delta_1',\dots,\Delta_s'\}$ be ordered as a ladder.
If $b_{k+1}+1<a_k$ then $\mathfrak{m}^t=(\mathfrak{m}')^t+\{\Delta_{k+1}\}^t$ and therefore
\[
\mathfrak{m}=\{\Delta_1',\dots,\Delta_s',\{b_{k+1}\},\dots,\{a_{k+1}\}\}
\]
is ordered as a ladder.
Otherwise, let $c=s-(b_{k+1}-a_k+2)$. Then $c\ge 0$ and
\begin{equation}\label{eq: z inv l}
\mathfrak{m}^t=\{\Delta_1',\dots, \Delta_c',{}^+\Delta_{c+1}',\dots,{}^+\Delta_s',\{a_k-2\},\dots,\{a_{k+1}\}\}
\end{equation}
where ${}^+[a,b]=[a-1,b]$.
In other words, in order to obtain the ladder $\mathfrak{m}^t$ from $(\mathfrak{m}')^t$ one has to perform the following steps.
If $\Delta_{k+1}\not\prec\Delta_k$ then add to $\mathfrak{m}'$ at the tail of the ladder, the ladder $\{\Delta_{k+1}\}^t$, i.e., the $\ell(\Delta_{k+1})$ length one segments consisting of elements of $\Delta_{k+1}$ in decreasing order. Otherwise, $b_{k+1}-a_k+2\ge 1$. Starting with $(\mathfrak{m}')^t$, replace $\Delta$ by ${}^+\Delta$ for each of the last $b_{k+1}-a_k+2$ segments of $(\mathfrak{m}')^t$ and then add at the tail of the resulting ladder, the $a_k-a_{k+1}-1$ length one segments consisting of elements of $[a_{k+1},a_k-2]$ in decreasing order.
\subsubsection{} Let $\mathfrak{m}_1,\dots,\mathfrak{m}_k\in\OO_\mathbb{Z}$ be ladders and let $\Delta_0$ be minimal in $\mathfrak{m}=\mathfrak{m}_1+\cdots+\mathfrak{m}_k$ with respect to $\le_b$. Let $\mathfrak{m}_i^\dag=\mathfrak{m}_i$ if $\Delta_0\not\in\mathfrak{m}_i$ and $\mathfrak{m}_i^\dag=\mathfrak{m}_i-\{\Delta_0\}$ otherwise.
\begin{lemma}\label{lem: t speh}
With the above notation, suppose that the ladder $(\mathfrak{m}_i^\dag)^t$ is of Speh type for all $i=1,\dots,k$ and that $\ell(\Delta_0)$ is even. If $\mathfrak{n}=\mathfrak{m}_1^t+\cdots+\mathfrak{m}_k^t$ is of Speh type then $\mathfrak{m}_i^t$ is of Speh type for all $i=1,\dots,k$.
\end{lemma}
\begin{proof}
If $\mathfrak{m}_i=\mathfrak{m}_i^\dag$ then $\mathfrak{m}_i^t$ is of Speh type. If $\mathfrak{m}_i\ne\mathfrak{m}_i^\dag$, let $\Delta_i$ be the minimal segment in $\mathfrak{m}_i^\dag$ with respect to $\le_b$. If $\Delta_0\not\prec\Delta_i$ then it follows from \S \ref{sss: MW alg} that $\mathfrak{m}_i^t=(\mathfrak{m}_i^\dag)^t+\{\Delta_0\}^t$. Since $\ell(\Delta_0)$ is even it follows that $\mathfrak{m}_i^t$ is of Speh type. If $\Delta_0 \prec \Delta_i$, then $\Delta_0\cup\Delta_i$ is a segment and by Lemma \ref{lem: lad Z} (applied once with $\mathfrak{m}=\mathfrak{m}_i^t$ and once with $\mathfrak{m}=(\mathfrak{m}_i^\dag)^t$), $\mathfrak{m}_i^t$ is of Speh type if and only if $\ell(\Delta_0\cap \Delta_i)$ is odd.
Let $A=\{1\le i\le k: \mathfrak{m}_i^t \ \text{is not of Speh type}\}$ and
$\mathfrak{n}=\mathfrak{n}_1+\mathfrak{n}_2$ where $\mathfrak{n}_1=\sum_{i\in A}\mathfrak{m}_i^t$.
Assume by contradiction, that $\mathfrak{n}$ is of Speh type and $\mathfrak{n}_1\ne 0$ (i.e. $A\ne \emptyset$).
Let $i\in A$ and $c_i=s_i-(e(\Delta_0)-b(\Delta_i)+2)$ where $s_i=\abs{(\mathfrak{m}_i^\dag)^t}$.
Note that by the above remarks, $s_i-c_i=e(\Delta_0)-b(\Delta_i)+2=\ell(\Delta_0\cap \Delta_i)+1\ge 1$ is odd. Since $(\mathfrak{m}_i)^t$ is of Speh type, $s_i$ is even, hence $c_i$ is also odd. Furthermore, $d_i=b(\Delta_i)-b(\Delta_0)-1$ is odd and by \eqref{eq: z inv l} $\mathfrak{m}_i^t$ ordered as a ladder has the form
\begin{multline}
\mathfrak{m}_i^t=\{\nu\Delta_1',\Delta_1',\dots, \nu\Delta_{(c_i-1)/2}',\Delta_{(c_i-1)/2}',\nu\Delta_{(c_i+1)/2}',{}^+\Delta_{(c_i+1)/2}',\\ \nu\Delta_1'',\Delta_1'',\dots,\nu\Delta_{(s_i-c_i-1)/2}'',\Delta_{(s_i-c_i-1)/2}'',\{x_0+d_i-1\},\dots,\{x_0+1\},\{x_0\}\}
\end{multline}
where $\ell(\Delta_i'')>1$ for $i=1,\dots,(s_i-c_i-1)/2$ and $x_0=b(\Delta_0)$.
Note further that $b({}^+\Delta_{(c_i+1)/2}'))=e(\Delta_0)$ is independent of $i\in A$. Thus, we may decompose
\[
\mathfrak{m}_i^t=\mathfrak{a}_i+\{\nu\Gamma_i,{}^+\Gamma_i\}+\mathfrak{b}_i+\{x_0\}
\]
where $\mathfrak{a_i}$ and $ \mathfrak{b}_i$ are of Speh type, $b(\Delta)>e(\Delta_0)+2$ for all $\Delta\in \mathfrak{a}_i$, $b(\Delta)<e(\Delta_0)$ for all $\Delta\in \mathfrak{b}_i$ and $\Gamma_i\in\operatorname{Sgm}$ is such that $b({}^+\Gamma_i)=e(\Delta_0)$ and therefore also $b(\nu\Gamma_i)=e(\Delta_0)+2$. In particular, $b(\Delta)\ne e(\Delta_0)+1$ for all $\Delta\in \mathfrak{m}_i^t$.
For $\ell\in \mathbb{N}$ and a multi-set $\mathfrak{a}\in \OO_\mathbb{Z}$ let $\mathfrak{a}(\ell)=\delta(\ell) \cdot\mathfrak{a}$ where $\delta(\ell)$ is the characteristic function of all segments of length $\ell$. Clearly, $\mathfrak{a}$ is of Speh type if and only if $\mathfrak{a}(\ell)$ is of Speh type for all $\ell\in\mathbb{N}$.
Fix $i_0\in A$ and let $\ell=\ell({}^+\Gamma_{i_0})>1$, $B=\{i\in A:\ell=\ell({}^+\Gamma_i)\}$ and $C=\{i\in A:\ell=\ell(\nu\Gamma_i)\}$. Since $\ell({}^+\Gamma_i)=\ell(\nu\Gamma_i)+1$, $B$ and $C$ are disjoint. We have
\[
\mathfrak{m}_i^t(\ell)=\begin{cases} \mathfrak{a}_i(\ell)+\{{}^+\Gamma_i\}+\mathfrak{b}_i(\ell) & i\in B \\ \mathfrak{a}_i(\ell)+\{\nu\Gamma_i\}+\mathfrak{b}_i(\ell) & i\in C \\ \mathfrak{a}_i(\ell)+\mathfrak{b}_i(\ell) & i\in A\setminus (B\cup C). \end{cases}
\]
Set $\Delta={}^+\Gamma_{i_0}$ and note further that for $i\in B$ we have ${}^+\Gamma_i=\Delta$ and for $i\in C$ we have $\nu\Gamma_i=\nu^2\Delta$. It follows that
\[
\mathfrak{n}(\ell)=\mathfrak{n}_1(\ell)+\mathfrak{n}_2(\ell)=(\abs{B}\{\Delta\}+\abs{C}\{\nu^2\Delta\})+\sum_{i\in A} ( \mathfrak{a}_i(\ell)+\mathfrak{b}_i(\ell) )+\mathfrak{n}_2(\ell).
\]
By assumption $\mathfrak{n}(\ell)$ and $\sum_{i\in A} ( \mathfrak{a}_i(\ell)+\mathfrak{b}_i(\ell) )+\mathfrak{n}_2(\ell)$ are both of Speh type and therefore, by Lemma \ref{lem: linear cond}, each of them satisfies the linear condition \eqref{eq: lin cond}. It follows that $\abs{B}\{\Delta\}+\abs{C}\{\nu^2\Delta\}$ satisfies the same linear condition, i.e., that $\abs{B}+\abs{C}=0$. But since $i_0\in B$ this is a contradiction.
\end{proof}
\section{On Distinction of representations induced from ladder}\label{s: dist ladder}
We now study distinction in the class of representations in ${\rm Irr}$ that are induced from ladder representations. For a product of more then two ladder representations, our results are only conditional on Hypothesis \ref{hyp**}.
\subsubsection{} We recall \cite[Lemma 5.17]{1411.6310}. It reduces the reducibility of a product of ladder representations to induction from a maximal parabolic.
\begin{lemma}\label{lem: red max}
Let $\pi_1,\dots,\pi_k$ be ladder representations and let $\pi=\pi_1\times\cdots\times \pi_k$. Then $\pi\in {\rm Irr}$ if and only if $\pi_i\times \pi_j\in{\rm Irr}$ for all $i< j$. \qed
\end{lemma}
\subsubsection{}\label{sss: NC} For a maximal parabolic, the following criterion is a combination of \cite[Proposition 5.21 and Lemma 5.22]{1411.6310}.
\begin{definition}\label{def: NC}
Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}$ and $\mathfrak{n}=\{\Delta_1',\dots,\Delta_s'\}$ be ladders in $\OO_\rho$.
We say that the condition $NC(\mathfrak{m},\mathfrak{n})$ is satisfied if
there exist
$k\ge 0$, $1 \le i \le t$ and $1 \le j \leq s$ such that $i+k\le t$, $j+k\le s$ and the following properties are satisfied:
\begin{enumerate}
\item\label{1} $\Delta_{i+l}\prec\Delta'_{j+l}$ for all $l=0,\dots,k$;
\item\label{2} $\nu^{-1}\Delta_{i-1}\not\prec \Delta'_j$ if $i>1$;
\item\label{3} $\nu^{-1}\Delta_{i+k}\not\prec\Delta'_{j+k+1}$ if $j+k+1\le s$.
\end{enumerate}
\end{definition}
\begin{proposition}\label{prop: irr max}
In the above notation $Z(\mathfrak{m})\times Z(\mathfrak{n})\in{\rm Irr}$ if and only if neither $NC(\mathfrak{m},\mathfrak{n})$ nor $NC(\mathfrak{n},\mathfrak{m})$ hold.
\end{proposition}
The main result of this section requires some preparation.
\subsubsection{}
For segments $\Delta,\,\Delta'\in \OO_\rho$, write $\Delta\le_b \Delta'$ if either $b(\Delta)<b(\Delta')$ or $b(\Delta)=b(\Delta')$ and $e(\Delta)\le e(\Delta')$.
Thus, $\le_b$ is a linear order on $\OO_\rho$.
\begin{lemma}\label{lem: excision}
Let $\mathfrak{m}_1,\dots,\mathfrak{m}_k\in\OO_\rho$ be ladders such that $Z(\mathfrak{m}_1)\times\cdots\times Z(\mathfrak{m}_k)\in{\rm Irr}$. Let $\Delta_0$ be a minimal segment in $\mathfrak{m}_1+\cdots+\mathfrak{m}_k$ with respect to $\le_b$ and let
\[
\mathfrak{m}_i^\dag=\begin{cases} \mathfrak{m}_i-\{\Delta_0\} & \Delta_0\in \mathfrak{m}_i \\ \mathfrak{m}_i & \text{otherwise.} \end{cases}
\]
Then $Z(\mathfrak{m}_1^\dag)\times\cdots\times Z(\mathfrak{m}_k^\dag)\in{\rm Irr}$.
\end{lemma}
\begin{proof}
By Lemma \ref{lem: red max}, it is enough to prove the lemma for the case $k=2$. Note that if $\Delta_{0}\notin \mathfrak{m}_{i},\ i=1,2$, then $\mathfrak{m}_{i}=\mathfrak{m}_{i}^\dag$ and we have nothing to prove. Thus we assume that $\Delta_{0}$ belongs to at least one of the two multi-sets. By the symmetry of the irreducibility criterion of Proposition \ref{prop: irr max}, we may assume without loss of generality that $\Delta_0\in \mathfrak{m}_1$. Write $\mathfrak{m}_1^\dag=\{\Delta_1,\dots,\Delta_t\}$ and $\mathfrak{m}_2^\dag=\{\Delta_1',\dots,\Delta_s'\}$ ordered as ladders.
Note that $\mathfrak{m}_1=\{\Delta_1,\dots,\Delta_t,\Delta_0\}$ is ordered as a ladder.
Assume by contradiction that $Z(\mathfrak{m}_1^\dag)\times Z(\mathfrak{m}_2^\dag)$ reduces. It follows from Proposition \ref{prop: irr max} that either $NC(\mathfrak{m}_1^\dag,\mathfrak{m}_2^\dag)$ or $NC(\mathfrak{m}_2^\dag,\mathfrak{m}_1^\dag)$ is satisfied.
We separate into two cases and show that in each case this implies that either $NC(\mathfrak{m}_1,\mathfrak{m}_2)$ or $NC(\mathfrak{m}_2,\mathfrak{m}_1)$ holds.
Since this is a contradiction to Proposition \ref{prop: irr max} the lemma will follow.
Consider first the case that $\Delta_0\not\in \mathfrak{m}_2$ (i.e., $\mathfrak{m}_2=\mathfrak{m}_2^\dag$) and $NC(\mathfrak{m}_1^\dag,\mathfrak{m}_2^\dag)$ holds. Let $i,\,j,\,k$ be the indices satisfying \eqref{1}-\eqref{3} of Definition \ref{def: NC} for $(\mathfrak{m}_1^\dag,\mathfrak{m}_2^\dag)$. Then the same indices $i,\,j,\,k$ show that $NC(\mathfrak{m}_1,\mathfrak{m}_2)$ holds.
Next consider the case $\Delta_0\in \mathfrak{m}_2$ or $NC(\mathfrak{m}_1^\dag,\mathfrak{m}_2^\dag)$ doesn't hold.
If $\Delta_0\not\in \mathfrak{m}_2$ then by assumption $NC(\mathfrak{m}_2^\dag,\mathfrak{m}_1^\dag)$ holds.
If $\Delta_0\in \mathfrak{m}_2$ then by symmetry between $\mathfrak{m}_1$ and $\mathfrak{m}_2$, without loss of generality, we may also assume that $NC(\mathfrak{m}_2^\dag,\mathfrak{m}_1^\dag)$ holds.
If the indices $i,\,j,\,k$ satisfy \eqref{1}-\eqref{3} of Definition \ref{def: NC} for $(\mathfrak{m}_2^\dag,\mathfrak{m}_1^\dag)$ then for the same indices \eqref{1} and \eqref{2} are automatic for $(\mathfrak{m}_2,\mathfrak{m}_1)$, while \eqref{3} is automatic unless $j+k=t$ in which case \eqref{3} reads $\nu^{-1}\Delta_{i+k}' \not\prec \Delta_0$. By the minimality of $\Delta_0$ and since $\Delta_0\not\in \mathfrak{m}_2^\dag$ it follows that $\nu^{-1}\Delta'\not\prec \Delta_0$ for all $\Delta'\in\mathfrak{m}_2^\dag$ and in particular for $\Delta'=\Delta_{i+k}'$. It therefore follows that $NC(\mathfrak{m}_2,\mathfrak{m}_1)$ holds.
\end{proof}
\subsubsection{} \label{sss: li}
Define an operation $\mathfrak{m}\mapsto \mathfrak{m}'$ on $\OO_\rho$ as follows. For $\mathfrak{m}\in\OO_\rho$, let $\Delta_0$ be minimal in $\mathfrak{m}$ with respect to $\le_b$ and $\mathfrak{m}'=\mathfrak{m}-\mathfrak{m}(\Delta_0)\{\Delta_0\}$.
\begin{proposition}\label{prop: li}
Let $\Omega\subseteq \OO_\rho$ be a subset closed under the operation $\mathfrak{m}\mapsto \mathfrak{m}'$ and such that Hypothesis \ref{hyp**} holds for $\mathfrak{m}^t$ for all $\mathfrak{m}\in\Omega$.
Let $\pi_1,\dots,\pi_k$ be ladder representations such that $\pi=\pi_1\times\cdots\times\pi_k\in {\rm Irr}$.
Let $\mathfrak{m}\in\OO_\rho$ be such that $\pi=Z(\mathfrak{m})$ and assume that $\mathfrak{m}\in\Omega$.
If $\pi$ is $\operatorname{Sp}$-distinguished then $\pi_i$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,k$.
\end{proposition}
\begin{proof}
Let $\mathfrak{m}_1,\dots,\mathfrak{m}_k$ be ladders such that $\pi_i=Z(\mathfrak{m}_i)$, $i=1,\dots,k$. Since $\pi$ is irreducible, we have $\mathfrak{m}=\mathfrak{m}_1+\cdots+\mathfrak{m}_k$.
The proof is by induction on $\abs{\mathfrak{m}}$. For $\mathfrak{m}=0$ there is nothing to prove.
Let $\Delta_0$ be the minimal segment in $\mathfrak{m}$ with respect to $\le_b$.
Let $\mathfrak{n}_0=\mathfrak{m}(\Delta_0)\{\Delta_0\}$ so that $\mathfrak{m}=\mathfrak{m}'+\mathfrak{n}_0$ and let $\mathfrak{m}_i^\dag=\min\{\mathfrak{m}_i,\mathfrak{m}'\}$, $i=1,\dots,k$. Note that
\[
\mathfrak{m}_i^\dag=\begin{cases}\mathfrak{m}_i' & \Delta_0\in \mathfrak{m}_i \\ \mathfrak{m}_i & \text{otherwise}\end{cases}
\]
and $\mathfrak{m}'=\mathfrak{m}_1^\dag+\cdots+\mathfrak{m}_k^\dag$.
By the definition of $\mathfrak{m}$, $\pi$ is the unique irreducible quotient of $\tilde\zeta(\mathfrak{m})$. Since no segment in $\mathfrak{m}'$ precedes $\Delta_0$ we have $\tilde\zeta(\mathfrak{m})=\zeta(\mathfrak{n}_0) \times \tilde\zeta(\mathfrak{m}')$ and therefore $\pi$ is also the unique irreducible quotient of $Z(\mathfrak{n}_0) \times Z(\mathfrak{m}')$.
Thus, by Lemma \ref{drmk: ist quot}, $Z(\mathfrak{n}_0) \times Z(\mathfrak{m}')$ is $\operatorname{Sp}$-distinguished. Apply the notation of \S \ref{General orbits} with $k=2$ for an orbit that is relevant to $Z(\mathfrak{n}_0) \otimes Z(\mathfrak{m}')$ (by Lemma \ref{lem: ex orb}).
It follows from Corollary \ref{cor: fine rel} and \eqref{eq: trans jm} that there exist irreducible components $\sigma_1$ of ${\bf r}_{M_{\beta_1},G_{n_1}}(Z(\mathfrak{n}_0))$ and $\sigma_2$ of ${\bf r}_{M_{\beta_2},G_{n_2}}(Z(\mathfrak{m}'))$ such that writing
\[
\sigma_i=\sigma_{i,1}\otimes\cdots\otimes\sigma_{i,k_i},\ i=1,2,\ \ \ \sigma_{i,j}\in{\rm Irr},\ j=1,\dots,k_i
\]
we have $\sigma_\imath=\nu\sigma_{\tau(\imath)}$ whenever $\imath\prec \tau(\imath)$ and $\sigma_\imath$ is $\operatorname{Sp}$-distinguished if $\tau(\imath)=\imath$.
By \cite[Theorem 4.2]{MR584084} we have
\[
Z(\mathfrak{n}_0)=\overbrace{\operatorname{Z(\Delta_0)\times\cdots\times Z(\Delta_0)}}\limits^{\mathfrak{m}(\Delta_0)-\text{times}}.
\]
In particular, it follows from the geometric lemma of Bernstein and Zelevinsky \cite[\S2.12]{MR0579172} and \S \ref{Z jm on M} that $\sigma_{1,1}=Z(\Delta)$ for some segment satisfying $b(\Delta)=b(\Delta_0)$. If $\tau(1,1)\ne (1,1)$ then, by \eqref{eq: stronger cond} and the fact that $k=2$, we must have $\tau(1,1)=(2,k_2)$ and therefore $\nu^{-1}\sigma_{1,1}=\sigma_{2,k_2}$. But since ${\rm Supp}(\sigma_{(2,k_2)})\subseteq {\rm Supp}(\sigma_2)\subseteq{\rm Supp}(Z(\mathfrak{m}'))$ (see \eqref{eq: supp jm}) we have $\nu^{-1}b(\Delta_0)\in {\rm Supp}(\nu^{-1}\sigma_{(1,1)})\setminus {\rm Supp}(\sigma_{(2,k_2)})$ which is a contradiction. It follows that $\tau(1,1)=(1,1)$ and since $k=2$, $\tau$ must be trivial.
In other words, both $Z(\mathfrak{n}_0)$ and $Z(\mathfrak{m}')$ are $\operatorname{Sp}$-distinguished.
It follows from Proposition \ref{prop: Z dist} that $\ell(\Delta_0)$ is even. Let $\pi_i'=Z(\mathfrak{m}_i^\dag)$, $i=1,\dots,k$. It follows from Lemma \ref{lem: excision} that $\pi'=\pi_1'\times\cdots\times \pi_k'\in{\rm Irr}$ and therefore $\pi'=Z(\mathfrak{m}')$. By the assumption on $\Omega$ and the induction hypothesis, $\pi_i'=L((\mathfrak{m}_i^\dag)^t)$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,k$.
By Theorem \ref{thm: dist lad}, $(\mathfrak{m}_i^\dag)^t$ is of Speh type for $i=1,\dots,k$. Since $\pi=L(\mathfrak{m}^t)$ is $\operatorname{Sp}$-distinguished and $\mathfrak{m}^t$ satisfies Hypothesis \ref{hyp**}, by Proposition \ref{prop: dist hyp**}, $\mathfrak{m}^t$ is of Speh type. It now follows from
Lemma \ref{lem: t speh} that $\mathfrak{m}_i^t$ is of Speh type and therefore, again by Theorem \ref{thm: dist lad}, that $\pi_i=L(\mathfrak{m}_i^t)$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,k$.
\end{proof}
\subsubsection{}
Let $\Omega_k$ be the set of all $\mathfrak{m}\in\OO_\rho$ that are obtained as sums of at most $k$ ladders, i.e. $\mathfrak{m}=\mathfrak{m}_1+\cdots+\mathfrak{m}_k$ where $\mathfrak{m}_i$ is either zero or a ladder, and such that $Z(\mathfrak{m}_1)\times\cdots \times Z(\mathfrak{m}_k)\in {\rm Irr}$.
Since both ladder representations and ${\rm Irr}$ are closed under Zelevinsky involution and in the Grothendick group it is multiplicative, it follows that $\Omega_k$ is closed under Zelevinsky involution. It further follow from Lemma \ref{lem: excision} that $\Omega_k$ is closed under the operation $\mathfrak{m}\mapsto \mathfrak{m}'$ defined in \S \ref{sss: li}.
For a product of two ladder representations this gives the following unconditional result.
\begin{corollary}\label{cor: prd two}
Let $\pi_1$ and $\pi_2$ be ladder representations such that $\pi=\pi_1\times\pi_2\in{\rm Irr}$.
If $\pi$ is $\operatorname{Sp}$-distinguished then $\pi_1$ and $\pi_2$ are $\operatorname{Sp}$-distinguished.
\end{corollary}
\begin{proof}
Note that $\abs{\mathfrak{m}[i]}\le 2$ for all $i$ and all $\mathfrak{m}\in\Omega_2$. Since, as remarked above, $\Omega_2$ is closed under Zelevinsky involution, it follows from Proposition \ref{prop: hyp 2} that $\mathfrak{m}^t$ satisfies Hypothesis \ref{hyp*} for all $\mathfrak{m}\in\Omega_2$. Since we also observed above that $\Omega_2$ is closed under the operation $\mathfrak{m}\mapsto \mathfrak{m}'$ defined in \S \ref{sss: li}, the statement follows from Proposition \ref{prop: li}.
\end{proof}
\subsubsection{}
We conclude this section with an example of a family of imprimitive, $\operatorname{Sp}$-distinguished representations that are not ladders.
\begin{definition}\label{def: non lad ex}
Let $\mathcal F$ denote a set of irreducible representations of the form $Z(\mathfrak{m})$ such that the multi-set $\mathfrak{m}=\{\Delta_{1},\Delta_{2},\Delta_{3}\}$ satisfies the following properties
\begin{enumerate}
\item $\ell(\Delta_{i})$ is even for all $i$,
\item $\Delta_{1}\subseteq \nu\Delta_{2}$ and $\Delta_{1}\subseteq \nu^{-1}\Delta_{2}$,
\item $\ell(\Delta_{3}\cap \Delta_{1})$ and $\ell(\Delta_{3}\cap \Delta_{2})$ are both odd.
\end{enumerate}
\end{definition}
Note that (2) implies that $\Delta_{1}\subseteq \Delta_{2}$ which in particular implies that none of these representations are ladders. Further note that $\mathcal F$ consists of only rigid representations and $\pi\in \mathcal F$ if and only if $\pi^{\vee}\in \mathcal F$. The conditions on the length of the segments in $\mathfrak{m}$ imply that the pairs $\{\Delta_{1},\Delta_{3}\}$ and $\{\Delta_{2},\Delta_{3}\}$ are linked. A simple example of a representation in $\mathcal F$ is $Z([\nu^{3},\nu^{4}],[\nu,\nu^{6}],[1,\nu^{3}])$.
The next lemma shows that indeed any representation in $\mathcal F$ has the desired properties. Before we proceed, recall that an elementary operation on an arbitrary multi-set $\mathfrak{m}$ is to choose a pair of linked segments in it and replace the pair by their union and their intersection. By \cite[Theorem 7.1]{MR584084} any irreducible subquotient of $\zeta(\mathfrak{m})$ is of the form $Z(\mathfrak{n})$ where $\mathfrak{n}$ is a multi-set obtained from $\mathfrak{m}$ by a sequence of elementary operations on it.
\begin{lemma}
Let $\pi\in \mathcal F$. Then $\pi$ is $\operatorname{Sp}$-distinguished and imprimitive.
\end{lemma}
\begin{proof}
Let $\mathfrak{m}=\{\Delta_{1},\Delta_{2},\Delta_{3}\}$ be a multi-set satisfying the conditions (1), (2) and (3) of Definition \ref{def: non lad ex} and $\pi=Z(\mathfrak{m})$. By Corollary \ref{cor: dist lad Z}, $Z(\Delta_i)$ is $\operatorname{Sp}$-distinguished, $i=1,2,3$ and therefore by Corollary \ref{open dist},
\[
{\rm I}(\mathfrak{m})=Z(\Delta_{1})\times Z(\Delta_{2})\times Z(\Delta_{3})
\]
is $\operatorname{Sp}$-distinguished.
Applying \cite[Theorem 1.9]{MR584084} to $\zeta(\mathfrak{m})$ we get that $\pi$ occurs as a subquotient of ${\rm I}(\mathfrak{m})$. We now analyze the other possible irreducible subquotients of ${\rm I}(\mathfrak{m})$ using \cite[Theorem 1.9 and Theorem 7.1]{MR584084}. Since $\Delta_{1}\subseteq \Delta_{2}$, any elementary operation on $\mathfrak{m}$ is performed on either $\{\Delta_{1},\Delta_{3}\}$ or on $\{\Delta_{2},\Delta_{3}\}$. The result will respectively contain either $\Delta_{1}\cap\Delta_{3}$ or $\Delta_{2}\cup\Delta_{3}$, which are of odd length.
Since the first is contained in and the second contains all three segments any further sequence of operations will result in a multi-set containing one of them. Thus by Proposition \ref{prop: Z dist} none of these subquotients are $\operatorname{Sp}$-distinguished. The $\operatorname{Sp}$-distinction of $\pi$ now follows from Lemma \ref{lem: dist comp}.
Next we show that $\pi$ is imprimitive. Assume, if possible, that it is not so. Then there exists indices $i,j,k$ such that $\{i,j,k\}=\{1,2,3\}$ and $\pi\cong Z(\Delta_{i})\times Z(\Delta_{j},\Delta_{k})$. By considering the multi-set $\mathfrak{m}^{\vee}$ instead of $\mathfrak{m}$ if required, assume further that $\Delta_{3}\prec \Delta_{2}$ and hence $\Delta_{3}\prec \Delta_{1}$. Note that $Z(\Delta_{1},\Delta_{2})\times Z(\Delta_{3})\cong {\rm I}(\mathfrak{m})$ which is reducible by \cite[Theorem 4.2]{MR584084}. Thus $\pi \cong Z(\Delta_{i},\Delta_{3})\times Z(\Delta_{j})$ where $\{i,j\}=\{1,2\}$. It follows from Proposition \ref{prop: irr max} and condition (2) of Definition \ref{def: non lad ex} that this product is reducible which is a contradiction.
\end{proof}
\section{On distinction by Klyachko subgroups}\label{s: Kl}
We continue the study of Klyachko models for representations of $\operatorname{GL}_n(F)$, following \cite{MR1078382}, \cite{MR2417789} and \cite{MR2414223}.
Over finite fields Klyachko models were introduced in \cite{MR691984}. In that case, it is a disjoint family of models and their direct sum contains every irreducible representation with multiplicity one \cite{MR1129515}.
Over a non-archimedean field, Heumos and Rallis observed that some representations do not admit a Klyachko model and classified those in the unitary dual that do in low rank cases (for $n\le 4$). The second and the third authors showed that the direct sum of all Klyachko models is multiplicity free and prescribed a model to any representation in the unitary dual.
In this section, we reduce the study of Klyachko models on the admissible dual to rigid representations and prove that models behave well with respect to parabolic induction.
\subsection{The Klyachko model setting}
\subsubsection{}
Let $G=G_n$. For a decomposition $n=2k+r$ let
\[
H_{2k,r}=\{\begin{pmatrix} h & X \\ 0 & u \end{pmatrix}: h\in \operatorname{Sp}_{2k}(F),\,X\in M_{2k\times r}(F),\,u\in N_r\}
\]
and $\psi=\psi_{2k,r}$ be defined by
\[
\psi(\begin{pmatrix} h & X \\ 0 & u \end{pmatrix})=\psi(u).
\]
(See \S \ref{sss: gen} for the definition of $N_r$ and its character $\psi$.)
For any $\pi\in\Pi$, being $(H_{2k,r},\psi)$-distinguished is independent of the choice of non-trivial character $\psi$ of $F$. Indeed, for any other character $\psi'\ne 1$ there is a diagonal matrix $a\in G$ normalizing $H_{2k,r}$ such that $\psi'_{2k,r}(h)=\psi_{2k,r}(ah a^{-1})$, $h\in H_{2k,r}$.
\subsubsection{}
Let $\tau$ be the involution on $G$ defined by $g^\tau=w^{-1}{}^tg^{-1}w$ where $w=\sm{0}{I_r}{I_{2k}}{0}$ and let
\[
H_{r,2k}'=H_{2k,r}^\tau=\{\begin{pmatrix} u & X \\ 0 & h \end{pmatrix}: h\in \operatorname{Sp}_{2k}(F),\,X\in M_{r\times 2k}(F),\,u\in N_r\}.
\]
In \cite{MR1078382}, \cite{MR2417789} and \cite{MR2414223} we studied distinction by $(H_{r,2k}',\psi)$. Clearly, $\pi\in\Pi$ is $(H_{2k,r},\psi)$-distinguished if and only if $\pi^\tau$ is $(H_{r,2k}',\psi)$-distinguished. If $\pi\in {\rm Irr}$ then $\pi^\tau\simeq\pi^\vee$, by \cite{MR0404534}, and we get a natural isomorphism
\begin{equation}\label{eq: 2 models}
\operatorname{Hom}_{H_{2k,r}}(\pi,\psi)\simeq \operatorname{Hom}_{H_{r,2k}'}(\pi^\vee,\psi).
\end{equation}
In particular, $\pi$ is $(H_{2k,r},\psi)$-distinguished if and only if $\pi^\vee$ is $(H_{r,2k}',\psi)$-distinguished.
More generally, if $\pi_1,\dots,\pi_t\in{\rm Irr}$ then $(\pi_1\times\cdots\times\pi_t)^\tau\simeq \pi_t^\vee\times \cdots \times \pi_1^\vee$
and therefore
\begin{equation}\label{eq: 2 models ind}
\operatorname{Hom}_{H_{2k,r}}(\pi_1\times\cdots\times\pi_t,\psi)\simeq \operatorname{Hom}_{H_{r,2k}'}(\pi_t^\vee\times \cdots \times \pi_1^\vee,\psi).
\end{equation}
In particular,
$\pi_1\times\cdots\times\pi_t$ is $(H_{2k,r},\psi)$-distinguished if and only if $\pi_t^\vee\times \cdots \times \pi_1^\vee$ is $(H_{r,2k}',\psi)$-distinguished.
\begin{remark}
We remark that the proof of \cite[Theorem 3.7]{MR2417789} applied \cite[Lemma 3.1]{MR2417789}, where \eqref{eq: 2 models} was mistakenly formulated for any representation $\pi$. The isomorphism \eqref{eq: 2 models ind} suffices to fill the gap. In any case, we provide in the sequel an independent generalization of \cite[Theorem 3.7]{MR2417789}.
\end{remark}
\subsubsection{}
Let $\pi\in{\rm Irr}\cap \Pi(G_n)$. If $\pi$ is $(H_{2k,r},\psi)$-distinguished for some decomposition $n=2k+r$ then by \eqref{eq: frob rec} it imbeds in $\operatorname{Ind}_{H_{2k,r}}^{G_n}(\psi)$ and we say that it admits a Klyachko model.
By the uniqueness and disjointness of Klyachko models, \cite[Theorem 1]{MR2414223}, both the imbedding (up to a constant multiple) and the decomposition $n=2k+r$ are uniquely determined by $\pi$. (Indeed, in the main result on distributions \cite[Proposition 1]{MR2414223} implying \cite[Theorem 1]{MR2414223} $H_{2k,r}$ and $H'_{r,2k}$ are in symmetry.) In that case we denote by
\[
r(\pi)=r
\]
the Klyachko type of $\pi$.
\subsubsection{}
The main tool in our study of Klyachko models is the theory of derivatives of representations of $G_n$ developed in \cite{MR0404534}, \cite{MR0425030}, \cite{MR0579172} and \cite{MR584084}. It allows a reduction of many of the problems concerned with Klyachko models to the study of $\operatorname{Sp}$-distinction and generic representations.
For $\pi\in \Pi(G_n)$ and any $r=0,1,\dots,n$ we denote by $\pi^{(r)}$ the $r$-th derivative of $\pi$ as defined in \cite[\S 3.5 and \S 4.3]{MR0579172}. It is a functor from $\Pi(G_n)$ to $\Pi(G_{n-r})$.
As a consequence of \cite[Lemma 4.7(a)]{MR0579172} we have
\begin{lemma}\label{high_der1}
Let $\pi\in \Pi(G_{n})$. Then, ${\rm Supp}(\pi^{(r)})\subseteq {\rm Supp}(\pi)$ for all $0\leq r \leq n$. \qed
\end{lemma}
\subsubsection{}
As observed in \cite[(3.2)]{MR2417789}, it follows from \cite[Proposition 3.7]{MR584084} that for $n=2k+r$ and $\pi\in\Pi(G_n)$ there is a natural linear isomorphism
\begin{equation}\label{eq: der frob}
\operatorname{Hom}_{H_{2k,r}}(\pi,\psi)\simeq \operatorname{Hom}_{\operatorname{Sp}_{2k}(F)}(\pi^{(r)},1).
\end{equation}
This is the reason that we prefer $H_{2k,r}$ to $H_{r,2k}'$.
Note that, in particular, $\pi$ is generic if and only if $\pi^{(n)}\ne 0$.
\subsection{Hereditary property of Klyachko models}
As we observe bellow, Klyachko models behave well with respect to parabolic induction.
\begin{proposition}\label{prop: herad}
Let $\pi_i\in\Pi(G_{n_i})$ and $n_i=2k_i+r_i$ be such that $\pi_i$ is $(H_{2k_i,r_i},\psi)$-distinguished for $i=1,\dots,t$. Then $\pi=\pi_1\times\cdots\times\pi_t$ is $(H_{2k,r},\psi)$-distinguished where $k=k_1+\cdots+ k_t$ and $r=r_1+\cdots+r_t$.
\end{proposition}
\begin{proof}
Induction on $t$ reduces the statement to the case $t=2$ that we now assume.
Let $\pi_s=\nu^s\pi_1\times\pi_2$ for $s\in \mathbb{C}$, so that $\pi=\pi_0$.
Recall that by the Leibnitz rule, \cite[Lemma 4.5]{MR0579172}, $\pi_s^{(r)}$ admits a filtration with factors $(\nu^s \pi_1)^{(i)}\times \pi_2^{(r-i)}\simeq\nu^s \pi_1^{(i)}\times \pi_2^{(r-i)}$, $i=0,\dots,r$.
Note, that there exists a small enough punctured neighborhood $U$ of $s=0$, so that for $i\ne j$ the central characters of the (finitely many) irreducible components of $\nu^s \pi_1^{(i)}\times \pi_2^{(r-i)}$ and of $\nu^s \pi_1^{(j)}\times \pi_2^{(r-j)}$ are disjoint. It follows that for $s\in U$ we have,
\[
\pi_s^{(r)}\simeq \oplus_{i=0}^r (\nu^s \pi_1^{(i)}\times \pi_2^{(r-i)}).
\]
In fact, when realizing all $\pi_s$ in the representation space of $\pi$, this direct sum decomposition is independent of $s\in U$. It follows, that there is a meromorphic map $P_s:\pi_s^{(r)}\rightarrow \nu^s \pi_1^{(r_1)}\times \pi_2^{(r_2)}$, that is surjective for $s\in U$.
By \eqref{eq: der frob}, $\pi_i^{(r_i)}$ is $\operatorname{Sp}_{2k_i}(F)$-distinguished, $i=1,2$ and therefore, by Lemma \ref{BD her}, there exists a non-zero, meromorphic family of linear forms $\ell_s$ such that in a possibly smaller punctured neighborhood $U_0$ of $s=0$ we have $\ell_s\in \operatorname{Hom}_{\operatorname{Sp}_{2k}(F)}(\nu^s \pi_1^{(r_1)}\times \pi_2^{(r_2)},1)$. Therefore, $\ell_s\circ P_s\in\operatorname{Hom}_{\operatorname{Sp}_{2k}(F)}(\pi_s^{(r)},1)$ is non-zero for $s\in U_0$. As in Corollary \ref{open dist}, a leading term argument implies that $\pi^{(r)}=\pi_0^{(r)}$ is $\operatorname{Sp}_{2k}(F)$-distinguished and therefore, by \eqref{eq: der frob}, $\pi$ is $(H_{2k,r},\psi)$-distinguished.
\end{proof}
\subsection{Reduction to cuspidal lines}
We reduce the study of $(H_{2k,r},\psi)$-distinguished representations in $\Pi$ to rigid representations, in fact, more generally to totally disjoint supports (Definition \ref{def: tot disj}).
\begin{proposition}\label{lem: Kly cusp}
Let $\pi_i\in\Pi$, $i=1,\dots,t$ be such that ${\rm Supp}(\pi_i)$ and ${\rm Supp}(\pi_j)$ are totally disjoint for all $i\ne j$. Then $\pi=\pi_1\times\cdots\times\pi_t$ admits a Klyachko model if and only if $\pi_i$ admits a Klyachko model for all $i=1,\dots,t$.
More precisely, $\pi$ is $(H_{2k,r},\psi)$-distinguished if and only if $\pi_i$ is $(H_{2k_i,r_i},\psi)$-distinguished, $i=1,\dots,t$ for some decomposition $r=r_1+\cdots+r_t$.
\end{proposition}
\begin{proof}
The `if' part is immediate from Proposition \ref{prop: herad}. Assume that $\pi\in\Pi(G_n)$, $n=2k+r$ and $\pi$ is $(H_{2k,r},\psi)$-distinguished.
By \eqref{eq: der frob}, $\pi^{(r)}$ is $\operatorname{Sp}_{2k}(F)$-distinguished. Therefore, by the Leibnitz rule, \cite[Lemma 4.5]{MR0579172}, and Lemma \ref{lem: dist comp}, there exists a decomposition $r=r_1+\cdots+r_t$ such that $\pi_1^{(r_1)}\times\cdots\times\pi_t^{(r_t)}$ is $\operatorname{Sp}_{2k}(F)$-distinguished. Since, by Lemma \ref{high_der1}, ${\rm Supp}(\pi_i^{(r_i)})\subseteq{\rm Supp}(\pi_i)$ it now follows from Lemma \ref{lem: useful cusp} that $\pi_i^{(r_i)}$ is $\operatorname{Sp}$-distinguished for all $i=1,\dots,t$. The lemma now follows from \eqref{eq: der frob}.
\end{proof}
\section{Klyachko models for ladder representations}\label{s: kl ladder}
We classify all ladder representations that admit, any given, Klyachko model.
\subsection{Klyachko models for Proper ladders}
\subsubsection{Proper Ladders}\label{sss: prop l}
\begin{definition} A ladder, $\mathfrak{m}=\{\Delta_1,\dots,\Delta_k\}\in\OO_\rho$ is called a \emph{proper ladder} if $\Delta_{i+1}\prec\Delta_i$, $i=1,\dots, k-1$. If $\mathfrak{m}$ is a proper ladder then $L(\mathfrak{m})$ is called a proper ladder representation.
\end{definition}
In fact, if $\mathfrak{m}\in\OO_\rho$ is a proper ladder then $\mathfrak{m}^t$ is also a proper ladder, hence $Z(\mathfrak{m})$ is a proper ladder representation, but this fact will not be used in the sequel.
\subsubsection{}\label{sss: prop}
Note that if $\mathfrak{m}\in\OO_\rho$ is a ladder then it can be decomposed uniquely (up to order) as a sum $\mathfrak{m}=\mathfrak{m}_1+\cdots+\mathfrak{m}_t$ where $\mathfrak{m}_i$ is a proper ladder for all $i=1,\dots,t$ and $\Delta\not\prec\Delta'$ for all $i\ne j$, $\Delta\in\mathfrak{m}_i$ and $\Delta'\in \mathfrak{m}_j$. Therefore, ${\rm Supp}(L(\mathfrak{m}_i))$ and ${\rm Supp}(L(\mathfrak{m}_j))$ are totally disjoint for all $i\ne j$ and, by Lemma \ref{lem; tot disj is irr},
\[
L(\mathfrak{m})=L(\mathfrak{m}_1)\times\cdots\times L(\mathfrak{m}_t).
\]
In other words, any ladder representation is a product of proper ladder representations uniquely determined up to order.
\subsubsection{Right aligned segments} We define the following relation on segments of cuspidal representations.
\begin{definition}\label{def: ra seg}
For segments $\Delta=[a,b]_{(\rho)}$ and $\Delta'=[a',b']_{(\rho)}$ we say that $\Delta'$ is right-aligned with $\Delta$ and write $\Delta' \vdash \Delta$ if
\begin{itemize}
\item $a\ge a'+1$ and
\item $b= b'+1$.
\end{itemize}
We label this relation by the integer $r=d(a-a'-1)$ where $\rho\in \Pi(G_d)$ and write $\Delta'\vdash_r \Delta$.
\end{definition}
Note, in particular, that $\Delta'\vdash_0\Delta$ means that $\Delta=\nu\Delta'$.
\begin{example}
Let $\Delta=[4,7]_{(\rho)}$ and $\Delta'=[0,6]_{(\rho)}$ be segments. Then
\[
\tiny{\xymatrix{
&&&&&&&\overset{4}\circ&\overset{5}\circ\ar@{-}[l]&\overset{6}\circ\ar@{-}[l]&\overset{7}\circ\ar@{-}[l] &\\
&&&\overset{0}\circ&\overset{1}\circ\ar@{-}[l]&\overset{2}\circ\ar@{-}[l]&\overset{3}\circ\ar@{-}[l] &\overset{4}\circ\ar@{-}[l]&\overset{5}\circ\ar@{-}[l]&\overset{6}\circ\ar@{-}[l] & }}
\]
illustrates the relation $\Delta'\vdash_{3d}\Delta$ if $\rho\in\Pi(G_d)$.
\end{example}
\subsubsection{} Before characterizing the proper ladder representations admitting Klyachko models we need the following technical result.
\begin{lemma}\label{lem: left-alg}
Let $d$ be such that $\rho\in \Pi(G_d)$ and $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}\in \OO_\rho$ a proper ladder. Write $\Delta_i=[a_i,b_i]_{(\rho)}$. Suppose that $c_1>\cdots>c_t$ are integers such that $a_i-1\le c_i\le b_i$, $i=1,\dots,t$ and let $\mathfrak{m}_1=\{[c_1+1,b_1]_{(\rho)},\dots,[c_t+1,b_t]_{(\rho)}\}$ and $\mathfrak{m}_2=\{[a_1,c_1]_{(\rho)},\dots,[a_t,c_t]_{(\rho)}\}$ be the associated ladders. If either $\mathfrak{m}_1=0$ or $L(\mathfrak{m}_1)$ is $\operatorname{Sp}$-distinguished and either $\mathfrak{m}_2=0$ or $L(\mathfrak{m}_2)$ is generic then $c_{t-2i}+1=c_{t-2i-1}=a_{t-2i-1}-1$ and $\Delta_{t-2i}\vdash_{r_i}\Delta_{t-2i-1}$ where $r_i=d(a_{t-2i-1}-a_{t-2i}-1)$ for all $i=0,\dots,\lfloor t/2\rfloor-1$. Moreover, if $t$ is odd then $c_1=b_1$.
\end{lemma}
\begin{proof}
If $t=1$ then the lemma follows from the fact that if $\mathfrak{m}_1\ne 0$ then $L(\mathfrak{m}_1)$ is generic and Lemma \ref{gen not sp}.
Assume that $t>1$.
Suppose that $1< i\le t$ is such that $c_i=b_i$ (in particular, $[a_{i},c_{i}]_{(\rho)}$ is not empty). Then, since $\mathfrak{m}$ is a proper ladder, we have $a_{i-1}-1\le b_i=c_i<c_{i-1}$ and therefore $[a_{i-1},c_{i-1}]_{(\rho)}$ is non-empty. But then $[a_i,c_i]_{(\rho)}\prec [a_{i-1},c_{i-1}]_{(\rho)}$ are both in $\mathfrak{m}_2$. By \cite[Theorem 9.7]{MR584084} this contradicts the assumption that $L(\mathfrak{m}_2)$ is generic. Therefore, $c_i<b_i$ for all $i>1$.
By the assumption that $L(\mathfrak{m}_1)$ is $\operatorname{Sp}$-distinguished and Theorem \ref{thm: dist lad}, $\mathfrak{m}_1$ is of Speh type. That is, $c_1<b_1$ if and only if $t$ is even and either way,
\[[c_{t-2i-1}+1,b_{t-2i-1}]_{(\rho)}=\nu[c_{t-2i}+1,b_{t-2i}]_{(\rho)}, \ \ \ i=0,\dots,\lfloor t/2\rfloor-1.
\]
To complete the proof it is only left to show that $c_{t-2i-1}=a_{t-2i-1}-1$ for all $i=0,\dots,\lfloor t/2\rfloor-1$. But if $c_{t-2i-1}\ge a_{t-2i-1}$ then $c_{t-2i}=c_{t-2i-1}-1\ge a_{t-2i-1}-1$, i.e., $[a_{t-2i},c_{t-2i}]_{(\rho)}\prec [a_{t-2i-1},c_{t-2i-1}]_{(\rho)}$ in $\mathfrak{m}_2$ which, again by \cite[Theorem 9.7]{MR584084}, is a contradiction. The lemma follows.
\end{proof}
\subsubsection{}
We now determine the proper ladder representations that admit any particular Klyachko model.
\begin{proposition}\label{prop: prop lad}
Let $\mathfrak{m}=\{\Delta_1,\dots,\Delta_t\}\in\OO_\rho$ be a proper ladder, so that $L(\mathfrak{m})\in \Pi(G_n)$ and let $n=2k+r$. If $t$ is odd, let $s$ be such that $L(\Delta_1)\in \Pi(G_s)$, otherwise, set $s=0$. Then $L(\mathfrak{m})$ is $(H_{2k,r},\psi)$-distinguished if and only if $\Delta_{t-2i}\vdash_{r_i}\Delta_{t-2i-1}$ for some $r_i$, $i=0,\dots,\lfloor t/2\rfloor-1$ and $r=r_0+\cdots +r_{\lfloor t/2\rfloor-1}+s$.
\end{proposition}
\begin{proof}
Let $\pi=L(\mathfrak{m})$ and note that $\operatorname{Ind}_{\operatorname{Sp}_{2k}(F) \times N_{r}}^{M_{(2k,r)}}(1\otimes \psi)=\operatorname{Ind}_{H_{2k,r}}^{P_{(2k,r)}}(\psi)|_{M_{(2k,r)}}$. By \eqref{eq: frob rec}, \eqref{eq: 1st adj} and transitivity of induction we have
\begin{multline}\label{eq: frobs}
\operatorname{Hom}_{H_{2k,r}}(\pi,\psi)\simeq\operatorname{Hom}_{G_n}(\pi,\operatorname{Ind}_{H_{2k,r}}^{G_n}(\psi)) \simeq \\ \operatorname{Hom}_{M_{(2k,r)}}({\bf r}_{M_{(2k,r)},G_n}(\pi),\operatorname{Ind}_{\operatorname{Sp}_{2k}(F) \times N_{r}}^{M_{(2k,r)}}(1\otimes\psi))\simeq \operatorname{Hom}_{\operatorname{Sp}_{2k}(F)\times N_r}({\bf r}_{M_{(2k,r)},G_n}(\pi),1\otimes \psi).
\end{multline}
Assume first that $\pi$ is $(H_{2k,r},\psi)$-distinguished. By \eqref{eq: frobs} and Lemma \ref{lem: dist comp} there is an irreducible component $\sigma_1\otimes\sigma_2$ of ${\bf r}_{M_{(2k,r)},G_n}(\pi)$, (where $\sigma_1\in\Pi(G_{2k})$ and $\sigma_2\in \Pi(G_r)$) so that $\sigma_1$ is $\operatorname{Sp}$-distinguished and $\sigma_2$ is generic.
If $\Delta_i=[a_i,b_i]_{(\rho)}$ then it follows from \cite[Theorem 2.1]{MR2996769} that there exist $c_1>\cdots>c_t$ such that $\sigma_1=L(\mathfrak{m}_1)$ and $\sigma_2=L(\mathfrak{m}_2)$ where $\mathfrak{m}_1=\{[c_1+1,b_1]_{(\rho)},\dots,[c_t+1,b_t]_{(\rho)}\}$ and $\mathfrak{m}_2=\{[a_1,c_1]_{(\rho)},\dots,[a_t,c_t]_{(\rho)}\}$. The `only if' part of the proposition therefore follows from Lemma \ref{lem: left-alg}.
Assume that $\Delta_{t-2i}\vdash_{r_i}\Delta_{t-2i-1}$, $i=0,\dots,\lfloor t/2\rfloor-1$ and $r=r_0+\cdots +r_{\lfloor t/2\rfloor-1}+s$.
Let $c_{t-2i}+1=c_{t-2i-1}=a_{t-2i-1}-1$, $i=0,\dots,\lfloor t/2\rfloor-1$. If $t$ is odd, further let $c_1=b_1$. Let
$\sigma_1=L(\mathfrak{m}_1)$ and $\sigma_2=L(\mathfrak{m}_2)$ where $\mathfrak{m}_1=\{[c_1+1,b_1]_{(\rho)},\dots,[c_t+1,b_t]_{(\rho)}\}$ and $\mathfrak{m}_2=\{[a_1,c_1]_{(\rho)},\dots,[a_t,c_t]_{(\rho)}\}$. Note that $\Delta\not\prec\Delta'$ for any two segments in the ladder $\mathfrak{m}_2$ and therefore $\sigma_2$ is generic by \cite[Theorem 9.7]{MR584084}. It is also clear from the above definitions that $\mathfrak{m}_1$ is of Speh type and therefore $\sigma_1$ is $\operatorname{Sp}$-distinguished by
Theorem \ref{thm: dist lad}. By \cite[Corollary 2.2]{MR2996769}, $\sigma_1\otimes\sigma_2$ is a direct summand (and in particular a quotient) of ${\bf r}_{M_{(2k,r)},G_n}(\pi)$. Therefore \eqref{eq: frobs} and Lemma \ref{drmk: ist quot} complete the proof of the proposition.
\end{proof}
\begin{remark}
If $\pi$ is a proper ladder representation then Proposition \ref{prop: prop lad} provides a recipe for computing $r(\pi)$ and in particular, directly implies the uniqueness of $r(\pi)$. The same is true more generally for ladder representations.
\end{remark}
\subsection{Klyachko models of ladder representations}
\begin{theorem}\label{thm: kly lad}
Let $\pi$ be a ladder representation and assume that $\pi=\pi_1\times\cdots\times\pi_t$ is the unique decomposition of $\pi$ as a product of proper ladder representations (see \S \ref{sss: prop}). Then $\pi$ admits a Klyachko model if and only if $\pi_i$ admits a Klyachko model for all $i=1,\dots,t$. Furthermore, in that case $r(\pi)=r(\pi_1)+\cdots+r(\pi_t)$.
\end{theorem}
\begin{proof}
This is immediate from Proposition \ref{lem: Kly cusp} and \S \ref{sss: prop}.
\end{proof}
\begin{remark}
Based on \eqref{eq: 2 models}, the classification of ladder representations that are $(H_{r,2k}',\psi)$-distinguished is obtained by `reflecting all segments along the origin of their $\mathbb{Z}$-line'.
\end{remark}
\def$'${$'$} \def$'${$'$} \def\leavevmode\lower.6ex\hbox to 0pt{\hskip-.23ex \accent"16\hss}D{\leavevmode\lower.6ex\hbox to
0pt{\hskip-.23ex \accent"16\hss}D}
\def\cftil#1{\ifmmode\setbox7\hbox{$\accent"5E#1$}\else
\setbox7\hbox{\accent"5E#1}\penalty 10000\relax\fi\raise 1\ht7
\hbox{\lower1.15ex\hbox to 1\wd7{\hss\accent"7E\hss}}\penalty 10000
\hskip-1\wd7\penalty 10000\box7}
\def\cfudot#1{\ifmmode\setbox7\hbox{$\accent"5E#1$}\else
\setbox7\hbox{\accent"5E#1}\penalty 10000\relax\fi\raise 1\ht7
\hbox{\raise.1ex\hbox to 1\wd7{\hss.\hss}}\penalty 10000 \hskip-1\wd7\penalty
10000\box7} \def\cftil#1{\ifmmode\setbox7\hbox{$\accent"5E#1$}\else
\setbox7\hbox{\accent"5E#1}\penalty 10000\relax\fi\raise 1\ht7
\hbox{\lower1.15ex\hbox to 1\wd7{\hss\accent"7E\hss}}\penalty 10000
\hskip-1\wd7\penalty 10000\box7} \def$'${$'$}
\def\leavevmode\lower.6ex\hbox to 0pt{\hskip-.23ex \accent"16\hss}D{\leavevmode\lower.6ex\hbox to 0pt{\hskip-.23ex \accent"16\hss}D}
\def\cftil#1{\ifmmode\setbox7\hbox{$\accent"5E#1$}\else
\setbox7\hbox{\accent"5E#1}\penalty 10000\relax\fi\raise 1\ht7
\hbox{\lower1.15ex\hbox to 1\wd7{\hss\accent"7E\hss}}\penalty 10000
\hskip-1\wd7\penalty 10000\box7}
\def\polhk#1{\setbox0=\hbox{#1}{\ooalign{\hidewidth
\lower1.5ex\hbox{`}\hidewidth\crcr\unhbox0}}} \def\dbar{\leavevmode\hbox to
0pt{\hskip.2ex \accent"16\hss}d}
\def\cfac#1{\ifmmode\setbox7\hbox{$\accent"5E#1$}\else
\setbox7\hbox{\accent"5E#1}\penalty 10000\relax\fi\raise 1\ht7
\hbox{\lower1.15ex\hbox to 1\wd7{\hss\accent"13\hss}}\penalty 10000
\hskip-1\wd7\penalty 10000\box7}
\def\ocirc#1{\ifmmode\setbox0=\hbox{$#1$}\dimen0=\ht0 \advance\dimen0
by1pt\rlap{\hbox to\wd0{\hss\raise\dimen0
\hbox{\hskip.2em$\scriptscriptstyle\circ$}\hss}}#1\else {\accent"17 #1}\fi}
\def\bud{$''$} \def\cfudot#1{\ifmmode\setbox7\hbox{$\accent"5E#1$}\else
\setbox7\hbox{\accent"5E#1}\penalty 10000\relax\fi\raise 1\ht7
\hbox{\raise.1ex\hbox to 1\wd7{\hss.\hss}}\penalty 10000 \hskip-1\wd7\penalty
10000\box7} \def\lfhook#1{\setbox0=\hbox{#1}{\ooalign{\hidewidth
\lower1.5ex\hbox{'}\hidewidth\crcr\unhbox0}}}
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{Introduction}
How do molecules form? This has been recognized as one of the ten unsolved
mysteries of Chemistry, enumerated in 2013 for the Year of Chemistry
Celebration \cite{SciAm}. Indeed, a new entity emerges when two identical
atoms meet. The reciprocal is also true: as a dimer approaches a catalyst's
surface, it may break down. But when and how does this break down precisely
happen? What distinguishes these two different quantum objects, i.e. the
molecule and the two independent atoms? It is natural to think that as some
control parameter move, e.g. an inter-atomic distance, a sort of
discontinuity or phase transition should happen. While a quantum calculation
can be set up to simulate such reaction, the calculations of an increasingly
realistic system quickly begin to overwhelm even the most powerful computer.
While DFT calculations hint a change in chemical bonds as the molecule-catalyst interaction
increases when the molecule approaches to the
surface \cite{BONDBREACK-Santos-2011-Diatomic-molecules}, this is confronted
with the fact that in a finite system no actual discontinuities can happen.
The key for the molecule formation/dissociation mystery can be found in P.
W. Anderson's inspiring paper \textquotedblleft More is
Different\textquotedblright\ \cite{anderson1972more}. There, Anderson
recalled that the inversion oscillations in ammonia-like molecules suffer a
sort of transition into a non-oscillating mode as the masses are increased.
Much as in a classical oscillator transition to an over-damped regime, the
crucial ingredient is the infinite nature of the environment which induces
dissipation while preventing the occurrence of Poincare's recurrences and
enable a dynamical phase transition. These concepts were formalized in the
context of the Rabi oscillations in a quantum system: a spin dimer immersed
in an environment of spins. Since this is solved in the thermodynamic limit
of infinitely many spins which provide the crucial continuum spectrum. \cite%
{w1954mathematical,Alvarez-LevsteinJCP2006}. In this case, the finite Rabi
frequency undergoes a non-analytic transition into a non-oscillatory mode as
the interaction with the environment increases \cite{sachdev2007quantum,
RedutionismBerry,leggett1987dynamics}. This mathematical discontinuity was
termed Quantum Dynamical Phase Transition (QDPT) \cite{Pastawski2007278}.
While the application of these ideas to molecular dissociation/formation is
not completely straightforward, in a previous paper we succeeded in
describing H$_{2}$ molecule formation/dissociation in the presence of a
catalyst as a QDPT \cite{grado1}. This description was achieved using a
variant of the model introduced by D. M. Newns for hydrogen adsorption in a
metallic surface \cite{NewnsHydrogen-and-dband}. However, that analysis was
restricted to the case when \textit{the molecular axis is perpendicular to
the catalyst surface}. In \cite{grado1} the environment provides the
infinitely many catalyst orbitals whose influence had to be treated beyond
linear response. Indeed, the interaction among the crystal states and the
dimer orbitals dramatically perturb each other and has to be obtained
through a self-consistent Dyson equation. In particular, the substrate
induces imaginary corrections to the molecular energies, accounting for
their finite lifetime. These complex energies, as those obtained from the
Fermi Golden Rule, represent resonances and are accounted by a non-Hermitian
Hamiltonian \cite{rotter2015review}. Our main result was that two \textit{%
resonances are formed inside} the $d$ band and that they present analytical
discontinuities as a function of the molecule-substrate interaction
(distance) \cite{Berry,Rotter1,Rotter2}. Thus, the \textit{molecular
dissociation/formation} was identified as the non-analytic \textit{%
collapse/splitting of these resonances}.
In this paper, we address another reaction mechanism that is known to occur
when a H$_{2}$ molecule approaches a catalyst with its \textit{molecular
axis parallel to the surface}. It is found that molecular dissociation is
also a phase transition associated to an analytic discontinuity, but of
different and unusual nature: the molecule is destabilized by the transition
from non-physical virtual states into actual localized states. For the rest
of the article we will be dealing with the same model and tools introduced
in our previous work \cite{grado1} which, in this case, provide
substantially new perspective into the molecular dissociation/formation
problem.
\section{The model}
Given a homonuclear molecule AB and a metal electrode with a
half filled $d$ band, two independent geometries arise to describe the
interaction. The particular configuration of a molecule approaching with its
axis perpendicular to the metal surface, was previously investigated in
reference \cite{grado1}. A fully different problem arises when the axis
along the molecule lies parallel to the surface. In this configuration the
distances between a given atom belonging to the metal surface and both atoms
forming the molecule remain equal, i.e. $d_{\text{A}}=d_{\text{B}}=d$ ( see
Fig. \ref{esquema}). Therefore, both atoms interact identically with the
metal, resulting in a completely different Hamiltonian respect to the
perpendicular case, and hence yielding a dissimilar kind of transition.
\begin{figure}[tbh!]
\centering{\ \includegraphics[width=5cm]{fig1.pdf} }
\caption{Homonuclear molecule interacting with a metallic surface. The
principal axis of the molecule is parallel to the surface and the distance
of each atom to the substrate are the same.}
\label{esquema}
\end{figure}
To set up the model Hamiltonian for the interaction between the molecule and
the metal, we write the molecule's Hamiltonian as:
\begin{equation*}
\hat{H}_{\mathrm{mol}}=E_{A}\left\vert A\right\rangle \left\langle
A\right\vert +E_{B}\left\vert B\right\rangle \left\langle B\right\vert
-V_{AB}\left( \left\vert A\right\rangle \left\langle B\right\vert
+\left\vert B\right\rangle \left\langle A\right\vert \right) .
\end{equation*}%
The atomic energies $E_{A}$ and $E_{B}$ are identical and their degeneracy
is broken by the mixing element $-V_{AB}$ that leads to the bonding and
antibonding states, i.e. the Highest Occupied Molecular Orbital (HOMO) and
Lowest Unoccupied Molecular Orbital (LUMO), respectively. In this
orientation, the molecule only can have substantial overlap with the metal $%
d_{z^{2}}$ and $d_{xz}$ orbitals of the underlying metallic atom. Therefore,
$z$ is considered to be perpendicular to the surface and $x$ is chosen
parallel to the molecular axis. Both orbitals interact with the target
molecule in different ways \cite{Hoffman}, as depicted in Fig. \ref{sitios}.
On one side, the overlap of the $d_{z^{2}}$ with the atomic orbitals A and B
have the same the sign and magnitude, resulting in a Hamiltonian coupling
element $-V_{0}$. On the other side, the molecule also interacts with the $%
d_{xz}$ orbital of the metal. In this case, while having equal strengths a
different sign appears for each atomic orbital. Taking these considerations
into account, there are two concurrent mechanisms for molecule-metal
interaction :
\begin{equation*}
\hat{V}_{\mathrm{int}}=V_{0}(\left\vert A\right\rangle \left\langle
d_{z^{2}}\right\vert +\left\vert B\right\rangle \left\langle
d_{z^{2}}\right\vert )+\lambda V_{0}(-\left\vert A\right\rangle \left\langle
d_{xz}\right\vert +\left\vert B\right\rangle \left\langle d_{xz}\right\vert
),
\end{equation*}%
where $\left\vert d_{z^{2}}\right\rangle $ and $\left\vert
d_{xz}\right\rangle $ are the metallic orbitals that interact with the
molecule. Furthermore, we have included a $\lambda $ factor to account for
the difference among the interaction strengths with the two $d$ orbitals.
\begin{figure}[tbh!]
\centering{\ \includegraphics[width=8cm]{fig2.pdf} }
\caption{Different signs for the interaction between the molecule and the
metallic atomic orbital, due to the lobe phase shift for the atomic orbital
functions $d_{z^2}$ and $d_{xz}$. The $\protect\lambda$ factor accounts for
the different strength interaction between the molecule and the orbitals $%
d_{z^2}$ and $d_{xz}$.}
\label{sitios}
\end{figure}
In Fig. \ref{sitios} we represent explicitly two, assumed independent, sets
of metallic $d$ orbitals associated with each symmetry of the surface
orbitals (i.e. $\left\vert d_{z^{2}}\right\rangle $ and $\left\vert
d_{xz}\right\rangle $). Therefore, the relevant part of the metal
Hamiltonian can be represented using a narrow band model. This approximation
was first proposed by Newns \cite{NewnsHydrogen-and-dband}, who stated that
the projection of the $d$ band Local Density of States (LDoS) over the
specific orbital (either $d_{z^{2}}$ or $d_{xz}$) could be schematized as a
semielliptic energy band that strongly interacts with the molecule \cite%
{xin2014effects}. This picture is validated by appealing to a Lanczos's
transformation \cite{grado1,lanczos1950iteration,haydock1972electronic} to
obtain this simple electronic structure for the $d$ band. The basic
procedure is visualised in Fig. \ref{lanczos} for a two dimensional metal
represented as two distinct collections of orthonormal $d$ orbitals. By
choosing one of the interacting metallic orbitals as a reference, the
intermetallic interactions provide (through the Lanczos's procedure) for
combination of atomic $d$ orbitals consistent with the initial symmetry.
Typically, these are progressively included according to their distance to
the initial orbital. These \textquotedblleft collective\textquotedblright\
substrate orbitals are naturally arranged in the Hilbert space in order to
evidence the tridiagonal nature of the Hamiltonian in the new basis. By
means of this procedure, the general three dimensional geometry of a
catalyst is reduced to a effective linear chain. The same reasoning applies
for both symmetries. Then, we can write the metal $d_{z^{2}}$ Hamiltonian as:
\begin{equation}
\hat{H}_{\mathrm{met}}^{z^{2}}=\sum_{n=1}^{\infty }E_{n}^{z^{2}}\left\vert
n\right\rangle \left\langle n\right\vert -\sum_{n=1}^{\infty
}V_{n,n+1}^{z^{2}}\left( \left\vert n\right\rangle \left\langle
n+1\right\vert +\left\vert n+1\right\rangle \left\langle n\right\vert
\right) , \label{emet}
\end{equation}%
where $\left\vert n\right\rangle $ and $E_{n}^{z^{2}}$ are the $n$-th
collective metal orbital obtained by the Lanczos's transformation and the
energy corresponding to that orbital, respectively. For the sake of
simplicity, all the hopping elements $V_{n,n+1}^{z^{2}}$ are considered to
be equal to $V$. This is consistent with the fast convergence of the hopping
elements, first addressed by Haydock et al. \cite{haydock1972electronic}. A
similar Hamiltonian $\hat{H}_{\mathrm{met}}^{xz}$ is obtained for the $xy$
symmetry. Thus,
\begin{equation*}
\hat{H}_{\mathrm{met}}^{{}}=\hat{H}_{\mathrm{met}}^{z^{2}}+\hat{H}_{\mathrm{%
met}}^{xz}.
\end{equation*}
In order to obtain an optimal configuration for our discussion on the
dissociation process \cite{Hush}, we make the $d$ band to be centered around
the Fermi energy $E$ by making $E_{A}=E_{B}=E_{n}=E$. Then, the bonding and
antibondig molecular states, i.e. HOMO and LUMO, fall outside the band as $%
2|V_{AB}|>4|V|$ \cite{Santos2011314}. This choice is consistent with the
standard knowledge of the Markus-Hush theory for optimal conditions of
electron transfer and molecular dissociation. In this work, we used $%
V_{AB}/V=2.5$ which is typical for H$_{2}.$
\begin{figure}[tbh!]
\centering{\ \includegraphics[width=8cm]{fig3.pdf} }
\caption{Effective non-Hermitian Hamiltonian due Lanczos's transformation
from a molecule A-B (in blue), interacting with a 2D metal substrate
composed of two distinct collections of $d$ orbitals. The transformation
implies combining each layer of orbitals at the same distance of the
interacting atom. The decimation process results in a four dimensional
Hamiltonian with the metal represented as two effective self-energies.}
\label{lanczos}
\end{figure}
The main features of the system, i.e. energy spectrum and relevant
eigenvalues properties, could be obtained using a decimation procedure \cite%
{Levstein-Damato,pastawski-medina}. This formulation deploys an infinite
order perturbation theory for the interaction $\hat{V}_{\mathrm{int}}$ to
dress the molecular Hamiltonian $\hat{H}_{\mathrm{mol.}}$ into an effective
molecular Hamiltonian that accounts for the presence of the catalyst, and
yields a complex correction, $\Sigma $, to the molecular bonding and
antibonding energies. This is sketched in the bottom panel of Fig. \ref%
{lanczos}. This precisely defined procedure resorts to the Green's function
matrix associated with the total Hamiltonian $\hat{H}=\hat{H}_{\mathrm{mol}}+%
\hat{H}_{\mathrm{met}}+\hat{V}_{\mathrm{int}}$,
\begin{equation}
\mathbb{G(}\varepsilon )=(\varepsilon \mathbb{I}-\mathbb{H})^{-1}.
\label{green}
\end{equation}%
We are going to profit from the fact that the poles of the Green's function
are the eigenvalues of the system. At this point, a brief introduction to
the decimation technique is convenient for the sake of clarity. Let us first
consider the molecular Hamiltonian without the presence of the metal:
\begin{equation}
\mathbb{H}_{\mathrm{mol}}=\left[
\begin{array}{cc}
E_{A} & -V_{AB} \\
-V_{AB} & E_{B}%
\end{array}%
\right] . \label{e1}
\end{equation}%
Then, the Green's function matrix adopts the form:
\begin{equation}
\mathbb{G}_{\mathrm{mol}}=\dfrac{1}{(\varepsilon -E_{A})(\varepsilon
-E_{B})-|V_{AB}|^{2}}\left[
\begin{array}{cc}
\varepsilon -E_{B} & V_{AB} \\
V_{AB} & \varepsilon -E_{A}%
\end{array}%
\right] . \label{molg}
\end{equation}%
The Green's function for atom A, the first diagonal element of $\mathbb{G}_{%
\mathrm{mol}}$, can be written as
\begin{equation*}
G_{\mathrm{mol}}^{AA}=(\varepsilon -E_{A}-\Sigma _{A})^{-1}
\end{equation*}%
Therefore, the energy of atom A is modified by the presence of the atom B
through the self-energy
\begin{equation*}
\Sigma _{A}=|V_{AB}|^{2}/(\varepsilon -E_{B}).
\end{equation*}%
This decimation procedure can be extended to the full semi-infinite chain
that describes the components of the $d$ band that couple with the HOMO and
LUMO according to their symmetry. The procedure consists on
\textquotedblleft dressing\textquotedblright\ the successive
\textquotedblleft Lanczos's orbitals\textquotedblright\ with the
corresponding self-energies to account for the interaction with the
neighbour atom at the right. In a finite system of $N+2$ orbitals, $\Sigma
_{A}$ is written in terms of $N+1$ levels of a continued fraction until one
reaches the last level. To simplify the study of the spectral density, the
energies of the system can be renormalized by introducing an imaginary small
quantity $-\mathrm{i}\eta $, thus $E\rightarrow E-\mathrm{i}\eta $. This
energy correction can be seen as a weak environmental interaction, a role
that could be assigned to the $sp$ band states \cite{CattenaBustosPRB2010}.
Thus, in the thermodynamic limit of a semi-infinite chain ($N\rightarrow
\infty $), the self-energy correction due to the metal becomes:
\begin{equation}
\Sigma (\varepsilon )=\dfrac{|V|^{2}}{\varepsilon -\left( E-\mathrm{i}\eta
\right) -\Sigma (\varepsilon )}=\Delta (\varepsilon )-\mathrm{i}\Gamma
(\varepsilon ), \label{sigma}
\end{equation}%
By setting $E=0$ in the whole system (i.e. setting down the Fermi level as
the energy reference) the analysis is further simplified. Equation \ref%
{sigma} has two solutions with different signs. The solution with physical
meaning provides a retarded response and results:
\begin{equation}
\Sigma (\varepsilon )=\dfrac{\varepsilon +\mathrm{i}\eta }{2}-\mathrm{sgn}%
(\varepsilon )\times \left( \sqrt{\dfrac{r+x}{2}}+\mathrm{i\times sgn}%
(y)\times \sqrt{\dfrac{r-x}{2}}\right) , \label{e4}
\end{equation}%
with $x=\dfrac{\varepsilon ^{2}-\eta ^{2}}{2}-V^{2}$, $y=\dfrac{\varepsilon
\eta }{2}$ and $r=\sqrt{x^{2}+y^{2}}$.
Then, the restriction to the first four orbitals of the total Hamiltonian
can be written in a simple way:
\begin{equation}
\widetilde{\mathbb{H}}=\left[
\begin{array}{cccc}
\Sigma ^{z^{2}}(\varepsilon ) & -V_{0} & -V_{0} & 0 \\
-V_{0} & -\mathrm{i}\eta & -V_{AB} & +\lambda V_{0} \\
-V_{0} & -V_{AB} & -\mathrm{i}\eta & -\lambda V_{0} \\
0 & +\lambda V_{0} & -\lambda V_{0} & \Sigma ^{xz}(\varepsilon )%
\end{array}%
\right] . \label{e5}
\end{equation}
Now, a basis change can be made to a molecular \textit{bonding} and \textit{%
antibonding} representation. Equation \ref{e6} shows the Hamiltonian in the
new basis. Notice that, the bonding state (second diagonal element) does not
interact with $\Sigma ^{xz}(\varepsilon )$ (fourth diagonal element) and the
antibonding state (third diagonal element) does not interact with $\Sigma
^{z^{2}}(\varepsilon )$ (first diagonal element):
\begin{equation}
\widetilde{\mathbb{H}}^{\prime }=\widetilde{\mathbb{H}}_{\mathrm{+}}\otimes
\widetilde{\mathbb{H}}_{\mathrm{-}}=\left[
\begin{array}{cccc}
\Sigma ^{z^{2}}(\varepsilon ) & -\sqrt{2}V_{0} & 0 & 0 \\
-\sqrt{2}V_{0} & -V_{AB}-\mathrm{i}\eta & 0 & 0 \\
0 & 0 & V_{AB}-\mathrm{i}\eta & \sqrt{2}\lambda V_{0} \\
0 & 0 & \sqrt{2}\lambda V_{0} & \Sigma ^{xz}(\varepsilon )%
\end{array}%
\right] . \label{e6}
\end{equation}%
Therefore, the system is naturally detached in two portions in which the
Green's function matrices can be solved independently. For the bonding
subspace, i.e. the bonding molecular orbital interacting with $\Sigma
^{z^{2}}(\varepsilon )$, the Green's function takes the form:
\begin{equation}
\mathbb{G}_{+}=\dfrac{1}{(\varepsilon +V_{AB}+\mathrm{i}\eta )(\varepsilon
-\Sigma ^{z^{2}}(\varepsilon ))-2V_{0}^{2}}\left[
\begin{array}{cc}
\varepsilon +V_{AB}+\mathrm{i}\eta & -\sqrt{2}V_{0} \\
-\sqrt{2}V_{0} & \varepsilon -\Sigma ^{z^{2}}(\varepsilon )%
\end{array}%
\right] , \label{e7}
\end{equation}%
while, for the antibonding molecular orbital interacting with $\Sigma
^{xz}(\varepsilon )$, there is a subspace where
\begin{equation}
\mathbb{G}_{-}=\dfrac{1}{(\varepsilon -V_{AB}+\mathrm{i}\eta )(\varepsilon
-\Sigma ^{xz}(\varepsilon ))-2(\lambda V_{0})^{2}}\left[
\begin{array}{cc}
\varepsilon -\Sigma ^{xz}(\varepsilon ) & \sqrt{2}\lambda V_{0} \\
\sqrt{2}\lambda V_{0} & \varepsilon -V_{AB}+\mathrm{i}\eta%
\end{array}%
\right] . \label{e8}
\end{equation}
For the rest of the article $\lambda$ will be set $\lambda \sim 1$ and $%
\Sigma ^{xz}=\Sigma ^{z^{2}}$. The eigenenergies and resonances of the
system are obtained by finding the poles of Eqs. \ref{e6} and \ref{e7}. This
is achieved solving the equations:
\begin{eqnarray}
\label{Eq-CharacBonding}
\varepsilon +V_{AB}-2\alpha \Sigma (\varepsilon )=0, \\
\label{Eq-CharacAntibonding}
\varepsilon -V_{AB}-2\alpha \Sigma (\varepsilon )=0.
\end{eqnarray}
Equation \ref{Eq-CharacBonding} accounts for the poles corresponding to the
bonding state interacting with the $d_{z^{2}}$ band and Eq. \ref%
{Eq-CharacAntibonding} for the poles of the antibonding state interacting
with the $d_{xz}$ band.
\section{Molecular dissociation}
A first hint for molecular dissociation arises from analysing the molecular
bonding orbital that intercats with the $d$ band through the $d_{z^{2}}$
orbital, Fig. \ref{polo+}. In this case, Eq. \ref{Eq-CharacBonding} provides
two poles which are bellow the $d$ band at the molecular bonding energy $%
\varepsilon =-V_{AB}$. One is a physical localized pole (green line in Fig. %
\ref{polo+}) which corresponds to the bonding state $\left\vert
AB\right\rangle $. As the interaction increases, $\left\vert AB\right\rangle
$ evolves to a bonding combination between the bonding state of the molecule
and the metal, i.e. $\left\vert (AB)d_{z^{2}}\right\rangle $, becoming more
localized and its energy lying well below the Fermi level. The other pole
corresponds to a non-physical virtual state which, as the interaction
increases, escapes to negative energies and reappears at positive values
(red dots in Fig. \ref{polo+}). As the non-physical pole gets closer to the $%
d$ band, it finally meets the band-edge and suffers a transition into a
physical localized state. This is an antibondig combination between the
molecular bonding state and the metal $\left\vert ((AB)d_{z^{2}})^{\ast
}\right\rangle $ (blue line). In this scenario, bound weakening occurs
because occupying the $\left\vert (AB)d_{z^{2}}\right\rangle $ state implies
diminish the occupation of the bonding $\left\vert AB\right\rangle $ from
100\% into a final 50\%. Indeed, the molecular bonding state now has 50\%
participation in the unoccupied $\left\vert ((AB)d_{z^{2}})^{\ast
}\right\rangle $ localized orbital that emerged from the upper top of the $d$
band.
\begin{figure}[tbh!]
\centering{\ \includegraphics[width=8cm]{fig4.pdf} }
\caption{Poles of the Green's function for the parallel configuration when
the molecule interacts with the $d_{z^2}$ orbital. }
\label{polo+}
\end{figure}
The previous discussion has a precise equivalence in the analysis of the
states that evolve from the molecular antibonding state. However, the same
formulation has now completely different meaning. The molecular antibonding
state interacts with the $d$ band through $d_{xz}$. The poles resulting from
Eq. \ref{Eq-CharacAntibonding}, are shown in Fig. \ref{polo-}. At the
antibonding energy $\varepsilon =V_{AB}$, two poles appear. A physical
localized state, related to the molecular antibonding state $\left\vert
(AB)^{\ast }\right\rangle $ (blue line in Fig. \ref{polo-}), whose energy
increases as $V_{0}$ increases and becomes an antibonding combination
between the molecular antibonding state and the metal site $%
\left\vert((AB)^{\ast }d_{xz})^{\ast }\right\rangle $. The other pole at $%
\varepsilon =V_{AB}$ is a virtual state \cite{Dente,bustos2010buffering}
(red dots in Fig. \ref{polo-}) which diverges as $V_{0}$ increases and
appears again from $-\infty$ until its energy touches the $d$ band. At this
critical value, the virtual state suffers a transition and becomes a
localized state (green line in Fig. \ref{polo-}) which is a bonding
combination between the molecular antibonding state and the metal band $%
\left\vert (AB)^{\ast }d_{xz}\right\rangle $. Therefore, molecular
dissociation can be interpreted as occurring at the precise value when the
virtual pole touches the $d$ band and becomes the localized, and occupied,
state$\left\vert(AB)^{\ast }d_{xz}\right\rangle$. Thus, molecular
dissociation occurs at a non-analytical point of the physical observables,
e.g. total energies. At this point the molecular electrons have a transition
from an increasingly occupied bonding state that participates of the
delocalized band into a localized combination between the $d$ states and
antibonding molecular orbital. This is a form of Quantum Dynamical Phase
Transition which, to the best of our knowledge, has not been identified
before in the context of molecular dissociation.
\begin{figure}[tbh]
\centering{\ \includegraphics[width=8cm]{fig5.pdf} }
\caption{Poles of the Green's function for the parallel configuration when
the molecule interacts with the $d_{xz}$ orbital. The molecule dissociation
as a QDPT can be observed when the interaction is with the $d_{xz}$ band.}
\label{polo-}
\end{figure}
From the results it becomes evident that the most interesting situation is
when the antibonding molecular orbital interacts with the $d_{xz}$. From Eq. %
\ref{e8} we get the diagonal Green's function at the $d_{xz}$ metallic
orbital:
\begin{equation}
G_{d_{xz}}(\varepsilon )=\dfrac{1}{\varepsilon +\mathrm{i}\eta -\Sigma
(\varepsilon )-\dfrac{2(\lambda V_{0})^{2}}{\varepsilon +\mathrm{i}\eta
+V_{AB}}}. \label{e11}
\end{equation}%
The LDoS for the $d$ band can be obtained from Eq. \ref{e11},
\begin{equation}
N_{d_{xz}}(\varepsilon )=-\dfrac{1}{\pi }\lim_{\eta \rightarrow 0^{+}}%
\mathrm{Im}\left[ G_{d_{xz}}(\varepsilon )\right] , \label{e12}
\end{equation}%
which becomes of great help to reinforce and extend the previous discussion.
This LDoS is shown in Fig. \ref{dens} for $\lambda V_{0}/V$ between 0 and
3.6 for $\lambda=1$. When $V_{0}\sim 0$ the shape of the LDoS corresponds to
a non interacting $d_{xz}$ band. As $\lambda V_{0}$ increases the $d_{xz}$
band starts to mix with the antibonding state of the dimer. The energy of
this antibonding combination $\left\vert ((AB)^{\ast }d_{xz})^{\ast
}\right\rangle$, progresses toward increasingly positive values as the
interaction grows. Meanwhile, the virtual state approaches the $d_{xz}$ band
from negative energies while it produces an \textquotedblleft
attraction\textquotedblright\ that increases the LDoS near the band edge. As
the virtual state meets the band a localized state \textit{emerges} from the
band edge and gains weight. A similar issue was recently discussed in the
context of engineered plasmonic excitations in metallic nanoparticle arrays
\cite{bustos2010buffering}. There, it was shown analytically that the
distorted band is the product among the original semi-elliptic band and a
Lorentzian centered in the virtual state. This concentrates a density of
states near band edge until it becomes a divergence and a localized state is
expelled at a critical interaction strength, shown as a dot in Fig. \ref%
{polo-}.
\begin{figure}[tbh]
\centering{\ \includegraphics[width=10cm]{fig6.pdf} }
\caption{LDoS of the $d$ band. As $V_{0}$ increases a state is expelled from
the band and, after the transition point, forms the localized state $%
\left\vert (AB)^{\ast }d_{xz}\right\rangle $, $\protect\eta =0.01$ eV.}
\label{dens}
\end{figure}
The previous conclusion is reinforced by the analysis of LDoS at the
antibonding orbital. Figure \ref{dens2} shows how the unoccupied antibonding
state $\left\vert (AB)^{\ast }\right\rangle $ looses its weight towards a
participation on combination with the $d_{xz}$ band which finally emerges as
an \textit{occupied localized state}. This is a crucial contribution to
molecular destabilization. As in the first part of this work \cite{grado1}
the new transition can be seen as a successful implementation of a
non-Hermitian Hamiltonian \cite{rotter2015review} in a well defined model.
\begin{figure}[tbh]
\centering{\ \includegraphics[width=10cm]{fig7.pdf} }
\caption{LDoS of the molecular antibonding state $\left\vert (AB)^{\ast
}\right\rangle$, interacting with the metallic orbital $d_{xz}$, as $V_{0}$
increases, $\protect\eta =0.05$ eV.}
\label{dens2}
\end{figure}
Notice that Figs. \ref{dens} \ and \ref{dens2} also serve to discuss the
interaction between the bonding molecular state $\left\vert AB\right\rangle $
and the $d_{z^{2}}$ band by exchanging the sign of the energy. Thus, in this
case, the $\left\vert ((AB)d_{z^{2}})^{\ast }\right\rangle $ emerges as an
unoccupied localized state above the $d_{z^{2}}$ band, while $\left\vert
AB\right\rangle $ state loses occupation as the $\left\vert
(AB)d_{z^{2}}\right\rangle $ state forms with increasing interaction.
\section{Conclusions}
As a H$_{2}$ molecule approaches a catalyst with its axis parallel to the
surface, the interaction creates two independent collective orbitals which
are superpositions with different surface $d$ orbitals that are part of
their corresponding metallic bands. The molecular bonding state becomes
mixed with the $d_{z^{2}}$ band while the molecular antibonding state
interacts with the $d_{xz}$ band. This gives rise to two processes described
by the same algebra, but with different physical meanings as their energies
are the reverse of each other.
On one side, the mixing of the molecular bonding state produces a decrease
of its occupation. While this occurs, the LDoS of the $d_{z^{2}}$ band is
distorted at its upper edge much as if it were \textquotedblleft
attracted\textquotedblright\ upwards. Finally, at a critical interaction
strength the divergent peak is expelled as a localized state emerging from
the upper (i.e. unoccupied) part of the $d_{z^{2}}$ band. This new \textit{%
unoccupied} state is an antibonding combination among the surface $d_{z^{2}}$
orbital and the \textit{bonding} state of the dimer.
On the other side, a fraction of the molecular antibonding state gets
increasingly mixed with the $d_{xz}$ metallic band. This produces a decrease
of the dimer participation on its unoccupied antibonding combination.
Simultaneously, the $d_{xz}$ \ LDoS is \textquotedblleft
attracted\textquotedblright\ towards its lower edge until it finally emerges
as an \textit{occupied localized state} build as a bonding combination among
the molecular \textit{antibonding} state and the $d_{xz}$ band.
These simultaneous mixing processes, i.e. the depopulation of the molecular
bonding state and the occupation of the molecular antibonding state, both
schematized in Fig. \ref{figura7}, are responsible for the dimer
destabilization that leads to its breakdown.
\begin{figure}[tbp]
\begin{equation*}
\xymatrix{ &&&\left\vert((AB)d_{z^{2}})^*\right\rangle&\\ \left\vert
AB\right\rangle\ar @{<->} [r] &d_{z^2}\ar[urr]\ar[drr]&&&\txt{$\left\vert
AB\right\rangle$ 50 \% unoccupied}\\ &&&\left\vert(AB)d_{z^{2}}\right\rangle
& \\ &&&\left\vert((AB)^*d_{xz})^*\right\rangle& \\
\left\vert(AB)^*\right\rangle \ar @{<->} [r]&
d_{xz}\ar[urr]\ar[drr]&&&\txt{$\left\vert(AB)^*\right\rangle$ 50 \%
occupied}\\ &&&\left\vert(AB)^*d_{xz}\right\rangle& }
\end{equation*}%
\caption{The interaction of the bonding molecular orbital with the $%
d_{z^{2}} $ band shields an antibondig combination that depopulates this
molecular orbital, while the occupied fraction losses weight towards the $%
d_{z^{2}}$ band. Simultaneously, the interaction of the antibonding
molecular orbital with $d_{xz}$ band enforces this molecular state to split
among an antibonding combination and an emergent bonding one that is
interpreted as the molecular breakdown. }
\label{figura7}
\end{figure}
While the essence of the molecule dissociation mechanisms are already hinted
by the resolution of toy models for the catalyst such as small metallic
clusters or even a single metal atom, the criticality of the dissociation
transition would not be readily captured. Indeed, as in the first part of
this work \cite{grado1}, the quasi-continuum nature of a metallic substrate
is crucial to describe dissociation as an analytical discontinuity. In this
case, we interpreted dissociation as the emergence of the localized state
from the band edges as the interaction strength increases. This is an actual
quantum dynamical phase transition. Remarkably, the elusive virtual states
(i.e. states that are non-physical poles of $\mathbb{G}(\varepsilon )$ \cite%
{landau1965course, moiseyev}) acquire a physical meaning as
\textquotedblleft attractors\textquotedblright\ of a distortion of the
continuum band creating a LDoS divergence that finally expels a localized
state. This is, a non-analytical transition.
\section*{Acknowledgements}
We acknowledge the financial support from CONICET (PIP 112-201001-00411),
SeCyT-UNC, ANPCyT (PICT-2012-2324) and DFG (research network FOR1376). We
thank P. Serra and W. Schmickler for discussions and references.
\newpage
\bibliographystyle{jpc}
|
\section{Introduction}
Topological data analysis (TDA) has emerged as a useful tool to analyze underlying properties of data before any contextual modeling assumptions kick in. Generally speaking, TDA seeks to characterize the \emph{shape} of high dimensional data (viewed as a point-cloud in some metric space), by quantifying various topological constructs such as connected components, high-dimensional holes, level-sets and monotonic regions of functions defined on the data \cite{edelsbrunner2002topological}. In particular, the number of $d$-dimensional holes in a data, also known as the \textit{Betti}-$d$ number, corresponds to the rank of the $d-$dimensional homology group. A commonly used and powerful topological feature, the \textit{persistence diagram} (PD) summarizes the \textit{persistence} of the Betti numbers with respect to the \textit{scale} parameter used to analyze the data. A typical machine learning pipeline using TDA features would first estimate the PDs from the given point cloud, and define a metric on them to compare different point clouds. The Wasserstein distance measure has become ubiquitous as a metric between such PDs, as it is stable and has a well-defined metric space associated with it \cite{turner2014frechet, cohen2010lipschitz}. However, computation of the Wasserstein distance involves finding a one-to-one map between the points in one persistence diagram to those in the other, which is a computationally expensive operation.\let\thefootnote\relax\footnotetext{This work was supported by NSF CAREER grant 1452163.}
In this paper, we propose a novel representation for persistence diagrams as points on a hypersphere, by approximating them as 2D probability density functions (pdf). We perform a square-root transform of the constructed pdf, wherein the Hilbert sphere becomes the appropriate space for defining metrics \cite{Srivastava2007}. This insight is used to construct closed form metrics -- geodesics on the Hilbert sphere -- which can be computed very efficiently, bypassing the correspondence problem entirely. The overall pipeline of operations for computing the proposed representation is given in Figure \ref{fig:pipeline}.
\begin{figure*}[!tb]
\centering
\includegraphics[clip = true,trim=0mm 88mm 0mm 50mm,width = 0.99\linewidth]{pipeline.pdf}
\caption{\small{The overall sequence of operations leading to the proposed representation, illustrated for the application of activity analysis. Joint position data from motion capture systems are collected as 1D time series. The phase space is reconstructed from these time series via Takens' embedding theorem. We represent the topological properties of this phase space using the persistence diagram (PD). Next, we use kernel density estimation to represent the PD itself as a 2D probability density function (pdf). Finally, we use the square-root framework to interpret these pdfs as points on a Hilbert sphere.}}
\label{fig:pipeline}
\end{figure*}
The biggest advantage of the proposed representation is that it completely works around the computationally expensive step of obtaining one-to-one correspondences between points in persistence diagrams, thereby making the distance computation between PDs extremely efficient. We show that approximating PDs as pdfs results in comparable performances to the popular and best performing $L_1$-Wasserstein metric, while at the same time being orders of magnitude faster. We also provide a theoretically well-grounded understanding of the geometry of the pdf representation. Additionally, the availability of closed form expressions for many important tools such as the geodesics, exponential maps etc. enables us to adapt powerful statistical tools -- such as clustering, PCA, sample mean etc. -- opening new possibilities for applications involving large datasets. To the best of our knowledge we are the first to propose this representation for persistence diagrams.
\paragraph{Contributions}
\begin{enumerate}
\item We present the first representation of persistence diagrams that are approximated as points on a Hilbert sphere resulting in closed-form expressions to compare two diagrams. This also completely avoids the correspondence problem, which is typically a computational bottleneck.
\item We demonstrate the ability of the proposed representation for statistical analyses, such as computing the mean persistence diagram, principal geodesic analysis (PGA), and classification using SVMs.
\item We show promising results for supervised tasks such as action recognition, and assessment of movement quality in stroke rehabilitation.
\item The space of the representation -- the Hilbert sphere -- is a geometrically well-understood and intuitive space, which may further promote its use in topological machine learning applications.
\end{enumerate}
The rest of the paper is organized as follows. Section \ref{sec:rel_work} discusses related work in more detail. Section \ref{sec:math_prelim} provides the necessary background on persistent homology, the space of persistence diagrams, and the square-root framework on the Hilbert space. Section \ref{sec:prop_framework} provides details about the proposed framework for using the new representation of persistence diagrams in a functional Hilbert space for statistical learning tasks. Section \ref{sec:expts} describes the experiments and results. Section \ref{sec:disc} concludes the paper.
\section{Related Work}
\label{sec:rel_work}
Persistence diagrams elegantly summarize the topological properties of point clouds, and the first algorithm for computing topological persistence was proposed in \cite{edelsbrunner2002topological}. Ever since, there has been an explosion in understanding the properties of PDs, comparing PDs, and computing statistics on them. Since a PD is a multiset of off-diagonal points along with infinite number of points on the diagonal, the cost of optimal bijective assignment, also known as the Wasserstein distance, between the individual points in the PDs is a valid distance measure. The time complexity of computing the distance between two PDs with $n$ points each is $\mathcal{O}(n^3)$ \cite{bertsekas1981new}. It has been shown in \cite{cohen2010lipschitz} that the persistence diagrams of Lipschitz functions are stable with respect to $p$-Wasserstein distance. However, the bottleneck and $p$-Wasserstein distance do not allow for easy computation of statistics. Hence, $L_p$-Wasserstein \textit{metrics} have been used to develop approaches for computing statistical summaries such as the Fr\'{e}chet mean \cite{turner2014frechet}. Computing Fr\'{e}chet mean of PDs involves iteratively assigning points from the individual diagrams to the candidate mean diagram, recomputing the mean, and repeating till convergence. Since the mean PD obtained is not guaranteed to be unique, the authors in \cite{munch2013probabilistic} proposed the probabilistic Fr\'{e}chet mean which itself is a probability measure on the space of PDs. An investigation of the properties of the space of PDs that allow for definition of various probability measures is reported in \cite{mileyko2011probability}, and a derivation of confidence sets that allow the separation of topological signal from noise is presented in \cite{fasy2014confidence}.
Since operations with PDs can be computationally expensive, Bubenik proposed a new topological summary - known as the persistence landscape - derived from the PD \cite{bubenik2015statistical}. It can be thought of as a collection of envelope functions on the points in PD based on the order of their importance. Persistence landscapes allow for fast computation of statistics since they lie in a vector space. However their practical utility has been limited since they provide decreasing importance to secondary and tertiary features in PDs that are usually useful in terms of discriminating between data from different classes. Recently, an approach that defines a stable multi-scale kernel between persistence diagrams has been proposed \cite{reininghaus2015stable}. The kernel is obtained by creating a surface with a Gaussian centered at each point in the PD along with a negative Gaussian centered at its reflection across the diagonal. In addition, the authors in \cite{adams2015persistent} propose to compute a vector representation - the persistence image - by computing a kernel density estimate of the PD and integrating it locally. The kernel density estimate is weighted such that the points will have increasing weights as they move away from the diagonal. A similar weighting approach is used in \cite{kusano2016persistence} to create a persistence-weighted Gaussian kernel.
The square-root representation for 1D probability density functions (pdfs) was proposed in \cite{Srivastava2007} and was used for shape classification. It has since been used in several applications, including activity recognition \cite{VeeraraghavanPAMI2009} - where the square-root velocity representation is used to model the space of time warping functions. This representation extends quite easily to arbitrary high-dimensional pdfs as well.
\section{Mathematical Preliminaries}
\label{sec:math_prelim}
In this section we will introduce the concepts of a) persistence homology, b) persistent diagrams, c) our proposed representation of PDs as a 2D pdf, and the square-root framework resulting in the Hilbert sphere geometry.
\subsection{Persistent Homology}
\label{sec:persistent homology}
The homology of point cloud $\mathbf{X}$ can be computed by first constructing a \textit{shape} out of the point cloud and estimating its homology. A simplex is imposed on every set of points in the point cloud that satisfy a neighborhood condition based on a scale parameter $\epsilon$. The collection of such simplices is known as the \textit{simplicial complex} $\mathcal{S}$. $\mathcal{S}$ can be thought of as a representation of the underlying shape of the data, and its homology can be inferred. However, homology groups obtained from $\mathcal{S}$ depend on the scale or time parameter $\epsilon$ based on which the complexes are constructed \cite{edelsbrunner2002topological}. The homological features of the simplicial complex constructed from the point cloud that are stable across scales, i.e., that are \textit{persistent}, are the ones that provide information about the underlying shape. Topological features that do not persist are considered to be noise. This information is represented using \textit{persistence diagrams} as a 2D plot of birth versus death times of each homology cycle corresponding to the homology group $\mathbb{H}_k, k = \{0, 1, \ldots\}$. The birth time is the scale at which the homological feature is born and the death time is the scale at which it ceases to exist. The homology cycle of dimension $d$ is also referred to as a $d-$dimensional hole. Therefore, the PD can be considered as an object that represents the number of holes in terms of the range of scales at which they appear. Typically, PDs of the point cloud data are obtained using \textit{filtration} of simplicial complexes. A well-known filtration is the Vietoris-Rips (VR) filtration, where a simplex is induced between a group of points whenever the distance between each pair of points is less than the given scale $\epsilon$ \cite{zomorodian2010fast}. An example of point clouds and their corresponding persistence diagrams for homology groups 0 and 1 are provided in Figure \ref{fig:PhaseSpace}.
\begin{figure}[!htb]
\captionsetup[subfigure]{labelformat=empty}
\centering
\subfloat[]{
\includegraphics[width=0.45\linewidth]{PH1.pdf}
}
\subfloat[]{
\includegraphics[width=0.45\linewidth]{Lorenz_PD.pdf}
}
\subfloat[Point cloud data]{
\includegraphics[width=0.45\linewidth]{PH2.pdf}
}
\subfloat[Persistence diagrams]{
\includegraphics[width=0.45\linewidth]{Rossler_PD.pdf}}
\caption{\small{The above example illustrates the topological features extracted from the point cloud data with two and one one-dimensional holes. These properties are reflected well in their corresponding persistence diagrams on the right.}}
\label{fig:PhaseSpace}
\end{figure}
\subsection{The Space of Persistence Diagrams}
\label{sec:pers_diag_space}
Every PD is a multiset of 2D points, where each point denotes the birth and death time of a homology group. Furthermore, the diagonal is assumed to contain an infinite number of points with the same birth and death times. For any two PDs $X$ and $Y$, the distance between the diagrams is usually quantified using the bottleneck distance or the $L_q$-Wasserstein metric \cite{kerbergeometry}. In this paper, we consider only the $L_2$- and $L_1$-Wasserstein ($d_{L_2}$ and $d_{L_1}$) metrics given respectively as,
\begin{equation}
d_{L_2}(X,Y) = \left( \inf_{\eta:X \to Y} \sum_{x \in X} ||x-\eta(x)||_2^2 \right)^{\frac{1}{2}},
\end{equation} and
\begin{equation}
d_{L_1}(X,Y) = \inf_{\eta:X \to Y} \sum_{x \in X} ||x-\eta(x)||_1.
\end{equation} Since each diagram contains an infinite number of points in the diagonal, this distance is computed by pairing each point in one diagram uniquely to another non-diagonal or diagonal point in the other diagram, and then computing the distance. This correspondence can be obtained via the Hungarian algorithm or its more efficient variants \cite{kerbergeometry}.
The space of PDs with respect to the $L_2$-Wasserstein metric is given by
\begin{equation}
\mathcal{D}_{L_2} = \{X | d_{L_2}(X,\varnothing) < \infty\}.
\end{equation} Turner {\em et. al.} \cite{turner2014frechet} show this is a non-negatively curved Alexandrov space. Furthermore, the diagram on the geodesic between the PDs $X$ and $Y$ in this space is given as
\begin{equation}
\gamma(s) = (1-s) x + s \phi (x),
\end{equation} where $x$ is a point in the diagram $X$, $\phi(x)$ is a corresponding point in the diagram $Y$ and $s \in [0,1]$ parametrizes the geodesic. Clearly, the points in the diagram on the geodesic can be simply obtained by linearly interpolating between the corresponding points on the candidate points $X$ and $Y$. Furthermore, the Riemannian mean between two PDs is easily computed as $\gamma(0.5)$.
\subsection{Square-root Framework and the Hilbert Sphere}
\label{sec:sqrt_fmwk}
We treat the points in a persistence diagram as samples from an underlying probability distribution function. This representation is inspired from recent work dealing with defining Mercer kernels on PDs \cite{reininghaus2015stable}, where the feature embedding obtained from the Mercer kernel closely resembles a kernel-density estimator from the given points, with an additional reflection term about the diagonal to account for boundary effects while solving a heat-equation. The fact that the feature embedding resembles a kernel density estimate is not further exploited in \cite{reininghaus2015stable}.
In our work, we more directly exploit this pdf interpretation of PDs, further endowing it with a square-root form \cite{Srivastava2007} -- and then utilizing the Hilbert spherical geometry that results.
Without loss of generality, we will assume that the supports for each 2D pdf is $[0,1]^2$. The space of pdfs that we will consider is
\begin{multline}
\mathcal{P} = \{p:[0,1]\times [0,1]\rightarrow \mathbb{R} ~~ \forall x,y| p(x,y)\geq 0, \\
\mbox{ and } \int_0^1\int_0^1p(x,y)dxdy = 1\}
\end{multline} It is well-known that $\mathcal{P}$ is not a vector space \cite{Srivastava2007}. Instead, it is a Riemannian manifold with the Fisher-Rao metric as the natural Riemannian metric. Geodesics under the Fisher-Rao metric are quite difficult to compute. Instead, we adopt a square-root form proposed by Srivastava {\em et. al.} \cite{Srivastava2007} which simplifies further analysis enormously. In other words we consider the space,
\begin{multline}
\Psi = \{\psi:[0,1]\times[0,1] \rightarrow \mathbb{R}| \psi\geq 0,\\
\int_0^1\int_0^1 \psi^2(x,y)dxdy = 1\}.
\end{multline} For any two tangent vectors $v_1,v_2 \in T_\psi(\Psi)$, the Fisher-Rao metric is given by,
\begin{equation}
\langle v_1,v_2\rangle = \int_0^1\int_0^1v_1(x,y)v_2(x,y)~dx dy.
\end{equation}
The above two pieces of information imply that the square-root form $\psi = \sqrt{p}$ results in the space becoming a unit Hilbert-sphere, endowed with the usual inner-product metric. Geodesics on the unit-Hilbert sphere under the above Riemannian metric are known in closed form. In fact, the differential geometry of the Hilbert sphere results in closed form expressions for computing geodesics, exponential and inverse exponential maps \cite{Srivastava2007}. Further, the square-root form with the above metric has additional favorable properties such as invariance to domain re-parameterization.
Given two points $\psi_1, \psi_2$ the geodesic distance between them is given by
\begin{equation}
\label{eqn:hilb_dist}
d_H(\psi_1,\psi_2) = \cos^{-1}(\langle \psi_1,\psi_2\rangle),
\end{equation}
where $\langle \psi_1,\psi_2\rangle_\psi = \int_0^1 \psi_1(t)\psi_2(t) dt$. The exponential map is defined as,
\begin{equation}
\mathrm{ exp}_{\psi_i}(\upsilon) = \mathrm{cos}(||{\upsilon}||_{\psi_i})\psi_i + \mathrm{sin}(||\upsilon||_{\psi_i})\frac{\upsilon}{||\upsilon||_{\psi_i}},
\end{equation} where $\upsilon \in T_{\psi_i}( \Psi)$ is a tangent vector at $\psi_i\mbox{ and }||\upsilon||_{\psi_i}$ = $\sqrt{\left<\upsilon,\upsilon\right>_{\psi_i}}$.
The logarithmic map from $\psi_i$ to $\psi_j$ is given by:
\begin{equation}
\exp^{-1}_{\psi_i}(\psi_j) = \frac{u}{\sqrt{\langle u,u\rangle}} \mathrm{cos}^{-1}\left< \psi_i, \psi_j\right>,
\end{equation}
with $u = \psi_i - \left< \psi_i, \psi_j\right>\psi_j.$ The geodesic on the sphere is given in closed form as
\begin{equation}
\label{eqn:geod1}
\pi(s) = \frac{(1-s)\psi_1+s \psi_2}{s^2+(1-s)^2+2s(1-s)(\langle \psi_1, \psi_2\rangle)},
\end{equation} or equivalently as
\begin{equation}
\label{eqn:geod2}
\pi(s) = \cos(s||v||) \psi + \sin(s||v||)\frac{v}{||v||}.
\end{equation}
A comparison of sampling of the geodesics in the Alexandrov space induced by the $L_2$-Wasserstein metric and the proposed Hilbert sphere are provided in Figure \ref{fig:sampling}.
\begin{figure*}[!htb]
\centering
\includegraphics[clip = true,trim=0mm 5mm 0mm 0mm,width = 0.95\linewidth]{geodesic_sampling_fig}
\caption{\small{Comparing geodesics in the Alexandrov space (top) and the proposed Hilbert sphere (bottom), for a simple persistence diagram containing just $3$ points. The Alexandrov space requires computing correspondence between points, and the geodesic involves linearly interpolating between corresponding points. The persistence pdf avoids correspondence estimation, hence the geodesics correspond to new local modes appearing and gradually increasing in intensity as the original modes decrease in intensity.}}
\label{fig:sampling}
\end{figure*}
\section{Algorithmic Details}
\label{sec:prop_framework}
Our proposed framework consists of reconstructing the phase space of the time series data, computing the PDs of the reconstructed phase space, transforming each PD to a 2D pdf by kernel density estimation, using the 2D pdfs in the square-root framework to estimate distances or obtain statistical summaries. The distances computed and the features obtained will be used in inference tasks. Additional details for some of the above steps are given below.
\paragraph{Phase-space Reconstruction from Activity Data}
In dynamical system understanding, we aim to estimate the underlying states, but we measure functions -- usually projections -- of the original state-space. e.g. while human movement is influenced by many joints, ligaments, muscles of the body, we might measure the location of only a few joints. One way to deal with this issue is to reconstructing the `phase space' by the method of delays used commonly in non-linear dynamical analysis \cite{Takens}. Reconstructing the phase-space in this manner only preserves important topological properties of the original dynamical system, and does not recover the true state-space. For a discrete dynamical system with a multidimensional phase-space, time-delay vectors (or embedding vectors) are obtained by the concatenation of delayed versions of the data points,
\begin{equation}
\mathbf{x}_t = [x(t),x(t+\tau),\cdots,x(t+(m-1)\tau)]^T.
\label{EmVec}
\end{equation} where $m$ is the embedding dimension, $\tau$ is the embedding delay. For a sufficiently large $m$, the important topological properties of the unknown multidimensional system are reproduced in the reconstructed phase-space \cite{abarbanel1996analysis}. The collection of $n$ time-delay vectors is the point cloud data under further consideration, and this is represented as $\mathbf{X} = [\mathbf{x}_t]_{t=0}^n$.
\paragraph{Estimating the PDs}
After estimating the point cloud in the reconstructed phase space, we will construct a simplicial complex $\mathcal{S}$ using the point cloud data $\mathbf{X}$ to compute the persistence diagrams for the Betti numbers using the VR filtration. However, this approach considers only spatial adjacency between the points and ignores the temporal adjacency. We improve upon the existing VR filtration approach by explicitly creating temporal links between $\mathbf{x}_{t-1}$, $\mathbf{x}_{t}$, and $\mathbf{x}_{t+1}$ in the one-skeleton of $\mathcal{S}$, thereby creating a metric space which encodes adjacency in both space and time \cite{venkataraman2016persistent}. The persistence diagrams for homology groups of dimensions $0$ and $1$ are then estimated.
\paragraph{Square-root Framework for Distance Estimation between PDs}
Since each PD is a multiset of points in the 2D space, we start by constructing a 2D pdf from each of them using kernel density estimation using a Gaussian kernel of zero mean and variance $\sigma^2$. For each PD, we compute the square-root representation $\psi$ using the approach provided in Section \ref{sec:sqrt_fmwk}, and the distance between two PDs can be computed using (\ref{eqn:hilb_dist}).
\paragraph{Dimensionality Reduction with Principal Geodesic Analysis (PGA)}
We are able to extract principal components using PGA \cite{Fletcher2004}, adapted to our Hilbert sphere -- which essentially performs classical PCA on the tangent space of the mean-point on the sphere. Given a dataset of persistence diagrams in their square-root form $\{\psi_1,\psi_2,\ldots, \psi_N\}$, we first compute the Riemannian center of mass \cite{GroveK72}. We use a simple approximation using an extrinsic algorithm that computes first the Euclidean mean and maps it on to the closest point on the sphere. Next, we represent each $\psi_i$ by its tangent vector from the mean. We then compute the principal components on the tangent space and represent a persistence diagram as a low-dimensional vector.
\section{Experiments}
We perform experiments on two datasets for human action analysis. First we perform action recognition on the MoCap dataset \cite{ali2007chaotic}, next we show the use of the framework in quality assessment of movements in stroke rehabilitation \cite{Chen2011Stroke}. We will describe the datasets next, followed by the evaluation settings and parameters used. In all our experiments, we performed a discretization of the 2D pdf into a $K\times K$ grid. The choice of $K$ indirectly affects classification performance as expected, i.e. -- a smaller value of $K$ results in a reduced ability to identify between nearby points in the PD due to lower resolution. On the other hand, a larger value of $K$ improves resolution, but also increases computational requirements. Typical values of $K$ used in experiments range from $50$ to $100$ at most, whereas typical values of $n$ -- the number of points in a PD -- ranges from $20$ to $50$.
\label{sec:expts}
\subsection{Motion Capture Data}
We evaluate the performance of the proposed framework using $3$-dimensional motion capture sequences of body joints \cite{ali2007chaotic}. The dataset is a collection of five actions: \textit{dance, jump, run, sit} and \textit{walk} with $31, 14, 30, 35$ and $48$ instances respectively.
\begin{table}[!htb]
\centering
\begin{tabular}{|p{1in}|c|p{1.0in}|}
\hline
\textbf{Method} &\small{\textbf{Accuracy (\%)}} & \textbf{Time (sec)} \\ \hline \hline
Chaos \cite{ali2007chaotic} & 52.44 & - \\
VR-Complex \cite{zomorodian2010fast} & 93.68 & - \\
DT2 \cite{vinay_PAMI} & 93.92 & - \\
T-VR Complex (L1) \cite{venkataraman2016persistent} & 96.48 & $(1.2 \pm 1.23) \times 10^3 $ \\ \hline
\textbf{Proposed} - 1NN & {89.87} & $\mathbf{(0.059 \pm 0.044)}$ \\
\textbf{Proposed} - PGA +SVM & {91.68}& - \\ \hline
\end{tabular}
\vspace{10pt}
\caption{\small{Comparison of classification rates for different methods using nearest neighbor classifier on the motion capture dataset. It is observed that on an average, the proposed metric is $10^5$ times faster than the traditional Wasserstein metric, while achieving a comparable recognition accuracy.}}
\label{tab:results1}
\end{table}
We generate $100$ random splits having $5$ testing examples from each action class and use an SVM classifier on the vector features computed with PGA, we get a performance of $91.68\%$. The mean recognition rates for the different methods are given in Table \ref{tab:results1}. We also compare with a 1-nearest neighbor classifier computed on the Hilbert sphere, which gives a performance of $89.87\%$. This is clearly competitive, when compared to approaches that use the $L_1$-Wasserstein metric as shown in table \ref{tab:results1}. Clearly, the proposed metric for persistence diagrams is able to capture most of the important information, with the added benefit being free from expensive correspondence estimation. To compute the times taken using the Wasserstein metric and the proposed metric, we average across 3 samples from 5 action classes, across $57$ (19 joint trajectories along 3 axes) trajectories -- a total of $855$ persistence diagrams. We computed a pairwise distance matrix for all these PDs, and computed the average time taken in Matlab 2014 on a standard Intel i7 PC.
\begin{figure}[!htb]
\centering
\includegraphics[clip = true,trim=150mm 0mm 140mm 0mm,width = 0.95\linewidth]{SigvRecog2.eps}
\caption{\small{Recognition accuracy vs $\sigma$, the standard deviation of the 2D Gaussians used in the kernel density estimates.}}
\label{fig:sigma}
\vspace{-0.1in}
\end{figure}
In Figure \ref{fig:sigma}, we compare the recognition accuracy with the choice of $\sigma$, the standard deviation used during kernel density estimation. The accuracy is generally higher for a small $\sigma$ and drops as the $\sigma$ increases. We note that a similar trend is also reported in \cite{reininghaus2015stable}.
\begin{figure*}[!htb]
\centering
\subfloat[\footnotesize{Unimpaired Subjects}]{
\includegraphics[clip = true,trim=5mm 5mm 10mm 10mm,width = 0.45\linewidth]{unimp_HT.eps}
\label{fig:unimp_HT}}
\subfloat[\footnotesize{Impaired Subjects}]{
\includegraphics[clip = true,trim=5mm 5mm 10mm 10mm,width = 0.45\linewidth]{imp_HT.eps}
\label{fig:imp_HT}}
\caption{\small{The mean persistence pdfs visualized as heat maps for unimpaired subjects (left) and stroke survivors (right), for the reach-and-grasp action, are shown. These means are computed as the extrinsic means (we choose extrinsic mean here for simplicity) on the Hilbert sphere using the proposed representation. The locations of the peaks are exactly the same, since the subjects perform the same general movement, however the intensity differs significantly indicating that the \emph{quality} of movement is captured using topological tools well}.}
\label{fig:stroke_mean}
\end{figure*}
\subsection{Stroke Rehabilitation Dataset}
Our aim in this experiment is to quantitatively assess the quality of movement performed by the impaired subjects during repetitive task therapy. The experimental data was collected using a heavy marker-based system ($14$ markers on the right hand, arm and torso) in a hospital setting from $15$ impaired subjects performing reach and grasp movements to a defined target. The stroke survivors were also evaluated by the Wolf Motor Function Test (WMFT) \cite{wolf2001assessing} on the day of recording, which evaluates a subject's functional ability on a scale of $1-5$ (with $5$ being least impaired and $1$ being most impaired) based on predefined functional tasks. In our experiments, we only use the data corresponding to the single marker on the wrist from the heavy marker-based hospital system, which allows us to evaluate the framework in a home-based rehabilitation system setting. We utilize the WMFT scores as an approximate high-level quantitative measure for movement quality (ground-truth labels) of impaired subjects performing the reach and grasp movements.
We compute explicit features as described earlier using PGA. This allows us to represent each movement by a low-dimensional feature vector. We use the WMFT scores and split the dataset into train and test to perform regression. Since the dataset is small, we use a leave-one-out approach, where we train on all but one sample, which is used for testing, and repeat this over all the samples in the dataset. The correlation performance with the WMFT score is shown in Table \ref{tab:strkcomptab}, and it can be seen that we are comparable to the state of the art, and much better than traditional features. The predicted scores are shown in Figure \ref{fig:stroke} compared to the original scores.
The dynamical features proposed in \cite{vinay_PAMI} and \cite{ali2007chaotic}, depend on describing the phase space for each movement. We compute the topological features, which are expected to be much weaker than other characteristics such as shape. However, we are still able to capture subtle information regarding movement quality. This is illustrated in Figure \ref{fig:stroke_mean}, where we see the 3D peaks associated with the average of persistence diagrams across subjects impaired with stroke and those who are unimpaired. It is clearly seen that since they are performing the same kind of movements, the peaks occur at the same location. However, the intensity is significantly different across these diagrams, perhaps indicating information regarding movement quality.
\begin{table}[!htb]
\centering
\begin{tabular}{|c|P{0.8in}|}
\hline
\textbf{Method} & \textbf{Correlation with WMFT} \\ \hline
KIM (14 markers) \cite{chen2011computational} & 0.85 \\
Lyapunov exponent (1 marker) \cite{lyaprosen} & 0.60 \\ \hline
\textbf{Proposed method (1 marker)} & \textbf{0.80} \\ \hline
\end{tabular}
\vspace{10pt}
\caption{\small{Comparison of correlation coefficient for different methods using leave-one-subject-out cross-validation scheme and SVM regressor on the stroke rehabilitation dataset. Even with just a single marker, We obtain comparable results to a clinical 14-marker system. }}
\label{tab:strkcomptab}
\vspace{-15pt}
\end{table}
\begin{figure}[!htb]
\centering
\includegraphics[clip = true,trim=130mm 0mm 140mm 10mm,width = 0.85\linewidth]{stroke_predict.eps}
\caption{\small{Predicting movement quality scores for reach-and-grasp in stroke survivors using topological features. We obtain a \textbf{0.8} correlation score with a p-value of $5.46 \times 10^{-4}$}}
\label{fig:stroke}
\vspace{-0.2in}
\end{figure}
\section{Discussion and Conclusion}
\label{sec:disc}
Based on the theory and experiments presented so far, it is instructive to compare the space of PDs with respect to the different distance metrics. We will consider only the $L_2$-Wasserstein metric ($d_{L_2}$) and the proposed Hilbert sphere metric $d_{H}$. The interpretation of PDs with respect to these two metrics is very different. $d_{L_2}(X,Y)$ is the Earth Mover's Distance between the PDs $X$ and $Y$ considered as scaled discrete probability mass functions (pmfs). However, $d_{H}(\psi_X,\psi_Y)$ is the geodesic distance between the square root of kernel density estimates of $X$ and $Y$. Furthermore, the ``length'' of the persistence diagrams induced by these metrics are very different. For the Wasserstein metric, this is given by $d_{L_2}(X,\varnothing)$ which can be arbitrarily high, whereas with the proposed metric it is constrained as the persistence diagrams live on a unit sphere. The most important distinction arises when the pairwise distances between all the points in the two PDs are sufficiently high. When the 2-dimensional pdfs obtained from kernel density estimates of the two PDs do not overlap anywhere, $d_{H}(X,Y) \rightarrow 1$. This implies that if the variance of the kernel is sufficiently small, two PDs with non-overlapping points will always have $d_H$ close to $1$. This problem can be alleviated by using kernels with multiple scales for smoothing the PDs to obtain and combining the distances obtained at each scale.
The comparison between the geodesics in the persistence space is also illuminating (Figure \ref{fig:sampling}). Sampling of the geodesic in the Alexandrov space shows that the points in one diagram move towards the corresponding points in the other, as we move along the geodesic. Whereas, a similar sampling in the Hilbert sphere shows that the PDs in the middle of the geodesic contain the modes from pdfs corresponding to both the candidate PDs. This results in markedly different interpretations of the means computed from $d_{L_2}$ and $d_H$ even for the case of two persistence diagrams.
Several directions of future work emerge from this endeavor. On the theoretical side a more formal understanding of the variety of metrics and their relationship to the proposed one can be considered. On the applied side, the favorable computational load reduction should open new applications of TDA for massive datasets.
{\small
\bibliographystyle{ieee}
|
\section{Introduction}
The Landau-Lifshitz equation was developed to model the behaviour of domain walls in magnetic regions within ferromagnetic structures \cite{Landau1935}. For example, the one-dimensional Landau-Lifshitz equation can be used to describe ferromagnetic nanowires, which are often found in memory storage devices such as hard disks, credit cards or tape recordings. Each set of data stored in a memory device is uniquely assigned to a specific stable magnetic state of the ferromagnet. This can be difficult to achieve due to the presence of hysteresis. Hysteresis is characterized by the presence of multiple equilibria, and looping in the input-output map is typical \cite{Chow2013ACC,Morris2011}. Consequently, a particular input can lead to different magnetizations. Therefore, it is desirable to control magnetization between different stable equilibria.
The Landau-Lifshitz equation is known to exhibit hysteretic behaviour. For example, \cite{Cowburn1999,Suess2002} investigated via experiments the shape change of the hysteresis loop as the structure of the nanomagnet is varied. Experiments conducted on nanowires also demonstrate hysteresis loops \cite{Noh2012}. Numerical simulations illustrating hysteresis loops are found in \cite{Wiele2006,Yang2011}. The dynamics of hysteresis in the Landau-Lifshitz equation has also been represented by a hysteresis operator \cite{CarbouEfendiev2009,Visintin1997}. In much of the aforementioned literature, the presence of hysteresis in the Landau-Lifshitz equation is identified by the fact that input--output curves exhibit a looping behaviour. This alone is not enough to characterize hysteresis \cite{Chow2013ACC,Morris2011,Bernstein2005}. A looping behaviour must persist with low frequency periodic inputs.
\begin{defn}\label{defBernstein}{\em \cite{Bernstein2005}}
A system exhibits {\it hysteresis} if a nontrivial closed curve in the input--output map persists for a periodic input as the frequency component of the input signal approaches zero.
\end{defn}
Another approach to hysteresis is based on the existence of multiple stable equilibria, which are present in the (uncontrolled) Landau-Lifshitz equation \cite[Chapter~6]{Guo2008}.
\begin{defn}{\em \cite[Definition~3]{Morris2011}} \label{defmultiequilibrium}
A system exhibits {\it hysteresis} if it has \\
(a) multiple stable equilibrium points and\\
(b) dynamics that are considerably faster than the time scale at which inputs are varied.
\end{defn}
Note that condition (b) is relative to the speed at which a controlled input is changed. In many cases, hysteresis is present but is rate-dependent \cite{Morris2011}.
There is now an extensive body of results on control and stabilization of linear partial differential equations (PDEs); see for instance the books \cite{Bensoussan-book,Curtain1995,LT00_1,LT00_2} and the review paper \cite{Morris_control_handbook_rev2010}. There are fewer results on control and stabilization of nonlinear partial differential equations and the Landau-Lifshitz equation is particularly problematic. Stability of the Landau-Lifshitz equation is often based on linearization \cite{CarbouLabbe2006,Carbou2006,Carbou2011,Jizzini2011,Labbe2012}. Local asymptotic stability is shown in \cite{Chow2015} for the controlled linearized Landau-Lifshitz equation. However, because the Landau-Lifshitz equation is not quasi-linear, analysis based on a linearization may not predict stability of the original system; see \cite{CoronNguyen2015,alJamal2013,AM2014}. Also, when the goal is to steer between equilibria, global stability results are needed. Experiments and numerical simulations on the control of domain walls in a nanowire are presented in \cite{Noh2012,Wieser2011}. In \cite{Carbou2008,Carbou2009}, solutions to the Landau-Lifshitz equation are shown to be arbitrarily close to domain walls given a constant control.
In the next section, the uncontrolled Landau-Lifshitz equation and its equilibrium points are described. In \cite{Chow2013ACC}, simulations were used to show the Landau-Lifshitz and the linearized Landau-Lifshitz equation exhibit hysteresis. This suggests hysteresis is not due entirely to nonlinearity. These results are reviewed in Section~\ref{secLL}. Theorem~\ref{thmzeroeigenvalue} demonstrates the linearized uncontolled Landau-Lifshitz equation has a zero eigenvalue. This suggests use of a proportioonal controller to stabilize the equation about a given point. It is then proven in Section~\ref{seccontrol} that stabilization of the full Landau-Lifshitz equation is achieved with a proportional affine control. Proportional control can be used to steer the system to an arbitrary equilibrium point of the uncontrolled equation; in fact, the system can be steered between these points. Simulations illustrating these results are presented in Section~\ref{secExample}.
Furthermore, simulations indicate that hysteresis is absent in the controlled system.
\section{Landau-Lifshitz Equation and Hysteresis}\label{secLL}
Consider the magnetization
$$
\mathbf m(x,t)=(m_1(x,t),m_2(x,t),m_3(x,t)),
$$
at position $x \in [0,L]$ and time $t \geq 0 $
in a long thin ferromagnetic material of length $L>0$. If only the exchange energy term is considered,
the magnetization
is modelled by the one--dimensional (uncontrolled) Landau-Lifshitz equation
\cite{Brown1963},\cite[Chapter~6]{Guo2008}
\begin{subequations} \label{eqLLcomplete}
\begin{align}\label{eqLLGuoDing}
\frac{\partial \mathbf m}{\partial t} &= \mathbf m \times \mathbf m_{xx}-\nu\mathbf m\times\left(\mathbf m\times\mathbf m_{xx}\right)\\
\mathbf m(x,0)&=\mathbf m_0(x)\\
\mathbf m_x(0,t)&= \mathbf m_x(L,t)=\mathbf 0.\label{eqboundarycondition}
\end{align}
\end{subequations}
where $\times$ denotes the cross product and $\nu\geq 0$ is the damping parameter, which depends on the type of ferromagnet. The notation $\mathbf m_{x}$ and $\mathbf m_{xx}$ means the magnetization is differentiated with respect to $x$ once and twice, respectively. The gyromagnetic ratio multiplying $ \mathbf m \times \mathbf m_{xx}$ has been normalized to $1$ for simplicity. For details on the damping parameter and gryomagnetic ratio, see \cite{Gilbert2004}. It is assumed there is no magnetic flux at the boundaries and so Neumann boundary conditions are appropriate.
Define $\mathcal L_2^3 = \mathcal L_2 ([0,L]; \mathbb R^3)$ with the usual inner product and norm, denoted $\|\cdot\|_{\mathcal L_2^3}$, and the operator
\begin{equation}
f(\mathbf m)=\mathbf m \times \mathbf m_{xx}-\nu\mathbf m\times\left(\mathbf m\times \mathbf m_{xx}\right),
\label{defn:f}
\end{equation}
and its domain
\begin{align}
\label{setDforfullLL}
D=\{ \mathbf m\in \mathcal L_2^3 : \mathbf m_x \in \mathcal L_2^3, \mathbf m_{xx} \in \mathcal L_2^3,
\mathbf m_x(0)=\mathbf m_x(L) = \mathbf 0 \}.
\end{align}
The following theorem is a consequence of the existence and uniqueness results in \cite{Carbou2001,Alouges1992}.
\begin{thm} \label{thmuncontrolledLLsemigroup} If $m(x,0) \in \mathcal L_2^3,$ then the operator $f(\mathbf m)$ with domain $ D$ generates a nonlinear contraction semigroup on $\mathcal L_2^3.$
\end{thm}
Ferromagnets are magnetized to saturation \cite[Section~4.1]{Cullity2009}; that is
\[
|| \mathbf m_0 (x) ||_{2} =M_s
\]
where $||\cdot||_{2}$ is the Euclidean norm and $M_s$ is the magnetization saturation. Physically, this means that at each point, $x$, the magnitude of $\mathbf m_0(x)$ equals the magnetization saturation. In much of the literature, $M_s$ is set to $1$; see for example, \cite[Section~6.3.1]{Guo2008}, \cite{Carbou2001,Alouges1992,Lakshmanan2011}. This convention is used here. The magnitude of the magnetization does not change with time.
\begin{thm}{\em \cite[Lemma~6.3.1]{Guo2008}}
If $||\mathbf m_0(x)||_{2}=1$, then for all $t\geq0$ the solution to (\ref{eqLLGuoDing}) satisfies
\begin{equation}\label{eqconstraint}
|| \mathbf m(x,t)||_{2} =1 .
\end{equation}
\end{thm}
\noindent
The initial condition $\mathbf m_0(x)$ is assumed to be real--valued, which implies $m(x,t)$ is real-valued for all time.
The set of equilibrium points of (\ref{eqLLcomplete}) is \cite[Theorem~6.1.1]{Guo2008}
\begin{align}\label{equilibriumset}
E=&\{\mathbf a=(a_1,a_2,a_3) : a_1,a_2,a_3 \mbox{ constants and }\mathbf a^\mathrm{T}\mathbf a=1 \}.
\end{align}
In \cite[Proposition~6.2.1]{Guo2008}, the stability of the equilibrium points is established using Lyapunov's Theorem and the Lyapunov function
\[
V(\mathbf m )=\frac{1}{2}\left| \left| \mathbf m_{x}\right|\right|_{\mathcal L_2^3}^2.
\] Furthermore, $E$ is an asymptotically stable equilibrium set, as stated in the following theorem.
The proof is the same as that in \cite[Proposition~6.2.1]{Guo2008} except it is for equilibrium sets, rather than equilibrium points. However, individual equilibrium points are only stable, not asymptotically stable. Control is needed to obtain asymptotic stability as illustrated in Section~\ref{seccontrol}.
\begin{thm}\label{thmlyapunov}
The equilibrium set in (\ref{equilibriumset}) is asymptotically stable in the $\mathcal L_2^3$--norm.
\end{thm}
The existence of multiple stable equilibria indicates the presence of hysteresis in the Landau-Lifshitz equation (care of Definition~\ref{defmultiequilibrium}).
Definition~\ref{defBernstein} is used to establish hysteresis in simulations of the Landau-Lifshitz equation. For the simulations, a Galerkin approximation for the Landau-Lifshitz equation using linear spline elements is used. The number of elements is 5 and a periodic input, $\hat{\mathbf u}=\left(0,0.001\cos(\omega t),0\right)$, is applied to the Landau-Lifshitz equation to construct the input-output map.
Plots of $\mathbf m(x,t)$ with $x$ fixed against the periodic input are illustrated in Figure~\ref{figLoopFrequencyVaried} for varying frequencies $\omega$. It is clear from Figure~\ref{figLoopFrequencyVaried} the input--output curves exhibit persistent looping behaviour as the frequency of the input approaches zero. The continuum of equilibrium points explains the absence of sharp jumps that often appear in hysteresis loops. The similar appearance of the loop shapes between $m_1(x,t), m_2(x,t), m_3(x,t)$ is due to the symmetric structure of the Landau-Lifshitz equation.
\begin{figure}\hspace*{-0.3cm}
\centering\scalebox{0.545}{
\subfloat[{\Large $\omega=1$}]{ \includegraphics{LoopFrequencyw1.pdf}}
\subfloat[{\Large$\omega=0.1$}]{\includegraphics{LoopFrequencyw2.pdf}}
\subfloat[{\Large$\omega=0.01$}]{ \includegraphics{LoopFrequencyw3.pdf}}
\subfloat[{\Large$\omega=0.001$}]{ \includegraphics{LoopFrequencyw4.pdf}}}
\hspace*{-0.2cm}\scalebox{0.545}{
\subfloat[{\Large$\omega=1$}]{ \includegraphics{LoopFrequencyM2w1.pdf}}
\subfloat[{\Large$\omega=0.1$}]{\includegraphics{LoopFrequencyM2w2.pdf}}
\subfloat[{\Large$\omega=0.01$}]{ \includegraphics{LoopFrequencyM2w3.pdf}}
\subfloat[{\Large$\omega=0.001$}]{ \includegraphics{LoopFrequencyM2w4.pdf}}}
\hspace*{-0.2cm}\scalebox{0.545}{
\subfloat[{\Large$\omega=1$}]{ \includegraphics{LoopFrequencyM3w1.pdf}}
\subfloat[{\Large$\omega=0.1$}]{\includegraphics{LoopFrequencyM3w2.pdf}}
\subfloat[{\Large$\omega=0.01$}]{ \includegraphics{LoopFrequencyM3w3.pdf}}
\subfloat[{\Large$\omega=0.001$}]{ \includegraphics{LoopFrequencyM3w4.pdf}} }
\caption{\label{figLoopFrequencyVaried} \small Input--output curves for the (nonlinear) Landau-Lifshitz equation demonstrate persistent looping behaviour as the frequency of the periodic input, $\hat{\mathbf u}$, approaches zero and hence indicates the presence of hysteresis. (a)--(d) Input--output curves for $m_1(x,t)$ with $\hat{\mathbf u}=\left(0.001\cos(\omega t),0,0\right)$ and $\mathbf m_0(x)=\left(1,0,0\right)$. (e)--(h) Input--output curves for $m_2(x,t)$ with $\hat{\mathbf u}=\left(0,0.001\cos(\omega t),0\right)$ and $\mathbf m_0(x)=\left(0,1,0\right)$. (i)--(l) Input--output curves for $m_3(x,t)$ with $\hat{\mathbf u}=\left(0,0,0.001\cos(\omega t)\right)$ and $\mathbf m_0(x)=\left(0,0,1\right)$. ($L=1$, $\nu=0.02$, $x=0.6$)}
\end{figure}
To obtain the linear uncontrolled Landau-Lifshitz equation, equation (\ref{eqLLGuoDing}) is first rewritten in semilinear form,
\begin{equation}\label{eqLLGuoDingTwo}
\frac{\partial \mathbf m}{\partial t} =\nu \mathbf m_{xx}
+\mathbf m \times \mathbf m_{xx}
+\nu\left| \left| \mathbf m_{x}\right| \right|_{2}^2 \mathbf m,
\end{equation}
using equation~(\ref{eqconstraint}) and properties of cross products, and then $\mathbf m (x,t) = {\mathbf a}+ \mathbf z (x,t)$ is substituted into (\ref{eqLLGuoDingTwo})
where $\mathbf a \in E$ is an equilibrium of (\ref{eqLLcomplete}) and $\mathbf z \in \mathcal L_2^3$ is a small perturbation. The Landau-Lifshitz equation linearized about an equilibrium $\mathbf a$ is
\begin{subequations}\label{eqstatespaceform}
\begin{align}\dot{ \mathbf z}&=A\mathbf z,\quad \mathbf z(0)=\mathbf z_0 \\
\mathbf z_{x}(0,t)&=\mathbf 0=\mathbf z_{x}(L,t)
\end{align}
\end{subequations}
where $A$ is the linear operator, $$A\mathbf z= \nu\mathbf z_{xx}+\mathbf a \times \mathbf {z}_{xx},$$
and the domain is
\begin{align*
D(A)= \{ \mathbf z: \mathbf z\in \mathcal L_2^3, \,\, \mathbf z_{x} \in \mathcal L_2^3, \, \, \mathbf z_{xx}^{} \in \mathcal L_2^3, \mathbf z_{x}(0,t)=\mathbf 0=\mathbf z_{x}(L,t)\},
\end{align*}
Using \cite[Theorem~6.2]{Banks2012}, the linear operator $A$ can be shown to generate an analytic semigroup; for details, see \cite[Theorem~4.16]{Amenda_thesis}.
\begin{thm}\label{thmzeroeigenvalue}\cite[Theorem~5]{Chow2013ACC}
Any constant $\mathbf c \in \mathbb R^3$ is a stable equilibrium of (\ref{eqstatespaceform}).
\end{thm}
\begin{pf}
For completeness, the proof is included here.
Since $A$ generates an analytic semigroup, the spectrum determined growth assumption is satisfied and so the eigenvalues of $A$ determine the stability of the linear system (\ref{eqstatespaceform}) \cite[Section~5.1]{Curtain1995}, \cite[Section~3.2]{Luo1999}.
It is clear that any constant function $\mathbf c$ is an equilibrium of (\ref{eqstatespaceform}). Let $\lambda \in \mathbb C$. The eigenvalue problem of (\ref{eqstatespaceform}) is $\lambda\mathbf v=A\mathbf v$ and
boundary conditions $\mathbf v_x(0)=\mathbf v_x(L)=\mathbf 0$ where $\mathbf v \in \mathcal L_2^3$.
Solving, the eigenvalues of (\ref{eqstatespaceform}) are the zero eigenvalue, $\lambda_1=0,$ which is associated to a nonzero constant eigenvector, and the remaining eigenvalues are of the form
\begin{align*}
\lambda_2^{+,-} &= \frac{-(1+2n)^2\pi^2\nu}{L^2}\pm i\frac{(1+2n)^2\pi^2}{L^2}, \qquad
& \lambda_3 = \frac{-(1+2n)^2\pi^2\nu}{L^2}, \\
\lambda_4^{+,-} &= \frac{-(2n)^2\pi^2\nu}{L^2}\pm i\frac{(2n)^2\pi^2}{L^2}, \qquad
& \lambda_5 = \frac{-(2n)^2\pi^2\nu}{L^2}
\end{align*}
where $n\in \mathbb Z$. Since all the eigenvalues have nonpositive real part, the equilibria of (\ref{eqstatespaceform}) are stable. \qed
\end{pf}
Using Theorem~\ref{thmzeroeigenvalue} and Definition~\ref{defmultiequilibrium} indicates the linear Landau-Lifshitz equation exhibits hysteresis. Furthermore, simulations with periodic inputs were performed to determine whether persistent loops exist and hence Definition~\ref{defBernstein} is satisfied. Again, $\nu=0.02$, $L=1$, and the same periodic input is applied as for the nonlinear Landau-Lifshitz equation. Figure~\ref{figlinearhysteresisloopM1} shows the input-output curves for the first component of the solution to the linear Landau-Lifshitz equation with $\mathbf a=(1,0,0)$ and initial condition $\mathbf z_0(x)=\left( 1, 0,0 \right)$. From the figure, it is clear a loop persists as the frequency of the input approaches zero. Similar plots are obtained when the control is on the second and third components. The hysteresis loops in Figure~\ref{figlinearhysteresisloopM1} are similar in shape to the nonlinear Landau-Lifshitz equation depicted in Figure~\ref{figLoopFrequencyVaried}, both of which have a continuum of equilibria.
\begin{figure}[h]\hspace*{-0.7cm}
\centering\scalebox{1}{
\subfloat[$\omega=1$]{ \includegraphics[trim = 0mm 60mm 0mm 60mm, clip,width=0.26\textwidth]{LinearLoopFrequencyw1.pdf}}
\subfloat[$\omega=0.1$]{\includegraphics[trim = 0mm 60mm 0mm 60mm, clip,width=0.26\textwidth]{LinearLoopFrequencyw2.pdf}}
\subfloat[$\omega=0.01$]{ \includegraphics[trim = 0mm 60mm 0mm 60mm, clip,width=0.26\textwidth]{LinearLoopFrequencyw3.pdf}}
\subfloat[$\omega=0.001$]{ \includegraphics[trim = 0mm 60mm 0mm 60mm, clip,width=0.26\textwidth]{LinearLoopFrequencyw4.pdf}}}
\caption{\label{figlinearhysteresisloopM1} \small Hysteresis loops for $z_1(x,t)$ of the linear Landau-Lifshitz equation with $x=0.6$ and $\nu=0.02$. The linearization is at $\mathbf a=(1,0,0)$. The input is $\mathbf u(t)=\left(0.001\cos(\omega t),0,0\right)$ and the initial condition is $\mathbf z_0(x)=\left( 1, 0,0 \right)$.}
\end{figure}
\section{Controller Design}\label{seccontrol}
A control, $ \mathbf u(t)$, is introduced into the Landau-Lifshitz equation~(\ref{eqLLGuoDing}) as follows
\begin{align}
\label{eqcontrolledLL}
\frac{\partial \mathbf m}{\partial t} & = \mathbf m \times \mathbf m_{xx}-\nu\mathbf m\times\left(\mathbf m\times \mathbf m_{xx}\right)+ \mathbf u(t)\\
\mathbf m(x,0)&=\mathbf m_0(x)\nonumber\\
\mathbf m_x(0,t)&=\mathbf m_x(L,t)=\mathbf 0. \nonumber
\end{align}
The goal is to choose a control $\mathbf u(t)$ so the system governed by the Landau-Lifshitz equation moves from an arbitrary initial condition, possibly an equilibrium point, to a specified equilibrium point $\mathbf r$, where $\mathbf r \in E$ and $E$ is defined in (\ref{equilibriumset}). The control function needs to be chosen so that $\mathbf r $ becomes an asymptotically stable equilibrium point of the controlled system.
Theorem~\ref{thmzeroeigenvalue} implies zero is an eigenvalue of the uncontrolled linearized Landau-Lifshitz equation. For finite-dimensional linear systems, simple proportional control of a system with a zero eigenvalue yields asymptotic tracking of a specified state and this motivates choosing the control
\begin{equation}\label{equ}
\mathbf u(x,t)=k(\mathbf r -\mathbf m(x,t) )
\end{equation}
where $k$ is a positive constant control parameter for equation~(\ref{eqcontrolledLL}). It is clear that $\mathbf r$ is an equilibrium point of (\ref{eqcontrolledLL}) with the control in (\ref{equ}). Figure~\ref{figblockdiagram} is a block diagram representation of (\ref{eqcontrolledLL}) with control (\ref{equ}).
\begin{figure}[h]
\scalebox{1}{\begin{picture}(450,140)(-120,-120)
\put(-30,0){\vector(1,0){50}}
\put(-35,0){\circle{10}}
\put(-70,0){\vector(1,0){30}}
\put(20,-20){\framebox(20,40)[c]{ $k$ }}
\put(40,0){\vector(1,0){70}}
\put(110,-20){\framebox(140,40)[c]{Landau-Lifshitz Equation}}
\put(250,0){\vector(1,0){40}}
\put(270,0){\line(0,-1){40}}
\put(270,-40){\line(-1,0){305}}
\put(-35,-40){\vector(0,1){35}}
\put(-45,-10){-}
\put(-90,-2){\small{ $\mathbf r$}}
\put(60,6){\small{ $\mathbf u(x,t)$}}
\put(295,-2){\small{$\mathbf m(x,t)$}}
\end{picture}}\vspace*{-2.25cm}
\caption{\label{figblockdiagram} \small Closed-loop control of the Landau-Lifshitz equation.}\end{figure}
The following theorem establishes well-posedness of the controlled equation. In particular, for any initial condition $\mathbf m_0,$ the solution to (\ref{eqcontrolledLL}) with control $u(t) =k(\mathbf r-{\mathbf m}) $ satisfies
$\| \mathbf m(\cdot, t)\|_{\mathcal L_2^3} \leq 1.$
\begin{thm}\label{thmstrongsolutionLLcontrol}
For any $\mathbf r \in E,$ define the operator
\begin{equation}
B\mathbf m =k(\mathbf r-{\mathbf m}) \label{eqlinearAforcontrolledLL}.
\end{equation}
If $k>0$, the nonlinear operator $f+B$ with domain $D$, where $f$ and $D$ are defined in (\ref{defn:f}), (\ref{setDforfullLL}) respectively, generates a nonlinear contraction semigroup on ${\mathcal L_2^3}$.
\end{thm}
\begin{pf}
(i) For any $\mathbf m, \mathbf y \in D$,
\begin{align*}
&\langle f(\mathbf m)+B\mathbf m - (f(\mathbf y) +B\mathbf y),\mathbf m -\mathbf y \rangle_{\mathcal L_2^3} \\
&=\langle f(\mathbf m)-f(\mathbf y), \mathbf m -\mathbf y \rangle_{\mathcal L_2^3} +\langle B\mathbf m-B\mathbf y,\mathbf m -\mathbf y \rangle_{\mathcal L_2^3}.
\end{align*}
Since $f$ generates a nonlinear contraction semigroup (Theorem~\ref{thmuncontrolledLLsemigroup}), then $f$ is dissipative \cite[Proposition~2.98]{Luo1999}; that is,
\[
\langle f(\mathbf m)-f(\mathbf y), \mathbf m -\mathbf y \rangle_{\mathcal L_2^3}\leq 0.
\]
It follows that
\begin{align*}
\langle f(\mathbf m)+B\mathbf m - (f(\mathbf y) +B\mathbf y),\mathbf m -\mathbf y \rangle_{\mathcal L_2^3}
&\leq \langle B\mathbf m-B\mathbf y,\mathbf m -\mathbf y \rangle_{\mathcal L_2^3}\\
&=\langle -k\mathbf m+k\mathbf y,\mathbf m -\mathbf y \rangle_{\mathcal L_2^3}\\
&=-k\langle \mathbf m-\mathbf y,\mathbf m -\mathbf y \rangle_{\mathcal L_2^3}\\
&\leq 0
\end{align*}
and hence $f+B$ is dissipative.
(ii) Since $f$ generates a nonlinear contraction semigroup (Theorem~\ref{thmuncontrolledLLsemigroup}), $\mathrm{ran}(\mathrm{I}-\hat \alpha f)=\mathcal L_2^3$ for any $\hat\alpha >0$ \cite[Lemma~2.1]{Kato1967}.
This means that for any $\mathbf y_2 \in \mathcal L_2^3$ there exists $\mathbf m \in D$ such that $\mathbf{m}-\hat{\alpha}f(\mathbf m)=\mathbf y_2. $
Choose any $\mathbf y_1 \in \mathcal L_2^3$, $\alpha >0$ and define
\[
\mathbf y_2 = \frac{\mathbf y_1}{1+\alpha k}+\frac{\alpha k\mathbf r }{1+\alpha k}
\]
and
\[
\hat \alpha=\frac{\alpha }{1+\alpha k}.
\]
There exists $\mathbf m \in D$ such that
\[
\mathbf{m}-\frac{\alpha }{1+\alpha k}f(\mathbf m)=\mathbf y_2= \frac{\mathbf y_1}{1+\alpha k}+\frac{\alpha k\mathbf r }{1+\alpha k}.
\]
Solving for $\mathbf y_1$ leads to
\[
\mathbf y_1 = \mathbf m-\alpha (k(\mathbf r-\mathbf m)+ f(\mathbf m)).
\]
Thus, for any $\mathbf y_1 \in \mathcal L_2^3$, there exists $\mathbf m \in D$ such that $
\mathbf y_1 = (\mathrm I-\alpha (B+f))\mathbf m$ and hence $\mathrm{ran}\left(\mathrm I-\alpha (B+f)\right)=\mathcal L_2^3$ for some $\alpha >0$. It follows that the range is $\mathcal L_2^3$ for all $\alpha>0$ \cite[Lemma~2.1]{Kato1967}.
Thus, since $B+f$ is dissipative and the range of $(\mathrm I-\alpha (B+f))$ is $\mathcal L_2^3$, then $B+f$ generates a nonlinear contraction semigroup \cite[Proposition~2.114]{Luo1999}.
\qed
\end{pf}
The following lemmas are needed in the proof of the main results in Theorem~\ref{thmr0isasymstable} and Theorem~\ref{thmr0isexpstable}. The first theorem demonstrates the control in (\ref{equ}) can steer the dynamics in the Landau-Lifshitz to an asymptotically stable state in the $L_2^3$-norm, while the latter theorem establishes exponential stability in the $H_1$-norm, $||\mathbf m ||^2_{H_1}=||\mathbf m||_{\mathcal L_2^3}^2+||\mathbf m_x||_{\mathcal L_2^3}^2.$
\begin{lem}\label{lemmaboundonacrossz}
If $\mathbf a \in E$ where $E$ is defined in (\ref{equilibriumset}), then $||\mathbf a \times \mathbf m_{}||_{\mathcal L_2^3} \leq || \mathbf m_{}||_{\mathcal L_2^3}$ for all $\mathbf m \in \mathcal L_2^3$.
\end{lem}
\begin{pf}
Since $||\mathbf a \times \mathbf m||_2 =||\mathbf a||_2||\mathbf m||_2\sin(\theta)$ where $\theta$ is the angle between $\mathbf a$ and $\mathbf m$, and $||\mathbf a||_2=1$, then
$||\mathbf a \times \mathbf m||_2 \leq ||\mathbf m||_2$. Extending to the $L_2^3$-norm, the desired result is obtained.
\end{pf}
Lemmas ~\ref{thmderivativemcrossmprime} and~\ref{thmderivativemcrossmprimedotproduct} are simple consequences of the product rule.
\begin{lem}\label{thmderivativemcrossmprime}
For $\mathbf m\in \mathcal L_2^3$, the derivative of $\mathbf g=\mathbf m \times \mathbf m_x$ is $\mathbf g_x=\mathbf m \times \mathbf m_{xx}$.
\end{lem}
\begin{lem}\label{thmderivativemcrossmprimedotproduct}
For $\mathbf m \in \mathcal L_2^3$, the derivative of $f=\left(\mathbf m \times \mathbf m_x\right)^{\mathrm T}\left(\mathbf m \times \mathbf m_x\right)$ is $f_x=2\left(\mathbf m \times \mathbf m_{x}\right)^{\mathrm T}\left(\mathbf m \times \mathbf m_{xx}\right).$
\end{lem}
\begin{lem}
\label{lemmazerointegral}
For $\mathbf m \in \mathcal L_2^3$ satisfying (\ref{eqboundarycondition}),
\[
\int_0^L (\mathbf m-\mathbf r)^{\mathrm T}(\mathbf m \times \mathbf m_{xx})dx=0.
\]
\end{lem}
\begin{pf}
Integrating by parts, and applying Lemma~\ref{thmderivativemcrossmprime} and the boundary conditions (\ref{eqboundarycondition}) implies
\[
\int_0^L (\mathbf m-\mathbf r)^{\mathrm T}(\mathbf m \times \mathbf m_{xx})dx = -\int_0^L \mathbf m_x^{\mathrm T}(\mathbf m \times \mathbf m_{x})dx.
\]
From properties of cross products,
$
\mathbf m_x^{\mathrm T}(\mathbf m \times \mathbf m_{x}) = \mathbf m^{\mathrm T}(\mathbf m_x \times \mathbf m_{x})=0,
$
and hence the integral is zero. \qed
\end{pf}
\begin{lem}\label{lemmaPoincareInequalityforCrossProducts}
For $\mathbf m \in \mathcal L_2^3$ satisfying (\ref{eqboundarycondition}),
\[
|| \mathbf m \times \mathbf m_x ||_{\mathcal L_2^3} \leq 4L^2|| \mathbf m \times \mathbf m_{xx} ||_{\mathcal L_2^3}
\]
\end{lem}
\begin{pf}
Integrating by parts, using Lemma~\ref{thmderivativemcrossmprimedotproduct} and the boundary conditions (\ref{eqboundarycondition}) leads to
\[
|| \mathbf m \times \mathbf m_x ||_{\mathcal L_2^3}^2 =-\int_0^L2\left(\mathbf m \times \mathbf m_{x}\right)^{\mathrm T}\left(\mathbf m \times \mathbf m_{xx}\right)xdx.
\]
It follows from Young's inequality that
\begin{align*}
|| \mathbf m \times \mathbf m_x ||_{\mathcal L_2^3}^2
\leq \frac{1}{2}\int_0^L\left(\mathbf m \times \mathbf m_{x}\right)^{\mathrm T}\left(\mathbf m \times \mathbf m_{x}\right)dx+\int_0^L2\left(\mathbf m \times \mathbf m_{xx}\right)^{\mathrm T}\left(\mathbf m \times \mathbf m_{xx}\right)x^2dx .
\end{align*}
Since $x \in [0,L],$
\begin{align*}
|| \mathbf m \times \mathbf m_x ||_{\mathcal L_2^3}^2
\leq \frac{1}{2}||\mathbf m \times \mathbf m_{x}||_{\mathcal L_2^3}^{2}+2L^2||\mathbf m \times \mathbf m_{xx}||_{\mathcal L_2^3}^2.
\end{align*}
Rearranging gives the desired inequality. \qed
\end{pf}
\begin{thm}\label{thmr0isasymstable}
Let $\mathbf r$ be an equilibrium point of (\ref{eqcontrolledLL}) with control defined in (\ref{equ}). For any positive constant $k$ such that $k> 8\nu L^4$, $\mathbf r$ is a globally asymptotically stable point of (\ref{eqcontrolledLL}) in the $\mathcal L_2^3$--norm.
\end{thm}
\begin{pf}
The Lyapunov candidate is
\[
V(\mathbf m)=\frac{1}{2}\left| \left| \mathbf m-\mathbf r\right|\right|_{\mathcal L_2^3}^2+\frac{1}{2}\left| \left| \mathbf m_x\right|\right|_{\mathcal L_2^3}^2
\]
which is clearly nonegative. Furthermore, $V=0$ if and only if $\mathbf m=\mathbf r$.
Taking the derivative of $V$
\begin{align*}
\frac{dV}{dt}&=\int_0^L(\mathbf m -\mathbf r)^{\mathrm T}\dot{{\mathbf m}} dx+\int_0^L\mathbf m_x^{\mathrm T} \dot{{\mathbf m}}_xdx \\
&=\int_0^L(\mathbf m -\mathbf r)^{\mathrm T}\dot{\mathbf m} dx-\int_0^L\mathbf m_{xx}^{\mathrm T} \dot{\mathbf m}dx
\end{align*}
where the dot notation means differentiation with respect to $t$. Substituting in (\ref{eqcontrolledLL}) to eliminate $\dot{\mathbf m}$,
\begin{align*}
\frac{dV}{dt}
&= \int_0^L(\mathbf m -\mathbf r)^{\mathrm T} \left(\mathbf m \times \mathbf m_{xx}\right)dx-\nu\int_0^L (\mathbf m -\mathbf r)^{\mathrm T}\left(\mathbf m\times\left(\mathbf m\times \mathbf m_{xx}\right)\right)dx \\
&+k\int_0^L(\mathbf m -\mathbf r)^{\mathrm T} (\mathbf r -\mathbf m)dx -\int_0^L\mathbf m_{xx}^{\mathrm T} \left(\mathbf m \times \mathbf m_{xx}\right)dx\\
&+\nu\int_0^L \mathbf m_{xx}^{\mathrm T}\left(\mathbf m\times\left(\mathbf m\times \mathbf m_{xx}\right)\right)dx -k\int_0^L\mathbf m_{xx}^{\mathrm T} (\mathbf r -\mathbf m)dx.
\end{align*}
From Lemma~\ref{lemmazerointegral}, the first integral is zero. Furthermore, from properties of cross products,
\[
\mathbf m_{xx}^{\mathrm T} \left(\mathbf m \times \mathbf m_{xx}\right)=\mathbf m^{\mathrm T} \left(\mathbf m_{xx} \times \mathbf m_{xx}\right)=0,
\]
and hence
\[
\int_0^L\mathbf m_{xx}^{\mathrm T} \left(\mathbf m \times \mathbf m_{xx}\right)dx=0.
\]
It follows that
\begin{align*}
\frac{dV}{dt}=&-\nu\int_0^L (\mathbf m -\mathbf r)^{\mathrm T}\left(\mathbf m\times\left(\mathbf m\times \mathbf m_{xx}\right)\right)dx -k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2\\
&-\nu||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2 -k\int_0^L\mathbf m_{xx}^{\mathrm T} (\mathbf r -\mathbf m)dx.
\end{align*}
Applying integration by parts to the last integral leads to
\begin{align}
\frac{dV}{dt}&=-\nu\int_0^L (\mathbf m -\mathbf r)^{\mathrm T}\left(\mathbf m\times\left(\mathbf m\times \mathbf m_{xx}\right)\right)dx -k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2-\nu||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2 -k||\mathbf m_{x}||_{\mathcal L_2^3}^2 \nonumber\\
&=-\nu\int_0^L\left( (\mathbf m -\mathbf r)\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx -k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2-\nu||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2 -k||\mathbf m_{x}||_{\mathcal L_2^3}^2 \nonumber\\
&=\nu\int_0^L\left( \mathbf r\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx -k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2-\nu||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2 -k||\mathbf m_{x}||_{\mathcal L_2^3}^2. \label{eqderivativeofVlastintegral}
\end{align}
Applying integration by parts with Lemma~\ref{thmderivativemcrossmprime} to the integral implies
\begin{align*}
\int_0^L\left( \mathbf r\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx&=\left[\left(\mathbf r \times \mathbf m\right)^{\mathrm T}\left(\mathbf m \times \mathbf m_x\right)\right]_0^L-\int_0^L\left( \mathbf r\times \mathbf m_x\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{x}\right)dx
\end{align*}
and substituting in the boundary conditions in (\ref{eqboundarycondition}) leads to
\begin{align*}
\int_0^L\left( \mathbf r\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx&=-\langle \mathbf r\times \mathbf m_x, \mathbf m\times \mathbf m_{x}\rangle_{\mathcal L_2^3}.
\end{align*}
Then from Cauchy-Schwarz and Lemma~\ref{lemmaPoincareInequalityforCrossProducts},
\begin{align*}
\int_0^L\left( \mathbf r\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx&\leq|| \mathbf r\times \mathbf m_x||_{\mathcal L_2^3} ||\mathbf m\times \mathbf m_{x}||_{\mathcal L_2^3}
\\
&\leq 4L^2 || \mathbf r\times \mathbf m_x||_{\mathcal L_2^3} ||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}.
\end{align*}
It follows from Young's Inequality that
\begin{align*}
\int_0^L\left( \mathbf r\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx&\leq 8L^4 || \mathbf r\times \mathbf m_x||_{\mathcal L_2^3} ^2+\frac{1}{2}||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2
\end{align*}
and from Lemma~\ref{lemmaboundonacrossz},
\begin{align*}
\int_0^L\left( \mathbf r\times \mathbf m\right)^{\mathrm T}\left(\mathbf m\times \mathbf m_{xx}\right)dx&\leq 8L^4 || \mathbf m_x||_{\mathcal L_2^3} ^2+\frac{1}{2} ||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2
\end{align*}
Substituting this result into (\ref{eqderivativeofVlastintegral}) leads to
\begin{equation}
\label{eqderivativeofVLL}
\frac{dV}{dt}\leq -\left(k-8\nu L^4\right)||\mathbf m_{x}||_{\mathcal L_2^3}^2-\frac{\nu}{2}||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2-k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2.
\end{equation}
The derivative is negative if $k > 8 \nu L^4.$
It follows that
\[
\frac{dV}{dt}\leq-k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2.
\]
Therefore, $dV/dt <0$ for all $\mathbf m \neq \mathbf r$ and $dV/dt=0 $ if $\mathbf m =\mathbf r$. Since $V(\mathbf m)\geq \frac{1}{2}||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2,$ $V\rightarrow \infty$ as $||\mathbf m -\mathbf r||\rightarrow \infty$.
From Lyapunov's Theorem \cite[Theorem~6.2.13]{Michel1995}, $\mathbf r$ is a globally asymptotically stable equilibrium of (\ref{eqcontrolledLL}). \qed
\end{pf}
\begin{thm}\label{thmr0isexpstable}
Let $\mathbf r$ be an equilibrium point of (\ref{eqcontrolledLL}) with control defined in (\ref{equ}). For any positive constant $k$ such that ${k> 8\nu L^4}$, $\mathbf r$ is a globally exponentially stable equilibrium point of (\ref{eqcontrolledLL}) in the $H_1$--norm. That is, for any initial condition on $H_1$, $\mathbf m$ decreases exponentially in the $H_1$-norm to $\mathbf r.$
\end{thm}
\begin{pf}
In the proof of Theorem~\ref{thmr0isasymstable}, we have from (\ref{eqderivativeofVLL}) that the derivative of V satisfies
\begin{align*
\frac{dV}{dt}&\leq -\left(k- {8\nu L^4}\right)||\mathbf m_{x}||_{\mathcal L_2^3}^2-\frac{\nu}{2}||\mathbf m\times \mathbf m_{xx}||_{\mathcal L_2^3}^2-k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2
\end{align*}
and hence
\begin{align*}
\frac{dV}{dt}&\leq -\left(k-{8\nu L^4}\right)||\mathbf m_{x}||_{\mathcal L_2^3}^2-k||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2\\
&\leq -\left(k-{8\nu L^4}\right)\left(||\mathbf m_{x}||_{\mathcal L_2^3}^2+||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2\right)\\
&=-2\left(k-{8\nu L^4}\right)V.
\end{align*}
Integrating with respect to time
\begin{align*}
||\mathbf m_{x}||_{\mathcal L_2^3}^2+||\mathbf m -\mathbf r||_{\mathcal L_2^3}^2
& \leq e^{-2\left(k-{8\nu L^4}\right)t}\left(||\mathbf {m}_{x}(x,0)||_{\mathcal L_2^3}^2+||\mathbf m(x,0) -\mathbf r||_{\mathcal L_2^3}^2 \right)
\end{align*}
Noting that $\mathbf r$ does not depend on $x$, it follows that
\[
||\mathbf m-\mathbf r||_{H_1}^2\leq e^{-2\left(k-{8\nu L^4}\right)t}||\mathbf m(x,0) -\mathbf r||_{H_1}^2
\]
and since $k>{8\nu L^4}$, $\mathbf r$ is an exponentially stable equilibrium point of (\ref{eqcontrolledLL}). \qed
\end{pf}
A natural question is whether $\mathbf r$ is exponentially stable in the $\mathcal L_2^3$-norm.
Analysis of the linear Landau-Lifshitz equation provides insight to this question. For the control in (\ref{equ}), the linearized controlled Landau-Lifshitz equation is
\begin{align}\label{eqcontrolledlinearLL}
\frac{\partial \mathbf z}{\partial t} & =\nu \mathbf z_{xx}+\mathbf a \times {\mathbf z}_{xx} +k \left(\mathbf r-\mathbf z\right), \qquad \mathbf z(0)=\mathbf z_0
\end{align}
with the same boundary conditions $\mathbf z_{x}(0)=\mathbf z_{x}(L)=\mathbf 0 .$
Since the uncontrolled linear Landau-Lifshitz equation (\ref{eqstatespaceform}) generates a linear semigroup and $k \left(\mathbf r-\mathbf z\right)$ is a bounded linear (affine) operator, then the operator in (\ref{eqcontrolledlinearLL}) generates a semigroup \cite[Theorem~3.2.1]{Curtain1995}. Substituting $\mathbf z=\mathbf{r}$ into (\ref{eqcontrolledlinearLL}) leads to ${\partial \mathbf z}/{\partial t}=\mathbf 0$ and hence $\mathbf{r}$ is a stable equilibrium point of (\ref{eqcontrolledlinearLL}).
\begin{thm}\label{thmlinearr0isasymstable}
Let $\mathbf r \in E .$ For any positive constant $k$, $\mathbf r$ is an exponentially stable equilibrium of the linearized system (\ref{eqcontrolledlinearLL}) in
the $\mathcal L_2^3$--norm.
\end{thm}
\begin{pf}
For $\mathbf z \in D(A)$, where $D(A)=D$ as in equation~(\ref{setDforfullLL}), consider the Lyapunov candidate
\begin{equation*}
V(\mathbf z)=\frac{1}{2}\left| \left|\mathbf z-\mathbf r\right|\right|_{\mathcal L_2^3}^2.
\end{equation*}
It is clear that $V\geq 0$ for all $\mathbf z\in D(A)$ and furthermore, $V(\mathbf z)=0$ only when $\mathbf z=\mathbf r$. Therefore, $V(\mathbf z)>0$ for all $\mathbf z \in D(A)\backslash \{\mathbf r\}$.
Taking the derivative of $V(\mathbf z)$ implies
\begin{align*}
\frac{dV}{dt}&=\int_0^L(\mathbf z-\mathbf r)^{\mathrm T}\dot{ {\mathbf z}} dx.
\end{align*}
Substituting in (\ref{eqcontrolledlinearLL}) yields
\begin{align*}
\frac{dV}{dt}&=\nu\int_0^L(\mathbf z-\mathbf r)^{\mathrm T} {\mathbf z}_{xx}dx +\int_0^L(\mathbf r-\mathbf z)^{\mathrm T}\left({\mathbf a \times {\mathbf z_{xx}}}\right)dx +k\int_0^L(\mathbf z-\mathbf r)^{\mathrm T} ({\mathbf r-\mathbf z}) dx.
\end{align*}
By Lemma~\ref{lemmazerointegral}, the middle term is zero. Using integration by parts, the first term becomes
\[
-\nu\int_0^L\mathbf z_x^{\mathrm T} {\mathbf z}_{x}dx.
\]
It follows that
\begin{align*}
\frac{dV}{dt}&=-\nu||\mathbf z_x||_{\mathcal L_2^3}^2 -k||\mathbf z-\mathbf r||_{\mathcal L_2^3}^2
\end{align*}
and since $\nu\geq0$,
\begin{align*
\frac{dV}{dt}&\leq -k||\mathbf z-\mathbf r||_{\mathcal L_2^3}^2=-2kV.
\end{align*}
Solving yields
\[
||\mathbf z-\mathbf r||_{\mathcal L_2^3}^2 \leq e^{-2kt}||\mathbf z_0-\mathbf r||_{\mathcal L_2^3}^2.
\]
For $k>0$ the equilibrium point, $\mathbf r,$ of (\ref{eqcontrolledlinearLL}) is exponentially stable. \qed
\end{pf}
Theorem~\ref{thmlinearr0isasymstable} suggests the equilibrium point in the controlled nonlinear Landau-Lifshitz equation (\ref{eqcontrolledLL}) is exponentially stable in the $\mathcal L_2^3$--norm. However, since the nonlinearity in the Landau-Lifshitz equation is unbounded, stability of the linear equation does not necessarily reflect stability of the original nonlinear equation; see \cite{CoronNguyen2015,alJamal2013,AM2014}.
\section{Example}\label{secExample}
Simulations illustrating the stabilization of the (nonlinear) Landau-Lifshitz equation are constructed using a Galerkin approximation with 12 linear spline elements. For the following simulations, the parameters are $\nu=0.02$ and $L=1$ with initial condition $\mathbf {{m}_0} (x)=(\sin(2\pi x),\cos(2\pi x),0)$. Figure~\ref{figMagnetizationNoControl3d} illustrates the solution to the uncontrolled Landau-Lifshitz equation settles to $\mathbf{r_0}=(0,-0.6,0).$
Stabilization of the Landau-Lifshitz equation with affine control (\ref{eqcontrolledLL}) is illustrated in Figures~\ref{figMagnetization3d} and~\ref{figLLmcontrolTWICEICEQEQ3d} with control parameter $k=0.5$. In Figure~\ref{figMagnetization3d}, the system dynamics are steered from $\mathbf m_0$ to
the equilibrium point $\mathbf r_1=(-\frac{1}{\sqrt{2}},0,\frac{1}{\sqrt{2}})$. Figure~\ref{figLLmcontrolTWICEICEQEQ3d} depicts applying the control twice in succession, forcing the system from the equilibrium $\mathbf{r_0}$ to $\mathbf r_2=(1,0,0)$ and then to $\mathbf r_3 = (0,0,1)$. In each case, the state of the controlled system converges to the specified point $\mathbf r_i $ as predicted by the analysis.
Adding a feedback control so that there is only one equilibrium point also removes hysteresis from the system.
Consider the input--output dynamics of the controlled Landau-Lifshitz equation with periodic input $\hat{\mathbf u}(t)=\left(0.001\cos(\omega t),0,0\right)$. The initial condition is $\mathbf m_0(x)=\left(1,0,0\right)$ and the control parameters are $k=0.5$ and $\mathbf r=(1,0,0)$. Figure~\ref{figcontrolm1LLLooping} illustrates the input--output dynamics for $m_1(x,t)$ with $x$ fixed, $L=1$ and $\nu=0.02$. It is clear from the figure that persistent looping behaviour does not occur and hence, based on Definition~\ref{defBernstein}, the controlled Landau-Lifshitz equation in (\ref{eqcontrolledLL}) does not exhibit hysteresis. Similar behaviour is observed for $m_2(x,t)$ and $m_3(x,t)$.
\begin{figure}[h]
\centering\scalebox{0.7}{
\includegraphics{LLm1nocontrolMeshBone.pdf}
\includegraphics{LLm2nocontrolMeshBone.pdf}
\includegraphics{LLm3nocontrolMeshBone.pdf}}
\caption{\label{figMagnetizationNoControl3d}
Magnetization in the uncontrolled (nonlinear) Landau-Lifshitz equation moves from initial condition $\mathbf{{m}_0} (x)$, to the equilibrium $\mathbf {r_0}=(0,-0.6,0)$. }
\end{figure}
\begin{figure}[h]
\centering\scalebox{0.7}{
\includegraphics{LLm1controlFeedbackMeshBoneBetterScales.pdf}
\includegraphics{LLm2controlFeedbackMeshBoneBetterScales.pdf}
\includegraphics{LLm3controlFeedbackMeshBoneBetterScales.pdf}}
\caption{\label{figMagnetization3d} With a proportional control $(k=0.5)$, magnetization in the (nonlinear) Landau-Lifshitz equation with a linear control moves from the initial condition
$\mathbf{{m}_0} (x)$ to the specified equilibrium
$\mathbf{ r_1}=(-\frac{1}{\sqrt{2}},0,\frac{1}{\sqrt{2}}) .$ }
\end{figure}
\begin{figure}[h]
\centering\scalebox{0.7}{
\includegraphics{LLm1controlTwiceICEQEQMeshBone.pdf}
\includegraphics{LLm2controlTwiceICEQEQMeshBone.pdf}
\includegraphics{LLm3controlTwiceICEQEQMeshBone.pdf}}
\caption{\label{figLLmcontrolTwiceICEQEQ3d} Steering magnetization between specified equilibria with a linear control. The uncontrolled magnetization moves from initial condition $\mathbf{{m}_0}$ to $\mathbf {r_0}=(0,-0.6,0)$. Proportional control $(k=0.5)$ with two successive values of $\mathbf r$ first forces the magnetization to $\mathbf {r_2}= (1,0,0)$ and then to $\mathbf{r_3}= (0,0,1)$.
}
\end{figure}
\begin{figure}[h]\hspace*{-0.3cm}
\centering\scalebox{0.544}{
\subfloat[{\Large$\omega=1$}]{ \includegraphics{LoopFrequencym1Controlw1.pdf}}\hspace*{0.05cm}
\subfloat[{\Large$\omega=0.1$}]{\includegraphics{LoopFrequencym1Controlw2.pdf}}
\subfloat[{\Large$\omega=0.01$}]{ \includegraphics{LoopFrequencym1Controlw3.pdf}}
\subfloat[{\Large$\omega=0.001$}]{ \includegraphics{LoopFrequencym1Controlw4.pdf}} }
\caption{\label{figcontrolm1LLLooping} Input--output dynamics for $m_1(x,t)$ of the controlled (nonlinear) Landau-Lifshitz equation in (\ref{eqcontrolledLL}) with $x$ fixed demonstrate the absence of persistent looping behaviour as the frequency of the periodic input $\left(0.001\cos(\omega t),0,0\right)$ approaches zero. ($L=1$, $\nu=0.02$, $\mathbf m_0(x)=\left(1,0,0\right)$, $k=0.5$, $\mathbf r=(1,0,0)$)}
\end{figure}
\section{Conclusion}
The Landau-Lifshitz equation is a nonlinear system of partial differential equations with multiple equilibrium points. The presence of a zero eigenvalue in the linearized equation suggested a simple feedback proportional control can steer the system to an arbitrary equilibrium point.
It was then proven that proportional control of the Landau-Lifshitz equation does lead to an equilibrium point that is globally asymptotically stable in the $\mathcal L_2^3$-norm (Theorem \ref{thmr0isasymstable}) and exponentially stable in the $H_1$-norm (Theorem \ref{thmr0isexpstable}).
The fact the Landau-Lifshitz equation is not quasi-linear means linearization is not guaranteed, without further analysis,
to predict stability of the nonlinear equation \cite{alJamal2013}.
Moreover, since the objective of the control is to steer between equilibrium points, a linearized analysis, which only yields local results, would not predict stability of the controlled system. Results on preservation of linearized stability require exponential stability of the linearized system; see for example \cite[Theorem~3.3]{alJamal2013} \cite[Corollary~2.2]{Kato1995}\cite[Theorem~11.22]{Smoller1983}.
The fact the linearized system is exponentially stable in the $\mathcal L_2^3$-norm (Theorem~\ref{thmlinearr0isasymstable}) is encouraging, but further research is needed to determine whether the controlled nonlinear system is also exponentially stable.
\section*{Acknowledgements}
The research described in this article was supported by the Natural Sciences and Engineering Research Council of
Canada (NSERC) through grant RGPIN-6053-2015.
\section{References}
\bibliographystyle{elsarticle-num}
|
\section{Introduction}
Recall that an \emph{oriented graph} is a digraph obtained from a simple, undirected graph by giving each edge one of its two possible orientations. Recall, also, that if $G$ and $H$ are oriented graphs, then a \emph{homomorphism of $G$ to $H$} is a function $\phi$
from the vertices of $G$ to the vertices of $H$ such that~\begin{math}
\phi(x)\phi(y) \in E(H) \end{math} whenever~\begin{math} xy \in E(G)\end{math}. If $G$ and $H$ are oriented graphs such that there is a homomorphism $\phi$ of $G$ to $H$, then we write~\begin{math}
\phi: G \to H
\end{math}, or~\begin{math}
G \to H
\end{math} if the name of the function $\phi$ is not important.
Let $k$ and $t$ be positive integers, and let $G$ be an oriented graph. \cite{MIWE06} defined a \emph{$k$-dipath $t$-colouring} of $G$ to be an assignment of $t$ colours to the vertices of $G$ so that any two vertices joined by a directed path of length at most $k$ are assigned different colours.
A $1$-dipath $t$-colouring of an oriented graph $G$ is a $t$-colouring of the underlying undirected graph of $G$.
See Figure \ref{kdipath:example1} for an example of a $3$-dipath $4$-colouring of an oriented graph. The \emph{$k$-dipath chromatic number of $G$}, denoted by~\begin{math}
\chi_{k\text{-dip}}(G)
\end{math}, is the smallest positive integer $t$ such that there exists a $k$-dipath $t$-colouring of $G$. \cite{MIWE06} showed that any orientation of a Halin graph has $2$-dipath chromatic number at most $7$, and there are infinitely many such graphs $G$ with~\begin{math}
\chi_{2\text{-dip}}(G) = 7
\end{math}.
\begin{figure}
\begin{center}
\includegraphics[width=0.5\textwidth]{example.pdf}
\end{center}
\caption{A $3$-dipath $4$-colouring.}
\label{kdipath:example1}
\end{figure}
For a positive integer $t$, an \emph{oriented $t$-colouring} of an oriented graph $G$ is a homomorphism of $G$ to some oriented graph on $t$ vertices. Oriented colourings were first introduced by \cite{CO94}, and have been a topic of considerable interest in the literature since then; see the recent survey by \cite{SO15}, and also work by \cite{BO99}, \cite{HDS06b}, \cite{SW13} for related topics.
The $2$-dipath chromatic number is of interest, in part, because it gives a lower bound for the \emph{oriented chromatic number}~\begin{math}
\chi_o(G)
\end{math} -- the smallest positive integer $t$ such that $G$ admits an oriented $t$-colouring.
Since oriented graphs have no directed cycles of length two, the definition implies that any two vertices of $G$ joined by a directed path of length at most two are assigned different colours (\textit{i.e.}, they have different images) in an oriented colouring of $G$.
It follows from the definition that any oriented colouring of $G$ is a $2$-dipath colouring of $G$; hence~\begin{math}
\chi_{2\text{-dip}}(G) \leq \chi_o(G).
\end{math}
In her Master's thesis \cite{KMY09} gives a \emph{homomorphism model} for $2$-dipath $t$-colouring. For each positive integer $t$, she describes an oriented graph $\mathcal{G}_t$ with the property that an oriented graph $G$ has a $2$-dipath $t$-colouring if and only if there is a homomorphism of $G$ to $\mathcal{G}_t$.
As is common with such theorems, it is possible to use the homomorphism to $\mathcal{G}_t$
to find a $2$-dipath $t$-colouring of $G$: the colour assigned to a vertex is determined by
its image (but is not equal to it).
The existence of this model implies an upper bound for the oriented chromatic number as a function of the $2$-dipath chromatic number. It also leads to a proof that deciding whether a given oriented graph has a $2$-dipath $t$-colouring is Polynomial if the fixed integer~\begin{math}
t \leq 2
\end{math}, and NP-complete if~\begin{math}
t \geq 3
\end{math}.
A natural question is whether Sherk's results can be generalized to $k$-dipath $t$-colouring.
We seek a model similar to hers, where the homomorphism to the target oriented graph can be used to
find a $k$-dipath $t$-colouring of the given oriented graph.
In particular, vertices of the given oriented graph $G$ with the same image should
be assigned the same colour.
We suggest that such model will likely exist only for colouring oriented graphs with no directed cycles of
length $k$ or less.
Consider the case of $3$-dipath $t$-colouring, where~\begin{math} t \geq 3 \end{math}. Suppose there exists a digraph $H_{3, t}$ with the property that an oriented graph $G$ has a $3$-dipath $t$-colouring if and only if there is a homomorphism of $G$ to $H_{3, t}$. The digraph $H_{3, t}$ has no loops, otherwise all vertices of $G$ can be assigned the same colour.
By definition, the directed $3$-cycle has a $3$-dipath $3$-colouring: assign each vertex a different colour. Thus, there is a homomorphism of the directed $3$-cycle to $H_{3, t}$. Consequently, $H_{3, t}$ has a directed $3$-cycle.
But, now there is a homomorphism of a directed path of length three to $H_{3, t}$ in which the two end vertices have the same image. Since the ends of a directed path of length 3 must be assigned different colours in
a $3$-dipath $t$-colouring, this model will not have the desired property.
Similar considerations apply to $k$-dipath $t$-colouring for all pairs of positive integers $k$ and $t$ with~\begin{math} t \geq k \end{math}. Hence, a homomorphism model of the type we seek will not exist if the oriented graphs being coloured can have directed cycles of length $k$ or less. Finally, we note that Sherk's homomorphism model for $2$-dipath $t$-colouring is for oriented graphs. These have no directed cycles of length two or less.
The main result of this paper is the construction of a homomorphism model for $k$-dipath $t$-colouring of oriented graphs with no directed cycles of length $k$ or less. That is, for all positive integers $k$ and $t$ we describe an oriented graph~\begin{math}\mathcal{G}_{k, t}\end{math} with the property that an oriented graph $G$, with no directed cycle of length at most $k$, has a $k$-dipath $t$-colouring if and only if $G$ admits a homomorphism to~\begin{math}\mathcal{G}_{k, t}\end{math} and, further,
the homomorphism can be used to find a $k$-dipath $t$-colouring of $G$.
After presenting this result in Section \ref{kdipath:homsec}, in Section \ref{kdipath:comsec} we determine the complexity of deciding the existence of a $k$-dipath $t$-colouring for all pairs of fixed positive integers $k$ and $t$. When instances are restricted to oriented graphs with no directed cycles of length $k$ or less, it is shown that that this problem is NP-complete whenever ~\begin{math}
t>k\geq 3
\end{math}, and Polynomial if~\begin{math} t = k \end{math} or~\begin{math} k \leq 2 \end{math}. When there are no restrictions, it is shown that this problem is NP-complete whenever~\begin{math} k \geq 3 \end{math} and~\begin{math} t \geq 3 \end{math}, and Polynomial whenever~\begin{math} k \leq 2 \end{math} or~\begin{math} t \leq 2 \end{math}.
\section{Preliminaries}\label{kdipath:presec}
In this section we review relevant definitions, the homomorphism model for $2$-dipath colourings, and make small improvements to the known upper bounds on the oriented chromatic number of oriented graphs with $2$-dipath chromatic number $3$ or $4$. We also observe some straightforward extensions to $k$-dipath colouring of known results for $2$-dipath colouring.
Let $G$ be an oriented graph, and let~\begin{math} x, y \in V(G)\end{math}.
A \emph{directed walk} is a sequence of vertices ~\begin{math}W = v_0, v_1,v_2, \dots, v_{\ell-1}, v_\ell \end{math}, such that
\begin{math}v_iv_{i+1} \in E(G)\end{math} for~\begin{math} = 0, 1, \ldots, \ell -1\end{math}.
The integer $\ell$ is the \emph{length} of $W$.
If~\begin{math}v_0 = x \end{math} and~\begin{math}v_\ell = y\end{math}, then $W$ is a \emph{directed walk from $x$ to $y$}.
Note that the vertices belonging to $W$ need not be different.
If no two vertices of $W$ are the same, then $W$ is a \emph{directed path}.
If all vertices of $W$ are different except $v_0$ and $v_\ell$, then $W$ is a \emph{directed cycle}.
If there is a directed walk from $x$ to $y$, then the vertex $y$ is said to be \emph{reachable} from $x$.
The \emph{distance} from $x$ to $y$ is defined to be the smallest length of a directed walk from $x$ to $y$, or infinity of no
such walk exists.
The \emph{weak distance between $x$ and $y$}, denoted $d_{weak}(x,y)$, is
defined to be the minimum of the distance from $x$ to $y$ and the distance from $y$ to $x$.
This parameter is $\infty$ if neither of $x$ and $y$ is reachable from the other.
The \emph{weak diameter} of an oriented graph $G$ is the maximum of the weak distance between any
two distinct vertices of $G$.
A directed graph is called \emph{weakly connected} if its weak diameter is finite.
Let $G$ be a directed graph, and~\begin{math}x \in V(G)\end{math}.
The \emph{out-neighbourhood} of $x$ is~\begin{math}{N^+(x) = \{y: xy \in E(G)\}}\end{math},
and the \emph{in-neighbourhood} of $x$ is~\begin{math}N^-(x) = \{y: yx \in E(G)\}\end{math}.
The vertex $x$ is a \emph{source} if~\begin{math}N^-(x) = \emptyset\end{math},
and is a \emph{sink} if~\begin{math}N^+(x) = \emptyset \end{math}.
A \emph{universal source} is a source such that~\begin{math}N^+(x) = V(G) - \{x\}\end{math},
and a \emph{universal sink} is a sink such that~\begin{math}N^-(x) = V(G) - \{x\}\end{math}.
More generally, the \emph{$\ell$-out-neighbourhood} of $x$ is the set of all vertices reachable from $x$ by a directed
walk of length at most $\ell$, and the
\emph{$\ell$-in-neighbourhood} of $x$ is the set of all vertices which can reach $x$ by a directed
walk of length at most $\ell$.
The \emph{directed girth} of a directed graph $H$ is defined to be the minimum length of a directed cycle in $H$,
or infinity if $H$ has no directed cycle.
Our homomorphism model for $k$-dipath colouring applies only to the family of oriented graphs
with directed girth at least $k+1$.
Let $\mathcal{F}$ be a family of oriented graphs.
The \emph{oriented chromatic number of $\mathcal{F}$}, denoted by $\chi_o(\mathcal{F})$,
is the least integer $t$ so that~\begin{math}\chi_o(F) \leq t \end{math} for all~\begin{math}F \in \mathcal{F}\end{math}.
We say that an oriented graph $H$ is a \emph{universal target for $\mathcal{F}$} if~\begin{math}F \rightarrow H\end{math} for all~\begin{math}F \in \mathcal{F}\end{math}.
If $H$ is a universal target for $\mathcal{F}$, then~\begin{math}\chi_o(\mathcal{F}) \leq |V(H)|\end{math}.
For example, using the quadratic residue tournament on seven vertices as a universal target
for the family $\mathcal{O}$ of orientations of outerplanar graphs, \cite{SO97} shows~\begin{math}\chi_o(\mathcal{O}) \leq 7\end{math}.
Let $G$ be an oriented graph with directed girth at least $k+1$.
We say $G$ is a \emph{$k$-dipath clique} if~\begin{math}\chi_{k\text{-dip}}(G) = |V(G)|\end{math}.
The terminology arises by analogy with undirected graphs, where a clique is a graph
for which the chromatic number equals the number of vertices.
Let $G$ be an oriented graph. Define $G^k$ to be the simple graph formed from $G$ as follows:
\begin{itemize}
\item~\begin{math}V(G^k) = V(G)\end{math}, and
\item~\begin{math}E(G^k) = \{uv | 0 < d_{weak}(u,v) \leq k \}\end{math}.
\end{itemize}
It follows from the definitions
that there is an equivalence between $2$-dipath colourings of $G$ and proper vertex colourings of the simple graph $G^2$
(see \cite{MAYO12}).
This equivalence extends to $k$-dipath colourings of $G$ and proper vertex colourings of $G^k$.
\begin{obs}
If $G$ is an oriented graph with directed girth at least $k+1$, then there is a one-to-one correspondence between $k$-dipath colourings of $G$ and proper colourings of $G^k$.
\end{obs}
Using this observation we generalize a result of \cite{BDS17} for $2$-dipath cliques.
This result will be used in Section \ref{kdipath:comsec}.
\begin{prop} [\cite{BDS17}]
An oriented graph is a $2$-dipath clique if and only if it has weak diameter at most $2$.
\end{prop}
\begin{prop}
Let $k\geq 2$ be an integer. An oriented graph is a $k$-dipath clique if and only if it has weak diameter at most $k$.
\end{prop}
\begin{proof}
Let $G$ be an oriented graph with directed girth at least $k+1$.
We observe that $G^k$ is a complete graph if and only if for each pair of non-adjacent vertices,
say $u$ and $v$, there is a directed path of length at most $k$, in some direction, between $u$ and $v$.
Equivalently, $G$ has weak diameter at most $k$.
\end{proof}
We now review the homomorphism model for $2$-dipath colouring given by \cite{MAYO12}.
Let $t$ be a positive integer.
The oriented graph $\mathcal{G}_t$ is defined as follows.
Its vertices are $(t+1)$~-~tuples in which the position among~\begin{math}1, 2, \ldots, t \end{math} indicated by the 0-th entry is filled with the place-holder
``$\cdot$'', and the remaining positions among $1, 2, \ldots, t$ are filled with a 0 or a 1.
\begin{displaymath}V(\mathcal{G}_t) = \{ (u_0;u_1,u_2,\dots,u_t) : u_0 \in \{1,2,\dots,t\}, u_{u_0} = \cdot, u_i \in \{0,1\} \text{ for } 1 \leq i \leq t \text{ and } i\neq u_0 \} \end{displaymath}
\begin{displaymath} E(\mathcal{G}_t) = \{ (u_0;u_1,u_2,\dots,u_t)(x_0;x_1,x_2,\dots,x_t) : u_{x_0} = 1, x_{u_0} = 0\}. \end{displaymath}
It is then proved that an oriented graph $G$ admits a homomorphism to
$\mathcal{G}_t$ if and only if $G$ has $2$-dipath chromatic number at most $t$.
That is, $\mathcal{G}_t$ is a universal target for the family of oriented graphs
with $2$-dipath chromatic number at most $t$.
Further, suppose~\begin{math}\phi: G \to \mathcal{G}_t \end{math}.
For each vertex $x$, assigning colour $u_0$ to $x$ if and only if the mapping $\phi$,~\begin{math}\phi(x) = (u_0;u_1,u_2,\dots,u_t)\end{math},
is a 2-dipath $t$-colouring of $G$.
Since homomorphisms compose,
the oriented chromatic number of $\mathcal{G}_t$ is an upper bound for the oriented chromatic number of $G$.
\cite{MRS} show that~\begin{math}\chi_o(\mathcal{G}_t) \leq 2^{t}-1\end{math}, which gives the following.
\begin{thm} [\cite{MRS}] \label{kdipath:2bounds}
If $G$ is an oriented graph, then
\begin{displaymath}\chi_{2\text{-dip}}(G) \leq \chi_o(G) \leq 2^{\chi_{2\text{-dip}}(G)} -1. \end{displaymath}
\end{thm}
The topic of universal targets for $k$-dipath colourings is considered in Section \ref{kdipath:homsec}.
Here we offer improvements for the cases~\begin{math}t=3,4\end{math} of Theorem \ref{kdipath:2bounds}.
\begin{prop}
Let~\begin{math}t\geq 2 \end{math} be an integer and let $G$ be an oriented graph with~\begin{math}\chi_{2\text{-dip}} \leq t \end{math}.
\begin{itemize}
\item If $t \leq 3$, then $\chi_o(G) \leq 5$.
\item If $t \leq 4$, then $\chi_o(G) \leq 12$.
\end{itemize}
\end{prop}
\begin{proof}
Figure \ref{kdipath:3dptarget} shows $\mathcal{G}_{3}$,
except for arcs between the source vertices on the left and the sink vertices on the right.
By inspection, $\mathcal{G}_3$ admits a homomorphism to the tournament of order $5$ formed from a
copy of a directed $3$-cycle together with a universal source vertex and universal sink vertex.
This proves the first statement.
Let $H$ be he oriented graph obtained from
$\mathcal{G}_{4}$
by deleting all sources and all sinks.
Figure \ref{kdipath:4dptarget} gives a mapping of $H$ to the oriented graph of order $10$ shown.
Thus, this oriented graph, together with a universal source and universal sink vertex, is a homomorphic image of $\mathcal{G}_{4}$.
This proves the second statement.
\end{proof}
\begin{figure}
\begin{center}
\includegraphics[width=0.75\textwidth]{3dipathtarget}
\end{center}
\caption{The universal target for the family of oriented graphs with$\chi_{2\text{-dip}} \leq 3$.}
\label{kdipath:3dptarget}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.75\textwidth]{4dipathtarget}
\end{center}
\caption{A homomorphic image of the modified universal target for the family of oriented graphs with $\chi_{2\text{-dip}} \leq 4$.}
\label{kdipath:4dptarget}
\end{figure}
For~\begin{math}k \geq 2\end{math}, a $k$-dipath colouring of an oriented graph $G$
is a $2$-dipath colouring of $G$.
Thus, Theorem \ref{kdipath:2bounds} implies the following result for the $k$-dipath chromatic number.
\begin{cor} \label{kdipath:kdpupper}
If $G$ is an oriented graph with directed girth at least $k+1$, then
\begin{displaymath}\chi_o(G) \leq 2^{\chi_{k\text{-dip}}(G)} -1.\end{displaymath}
\end{cor}
Note, however, that by Theorem \ref{TargetHoms}, $\chi_{k\text{-dip}}$ can always be replaced by $\chi_{2\text{-dip}}$.
By contrast, the lower bound for $\chi_o(G)$ given in Theorem \ref{kdipath:2bounds} does not hold
if $\chi_{2\text{-dip}}$ is replaced by $\chi_{k\text{-dip}}$.
The oriented chromatic number of a directed path on $4$ vertices is $3$,
but its 4-dipath chromatic number is $4$.
\section{Homomorphisms and $k$-dipath Colouring} \label{kdipath:homsec}
We begin our study of homomorphisms and $k$-dipath colourings by considering oriented graphs without directed cycles.
The transitive tournament on $t$ vertices is denoted by $T_t$.
\begin{thm}[\cite{NP78}] \label{kdipath:acyclicmsw}
If $G$ is an acyclic oriented graph, then~\begin{math}G \rightarrow T_t \end{math} if and only if
there is no homomorphism of a directed path on $t+1$ vertices to $G$.
\end{thm}
A different way to phrase the condition in the theorem is that if~\begin{math}G \rightarrow T_t\end{math}, then $G$ has no directed cycles, and no
directed path of length $t$ or more.
\begin{cor} \label{kdipath:acyclicme}
If $G$ is an acyclic oriented graph and the longest directed path in $G$ has $t$ vertices, then~\begin{math}\chi_{k\text{-dip}}(G) = t \end{math} for all ~\begin{math}k \geq t > 1\end{math}.
\end{cor}
\begin{proof}
Suppose $k \geq t$.
Observe that since $G$ has a path on $t$ vertices, we have~\begin{math}\chi_{k\text{-dip}}(G) \geq t\end{math}. By Theorem \ref{kdipath:acyclicmsw}, there exists a homomorphism~\begin{math}\phi: G \to T_t\end{math}.
Since $T_t$ has no directed cycle,
any two vertices joined by a directed path must have different images.
If the vertices of $T_t$ are regarded as the colours $1, 2, \ldots, t$, then
$\phi$ is a $k$-dipath colouring of $G$.
Thus~\begin{math}\chi_{k\text{-dip}}(G) \leq t \end{math}.
\end{proof}
Since any oriented graph which admits a homomorphism to $T_t$ is acyclic, it follows that $T_t$ is a universal target
for $k$-dipath $t$-colouring of acyclic oriented graphs.
\begin{cor} \label{kdipath:acycliccor}
Let $k$ and $t$ be positive integers. An acyclic oriented graph $G$ has~\begin{math}\chi_{k\text{-dip}}(G) \leq t\end{math} if and only if $G$ admits a homomorphism to $T_t$.
\end{cor}
\begin{cor}
Let $k$ and $t$ be positive integers. An acyclic oriented graph $G$ has~\begin{math} \chi_{k\text{-dip}}(G) \leq t \end{math} if and only if $G$ has no directed path on at least $t+1$ vertices.
\end{cor}
Though not a direct analogue, Corollary \ref{kdipath:acyclicme} has a similar flavour to the early results on graph colourings of \cite{V62}, \cite{H65} , \cite{G67}, and \cite{R67}.
Observe that for $k \geq t$, the $k$-dipath chromatic number of $T_t$ equals $t$.
Thus, Corollary \ref{kdipath:acyclicme} states that the $k$-dipath chromatic number of $T_t$ is
an upper bound on the $k$-dipath chromatic number
of any oriented graph which admits a homomorphism to $T_t$.
The same statement holds if $T_t$ is replaced by any oriented graph with large enough directed girth.
\begin{thm} \label{kdipath:homgirth}
Let $G$ and $H$ be oriented graphs such that $H$ has directed girth at least $k+1$.
If~\begin{math}G \rightarrow H\end{math}, then~\begin{math}\chi_{k\text{-dip}}(G) \leq \chi_{k\text{-dip}}(H)\end{math}.
\end{thm}
\begin{proof}
Suppose~\begin{math}\phi: G \rightarrow H \end{math}.
Let $c$ be a $k$-dipath $t$-colouring of $H$.
Let~\begin{math}{c^\prime: V(G) \rightarrow \{1,2,3,\dots,t\}}\end{math}
be defined by~\begin{math}c^\prime(v) = c(\phi(v))\end{math}, for all ~\begin{math}v \in V(G)\end{math}.
We claim that this is a $k$-dipath $t$-colouring of $G$.
Let~\begin{math}u, v \in V(G)\end{math}.
Suppose that there is a directed path from $u$ to $v$ of length $\ell \leq k$.
Then there is a directed walk $W$ of length $k$ in $H$ from $\phi(u)$ to $\phi(v)$.
Since the directed girth of $H$ is $k+1$ or more, $W$ is a directed path.
Thus,~\begin{math}c(\phi(u)) \neq c(\phi(v))\end{math}.
Therefore~\begin{math}c^\prime(u) \neq c^\prime(v)\end{math}, and $c^\prime$ is a $k$-dipath colouring of $G$.
\end{proof}
We now describe a homomorphism model for $k$-dipath $t$-colouring of oriented graphs with directed girth at least $k+1$.
That is, for all positive integers $k$ and $t$, we will describe an oriented graph $\mathcal{G}_{k, t}$
with the property that an oriented graph $G$, with no directed cycle of length at most $k$,
has a $k$-dipath $t$-colouring if and only if
$G$ admits a homomorphism to $\mathcal{G}_{k,t}$ and, further,
the homomorphism can be used to find
a $k$-dipath $t$-colouring of $G$.
The graph $\mathcal{G}_{k,t}$ is a universal target for $k$-dipath $t$-colouring.
Once this model is in place, Theorem \ref{kdipath:homgirth} could be viewed as a direct consequence of
the fact that a composition of homomorphisms is a homomorphism; it arises in the proof for the model, however.
We begin by defining a special set of matrices. Let~\begin{math}k \geq 2 \end{math} be an integer.
Suppose that an oriented graph $G$, with directed girth at least $k+1$,
admits a $k$-dipath $t$-colouring, $c$.
For~\begin{math} x \in V(G)\end{math}, given $c(x)$, we want to encode information about the colours of vertices
in the $k$-in-neighbourhood of $x$ and $k$-out-neighbourhood of $x$.
We define the \emph{$k$-dipath colouring matrix of $x$ with respect to $c$} to be the~\begin{math}(2k-1) \times t \end{math}
zero-one matrix
$A_{x, c}(G)$
with rows indexed by
\begin{math}-(k-1), -(k-2), \dots,-1, 0, 1, \dots,(k-2), (k-1)\end{math}, columns indexed by~\begin{math}1, 2, 3, \dots, t \end{math}, and $(i, j)$-entry equal to 1
if and only if there exists a vertex~\begin{math}y \in V(G) \end{math} such that~\begin{math}c(y) = j \end{math}, and
\begin{itemize}
\item if~\begin{math}i \in \{-(k-1), -(k-2), \dots, -1\}\end{math}, then there is a directed path from $y$ to $x$ of length $i$;
\item if~\begin{math}i \in \{1, 2, \dots, (k-1)\}\end{math}, then there is a directed path from $x$ to $y$ of length $i$; and
\item if $i = 0$, then $x = y$.
\end{itemize}
When the graph $G$ is clear from the context, or unimportant in the discussion,
the $k$-dipath colouring matrix of $x$ with respect to $c$ is denoted by
$A_{x, c}$.
We illustrate the definition with an example.
Consider the colouring, $c$, given in Figure \ref{kdipath:example1}.
Let $x$ be the unique vertex such that~\begin{math}c(x) = 3 \end{math}. The $3$-dipath colouring matrix of $x$ is given by
\begin{center}
\begin{tabular}{c|cccc}
& $1$ & $2$ & $3$ & $4$ \\
\hline $-2$ &$0$ &$1$ &$0$ & $0$ \\
$-1$ & $1$ &$1$ &$0$ &$0$ \\
$0$& $0$ & $0$ &$1$ &$0$ \\
$1$ & $0$ & $0$ & $0$ & $1$ \\
$2$ & $0$ & $0$ & $0$ & $0$ \\
\end{tabular}
\end{center}
For example, the value, $1$, in
the entry~\begin{math}(-1, 2)\end{math}
arises because there is a vertex $w$ such that~\begin{math}c(w) = 2\end{math}
and there is a directed path of length $1$ from $w$ to $x$.
Let $\mathcal{A}_{k, t}$ denote the set of all possible $k$-dipath colouring matrices
over all $k$-dipath $t$-coloured oriented graphs with directed girth at least~\begin{math}k+1\end{math}. We note that $\mathcal{A}_{k, t}$ is necessarily finite, as each member is a~\begin{math}(2k-1)\times t\end{math} matrix with entries from $\{0,1\}$.
Since vertices at weak distance at most $k$ must be assigned different colours,
each element of $\mathcal{A}_{k, t}$ is a matrix that satisfies several conditions.
Suppose $A$ is the $k$-dipath colouring matrix of $x$ with respect to $c$.
If the colour of $x$ is $j$, then no vertex at weak distance at most $k$ from $x$ can
be assigned colour $j$.
Thus, the entry of $A$ in column $j$ and row 0 equals 1, and all other entries in column $j$ are zero.
If there is a directed path of length $i < k$ from $y$ to $x$, then no vertex
$w$ for which there is a directed path of length $k-i$ from $x$ to $w$ can have the same colour as $y$.
If the colour of $y$ is $j$, then the entry of $A$ in column $j$ and row $-i$ equals 1, and
the entries of $A$ in column $j$ and rows $0, 1, \ldots, k-i$
are all equal to zero.
Note that no assertion can be made about the entries of $A$ in column $j$ and rows~\begin{math}-(i-1), -(i-2), \ldots, -1\end{math}
because there is no guarantee that a vertex $v$ for which there is a directed path of length at most $i-1$
to $x$ lies on the directed path from $y$ to $x$.
Similar considerations apply to vertices $w$ for which there is a directed path of length $i < k$ from $x$ to $w$.
These conditions are formalized succinctly as follows.
\begin{obs} \label{kdipath:colstruc}
Let~\begin{math}A = [a_{ij}] \in \mathcal{A}_{k, t}\end{math}. Suppose ~\begin{math}a_{ij}=1\end{math}.
Then,
\begin{itemize}
\item if~\begin{math}i = 0 \end{math}, then~\begin{math}a_{\ell j} = 0\end{math} for all~\begin{math}\ell \neq i\end{math};
\item if~\begin{math}i < 0 \end{math}, then~\begin{math}a_{\ell j} = 0\end{math},~\begin{math} 0 \leq \ell \leq k-i \end{math}; and
\item if~\begin{math}i > 0 \end{math}, then~\begin{math}a_{-\ell j} = 0\end{math},~\begin{math}0 \leq \ell \leq k-i \end{math}.
\end{itemize}
\end{obs}
We now construct an oriented graph, $\mathcal{G}_{k, t}$, which is
a universal target for the family of $k$-dipath $t$-colourable oriented graphs of directed girth at least $k+1$.
The vertex set~\begin{math}V(\mathcal{G}_{k, t}) = \mathcal{A}_{k, t}\end{math}.
We describe the edge set informally at first, and then formally.
Suppose the matrices~\begin{math}A = A_{x, c}\end{math} and~\begin{math}B = B_{y, c^\prime}\end{math} are vertices of $\mathcal{G}_{k,t}$.
Let~\begin{math}c(x) = m_1\end{math} and~\begin{math}c^\prime(y) = m_2\end{math}.
Then~\begin{math}AB \in E(G)\end{math} if
\begin{enumerate}
\item $A$ encodes the fact that $x$ has an out-neighbour of colour $m_2$;
\item $B$ encodes the fact that $y$ has an in-neighbour of colour $m_1$;
\item if $A$ encodes the fact that a vertex of colour $m_3$ is joined to $x$ by a directed path of length~\begin{math}i < k\end{math},
then $B$ must encode the fact that a vertex of colour $m_3$ is joined to $y$ by a directed path of length~\begin{math}i+1\end{math};
\item if $B$ encodes the fact that $y$ is joined to a vertex of colour $m_3$ by a directed path of length~\begin{math}i < k\end{math},
then $A$ must encode the fact that $x$ is joined to a vertex of colour $m_3$ by a directed path of length~\begin{math}i + 1\end{math}.
\end{enumerate}
We now formally define the edge set of $\mathcal{G}_{k, t}$.
Let~\begin{math}A = [a_{ij}]\end{math} and~\begin{math}B = [b_{ij}]\end{math} be vertices of $\mathcal{G}_{k, t}$ (\textit{i.e.}, matrices in $\mathcal{A}_{k, t}$).
Then~\begin{math}AB \in E(\mathcal{G}_{k, t})\end{math} if, whenever~\begin{math}a_{0 m_1} = b_{0 m_2} = 1\end{math}, the following conditions all hold:
\begin{enumerate}
\item $a_{1 m_2} = 1$;
\item $b_{-1 m_1} = 1$;
\item if $0 < i < k-1$ and $a_{-i m_3} = 1$, then $b_{-(i-1) m_3} = 1$; and
\item if $0 < i < k-1$ and $b_{i m_3} = 1$, then $b_{(i+1) m_3} = 1$.
\end{enumerate}
We now establish some properties of $\mathcal{G}_{k, t}$.
\begin{lem} \label{kdipath:lem2}
The digraph $\mathcal{G}_{k, t}$ has directed girth at least $k+1$. In particular, it is an oriented graph.
\end{lem}
\begin{proof}
Let~\begin{math}A_1,A_2,\dots,A_\ell, A_1\end{math} be a directed cycle in $\mathcal{G}_{k, t}$. Suppose~\begin{math}\ell \leq k\end{math}, and $A_1$ has a $1$ in entry~\begin{math}(0,m_1)\end{math}. This implies $A_2$ has a $1$ in entry~\begin{math}(-1, m_1)\end{math} and a $1$ in entry~\begin{math}((\ell-1), m_1)\end{math}, contrary to Observation \ref{kdipath:colstruc}.
\end{proof}
\begin{lem} \label{kdipath:lem1}
For integers $t\geq 2$ and $k \geq 2$ the $k$-dipath chromatic number of $\mathcal{G}_{k, t}$ is at most $t$.
\label{AtMostT}
\end{lem}
\begin{proof}
Consider the colouring, $c$, given by $c(A) = m_1$,
where $m_1$ the is unique column of the matrix of $A$ for which the entry $(0,m_1)$ of $A$ is a $1$.
We claim that $c$ is a $k$-dipath colouring of $\mathcal{G}_{k,t}$.
Let~\begin{math} A_1A_2,\dots,A_\ell\end{math} be a directed path of length~\begin{math}\ell \leq k\end{math} in $\mathcal{G}_{k, t}$.
If there exists a pair of indices $i$ and $j$ such that~\begin{math}1 \leq i < j \leq k \end{math} and~\begin{math}c(A_i) =c(A_j)\end{math},
then $A_{i+1}$ has a $1$ in entry~\begin{math}(-1, c(A_i))\end{math} and a $1$ in entry~\begin{math}((j-(i+1)), c(A_i))\end{math}, contrary to Observation \ref{kdipath:colstruc}.
\end{proof}
\begin{thm}
Let~\begin{math}t\geq 2\end{math} and~\begin{math}k\geq 2\end{math} be integers. If $G$ is an oriented graph with directed girth at least $k+1$, then~\begin{math} \chi_{k\text{-dip}}(G) \leq t \end{math} if and only if~\begin{math}G \rightarrow \mathcal{G}_{k, t}\end{math}.
\label{HomModel}
\end{thm}
\begin{proof}
Let $G$ be an oriented graph with directed girth at least $k+1$. If~\begin{math}G \rightarrow \mathcal{G}_{k, t}\end{math}, then by Lemmas \ref{kdipath:lem2} and \ref{kdipath:lem1}, and by Theorem \ref{kdipath:homgirth},~\begin{math} \chi_{k\text{-dip}}(G) \leq t \end{math}.
Suppose ~\begin{math} \chi_{k\text{-dip}}(G) \leq t \end{math}.
Let $c$ be a $k$-dipath colouring of $G$ using $t$ colours.
Consider the mapping~\begin{math}\phi: V(G) \rightarrow V(\mathcal{G}_{k, t})\end{math},
where for all~\begin{math}v \in V(G)\end{math},~\begin{math}\phi(v)=A_v\end{math}, the $k$-dipath colouring matrix of $v$ with respect to $c$.
Let $uv$ be an arc of $G$.
We claim $A_uA_v$ is an arc of $\mathcal{G}_{k, t}$.
Suppose $A_u$ has a $1$ in entry~\begin{math}(0,m_1)\end{math}, and $A_v$ has a $1$ in entry~\begin{math}(0,m_2)\end{math} (\textit{i.e.},~\begin{math}c(u) = m_1\end{math} and~\begin{math}c(v) = m_2\end{math}).
To show that $A_u$ and $A_v$ are adjacent in $\mathcal{G}_{k, t}$ we must check that the four conditions
in the definition are satisfied:
\begin{enumerate}
\item Since~\begin{math}c(v) = m_2\end{math}, $A_u$ has a $1$ in entry~\begin{math}(1, m_2)\end{math}.
\item Since~\begin{math}c(u) = m_1\end{math}, $A_v$ has a $1$ in entry~\begin{math}(-1, m_1)\end{math}.
\item Suppose there exists $i>0$ such that $A_u$ has a $1$ in entry~\begin{math}(-i, m_3)\end{math}.
Since $c$ is a $k$-dipath colouring the entries of $A_u$ in column $m_3$ and rows~\begin{math} 0, 1, \ldots, k-i\end{math} are all equal to 0.
\item Suppose there exists $i>0$ such that $A_v$ has a $1$ in entry $(i, m_3)$.
Since $c$ is a $k$-dipath colouring, the entries of $A_v$ in column $m_3$ and rows~\begin{math}-(k-i), -(k-i+1), \ldots, -1, 0 \end{math} are all equal to zero.
\end{enumerate}
This proves the claim. Therefore~\begin{math}\phi:G \rightarrow \mathcal{G}_{k, t} \end{math} is a homomorphism.
\end{proof}
\begin{cor}
For integers $t\geq 2$ and $k\geq 2$,~\begin{math}\chi_{k\text{-dip}}(\mathcal{G}_{k, t}) = t \end{math}.
\end{cor}
\begin{proof}
We show ~\begin{math}\chi_{k\text{-dip}}(\mathcal{G}_{k, t}) \geq t \end{math}. The result then follows from Lemma \ref{AtMostT}.
Consider the transitive tournament $T_t$.
Assigning each vertex a different colour is a $k$-dipath $t$-colouring of $T_t$.
Hence, by Theorem \ref{HomModel}, there is a homomorphism~\begin{math}T_t \to \mathcal{G}_{k, t}\end{math}.
Since $T_t$ does not have a $k$-dipath $m$-colouring for~\begin{math}m < t \end{math}, by Theorem \ref{HomModel}, there is
no homomorphism~\begin{math}T_t \to \mathcal{G}_{k, m}\end{math}.
Since a composition of homomorphisms is a homomorphism, it follows that
\begin{math}\mathcal{G}_{k, m} \not\to \mathcal{G}_{k, t}$ when $m < t\end{math}.
Therefore,~\begin{math}\chi_{k\text{-dip}}(\mathcal{G}_{k, t}) > t-1\end{math}.
The result now follows.
\end{proof}
Our final result of this section shows that the homomorphism model captures three facts.
First, if~\begin{math}t < k \end{math}, then no oriented graph with directed girth at least $k+1$ and a $k$-dipath $t$-colouring
can have a directed path of length greater than $t$. Any such digraph is $t$-dipath $t$-colourable (as the proof shows).
Second, if~\begin{math}t \geq k \end{math}, then any $k$-dipath $t$-colouring of an oriented graph is also a $k$-dipath $(t+1)$-colouring,
and there exist oriented graphs with $k$-dipath chromatic number $t+1$.
Third, every $(k+1)$-dipath $t$-colouring of an oriented graph is a $k$-dipath $t$-colouring.
\begin{thm}
Let ~\begin{math}k \geq 2\end{math} and~\begin{math}t \geq 2\end{math} be integers.
Then
\begin{enumerate}
\item if $t \leq k$, then~\begin{math}\mathcal{G}_{k, t} \to \mathcal{G}_{t, t}\end{math};
\item~\begin{math}\mathcal{G}_{k, t} \to \mathcal{G}_{k, (t+1)}\end{math} and~\begin{math} \mathcal{G}_{k, (t+1)} \not\to \mathcal{G}_{k, t} \end{math};
\item if $t > k$, then~\begin{math}\mathcal{G}_{(k+1), t} \to \mathcal{G}_{k, t}\end{math} and~\begin{math}\mathcal{G}_{k, t} \not\to \mathcal{G}_{(k+1), t} \end{math}.
\end{enumerate}\label{TargetHoms}
\end{thm}
\begin{proof}
We first prove (1).
Suppose~\begin{math}t \leq k\end{math}.
Then no oriented graph with directed girth at least $k+1$ and a $k$-dipath $t$-colouring has a
directed path of length greater than $t$.
By Theorem \ref{kdipath:acyclicmsw} there is a homomorphism of $G$ to the transitive tournament $T_t$.
In particular,~\begin{math}\mathcal{G}_{k, t} \to T_t \end{math}.
But, by Theorem \ref{HomModel},~\begin{math}T_t \to \mathcal{G}_{t, t}\end{math}.
Therefore,~\begin{math}\mathcal{G}_{k, t} \to \mathcal{G}_{t, t}\end{math}.
We now prove (2).
It is clear that the subgraph of $\mathcal{G}_{k, (t+1)}$ induced by the vertices (colouring matrices) in which
every entry in column $t+1$ is zero (\textit{i.e.}, $k$-dipath $(t+1)$-colourings in which colour $t+1$ is never used) is
isomorphic to $\mathcal{G}_{k, t}$.
Thus, $\mathcal{G}_{k, t} \to \mathcal{G}_{k, (t+1)}$.
On the other hand, the transitive tournament on $t+1$ vertices has a homomorphism to
$\mathcal{G}_{k, (t+1)}$ but not to $\mathcal{G}_{k, t}$.
Therefore, ~\begin{math} \mathcal{G}_{k, (t+1)} \not\to \mathcal{G}_{k, t} \end{math}.
Finally, we prove (3).
Suppose $t > k$.
Since a $(k+1)$-dipath $t$-colouring is a $k$-dipath $t$-colouring, it follows from Theorem \ref{HomModel} that
\begin{math}\mathcal{G}_{(k+1), t} \to \mathcal{G}_{k, t}\end{math}.
To see the second statement, note that a directed cycle of length $k+1$ has a homomorphism to $\mathcal{G}_{k, t}$ but,
by Lemma \ref{kdipath:lem2}, no homomorphism to $\mathcal{G}_{(k+1), t}$.
Therefore,~\begin{math}\mathcal{G}_{k, t} \not\to \mathcal{G}_{(k+1), t}\end{math}.
\end{proof}
Directed graphs, $G$ and $H$ are called \emph{homomorphically equivalent} if there are
homomorphisms $G \to H$ and $H \to G$.
If $G$ and $H$ are homomorphically equivalent, then a directed graph admits a
homomorphism to $G$ if and only if it admits a homomorphism to $H$.
A directed graph is a \emph{core} if it is not homomorphically equivalent to any proper subgraphs.
Every directed graph $G$ has a minimal subgraph $H$ which is a core, and to which $G$ is homomorphically equivalent; $H$ is unique up to isomorphism, and is called \emph{the core of $G$} (see \cite{F82} and \cite{W82}).
We note that the core, $H$, of $G$ is an induced subgraph of $G$ because, by minimality,
any homomorphism~\begin{math}G \to H\end{math} must map $H$ onto itself.
\begin{cor}
Let $t$ and $k$ be positive integers. If $t \leq k$, then the core of $\mathcal{G}_{k, t}$ is $T_t$.
\label{Core}
\end{cor}
\begin{proof}
Note that $T_t$ is a core. The existence of the required homomorphism is noted in the proof of statement (1) in Theorem \ref{TargetHoms}.
\end{proof}
\section{Complexity of $k$-dipath Colourings} \label{kdipath:comsec}
In this section we consider the following decision problem.
\noindent \underline{$k$-DIPATH $t$-COLOURING}\\
Instance: an oriented graph, $G$.\\
Question: does $G$ have a $k$-dipath $t$-colouring?
The dichotomy theorem stated below covers the cases where $k = 2$.
We shall find a similar result for all remaining cases.
\begin{thm} [\cite{MAYO12, KMY09}]
Let $t \geq 1$ be a fixed integer. If~\begin{math} t \leq 2 \end{math}, then $2$-DIPATH $t$-COLOURING is Polynomial. If~\begin{math} t \geq 3 \end{math}, then $2$-DIPATH $t$-COLOURING is NP-complete.
\label{2DPDichotomy}
\end{thm}
Given a simple graph $G$, we construct an oriented graph, $H_{k,t}$ (\begin{math}t > k \geq 3\end{math}), such that~\begin{math}\chi_{k\text{-dip}}(H_{k,t}) \leq t\end{math} if and only if~\begin{math}\chi(G) \leq t \end{math}. Let $G$ be a simple graph, and let $\tilde{G}$ be an arbitrary acyclic orientation of $G$.
Corresponding to each vertex~\begin{math}v \in V(\tilde{G})\end{math} the oriented graph $H_{k,t}$ contains:
\begin{enumerate}
\item vertices $v_\text{in}$ and $v_\text{out}$;
\item a transitive tournament on~\begin{math}t-k+1\end{math} vertices with source vertex $s_v$ and sink vertex $t_v$:
\item vertices~\begin{math}v^\prime_1, v^\prime_2, \ldots, v^\prime_{k-2}\end{math};
\item the directed path~\begin{math}t_v,v^\prime_1, v^\prime_2, \ldots, v^\prime_{k-2},v_\text{in}\end{math},
which has length $k-1$; and
\item an arc $v_\text{out}s_v$.
\end{enumerate}
For each arc~\begin{math}uv \in E(\tilde{G})\end{math}, the oriented graph $H_{k,t}$ is augmented by adding
the arc~\begin{math}u_\text{out}v_\text{in}\end{math}.
This completes the construction of $H_{k,t}$. It can clearly be carried out in polynomial time.
We note that since $\tilde{G}$ has no directed cycles, by construction the same is true of $H_{k,t}$.
Hence $H_{k,t}$ (formally) has directed girth at least $k+1$.
We also note that each vertex $v_\text{out}$ has in-degree 0 and and each vertex
$v_\text{in}$ has out-degree $0$.
\begin{obs}
\begin{math}\chi_{k\text{-dip}}(H_{k,t}) \geq t\end{math}.
\end{obs}
\begin{proof}
For any vertex~\begin{math}v \in V(G)\end{math}, observe that the subgraph of $H_{k,t}$ induced by the~\begin{math}t-k+1\end{math} vertices of the transitive tournament corresponding to $v$, together with the vertices~\begin{math}v^\prime_1, v^\prime_2, \ldots, v^\prime_{k-2},v_\text{in}\end{math}
is a $k$-dipath clique on $t$ vertices.
\end{proof}
\begin{obs} \label{kdipath:obslift2}
Let~\begin{math}t > k \geq 3\end{math}. If $H_{k,t}$ has a $k$-dipath $t$-colouring, then for every $v \in V(G)$ and every $k$-dipath $t$-colouring, $c$, of $H_{k,t}$, ~\begin{math}c(v_\text{out}) = c(v_\text{in})\end{math}.
\end{obs}
\begin{proof}
For any $v \in V(G)$, observe that the subgraph of $H_{k,t}$ induced by the vertex $v_\text{out}$, the~\begin{math}t-k+1\end{math} vertices of the transitive tournament corresponding to $v$ and the vertices~\begin{math}v^\prime_1, v^\prime_2, \ldots, v^\prime_{k-2}\end{math} is a $k$-dipath clique on $t$ vertices.
Since the subgraph of $H_{k,t}$ induced by the $t-k+1$ vertices of the transitive tournament corresponding to $v$, together with the vertices~\begin{math}v^\prime_1, v^\prime_2, \ldots, v^\prime_{k-2}\end{math} and $v_\text{in}$ is also $k$-dipath clique on $t$ vertices, and $c$ is a $k$-dipath $t$-colouring, it follows that~\begin{math}c(v_\text{out}) = c(v_\text{in})\end{math}.
\end{proof}
\begin{lem} \label{kdipath:nplift2}
If $G$ is a simple graph and $H_{k,t}$ is constructed from $G$ as above, then, for all~\begin{math}
t > k \geq 3
\end{math}, the graph $G$ is $t$-colourable if and only if $H_{k,t}$ has a $k$-dipath $t$-colouring.
\end{lem}
\begin{proof}
Suppose $H_{k,t}$ has a $k$-dipath $t$-colouring, $c$.
Let~\begin{math}
\phi: V(G) \rightarrow \{1,2,3, \dots, t\}
\end{math} be defined by~\begin{math}
\phi(v) = c(v_\text{in})
\end{math}.
We claim that $\phi$ is a proper colouring of the graph $G$.
Suppose~\begin{math}
ab \in E(\tilde{G})
\end{math}.
Then, ~\begin{math}
a_\text{out}b_\text{in} \in E(H_{k,t})
\end{math}, and so
~\begin{math}
c(a_\text{out}) \neq c(b_\text{in}).
\end{math}
By Observation~\ref{kdipath:obslift2},~\begin{math}
c(b_\text{out}) = c(b_\text{in}).
\end{math}
Therefore,~\begin{math}
c(a_\text{in}) \neq c(b_\text{in})
\end{math}, and~\begin{math}\phi(a) \neq \phi(b)\end{math}.
This proves the claim. Hence $G$ is $t$-colourable.
Suppose now that $G$ has a $t$-colouring, $\phi$.
We construct a $k$-dipath $t$-colouring, $c$, of $H_{k,t}$.
For each vertex~\begin{math}
v \in V(G)
\end{math}, set ~\begin{math}
c(v_\text{out}) = c(v_\text{in}) = \phi(v)
\end{math}.
We claim that $c$ can be completed to a $k$-dipath $t$-colouring of $H_{k,t}$.
Since, for every~\begin{math}v \in V(G)\end{math}, if~\begin{math}
\phi(v_\text{out})= i
\end{math}, then vertices of the transitive tournament corresponding to $v$ together with the vertices~\begin{math}
v^\prime_1, v^\prime_2, \ldots, v^\prime_{k-2}
\end{math} can be assigned colours from the set~\begin{math}\{1,2,3 \dots, t\} \setminus \{i\}\end{math} so that the resulting colouring has the property that no two vertices at weak distance at most $k$ are assigned the same colour.
This proves the claim.
\end{proof}
\begin{thm}
Let $t$ and~\begin{math}
k \geq 3
\end{math} be fixed positive integers.
When restricted to instances of directed girth at least~\begin{math}
k+1
\end{math}, the decision problem
\emph{$k$-DIPATH $t$-COLOURING} is NP-complete if~\begin{math}
t > k
\end{math}, and Polynomial if ~\begin{math}
t \leq k
\end{math}.
\label{Dichotomy1}
\end{thm}
\begin{proof}
The problem is clearly in NP.
If~\begin{math}
t > k
\end{math} then NP-completeness follows from Lemma \ref{kdipath:nplift2}.
Suppose~\begin{math}
t \leq k
\end{math}.
By Corollary \ref{Core}, an oriented graph $G$ with directed girth at least $k+1$ has a $k$-dipath $t$-colouring if
and only if it admits a homomorphism to the transitive tournament $T_t$.
Homomorphism to $T_t$ can be checked in polynomial time as shown by \cite{BHG88}.
The result now follows.
\end{proof}
We now show that if the girth restriction is removed, then the dichotomy changes.
\begin{thm}
Let $k$ and $t$ be positive integers.
If~\begin{math} t \leq 2 \end{math}, then \emph{$k$-DIPATH $t$-COLOURING} is Polynomial.
If~\begin{math} t \geq 3 \end{math}, then \emph{$k$-DIPATH $t$-COLOURING} is NP-complete.
\end{thm}
\begin{proof}
Suppose first that~\begin{math} t \leq 2 \end{math}.
If~\begin{math}
k = 1
\end{math}, then the condition that vertices joined by a directed path of length at most $k$ must be assigned different colours
is the same as the condition that adjacent vertices must be assigned different colours.
Hence, a digraph $G$ has a $k$-dipath $t$-colouring if and only if the underlying graph of $G$ has a $t$-colouring.
The latter problem is Polynomial.
If $k=2$, $k$-dipath $t$-colouring is Polynomial by Theorem \ref{2DPDichotomy}.
For~\begin{math}
k \geq 3
\end{math}, if $G$ has a directed path of length at least 3 or a directed 3-cycle (which is easy to check),
then, since~\begin{math}
t \leq 2
\end{math}, it has no $k$-dipath $t$-colouring.
Otherwise, ($G$ has no directed path of length greater than 2),
a $k$-dipath $t$-colouring is a $2$-dipath $t$-colouring, which is Polynomial.
Now suppose~\begin{math}
t \geq 3
\end{math}.
If~\begin{math}
k=2
\end{math}, then $k$-dipath $t$-colouring is NP-complete by Theorem \ref{2DPDichotomy}.
Hence, assume~\begin{math} k \geq 3 \end{math}.
If $t > k$, the result follows from Theorem \ref{Dichotomy1}.
Hence we may also assume $t \leq k$.
Similarly as above, if $t < k$ then
no oriented graph with a directed path of
length greater than $t$ (which is easy to check), can have a
a $k$-dipath $t$-colouring.
Otherwise ($G$ has no directed path of length greater than $t$),
a $k$-dipath $t$-colouring is a $t$-dipath $t$-colouring.
Thus, we may further assume that~\begin{math}
t = k \geq 3
\end{math}.
We have previously noted that the problem belongs to NP.
The transformation is from $t$-colouring.
Suppose an instance of $t$-colouring, a simple, undirected graph $G$ is given.
We will replace each edge of $G$ by the oriented graph which we now construct.
Let $C$ be the directed cycle of length $t$ with vertex sequence~\begin{math}
v_0, v_1, \ldots, v_{t-1}, v_0
\end{math}.
Add four new vertices~\begin{math}
x_0, x_1, y_0, y_1
\end{math}, and arcs to make the directed paths
\begin{math}
x_0, x_1, v_1
\end{math} and~\begin{math}
y_0, y_1, v_2
\end{math}.
This completes the construction of $F$.
Observe that $x_0$ is joined to every vertex of $C$ except $v_0$ by a directed path of length at most $t$.
Every vertex of $C$ is assigned a different colour in a $k$-dipath $t$-colouring of $F$.
Hence, the vertices $x_0$ and $v_0$ must be assigned the same colour.
Similarly, the vertices $y_0$ and $v_1$ must be assigned the same colour.
In particular, $x_0$ and $y_0$ must be assigned different colours.
Furthermore, any assignment of two different colours to $x_0$ and $y_0$ can be extended to a
$k$-dipath $t$-colouring of $F$.
Construct an oriented graph $G^\prime$ from $G$ as follows.
Replace each edge $xy$ of $G$ with a copy $F_{xy}$ of $F$ by
identifying $x_0$ with $x$ and $y_0$ with $y$.
The construction can clearly be carried out in polynomial time.
Observe that each vertex in~\begin{math}
V(G) \cap V(G^\prime)
\end{math} has in-degree 0.
We claim that $G$ is $t$-colourable if and only if $G^\prime$ has a $k$-dipath $t$-colouring.
Suppose that a $t$-colouring of $G$ is given.
For each copy of $F$ in $G^\prime$, this assignment gives different colours to $x_0$ and $y_0$.
By our earlier observation, this assignment can be extended to a $k$-dipath $t$-colouring of $F$.
Since each vertex in~\begin{math}
V(G) \cap V(G^\prime)
\end{math} has in-degree 0, there is no directed path of positive length
joining vertices in different copies of $F$, this results in a $k$-dipath $t$-colouring of $G^\prime$.
Now suppose a $k$-dipath $t$-colouring of $G^\prime$ is given.
By our observation on colourings of $F$, any two vertices which are adjacent in $G$ are assigned
different colours. Hence restricting this colouring to $V(G)$ gives a $t$-colouring of $G$.
This completes the proof.
\end{proof}
The problem of deciding whether the $k$-th power of a graph $G$ is $t$-colourable is known to be
NP-complete (see \cite{L95} and \cite{M83}).
The results above imply that, if~\begin{math}
t > k
\end{math}, the problem of deciding whether the underlying graph of the
$k$-th power
of an oriented graph $G$ (\textit{i.e.}, two vertices are adjacent if and only if they are at weak distance at most $k$)
is $t$-colourable is NP-complete, even when restricted to powers of oriented graphs with directed girth at least $k+1$,
and also that, if~\begin{math}
t = k \geq 3
\end{math},
the problem of deciding whether the underlying graph of the $k$-th power
of an oriented graph $G$ is $t$-colourable is NP-complete.
\bibliographystyle{abbrvnat}
|
\section{Introduction}
Transportation, electrical and telecommunication networks, to name a few, have become fundamental for the proper functioning of the modern world. For that reason it has become of extreme importance that these systems remain operative. However these systems are prone to failure due to malfunctions, catastrophes or attacks.
All these structures can be studied through complex networks by representing the components of the structure by nodes and interactions among the components by edges.
Since their correct functioning requires that the network is properly connected it is of great importance to study their ability to resist failures (either unintentional or targeted attacks). This ability is called robustness.
In this work, we focus on the scenario of targeted attacks by an adversary. We notice that this scenario corresponds to an upper bound on the damage of (unintentional or intentional) failures.
We consider that an ``adversary'' should plan a greedy strategy aiming to maximize the damage with minimum number of strikes.
Thus, in this article we discuss the performance of attacks based on the edge betweenness centrality metric \cite{bersano2012metrics} over the Internet Backbone (the network formed by Internet exchange points, IXP), and its correlation with what users perceive from such networks if they want to receive content from the major content provider (Goo\-gle), with a metric called Go-Index.
We consider the Go-index measure with contains different supply network measures whose provider is Goo\-gle.
Just like economy uses the price of the Big Mac as a way of measuring purchasing power parity for its wide availability,
here we measure the ability of the nodes to remain connected with Google for the same reason.
The idea to consider an IXP-based network as ``the Internet backbone'' is not new, previously has been used, for instance, as part of the ``internet core'' to study the inter-AS traffic patterns and an evolution of provider peering strategies \cite{labovitz2011internet}, for optimize the content delivery from Google via direct paths \cite{chiu2015we} (in this case filtering IXP ASs because they facilitate connectivity between peers), and the Internet Backbone Market \cite{buccirossi2005competition}. The importance of our study is based in with the use of the IXP network as a model for ``backbone Internet'', we can have a good approximation of Internet physical robustness. As far as the authors knowledge, this is the first time that the robustness of IXP network is studied.
The article is organized as follows, next section pre\-sents related work, followed by the methodology for building the IXP network, the attacking strategy using betweenness centrality and Go-Index (Sections \ref{bet} and \ref{go}). Conclusions are presented in Section \ref{conclusions}.
\section{Related work}
\label{relatedwork}
To study the robustness of a network its evolution against failure must be analyzed. On real world situations networks may confront random failures as well as targeted attacks.
For the latter, two main categories of attacking strategies have been defined: simultaneous and sequential attacks \cite{31holme2002attack}. Simultaneous attacks chose a set of nodes and remove them all at once while sequential attacks chose a node to remove and given the impact of this removal it chooses another node, proceeding in turns.
Targeted attacks have been thoroughly studied to analyze network robustness. Holme \textit{et al.} tested node degree and betweenness centrality strategies using simultaneous and sequential attacks. On \cite{33trajanovski2013robustness} simultaneous attacks based on network centrality measures and random attacks were studied.
The stability of scale-free network under degree-based attacks was studied on \cite{35yehezkel2012degree}. Experimental results are shown in \cite{iyer2013attack} who also consider sequential and simultaneous attacks as well as centrality measure strategies. To get closer to a real world strategy scenario on \cite{36gallos2006attack,37wu2007vulnerability} studied the resilience of scale-free networks to a variety of attacks with different amounts of information available to the attacker about the network.
On \cite{38booker2012effects} the impact of the effectiveness of the attack under observation error was studied. More recently \cite{ventresca2015network} studied sequential multi-strategy attacks using multiples robustness measures including the \textit{Unique Robustness Measure} ($R$-index)~\cite{schneider2011mitigation}.
The attacking strategies have been analyzed through the lens of an attacker, an adversary whose objective is to perform the most damage possible to the network. However the case of an adversary with a limited amount of strikes remain to be tested. Here this case is analyzed and an option to measure the robustness of a network in these circumstances is presented.
On \cite{32estrada2006network} was found that targeted attack can be more effective when they are directed to bottlenecks rather than hubs. On \cite{miuz} authors present partial values of $R$-index while nodes are disconnected, showing the importance of a well chosen robustness metric for performing the attacks.
The idea of planning a ``network attack'' using centrality measures has captured the attention of researchers and practitioners nowadays. For instance, Sterbenz et al.~\cite{sterbenz2011modelling} used bet\-ween\-ness-centrality (\textit{bcen}) for planning a network attack, calculating the \textit{bcen} value for all nodes, ordering nodes from higher to lower \textit{bcen}, and then attacking (discarding) those nodes in that order. They have shown that disconnecting only two of the top \textit{bcen}-ranked nodes, their packet-delivery ratio is reduced to $60\%$, which corresponds to~$20\%$ more damage than other attacks such as random links or nodes disconnections, tracked by link-centrality and by node degrees.
A similar approach and results were presented by {\c{C}}etin\-kaya et al.~\cite{ccetinkaya2013modelling}. They show that after disconnecting only $10$ nodes in a network of $100+$ nodes the packet-delivery ratio is reduced to $0\%$. Another approach, presented as an improved network attack \cite{rak2010survivability, sydney2010characterising}, is to recalculate the betweenness-centrality after the removal of each node \cite{holme2002attack,molisz2006end}. They show a similar impact of non-recalculating strategies but discarding sometimes only half of the equivalent nodes.
In the study of resilience after edge removing, Rosen\-kratz et al. \cite{rosenkrantz2009resilience} study backup communication paths for network services defining that a network is ``\textit{k-edge-failure-resilient if no matter which subset of k or fewer edges fails, each resulting subnetwork is self-sufficient}'' given that ``\textit{the edge resilience of a network is the largest integer k such that the network is k-edge-failure-resilient}''.
For a better understanding of network attacks and strategies, see \cite{holme2002attack,molisz2006end,rak2010survivability, sydney2010characterising}.
\section{Building the Internet backbone graph}
\label{graph}
\begin{figure}[ht!]
\centering
\includegraphics[width=\linewidth]{graph.png}
\caption{Peering Graph}
\label{fig:graph}
\end{figure}
Internet peering is the contract (formal or informal) between two autonomous systems (AS) that agree to exchange traffic (and traffic routes) through a physical link. In \cite{Dhamdhere:2010:IFM:1921168.1921196} authors present that ``\textit{The core of the Internet is a multi-tier hierarchy of Transit Providers (TPs). About 10-20 tier-1 TPs, present in many geographical regions, are connected with a clique of peering links. Regional (tier-2) ISPs are customers of tier-1 TPs. Residential and small business access (tier-3) providers are typically customers of tier-2 TPs}''. Therefore, it is natural to think that the peering network is a coarse grained approximation of the Internet itself. Thus, we used it to model Internet for studying its robustness.
From \url{peeringdb.com} we collected the autonomous systems (AS) from every Internet Exchange Point (IXP) and defined them as graph nodes. Therefore, an AS could belong to different IXPs and an IXP could have multiple ASs. Then, we connected the nodes if they fulfill at least one of the following rules:
\begin{itemize}
\item Physically linked ASs that exchange traffic
\item ASs belonging to the same IXP
\item ASs belonging to the same facility
\end{itemize}
Where we considered IXP as public peering and facility as private peering.
Figure~\ref{fig:graph} shows the resulting Graph, which has $522$ nodes and $14,294$ edges (orange edges are public peering, blue are private peering, and green are direct network connection). The resultant network has a well connected core network with some isolated nodes at the edges. In Figure \ref{fig:degree} we present a degree distribution of nodes for our IPX Graph.
\begin{figure}[ht!]
\centering
\includegraphics[width=\linewidth]{degree.png}
\caption{Degree distribution of IPX autonomous systems}
\label{fig:degree}
\end{figure}
\section{Studying the robustness of the Internet backbone}
\label{bet}
Betweenness centrality is a metric that determines the importance of an edge by looking at the shortest paths between all of the pairs of remaining nodes. Betweenness has been studied as a resilience metric for the routing layer~\cite{smith2011network} and also as a robustness metric for complex networks \cite{iyer2013attack} and for internet autonomous systems networks~\cite{mahadevan2006internet} among others. Betweenness centrality has been widely studied and standardized as a comparison base for robustness metrics, thus in this study it will be used for performance comparison.
If we plan a network attack by disconnecting edges with a given strategy, it is widely accepted to compare it against the use of betweenness centrality metric, because the latter reflects the importance of an edge in the network \cite{iyer2013attack}. These attack strategies are compared by means of the \textit{Unique Robustness Measure} ($R$-index)~\cite{schneider2011mitigation}, defined as:
\begin{equation}
R = \frac{1}{N}\sum_{Q=1}^{N} {s(Q)},
\end{equation}
where $N$ is the number of nodes in the network and $s(Q)$ is the fraction of nodes in the largest connected component after disconnecting edges using a given strategy. Therefore, the higher the $R$-index, the better in terms of robustness.
Instead of just comparing the robustness, after the removal of all of the edges, we would like to study the behavior of the attacks after only a few strikes. To do so, we define a variant of the $R$-index which takes into account only the first $n$ strikes of an attack. Thus, for a simultaneous attack (where the nodes are ranked by a metric only once at the beginning), the $R_n$-index is defined as:
\begin{equation}
R_n = \frac{1}{n}\sum_{Q=1}^{n} {s(Q)}.
\end{equation}
That is, the area under the curve produced by the largest connected component ratio (compared to the whole network) until $n$.
For a sequential attack, the order of node disconnection is recomputed after each disconnection. Similar to the $R$-index, notice that the lower the $R_n$-index, the more effective the attack is, since that gives us a higher reduction of robustness.
Results are shown in Figure \ref{fig:bet-attack}. We tested sequential attacks: At each strike, the next edge to disconnect was the one with the highest betweenness value. The figure shows the behavior of the $R_n$-index in our IPX Graph. The strategy proves to be very effective in attacks, disconnecting half of the network removing only 20\% of the edges, more than 30\% of the nodes after removing 10\% of edges, and 10\% of nodes after 1\% of edges.
\begin{figure}[th!]
\centering
\includegraphics[width=\linewidth]{bet-attack.png}
\caption{\% of the largest component size compared to the original network. In the plot, $R_{20\%}$-index is marked in cyan.}
\label{fig:bet-attack}
\end{figure}
\section{The Go-Index: If you cannot see Google, you are not connected}
\label{go}
Google has been reported as having the 40\% of Internet traffic in 2013\footnote{See the Forbes article at \url{http://goo.gl/aHdeiN}}. Thus, given the Google peering policies\footnote{See \url{https://peering.google.com/#/options/peering}.} and knowing the Google policies to interconnect their datacenters \cite{jain2013b4}, we can also study Internet as an information supply network, adapting to Internet the metrics presented in \cite{zhao2011achieving}:
\begin{enumerate}
\item \textbf{Supply Availability} (SAR): The percentage of ASs that have access to Google from at least one of its ASs.
\item \textbf{Network Connectivity} (NetCON): The number of ASs in the largest functional sub-network, in which there is a path between any pair of ASs and there exist at least one of the Google ASs.
\item \textbf{Best Delivery Efficiency} (BDE): The reciprocal of the average of each demand AS's shortest supply path length to its nearest Google AS. Values go from $1$ (everyone is connected directly with a Google AS) to $0$ (there are only Google ASs in the network).
\item \textbf{Average Delivery Efficiency} (ADE): The average inverse supply path length for all possible \{Network AS,Google AS\} pairs, adjusted by a weighting factor for each path (in our study all Google ASs have the same importance). In this case, values go from $2$ (everyone is connected to both Google ASs directly) to $0$ (nobody is connected with Google ASs).
\end{enumerate}
Notice that Google delegated some services at ISPs autonomous systems \cite{calder2013mapping}, nevertheless they must eventually connect with Google backbone for updating. We called the tuple \{1,2,3,4\} the \textit{Go-Index}, that is, the supply network measures whose provider is Google.
Using the same attack strategy from previous section, we calculated the Go-Index after edge removal (removing the edge with higher betweenness). The results are presented in Figure \ref{fig:sar} (Supply Availability), \ref{fig:netcon} (Network Connectivity), \ref{fig:bde} (Best Delivery Efficiency), and \ref{fig:ade} (Average Delivery Efficiency).
The first two metrics are very related with the largest connected component ratio, which in this study include at least one of the two Google ASs, that is, AS15169 and AS36040 (marked in pink at Figure \ref{fig:graph}, the former in the center and the latter in the edge of the network). Therefore, no new information are provided by those metrics.
The following two metrics are based in ``how far is Google from a given autonomous system'', but BDE considers only the connected component that includes Google ASs, by itself it has no information about the isolated portion of the network that cannot reach the Google ASs, therefore BDE improves when large subnets are disconnected from the core network that contains the Google ASs.
Note that for a user inside that core network the main content provider always exists and there are no indications that the network is falling apart (or losing half of its members, as produced by eliminating 20\% of its edges), this can be appreciated through BDE. Nevertheless, the big picture is different: after having only the 5\% of the network disconnected one of Google ASs is isolated, showing that from this point the supply network is only maintained by AS15169. The perception error is corrected when ADE is used because it includes all nodes in its calculus.
Then, the Go-Index accomplish its objectives, reflecting both infrastructure (SAR+NetCON+ADE) and user perception (BDE+ADE), for Internet robustness studies.
\begin{figure}[th!]
\centering
\includegraphics[width=\linewidth]{sar.png}
\caption{Supply Availability Ratio}
\label{fig:sar}
\end{figure}
\begin{figure}[th!]
\centering
\includegraphics[width=\linewidth]{netcon.png}
\caption{Network Connectivity}
\label{fig:netcon}
\end{figure}
\begin{figure}[th!]
\centering
\includegraphics[width=\linewidth]{bde.png}
\caption{Best Delivery Efficiency}
\label{fig:bde}
\end{figure}
\begin{figure}[th!]
\centering
\includegraphics[width=\linewidth]{ade.png}
\caption{Average Delivery Efficiency}
\label{fig:ade}
\end{figure}
\section{Conclusions and Future Work}
\label{conclusions}
In this paper we have presented how robust the Internet backbone (the peering AS network) would be
if an adversary can choose wisely which physical link he will cut (or if a very unlucky accident happen).
Following the recommendations, the chosen one would be the edge with higher
betweenness centrality value.
Using this strategy the adversary is capable of disconnecting half of the network by removing only 20\% of the edges, more than 30\% of the nodes after removing 10\% of the edges, and 10\% of nodes after 1\% of the edges as we have seen with the values of $R_{n}$ index in Section \ref{bet}.
Furthermore, we consider the Internet as a (information) supply network,
and considering that Google is the main Internet content provider,
and propose to study the Internet backbone with the Go-Index
(the adaptation of metrics presented in \cite{zhao2011achieving} for supply networks).
The metrics used were Supply Availability, Network Connectivity (both highly correlated with the larger connected component ratio),
Best Delivery Efficiency, and Average Delivery Efficiency.
If only Best Delivery Efficiency is considered, the network can be declared robust
because a user located inside the core network, which is always connected with a Google AS,
will not perceive that the network is being disconnected. The Go-index will correct that perception since it also contains the Average Delivery Efficiency which includes all nodes in its calculus.
As future work we plan to apply similar studies to other Internet infrastructures,
such as country-based fiber interconnection, submarine Internet cables, etc.
Also, we plan to improve the metrics for robustness reflecting both
the infrastructure part (such as Rn index) and the user perception
(such as Best Delivery Efficiency/Average Delivery Efficiency).
|
\section{Introduction}
\label{sec:introduction}
Mathematical models have always played an import role science. They are a key way of summarizing and exchanging complex information, evolve with the acquisition of new knowledge, and motivate starting points for future explorations. Driven in part by big data and expanding computational capabilities, models with many parameters (in some cases thousands or more) are increasingly common. These models create new challenges and raise new questions ranging from the technical to the profound. Large-scale models may introduce numerical instability, be computationally expensive, or be challenging to fit to data. Methods to manage model complexity are a central concern in modern science.
Model reduction is the problem of finding approximate, simpler models that capture the same behavior as the original model. Simple models have many motivations. They reduce computational cost and avoid technical numerical problems. They avoid statistical challenges associated with model calibration and over-fitting. They are easier to interpret and so may also help reveal the important mechanisms that drive a particular behavior.
Model reduction is a vast topic and a comprehensive review of the literature would be impossible. Although there is not a one-size-fits all solution, there are a host of available methods that have found consistent success within their appropriate scope. Of particular note are methods that exploit scale separation, particularly singular perturbation methods\cite{saksena1984singular,kokotovic1999singular,naidu2002singular}, lumping or clustering methods\cite{wei1969lumping,liao1988lumping,huang2005systematic}, and many others\cite{conzelmann2004reduction,antoulas2005approximation}. Of particular note are methods developed by the control and chemical kinetics communities focused on dynamics systems\cite{moore1981principal,saksena1984singular,kokotovic1999singular,naidu2002singular,antoulas2005approximation,lee2010multi,dullerud2013course}. In physics, techniques such as mean-field theory and the Renormalization Group (RG) are also powerful techniques\cite{goldenfeld1992lectures,zinn2007phase}.
Typically, choosing a model reduction algorithms is determined by the functional form of the model to be reduced and the motivation behind the reduction. Many methods are automatic methods that give ``black box'' approximations that are computationally efficient surrogates for the original model. One possible motivation for these ``black box'' approximations is to enable real-time control of a complex system in cases where evaluating the complete model would be too slow or otherwise infeasible.
In this work, we are particularly interested in models motivated by mechanistic information about a system. In that case, a good model reduction method preserves the relevant mechanistic information to produce a ``gray-box'' description that balances model complexity with mechanistic fidelity\cite{verghese2009getting,kourdis2010physical}. In this case, model reduction can help to identify the relevant, governing mechanism that governs a particular behavior, i.e., distinguish between the relevant and irrelevant details. For example, the Renormalization Group (RG) quantifies the stability of macroscopic observations to microscopic variations in structure, justifying the use of coarse-grained, simplified models. It also makes concrete the concepts of relevant and irrelevant degrees of freedom by identifying those control knobs that must be tuned to observe a particular behavior. These methods are not without their limitations; applications typically require that systems exhibit a high-degree of symmetry in the microscopic interactions.
Information theory has recently been invoked to justify why simple, effective models may be quantitatively predictive in many other complex systems\cite{machta2013parameter,transtrum2015perspective}. This approach uses the Fisher Information Matrix (FIM) to quantify the relative importance of parametric degrees of freedom in a complex model. Small eigenvalues of the FIM therefore correspond to irrelevant details that could, in principle, be discarded from the model. This approach complements and extends traditional arguments based on renormalization group or continuum limits. Where applicable, methods such as RG can construct simple macroscopic representations from microscopic models. However, these techniques break down on models of complex, heterogeneous systems. Nevertheless, these models (sometimes known as ``sloppy models'') will often have many small FIM eigenvalues\cite{brown2003statistical,brown2004statistical,waterfall2006sloppy,gutenkunst2007universally,chachra2012structural}, giving hope that simple, parsimonious representations of these systems may yet be found that transparently bridge microscopic mechanisms with macroscopic phenomenology.
A candidate reduction method that grew out of the information theory approach to modeling is the Manifold Boundary Approximation Method (MBAM)\cite{transtrum2014model}. The foundation for this approach is a geometric interpretation of statistics in which the FIM acts a Riemannian metric on the model's parameter space\cite{rao1949distance,bates1980relative,murray1993differential,amari2007methods,transtrum2010nonlinear,transtrum2011geometry}. The basis for this approach is the observation that sloppy models often correspond to manifolds that are bounded with a hierarchy of widths. The MBAM identifies the boundary oriented with the principal components of the manifold and uses this boundary as an approximation. The MBAM operates iteratively, removing one parameter a time, leading to a sequence of approximate models that connect microscopic components with systems-level behavior. It is therefore a promising method for identifying the mechanistic causes of system-behavior.
The MBAM is a general method that makes few assumptions about the mathematical form of the model. In principle it can be applied to models of dynamical systems as well as field theories, for example. However, the MBAM requires that the model have a particular \emph{global} structure: a model manifold with a hierarchical boundary structure with a sequence of faces, edges, corners, hyper-corners, etc.. Because this structure is invariant to diffeomorphisms of the model manifold, it has been described as the \emph{information topology} of the model manifold\cite{transtrum2014information}.
This paper has two primary objectives. First we elaborate on the details of the manifold boundary approximation method. Second, we use computational differential geometry to explore the boundary structure of several model classes and demonstrate how MBAM would be realized on these models. We show that the necessary hierarchical boundary structure is common to many model classes. In many cases, boundaries have been implicitly used by insightful modelers seeking effective approximations. In these cases, the MBAM would have identified these approximations in a semi-automatic way. The paper is organized as follows: In the next section, we give a review of the manifold boundary approximation method, followed by a simple illustrative example. In section~\ref{sec:Examples}, we consider several different model classes one-by-one, including models composed of elementary operations, dynamical systems, network models, log-linear distributions such as the Ising model, and models with symmetries. We discuss functional similarities to other model reduction methods in section~\ref{sec:relation}, as well as limitations of the MBAM in section~\ref{sec:limitations}. Finally we discuss some of the implications of MBAM in section~\ref{sec:Discussion}.
\section{The Manifold Boundary Approximation Method (MBAM)}
\label{sec:mbam}
\subsection{Algorithmic Description}
\label{sec:mbamalgorithm}
The Manifold Boundary Approximation Method (MBAM) is a model reduction scheme described in reference\cite{transtrum2014model}. As the name suggests, it is based on a geometric interpretation of information theory (known as information geometry\cite{rao1949distance,beale1960confidence,bates1980relative,murray1993differential,amari2007methods,transtrum2010nonlinear,transtrum2011geometry}) that aims to bridge underlying mechanisms with the system-behavior in a wide range of model types. In this section we give a more detailed presentation than what was originally described in reference\cite{transtrum2014model}. In-depth applications to biological models are given elsewhere\cite{transtrum2015bridging}.
We assume the existence of a model in the form of a probability distribution with parameter vector $\theta$. Since approximations such as the MBAM necessarily disregard pieces of the model, it is necessary to identify the objective in mind, i.e., which model behaviors the approximation should preserve. We refer to the particular system behaviors that should be preserved under model reduction as Quantities of Interest (QoIs) which we denote by $\xi$. The model $P(\xi, \theta)$ gives the probability of observing the QoIs given parameters $\theta$.
There is no general rule for choosing QoIs. In practice, the QoIs may often include predictions for which experimental data is available. The data will then be used to calibrate the reduced model. However, QoIs may also include predictions for which data is unavailable but for which the modeler would nevertheless like to make predictions. Alternatively, QoIs may include a very small subset of possible predictions in order to identify a minimal characterization of a system behavior. In other cases, the QoIs may be the probability of all possible predictions, for example as in statistical mechanics which we consider later. In any case, identifying appropriate QoIs for the application in mind is an important step in applying MBAM.
The underlying idea of the MBAM is that the function $P(\xi, \theta)$ is a vector in an inner-product space. If the model contains $N$ parameters, then this vector sweeps out an $N$-dimensional hyper-surface embedded in this space. This hyper-surface is known as the model manifold and denoted by \MM. For many systems, the model manifold is bounded with cross-sections forming an exponential hierarchy of widths. Consequently, \MM \ often has an \emph{effective dimensionality} that is much less than $N$. Our goal is to construct a low dimensional approximation to the model manifold by finding the boundaries of \MM. The procedure for doing this can be summarized as a four step algorithm.
First, from an estimate of the parameters $\theta_0$ calculate the Fisher Information Matrix (FIM)
\begin{equation}
\label{eq:FIM}
g_{\mu\nu} = - \left\langle \frac{\partial^2 \log P}{\partial \theta_\mu \theta_\nu} \right\rangle = \left\langle \frac{\partial \log P}{\partial \theta_\mu} \frac{\partial \log P}{\partial \theta_\nu} \right\rangle.
\end{equation}
The FIM acts as a Riemannian metric on \MM. Calculating the eigenvalues of this matrix reveals the ill-conditioned nature of the corresponding parameter inference problem. The eigenvectors with small eigenvalues correspond to the parameter combinations that would be unidentifiable from an observation of the QoIs and so have little explanatory value. We denote the eigenvector with smallest eigenvalue by $v_0$.
The second step is to calculate a parameterized path through parameter space $\theta(\tau)$ corresponding to the geodesic originating with parameters $\theta_0$ and direction $v_0$. This is found by numerically solving a differential equation:
\begin{equation}
\label{eq:geodesic}
\frac{d^2}{d \tau^2} \theta^\mu = \Gamma^\mu_{\alpha \beta} \frac{d \theta^\alpha}{d\tau} \frac{d \theta^\beta}{d\tau}
\end{equation}
where
\begin{equation}
\label{eq:Gamma}
\Gamma^\mu_{\alpha\beta} = g^{\mu\nu} \left\langle \frac{\partial \log P}{ \partial \theta_\nu} \left( \frac{\partial^2 \log P}{\partial \theta_\alpha \partial \theta_\beta} + \frac{1}{2} \frac{\partial \log P}{\partial \theta_\alpha} \frac{\partial \log P}{\partial \theta_\beta} \right) \right\rangle
\end{equation}
is the Riemann connection on the manifold and $g^{\mu\nu} = \left(g^{-1}\right)^{\mu\nu}$.
Computational cost is always a concern for methods targeted at large models. The most expensive part of this calculation is evaluating the FIM which requires calculating the derivative of the model with respect to all parameters and thus scales linearly with the number of parameters. These derivative calculations are trivially parallelized. In contrast, calculating $\Gamma^\mu_{\alpha\beta} \dot{\theta}^\alpha \dot{\theta}^\beta$ requires only an additional second directional derivative. This can be estimated with computational cost comparable to a single evaluation of the model. Thus, the additional overhead of calculating a geodesic beyond the FIM becomes negligible as models become large. However, the FIM must be evaluated repeatedly while solving Eq.~\eqref{eq:geodesic}. We discuss the implications of this in section~\ref{sec:limitations}.
The solution to Eq.~\eqref{eq:geodesic} is a parameterized curve through the parameter space. Along this curve, the modeler monitors the eigenvalues of the FIM. A boundary of the model manifold is identified by the smallest eigenvalue of $g_{\mu\nu}$ approaching zero. When the smallest eigenvalue becomes much less than the next smallest, then the corresponding eigendirection reveals a limiting approximation in the model.
The approximation will typically correspond to one or more parameters approaching zero or infinity in a coordinated way. The third step is to identify this limit and analytically evaluate it in the model. This step often requires some theoretical insight; several examples of how to do this are given later. The result of the process is a new model with one less parameter. We denote this reduced model by $\tilde{P}(\xi, \phi)$, where $\phi$ are the reduced set of parameters.
Finally, the values of the parameters $\phi$ in the approximate model are calibrated to the parameters $\theta_0$ by minimizing the information distance to the original model:
\begin{equation}
\label{eq:LSestimate}
\min_\phi \left\langle \log P(\xi, \theta_0) - \log \tilde{P}(\xi, \phi) \right\rangle
\end{equation}
where $\langle \cdot \rangle$ means expectation value with respect to the original distribution $P(\xi, \theta_0)$. Because the first term in Eq.~\eqref{eq:LSestimate} is a constant, calibrating the model is equivalent to maximizing the log-likelihood that the model $\tilde{P}(\xi, \phi)$ generated the data.
This four-step procedure is iterated, removing one parameter at a time, until the model becomes sufficiently simple.
The procedure just described requires a few comments. First, the MBAM requires a parameter estimate as a starting point $\theta_0$, which usually cannot be estimated accurately. The final reduced model is largely independent to these uncertainties. The reason for this is seen by considering a geometric argument given in reference\cite{transtrum2014model}. Huge variations in parameter values can result when fitting to data, but these variations all lie within the same statistical confidence region, which means they map to nearby points on the model manifold. Starting from any points within this confidence region will identify the same sequence of boundaries as the true parameters. In other words, choosing a $\theta_0$ is incidental to the procedure but unnecessary for the final result.
Because of its geometric motivation, the reduced models are invariant to changes in a model's parameterization, such as using rates vs. time constants. These transformations are equivalent to coordinate transforms on the manifold. In many applications, the microscopic parameters are restricted to positive values. In order to guarantee positivity, it is helpful to use log-transformed parameters in the model. This serves the dual purpose of non-dimensionalizing the parameters, that is important for the initial eigendirection of the FIM to point to the narrowest width of the \MM.
Finally, MBAM is a nonlocal approximation in the sense that the reduced model approximates not just the behavior at $\theta_0$ but at all other nearby behaviors, i.e., those behaviors along the long-axis of the manifold. In general, identifying all model behaviors using a brute-force exploration of parameter space is infeasible for models with more than a handful of parameter. MBAM is able to approximate large regions of parameter space without a direct exploration by exploiting the hierarchical boundary structure of many models. By choosing a boundary oriented with the principal axis of the manifold, the MBAM avoids the need to explicitly explore the entire parameter space.
\subsection{Illustrative Example: Biological Adaptation and Negative Feedback}
\label{sec:example}
We illustrate the manifold boundary approximation method with a simple example. Consider a simple two-parameter model that arises in the study of biological adaptation to the mean\cite{sontag2008remarks,ma2009defining,nemenman20124,transtrum2015bridging}:
\begin{eqnarray}
\label{eq:nfblb1}
\frac{dA}{dt} & = & (1 - A) - k_2 A B \\
\label{eq:nfblb2}
\frac{dB}{dt} & = & k_1 A (1 - B)
\end{eqnarray}
with $k_1$ and $k_2$ as parameters and $A$ and $B$ dynamical variables with initial conditions zero. We take as quantities of interest the time series for $A$ with additive Gaussian noise leading to a least-squares estimate when fit to data. Varying parameter values leads to different time series for this model as in Figure~\ref{fig:NFBLBMM} (top). Fitting this model to a single realization of the data (red dots in Figure~\ref{fig:NFBLBMM}, top) leads to a chi-squared cost surface (Figure~\ref{fig:NFBLBMM}, center). By considering all possible model predictions for the QoIs in data space, we generate the model manifold (Figure~\ref{fig:NFBLBMM}, bottom).
\begin{figure}
\includegraphics[width=3.5in]{NFBLBMM}
\caption{\label{fig:NFBLBMM} \textbf{Visualizing ranges of model behavior.} Top: Varying the parameters of the model leads to a family of time series. Center: Fitting these time series to data (red dots in top panel) generates a chi-squared cost surface in parameter space. Bottom: The Model Manifold \MM \ is generated by collecting all possible predictions of the model at specific time points (dashed, red vertical lines in top panel). Geodesics are the analogs of straight lines on curved surfaces and are depicted by the red curves in the center and bottom panels.}
\end{figure}
The red curves in Figure~\ref{fig:NFBLBMM} center and bottom panels are geodesics on \MM \ represented in parameter space and behavior space respectively. Notice how the geodesics locally align with the cost surface while globally exploring the model manifold. In this way geodesics connect local information, i.e., the Fisher Information, with the model's nonlocal structure, i.e. boundaries.
Manifold boundaries correspond to physically realistic models; the boundaries oriented with the principal axes of the manifold are natural approximations to the full model. Consider the geodesics moving toward the upper-left corner of Figure~\ref{fig:NFBLBMM}, middle panel. These curves ultimately terminate at the lower boundary of the manifold in Figure~\ref{fig:NFBLBMM}, bottom panel. This boundary corresponds to the limit that $k_1 \rightarrow 0$ while $k_2 \rightarrow \infty$. Evaluating this limit in the model requires some theoretical work. Inspecting Eqs.~\eqref{eq:nfblb1}-\eqref{eq:nfblb2}, we see that if $k_1 \rightarrow 0$, then $B \rightarrow 0$. However, the influence of $B$ on $A$ is scaled by the parameter $k_2$ which becomes very large. Therefore, we define a \emph{renormalized} variable $\tilde{B} = k_2 B$, and the model becomes
\begin{eqnarray}
\label{eq:nfblbrenorm1}
\frac{dA}{dt} & = & (1 - A) - A \tilde{B} \\
\frac{d \tilde{B}}{dt} & = & (k_1 k_2) A ( 1 - \tilde{B}/k_2) \nonumber \\
\label{eq:nfblbrenorm2}
& \rightarrow & \tilde{k} A,
\end{eqnarray}
where we have used the renormalized parameter $\tilde{k} = k_1 k_2$.
Eqs.~(\ref{eq:nfblbrenorm1}-\eqref{eq:nfblbrenorm2}) have a simple physical interpretation. The model behavior does not depend on absolute scale of $B$, but on $k_2 B$. Furthermore, the curves in Figure~\ref{fig:NFBLBMM} (top) are determined primarily by the product $k_1 k_2$ and not either parameter individually. Consequently, simultaneously estimating $k_1$ and $k_2$ from data would have led to large uncertainties (Figure~\ref{fig:NFBLBMM}, middle panel)). On the other hand, the parameter uncertainties in the reduced model would be much smaller. The renormalized expressions for $\tilde{B}$ and $\tilde{k}$ therefore characterize the family of physical systems with equivalent behavior. This ``gray-box'' representation preserves the negative feedback mechanism while discarding the information about the scale of $B$.
The aspect ratio of the model manifold in Figure~\ref{fig:NFBLBMM} is not particularly dramatic so the approximation in Eqs.~\eqref{eq:nfblbrenorm1}-\eqref{eq:nfblbrenorm2} is not very accurate for some values of the parameters. For sloppy models with more than a few parameters, aspect ratios greater than 1000:1 are not uncommon, in which case the resulting approximate models are very accurate. The width of the model manifold is a measure of the model error that is estimated automatically from the MBAM algorithm during the calibration step (Eq.~\eqref{eq:LSestimate}).
There is an inherent ambiguity in the MBAM because of the two possible boundaries in Figure~\ref{fig:NFBLBMM}. Alternatively, we could have chosen the opposite boundary. In this case, the geodesics in Figure~\ref{fig:NFBLBMM} indicate that $k_1 \rightarrow \infty$. In this limit we have $B\rightarrow 1$ and the approximate model becomes
\begin{eqnarray}
\label{eq:nfblbrenormalt1}
\frac{dA}{dt} & = & (1 - A) - k_2 A \\
\label{eq:nfblbrenormalt2}
B & = & 1.
\end{eqnarray}
Notice how this limit is dual to the first. In both cases, the physical unconstrained quantity is in the magnitude of $B(t)$. In the first case we employed the approximation $B\rightarrow 0$, while in the second we employ the approximation $B \rightarrow 1$. This illustrates a general principle: reduced models are not unique. In general, any model can be approximated by any other model in the same universality class. Manifold boundaries are convenient choices because structural simplifications occur at extreme values of the parameters.
When MBAM is applied to a model with many parameters, the four-step algorithm is iterated several times. At each step there may be several approximations to choose from. In our experience, the final model is largely independent of these choices. To understand why, image that Figure~\ref{fig:NFBLBMM} corresponds to a two-dimensional cross section of a high-dimensional model manifold. If the limit $k_1 \rightarrow \infty$ is selected first, then the second iteration is likely to identify the dual limit so that the approximate model corresponds to one of the two hyper-corners in Figure~\ref{fig:NFBLBMM}. That same hyper-corner would have been reached if the order of limits had been reversed. Ultimately, the only ambiguity is the order in which the boundaries are evaluated.
\section{Examples of Models with a Hierarchical Boundary Structure}
\label{sec:Examples}
Having illustrated the basic process of the MBAM, we now begin to enumerate a catalog of model classes that have a hierarchical boundary structure, i.e., faces, edges, corners, hyper-corners, etc. This list is not complete, and some counterexamples will be given section~\ref{sec:limitations}.
\subsection{Compositions of Elementary Functions}
\label{sec:compositions}
\textbf{Rational Functions:}
We first consider several functions composed from elementary operations. For example, consider the six parameter rational function
\begin{equation}
\label{eq:rational}
y(t, \theta) = \frac{t^2 + \theta_1 t + \theta_2}{\theta_3 t^3 + \theta_4 t^2 + \theta_5 t + \theta_6}.
\end{equation}
We take as QoIs the output of this model at several time points with additive Gaussian noise. The geodesic for this model is summarized in Figure~\ref{fig:RationalGeodesic}. The manifold boundary is identified by a singularity that occurs around $\tau = 13$. At this point, the absolute value of several of the parameters are growing without bound.
\begin{figure}
\includegraphics[width=3.5in]{RationalGeodesic}
\caption{\label{fig:RationalGeodesic} \textbf{Geodesics for a rational function.} Top: Parameter values along the geodesic for the model in Eq.~\eqref{eq:rational}. Notice the geodesic encounters a singularity around $\tau = 13$. The components of the geodesic velocity are shown in the next three panels. First the initial components of the velocity are determined by the eigenvector of the FIM with smallest eigenvalue. Next, the final components of the geodesic velocity near the singularity at $\tau = 13$. These components satisfy the equation $d\phi = 0$ as explained in the text. This is made clear by considering the velocity components in the log-parameters (bottom panel), in which case the components are equal.}
\end{figure}
In this case, knowing that the parameters become infinite (in absolute value) at the boundary is sufficient to identify the reduced model. Dividing the numerator and denominator by $\theta_6$, for example, gives
\begin{eqnarray}
\tilde{y}(t, \theta) & = & \frac{t^2/\theta_6 + (\theta_1/\theta_6) t + (\theta_2/\theta_6)}{(\theta_3/\theta_6) t^3 + (\theta_4/\theta_6) t^2 + (\theta_5/\theta_6) t + 1} \nonumber \\
\label{eq:rationalapprox}
& \rightarrow & \frac{ \phi_1 t + \phi_2}{\phi_3 t^3 + \phi_4 t^2 + \phi_5 t + 1} \nonumber \\
\end{eqnarray}
where the renormalized parameters are $\phi_\mu = \theta_\mu/\theta_6$.
Consider the hyper-surfaces in parameter space defined by $\phi_\mu$ = constant. These hyper-surfaces are very nearly parallel to the manifold boundary so that the unconstrained parameter combination must be perpendicular to this surface. Normals to these surfaces given by the differential form $d \phi_\mu = 0$. Since the geodesic velocity is approximately parallel to the unconstrained parameter combination, the components of the geodesic velocity are often a useful clue for identifying the boundary approximation and the reduced parameters as we now illustrate.
In the second and third panels of Figure~\ref{fig:RationalGeodesic} we give the initial and final) components of the geodesic velocity. In this case, the components of the final velocity all have different magnitudes. These magnitudes are determined by the requirement $d \phi_\mu = 0$ which gives $d \theta_\mu = (\theta_\mu / \theta_6) d \theta_6$. We confirm that the relative height of the peaks in Figure~\ref{fig:RationalGeodesic} middle panel are given by this relation. Indeed, considering log-transformed parameters, we have $d \phi_\mu = 0$ implies that $ d \log \vert \theta_\mu \vert = d \log \vert \theta_6 \vert$ as shown in Figure~\ref{fig:RationalGeodesic} bottom panel. This example demonstrates that the boundaries and corresponding limits identified by MBAM are invariant to reparameterization of the model.
By iterating the MBAM, we remove additional parameters. In the next iteration, $\phi_\mu \rightarrow \infty$, the model becomes:
\begin{equation}
\label{eq:rational2approx}
\tilde{\tilde{y}}(t) = \frac{ (\phi_1/\phi_5) t + (\phi_2/\phi_5)}{(\phi_3/\phi_5) t^3 + (\phi_4/\phi_5) t^2 + t}.
\end{equation}
\textbf{Sums of Exponentials:}
The model corresponding to
\begin{equation}
\label{eq:exponentials}
y(t, A, \lambda) = \sum_\mu A_\mu e^{-\theta_\mu t}
\end{equation}
with parameters $\theta = (A_\mu, \lambda_\mu) > 0$ was considered in reference\cite{transtrum2014model} in which case it was shown that the manifold boundary approximation was equivalent to reducing the number of terms in the sum and adding a constant term. This model is interesting, however, because the model manifold has unbounded directions. In particular, any of the linear parameters $A_\mu$ can be taken to infinity and the model predictions will similarly become infinite. Although the manifold is not bounded, it nevertheless has bounded cross sections that form a hierarchical structure. We discuss the possibility of unbounded manifolds in section \ref{sec:limitations}.
\textbf{Composing Rational Functions and Exponentials:}
We now elaborate on previous examples by composing rational and exponential functions:
\begin{equation}
\label{eq:partitionfunction}
y(T, g, \Delta E) = \frac{g e^{-\Delta E/k_B T}}{1 + g e^{-\Delta E/k_B T}}.
\end{equation}
This function arises in modeling the probability of observing a two-state system in a particular state in thermal equilibrium as a function of temperature\cite{hansen2015enzyme}. The parameters are $\theta = (g, \Delta E)$ where $g\geq 0$, $\Delta E$ is unbounded, and temperature $T$ is the independent variable.
This model has a few obvious limits. If $g \rightarrow \infty$ or if $\Delta E \rightarrow -\infty$ then $y(T) = 1$ and if $g \rightarrow 0$ or $\Delta E \rightarrow \infty$ then $y(T) = 0$. However, neither of these are boundaries of the model manifold. The reason is that these obvious limits remove both parameters. Except in unusual circumstances (as we discuss in section \ref{sec:limitations}), the edge of the two-parameter model is a one-parameter model.
We consider the simple case in which the QoIs consist of $y(T)$ at two different temperatures. In Figure~\ref{fig:partition} (left) we show the geodesics for this model originating from the origin in many different directions. In this case the model has four boundaries as can be identified by the four different slopes with which the geodesic path takes parameters to infinity.
\begin{figure}
\includegraphics[width=\linewidth]{PartitionFunctionGeodesic}
\caption{\label{fig:partition} \textbf{Geodesics for model in Eq.~\eqref{eq:partitionfunction}} The different slopes at which the geodesic approaches infinity correspond to different boundaries. Selecting two observation temperatures as QoIs leads to four boundaries (left) while three temperatures leads to six boundaries (right). }
\end{figure}
If $g\rightarrow \infty$ and $\Delta E \rightarrow \infty$, then the critical expression $\log g - \Delta E / k_B T$ will become either positive or negative infinity depending on the temperature. Therefore, there is a critical temperature, $T_c$ at which a transition occurs: $y(T < T_c) = 0$ and $y(T > T_c) = 1$. The reduced parameter $\phi = \log g - \Delta E / k_B T_C$ controls the value of $y(T_c)$. The two boundaries for which $g\rightarrow \infty$ and $\Delta E \rightarrow \infty$ therefore correspond to the two possible choices of critical temperatures. Indeed, the slopes of these geodesics paths are given by $d \phi = 0$ which gives $d \log g/ d\Delta E = 1/k_B T$.
The two boundaries for which $g\rightarrow -\infty$ and $\Delta E \rightarrow -\infty$ similarly correspond to the two possible choices of $T_c$, but with a reversed transitions: $y(T < T_c) = 1$ and $y(T > T_c) = 0$.
This analysis suggests that changing the QoIs to include $3$ temperatures would result in a model manifold with six faces, as we confirm numerically in Figure~\ref{fig:partition} (right).
In terms of the probability distribution of a two-state system, these limits have natural interpretations. In general, there will be a transition from low to high probability with temperature. The temperature over which this transition occurs depends on the balance between the relative multiplicities of the two states and the energy difference. Each of the boundaries above correspond to the approximation that this transition is abrupt.
\subsection{Dynamical Systems}
\label{sec:dynamicalsystems}
Many systems of physical interest are described by dynamical systems. Depending on the functional form of the dynamical system, there are many classes to consider. In the interest of space, several of these classes have been or will be discussed in more detail elsewhere, including network models of biochemical kinetics (including mass-action and Michaelis-Menten dynamics)\cite{transtrum2015bridging}, linear-time invariant systems (such as arise in control theory)\cite{pare2015unified}, models of transient dynamics in power systems, and neuroscience models of Hodgkin-Huxley neurons. For completeness, we here note that models in each of these classes also exhibit a hierarchical boundary structure.
Here we consider an example of a multi-compartment model that will serve as a segue into a discussion about network models in section~\ref{sec:networks}. A compartment model describes the flow of material or energy among compartments and is common in a variety of fields including epidemiology (such as the SIR model), pharmokinetics for describing drug delivery, ecology, and many others. Consider here for example a three-compartment model with compartments $A$, $B$, and $C$ connected in series so that material begins in compartment $A$ and then flows to compartment $B$ and then from $B$ to $C$: $A \rightarrow B \rightarrow C$. Assuming linear kinetics, the corresponding differential equations are
\begin{eqnarray}
\label{eq:compartmentA}
\frac{dA}{dt} & = & -k_1 A \\
\label{eq:compartmentB}
\frac{dB}{dt} & = & k_1 A - k_2 B \\
\label{eq:compartmentC}
\frac{dC}{dt} & = & k_2 B.
\end{eqnarray}
We take as QoIs, the values of $C$ at several times.
Geodesics reveal that one boundary of this model manifold is the limit $k_2 \rightarrow \infty$ (Figure~\ref{fig:ssaGeodesic}). This limit can be evaluated by noting that $B \rightarrow 0$ so that $k_2 B = k_1 A$:
\begin{eqnarray}
\label{eq:compartmentapproxA}
\frac{dA}{dt} & = & -k_1 A \\
\label{eq:compartmentapproxC}
\frac{dC}{dt} & = & k_1 A.
\end{eqnarray}
\begin{figure}
\includegraphics[width=3.5in]{SSAGeodesic}
\caption{\label{fig:ssaGeodesic} \textbf{Geodesic reveals a steady state approximation} The manifold boundary for the model in Eqs.~\eqref{eq:compartmentA}~\eqref{eq:compartmentC} is identified by the singularity in the geodesic near $\tau = 1.3$ and corresponds to the limit that $k_2\rightarrow\infty$. In this limit, $dB/dt = 0$, and is equivalent to a steady state approximation.}
\end{figure}
The steady state approximation is frequently used to reduce the order of many dynamical systems, such as the Michaelis-Menten approximation which is common in biochemical kinetics. In this case, it would typically be applied as the condition $dB/dt = 0$ which leads to the relation $k_1 A = k_2B$. The steady state approximation is therefore mathematically equivalent to the manifold boundary approximation.
Many other methods for deriving reduced mechanisms operate in a similar fashion. Just as the steady-state assumptions force the rate of production of some species to vanish, partial equilibrium assumptions similarly force other rates to extreme values and therefore decouple the relevant mechanisms\cite{maas1992simplifying}. The key difference between the is that the manifold boundary can be identified in a semi-automatic fashion without the need to have deep insights into the mechanisms driving complex systems.
In control theory, the controllability and observability of a system is quantified by the Hankel singular value\cite{dullerud2013course}. A common approximation is to disregard the states that are least-controllable and observable, known as balanced truncation\cite{moore1981principal}. If the Hankel singular values are treated as parameters, then balanced truncation is equivalent to the manifold boundary in which the smallest Hankel singular values become zero, i.e., the state becomes completely unobservable and controllable\cite{pare2015unified}.
\subsection{Networks Models}
\label{sec:networks}
Many models have an associated network structure. The problem of structure-preserving model reduction is an important one with extensive treatment in the controls community\cite{lall2003structure,sandberg2009model,yeung2011mathematical,trnka2013structured,ishizaki2014model,abad2014graph}. In these cases it is often desirable to approximate the model that retains some semblance of the original network. Within the context of the MBAM, if the parameters are directly related to the network structure, representing the strength of edge connections for example, then the resulting approximations have an approximate network structure directly linked to the original.
We have already seen this in the previous section for the case of a simple compartment model. The original network ($A \rightarrow B \rightarrow C$) is automatically condensed into $A \rightarrow C$ in which the indirect flow of material from $A$ to $C$ is replaced with an direct, effective link. Approximations like this are also important in fields for which meaning is attached to the network topology. Systems biology, for example, is particular concerned with the appearance of network motifs\cite{shen2002network,milo2002network,mangan2003structure,alon2007network}. In reference\cite{transtrum2015bridging} it was shown how repeated iteration of MBAM can condense a complex topology of protein reactions into a simple effective topology among clusters of proteins.
We now consider other types of network models that similarly have a hierarchical boundary structure such that iterating MBAM preserves an effective network structure.
\textbf{Bayesian Networks and Markov Chains:} As a first example, consider the case of a Bayesian Network model. Bayesian networks are probabilistic models formulated as directed acyclic graphs representing the conditional dependence of random variables. Because the parameters correspond to conditional probabilities, the model manifold inherits a hierarchical boundary structure from the probabilistic meaning of the parameters. Each parameter is positive and each set of conditional probabilities must sum to one. The parameter space, is therefore restricted to some high-dimensional simplex. The model manifold inherits the same hierarchical boundary structure from the region of allowed parameter values.
The arguments above generalize for network structures in which parameters are themselves probabilities. For example, Markov chains are often depicted as networks with nodes as states and edges representing transition probabilities. These probabilities are subject to similar constraints as those in Bayesian networks leading to a hierarchical boundary structure.
\textbf{Artificial Neural Networks:}
Another example is an artificial neural network, a common model for machine learning. Nodes represent artificial neurons, which are typically understood to represent a sigmoidal activation function of its inputs $\sigma(I)$. (There are many possible activation functions. We use $\sigma(I) = \tanh(I)$, but the following results are independent of this choice.) The network structure indicates which artificial neurons serve as inputs to other neurons. Parameters are weights associated with edges that indicate both the strength and type (promoting or inhibitory) of the connection.
It is instructive to first consider a single neural activated by two environmental inputs
\begin{equation}
\label{eq:ann}
y(x_1, x_2, \theta) = \tanh(\theta_1 x_1 + \theta_2 x_2 + \theta_3)
\end{equation}
where $x_1$ and $x_2$ are the environmental inputs that activate the neuron with weight $\theta_1$ and $\theta_2$ respectively. The third parameter $\theta_3$ is a bias term.
The boundaries of this model correspond to the limit in which $\vert \theta_\mu \vert \rightarrow \infty$ for $\mu$ = 1, 2, and 3 with reduced parameters $\phi_1 = \theta_2/\theta_1$ and $\phi_2 = \theta_3/\theta_1$. In this limit the output of the neuron is $\pm$1 for any values of $x_1$ and $x_2$. The reduced model is therefore
\begin{equation}
\label{eq:annred}
\tilde{y}(x_1, x_2, \phi) = \begin{cases}
1 & \mbox{if } x_1 > -\phi_1 x_2 - \phi_2 \\
-1 & \mbox{if } x_2 < -\phi_1 x_2 - \phi_2.
\end{cases}
\end{equation}
The two parameters therefore define a line through the input plane that partition inputs based on how it is classified by the reduced model.
This reduced model is equivalent to another machine learning model known as a perceptron. Perceptrons are binary, linear classifiers with a long history going back nearly half a century\cite{rosenblatt1958perceptron}. Although the relationship between artificial neurons (as in Eq.~\eqref{eq:ann}) and perceptrons is well-known (networks of these neural often known as multi-layer perceptrons), of particular note in this context is that this equivalence is naturally recovered by interpreting the perceptron as a boundary approximation to an artificial neuron.
Combining several artificial neurons into a network produces a richer boundary structure in the model. However, these boundaries can be understood as a generalization of the single-neuron boundary just considered. Several iterations of the MBAM effectively lead to a composition of several perceptrons. The resulting model is a binary classifier that can approximate more complex divisions of the input space, i.e., not just a single line. The resulting calculation is closely related to yet another machine learning algorithm: Support Vector Machines (SVMs). The relationship among SVMs, perceptrons, and artificial neural networks is also known\cite{freund1999large}. In the current context, the interesting result is that this relationships is naturally captured in the differential topological structure of the artificial neural network model and automatically recovered by iterative application of the MBAM.
\textbf{Markov Random Fields:}
The final network model class that we consider here is the Markov Random Field (MRF). Like Bayesian networks, MRFs aim to represent probabilistic dependence among random variables. While Bayesian networks are acyclic, MRFs may be cyclic. While the hierarchical boundary structure of Bayesian networks followed naturally from the probabilistic interpretation of the parameters, the ranges of allowed parameters of a MRF may be unbounded. The boundary structure of the model manifold in Markov random fields has a different explanation.
For concreteness, consider the network in Figure~\ref{fig:mrfexmaple} (left). Each node is a random variable that can take values $\pm$ 1. The probability of a particular configuration is the proportional to
\begin{equation}
\label{eq:mrfprob}
P(\mathbf{s}, \theta) \propto \exp \left( -\frac{1}{2} \, \mathbf{s}^T A(\theta) \mathbf{s} \right)
\end{equation}
where $\mathbf{s} = (s_1, s_2, s_3, s_4, s_5)^T$ and
\begin{equation}
\label{eq:MRFPrecisionMatrix}
A(\theta) = \left( \begin{array}{ccccc} 0 & \theta_1 & 0 & \theta_2 & 0 \\
\theta_1 & 0 & 0 & \theta_3 & 0 \\
0 & 0 & 0 & 0 & \theta_5 \\
\theta_2 & \theta_3 & 0 & 0 & \theta_4 \\
0 & 0 & \theta_5 & \theta_4 & 0
\end{array} \right).
\end{equation}
Notice that the nonzero entries of $A$ reflect the network structure in Figure~\ref{fig:mrfexmaple}. There is a missing normalization in Eq.~\eqref{eq:mrfprob} that depends on the parameters $\theta$ but not on the state vector $\mathbf{s}$.
\begin{figure}
\includegraphics[width=3.5in]{MRF}
\caption{\label{fig:mrfexmaple} \textbf{MBAM on a Markov Random Field} Left: Network structure for the Markov Random Field in Eq.~\eqref{eq:MRFPrecisionMatrix}. Right: The manifold boundary corresponding to $\theta_4\rightarrow\infty$ corresponds to a clustering of nodes 4 and 5.}
\end{figure}
For this model the manifold boundaries correspond to the limits $\theta_\mu \rightarrow \pm \infty$. These limits correspond to a type of ``clustering'' of network nodes. For example, if $\theta_4 \rightarrow \infty$, then the random variables $s_4$ and $s_5$ become perfectly correlated. This reduced model therefore corresponds to the condensed network given in Figure~\ref{fig:mrfexmaple} (right).
Interesting, is the random variables can be any real number, then the model no longer has a boundary structure. We discuss this further in section~\ref{sec:limitations}.
\subsection{Log-Linear Discrete Distributions}
\label{sec:loglinear}
We now consider probabilistic models with a finite number of outcomes such that the probability of outcome $i$ may be written as
\begin{equation}
\label{eq:loglinear}
P_i \propto \exp \left( \sum_\mu \Pi_{i\mu} \theta_\mu \right).
\end{equation}
The proportionality constant depends on the parameters but is the same for each outcome. Apart from an over-all normalization, models of this form are log-linear in the parameters. The Markov random field from the previous section falls into this category since the matrix $A(\theta)$ is linear in the parameters. Many standard statistical mechanics models falls into this category including the Ising model and generalizations such as the Potts model and cluster expansions of alloy formation enthalpies.
Models of this class can be shown to always have a hierarchical boundary structure that is closely related to the structure of the model's $\Pi$ matrix. In particular, we interpret each row of the $\Pi$ matrix as a vector in parameter space. The convex hull of this set of points will generally have a hierarchical structure that is diffeomorphic to, i.e., has the same boundary structure as, the model manifold.
To understand this correspondence, we first assume that the model has $N$ parameters and the points in $\Pi$ are not constrained to a linear subspace of the parameter space. This requirement means that the convex hull of points in $\Pi$ has non-zero volume. In this case the FIM for the model is not singular at any finite values of the parameters, and any manifold boundaries must correspond to infinite parameter values. Note that this requirement guarantees that $\Pi$ has $N$ non-zero singular values, although the converse is not true.
We now consider the behavior of the model at infinite parameter values. These limits are characterized by some outcomes occurring with zero probability in the model (effectively freezing out the highest-energy configurations). For any parameter values, the relative probability of observing two states is related to the difference in their $\Pi$ matrix rows:
\begin{equation}
\label{eq:relativeprob}
\frac{P_i}{P_j}= \exp \left( \sum_\mu \left( \Pi_{i\mu} - \Pi_{j\mu} \right) \theta_\mu \right).
\end{equation}
Now consider a parameterized path through parameter space $\theta(\tau)$. We assume that for large $\tau$ the path moves infinitely far from the origin and approaches a straight line, so that near the geodesic singularity we may write $\theta(\tau) = \theta_0 + v \tau$ for some vector $v$ that becomes very large but does not rotate. In this case, Eq.~\eqref{eq:relativeprob} becomes
\begin{equation}
\label{eq:relatieprob2}
\frac{P_i}{P_j} = e^{\Delta \Pi \theta_0} e^{\Delta \Pi v \tau}
\end{equation}
where $\Delta \Pi = \Pi_{i\mu} - \Pi_{j\mu}$.
As $\vert v \vert \rightarrow \infty$, Eq.~\eqref{eq:relatieprob2} suggests that $P_i/P_j$ will become either $0$ or $\infty$ depending on the sign of $\Delta \Pi v$ and corresponding to the cases $P_i \rightarrow 0$ or $P_j \rightarrow 0$ respectively. For $P_i$ to remain nonzero in the limit $\vert v \vert \rightarrow \infty$, the point $\Pi_i$ must have the largest projection onto the vector $v$ of any other points in the $\Pi$ matrix. Since $\vert v \vert \rightarrow \infty$ is the limit of zero temperature, the outcomes for which $P_i \neq 0$ correspond (typically) to ground states. This means that ground states of the system correspond to the extreme points in $\Pi$ as we illustrate in Figure~\ref{fig:convexhull} left. This argument was first given in reference\cite{seko2014efficient} and used to efficiently identify ground states of a cluster expansion.
Consider the limit $\vert v \vert \rightarrow \infty$ such that only the ground state(s) has nonzero probability (as in the previous paragraph). In this case, the model becomes infinitely insensitive to all parameters, i.e., the lowest energy outcomes all become equally probable independent of any variation in the parameters. This limit therefore corresponds to a vertex (i.e., a zero-dimensional boundary) of the model manifold.
In order to find the limits that correspond to higher-dimensional boundaries, we must consider limits in carefully chosen directions. As $\vert \theta \vert \rightarrow \infty$, it follows from Eq.~\eqref{eq:relatieprob2} that that $P_i / P_j$ remains nonzero and finite only if $\Pi_{i\mu} v_\mu = \Pi_{j\mu} v_\mu$ which means that the points $\Pi_i$ and $\Pi_j$ must have the same projection onto the vector $v$, i.e., $v$ is perpendicular to the line connecting points $\Pi_i$ and $\Pi_j$, as in Figure~\ref{fig:convexhull}, right.
\begin{figure}
\includegraphics[width=3.5in]{ConvexHull}
\caption{\label{fig:convexhull} \textbf{Convex Hull of $\Pi$ determines the boundary structure.} Rows of $\Pi$ in Eq.~\eqref{eq:loglinear} represent discrete outcomes in the model and correspond to points in parameter space (dark red and light blue dots). Here we choose a model with three outcomes. The two parameters control the relative probabilities $P_1/P_2$ and $P_1/P_3$. \textbf{Left}: Consider the path through parameter space given by $\theta(\tau) = \theta_0 + v \tau$ as shown. The relative probability of each outcome is determined by the the projection of the corresponding point onto the line connecting $\theta$ to the origin (the solid dark red and light blue lines indicate this projection onto the dashed red line). In the limit $v \rightarrow \infty$, only the outcome corresponding to $\Pi_1$ (the dark red dot) occurs with nonzero probability (the ground state). Therefore, in this limit the model becomes insensitive to all parameters and corresponds to a vertex (i.e., zero-dimensional boundary) of the model manifold. \textbf{Right}: Consider the direction depicted on the right that is perpendicular to the face of the convex hull. In the limit $v \rightarrow \infty$ the difference in the projections of $\Pi_1$ and $\Pi_2$ onto the line from $\theta$ to the origin becomes the same and both outcomes occur with nonzero probability. In this limit, the model remains sensitive to one parameter combination that controls the relative probability of these two outcomes while the third outcome is frozen out. This limit corresponds to an edge (i.e., one-dimensional boundary) of the model manifold. The model manifold has the same boundary structure as that of the convex hull of $\Pi$, in this case a triangle (represented by the black dashed line).}
\end{figure}
We now consider which vectors $v$ reduce the dimensionality of the parameter space by one. Assume the parameter space has $N$ dimensions and consider the limit $\vert v \vert \rightarrow \infty$. In this limit, only some of the states will remain with nonzero probability. As we just argued, the rows of the $\Pi$ matrix corresponding to these non-zero states must lie in a space perpendicular to the vector $v$. However, in order for this limit to be a manifold boundary of dimension $N-1$, these points must span a space of dimension $N-1$. Therefore $v$ must perpendicular to an $N-1$ dimensional face of the convex hull of $\Pi$ matrix points. This is illustrated in Figure~\ref{fig:convexhull} (right) for the case of $N = 2$.
From this argument, it follows that there is a one-to-one correspondence between $N-1$ dimensional faces of the convex hull of the points in $\Pi$ and the $N-1$ dimensional boundaries of the model manifold. By repeating this argument for $N-2$, $N-3$, etc. dimensional boundaries, we conclude that the model manifold has the same boundary structure as the convex hull of points in $\Pi$.
Geodesics will typically encounter a boundary of dimension $N-1$. (Only by fine-tuning the initial direction of a geodesic will it encounter a corner of the model manifold.) This suggests that geodesics in parameter space will asymptotically become perpendicular to the faces of the convex hull of the points in $\Pi$. We show this explicitly in Figure~\ref{fig:fccgeodesics} for the model whose parameters are the first two non-trivial terms of a cluster expansion of a binary alloy on an FCC lattice. The first parameter is the on-site energy for having an atom of one type. The second parameter is the nearest neighbor interaction. Notice that for large parameters values along the geodesic, the path orients itself to be perpendicular to the faces of the convex hull of points. In this case, the geodesics paths approach infinity with five distinct slopes, so the model manifold is a pentagon (notice the similarity to Figure~\ref{fig:partition}).
\begin{figure}
\includegraphics[width=3.5in]{FCCGeodesics}
\caption{\label{fig:fccgeodesics} \textbf{Convex Hull for cluster expansion on an FCC lattice.} The model manifold is has five edges and five corners (pentagon-like), revealed by the five limiting slopes of the geodesics in parameter space. This boundary structure is the same as that of the convex hull of the points in $\Pi$ (black dashed line).}
\end{figure}
\subsection{Models with Symmetries}
\label{sec:symmetries}
The final general class of models we consider are those with discrete symmetries so that the predictions of the model are invariant under some discrete transformation of the parameters. The model manifold is therefore isomorphic to a quotient of the parameter space. Consequently the model manifold will have boundaries associated with the boundaries of the parameter quotient space with boundary points corresponding to fixed points of the parameter transformation.
To make this concept more concrete, consider a model that is invariant to an operation on its parameter space $g$: $P(\theta) = P(g(\theta))$. Let $\theta^*$ be a fixed pointed of $g$ so that $g(\theta^*) = \theta^*$. Let $v$ be a vector denoting a direction in parameter space; we are interested in characterizing the circumstances under which the directional derivative of the model in the direction of $v$ will be zero. If $\partial P/\partial \theta^\mu v^\mu = 0$, then the Fisher Information matrix will be singular in the direction of $v$.
Let $J = \partial g/\partial \theta$ evaluated at $\theta^*$ and let $v$ be an eigenvector of $J$ with eigenvalue $\lambda \neq 1$. We can then write
\begin{equation}
\label{eq:direcderivPG}
\frac{\partial P}{\partial \theta^\mu} v^\mu \biggr\vert_{\theta = \theta^*}= \lim_{h\rightarrow 0} \frac{P(\theta^* + hv) - P( g(\theta^* + hv) )}{h (1 - \lambda)}.
\end{equation}
It is straightforward to check Eq.~\eqref{eq:direcderivPG} by expanding the terms numerator to lowest order in $h$:
\begin{eqnarray}
P(\theta^* + hv) & \approx & P(\theta^*) + h \frac{\partial P}{\partial \theta} v \\
P(g(\theta^* = hv)) & \approx & P\left( g(\theta^*) + h J v \right) \\
& = & P\left( \theta^* + h \lambda v\right) \\
& \approx & P(\theta^*) + h \lambda \frac{\partial P}{\partial \theta} v.
\end{eqnarray}
Invoking the invariance of the model under operation $g$ gives $P(\theta^* + hv) - P( g(\theta^* + hv) = 0$. We therefore conclude that at the fixed point of a symmetry operation, the manifold is non-singular in as many dimensions as $J$ has eigenvalues $\lambda = 1$.
As a concrete example consider the model
\begin{equation}
\label{eq:2exp}
y(t, \theta) = e^{-\theta_1 t} + e^{-\theta_2 t}
\end{equation}
with $\theta_\mu \geq 0$. This model is invariant to the permutation $(\theta_1, \theta_2) \rightarrow (\theta_2, \theta_1)$. The fixed point of this operation corresponds to the line $\theta_1 = \theta_2$. This line maps to a boundary on the model manifold as in Figure 1 in reference\cite{transtrum2011geometry} where it is described as a fold line because the model effectively ``folds'' the parameter space along this line.
To see how this geometry relates to the arguments above, consider the Jacobian of the permutation operation:
\begin{equation}
\label{eq:symmetryjacobian}
J = \left( \begin{array}{cc} 0 & 1 \\
1 & 0
\end{array} \right)
\end{equation}
which has two eigenvalues, $1$ and $-1$. The eigenvector with eigenvalue $\lambda = -1$ is $v = (1, -1)^T$ is the parameter space direction perpendicular to the fold line, i.e., perpendicular to the boundary of \MM. The direction $u = (1,1)$ is a nonsingular direction of the FIM.
Manifolds that locally look like quotient subspaces of $\mathbb{R}^N$ are known as orbifolds. Depending on the symmetries involved, orbifolds can have a rich boundary structure, a discussion of which is beyond the scope of this paper. For some orbifolds with more unusual singularity structures (such as a cone), it may not be possible to identify the boundaries using geodesics as we speculate in section \ref{sec:limitations}. Orbifolds are relevant for this discussion as they are another class of models for which singularities exist that contribute to a hierarchical boundary structure. Furthermore, there exists a rich mathematical theory to study the global, topological properties and singularity structures of such mappings.
\section{Relation to Other Approximation Methods}
\label{sec:relation}
As we have seen, many boundaries of model manifolds correspond to limiting approximations among its parameters. These typically correspond to parameters reaching the limit of their physically allowed values. Approximations of this nature have a long and venerable history in science. MBAM is an attempt to semi-automate this technique so that it may be applied in new contexts and reveal new insights into complex physical phenomena.
The converse of this observation is that the countless examples of limiting approximations that have historically been applied to mathematical models can be reinterpreted as special cases of the manifold boundary approximating method. A continuum limit is a typical example. A continuum theory, such as a field theory, may be arise when a microscopic length scale, such as a lattice constant, is much smaller than the quantities of interest. The field theory emerges in the limit that the lattice constant becomes zero, i.e., the boundary of the model manifold. Similar arguments hold for thermodynamic limits (limits of infinite system size) and various classical limits (limits in which Planck's constant become zero or the speed of light become infinite).
Although not as obvious, the MBAM is also related to the Renormalization Group (RG). In particular, note that the Ising model falls into the class of models described in section~\ref{sec:loglinear}, and the Markov random field in section \ref{sec:networks} is an Ising model on a network. Recall that the Kadanoff block-spin renormalization procedure involves an iterative clustering of spins, not unlike the clustering illustrated in Figure~\ref{fig:mrfexmaple}. Furthermore, the process of marginalizing a field theory over the highest energy field configurations is analogous to the approximation that those configurations occur with probability zero. In section~\ref{sec:loglinear}, we showed that models with Hamiltonians linear in the parameters have boundaries that similarly remove a sequence of high-energy configurations from the model. These superficial connections to RG methods will be explored in more depth elsewhere.
\section{Potential Limitations and Pitfalls}
\label{sec:limitations}
One of the potential limitations to practical application of the MBAM is the computational cost. Models with many parameters tend to be very complex and computationally expensive. The most expensive part of the MBAM calculation is repeated calculation of the FIM along the geodesic. In some models, calculating the derivatives with respect to all of the parameters may be prohibitive. In other models, it may not be possible to estimate the required expectation values. Where possible, the calculation of derivatives for the FIM is trivial to parallelize, so the overall calculation scales well with system size. In our experience, we have successfully applied MBAM to models based on differential equations with hundreds of parameters. Models of the form of Eq.~\eqref{eq:loglinear} can be explored with even more parameters.
Unlike other, automatic model reduction methods, MBAM is not fully algorithmic. While most of the steps can be automated, MBAM (at least in its current form) requires human intervention to identify and evaluate the limits in the model. MBAM is therefore only semi-automatic. This step is key to MBAM for at least two reasons. First, it allows MBAM to be very general, effectively adapting itself to the functional form of the model. Second, it generates theoretical insight into the behavior of the complex model. The need for human intervention may often be the limiting factor in the size of models tractable by MBAM. We believe that as the manifold boundaries of specific model classes are better understood, that this problem may be mitigated. For example, it would be straightforward to fully automate the method for the model class of Eq.~\eqref{eq:loglinear}. Similarly, we find that in models of chemical kinetics described by differential equations that similar types of limits are repeated so that it may be possible to automate their evaluation and minimize the need for human guidance.
Calculating a geodesic is incidental to the implementation MBAM. It is a useful tool for finding boundaries, but for some cases there may be other methods. For example, for models of the form in Eq.~\eqref{eq:loglinear}, convex hull algorithms applied to $\Pi$ could be used to construct reduced models. In other cases, it may be possible to theoretically identify relevant boundaries based on symmetries or other arguments. In some cases, the boundary approximation may be the theoretically desired result although it may be hard to find.
A key to managing the computational cost of the MBAM is identifying a global property of the model manifold (i.e., the boundary) using a sequence of local calculations (i.e., the FIM). MBAM therefore exploits a nontrivial relationship between global structure and local information that was first identified in reference\cite{transtrum2010nonlinear} (see Figure 3 specifically). In order for the eigenvalues of the FIM to reflect the model's global structure, the parameters of the model need to be cast in their natural units. In practice, we do this by transforming to log-parameters. In some cases, it may be difficult to find a convenient parameterization that makes the FIM eigenvalues useful. If parameters are poorly scaled, the procedure may encounter difficulties.
MBAM requires that the model manifold have a hierarchy of boundaries. A major part of this paper was devoted to exploring this structure for a wide variety of models. Although we have found this structure to be common, one can imagine scenarios that could be problematic for the procedure (i.e., geodesics may not easily identify the desired boundaries) or in which the boundaries do not exist at all.
It is possible that a model may have the desired structure but that it can't be identified by geodesics. Models with high curvature, for example, may divert the geodesic away from the desired boundary (analogous to a gravitational slingshot). Another possibility is that the manifold structure may break down at some points. For example, a cone is a two-dimensional surface that is bounded, not by a one-dimensional line, but by a point. A singularity of this type could easily arise as the consequence of a symmetry (as in section~\ref{sec:symmetries}). It may be desirable to approximate such a ``cone'' by its apex, but a geodesic will circle the cone indefinitely without finding the desired point.
In some cases a manifold may be bounded but not have a hierarchy of boundaries. For example, a model manifold may look more like a sphere than a polyhedron. In other cases, a manifold may have a ``soft'' boundary, e.g., if one cannot put a hard limit on physical range of a parameter.
Unbounded manifold are also potentially problematic. We have seen that it is not necessary for the model manifold to be strictly bounded. The model in Eq.~\eqref{eq:exponentials} is unbounded, but MBAM is still applicable because the cross-sections have the hierarchical structure we seek. However, if the model has no bounded cross sections, then MBAM will fail. The simplest example of this is a linear least squares model. The model manifold is a hyper-plane with no effective low-dimensional structure to approximate. Another example of an unbounded model manifold is a Markov Random Field (as in section~\ref{sec:networks}) with normally distributed random variables (so that $A$ in Eq.~\eqref{eq:MRFPrecisionMatrix} is the inverse covariance matrix).
Necessary and sufficient conditions to guarantee the hierarchical boundary structure remains an open problem; a major purpose of this paper is to demonstrate empirically that is shared by a wide variety of models.
\section{Discussion}
\label{sec:Discussion}
In this paper, we have presented a detailed description of the Manifold Boundary Approximation Method (MBAM). The primary motivation of this model reduction method is to provide a ``gray-box'' representation of a model that is able to bridge the gap between microscopic mechanisms and a system's collective behavior. It aims to exploit a low-effective dimensionality in the behavior space of an overly parameterized model, using geodesics to find a series of limiting approximations that minimally impact the behavior of the system.
One of the interesting aspects of the MBAM is its potential to guide theoretical studies of complex systems. Complex models are often difficult to interpret and ascribing a mechanistic origin to particular behaviors is challenging. It is often the case that experimental or engineering control knobs operate on a microscopic, mechanistic level of the system. In these cases (including much of material science, biology, neuroscience, etc), a way to effectively connect the complicated mechanistic description to a simple phenomenological description would be useful.
In these cases, effective model reduction does not simply fix a technical problem that arises from not having a big enough computer. Rather, it is an ongoing process that provides theoretical insight and refines one's view of the system. The whole process ought to provide feedback to the original, complicated model that incorporates the insights gained from the simple representation. Because of its ``gray-box'' structure, MBAM is an effective complement to other automatic model reduction methods that produce ``black-box'' approximations in a way that accommodates this sort of ongoing refinement.
One of the curious requirements of MBAM is that the model manifold have a hierarchical structure of boundaries. One of the primary purposes of this paper is to demonstrate that this structure is common to many models. Although there are counter examples (as in section \ref{sec:limitations}), it is surprising that this structure is so common. The reason for the ubiquity of this structure remains an open question.
It is also interesting to note that many common methods for constructing effective mechanisms (such as continuum limits, singular perturbation, steady-state and partial equilibrium approximations) implicitly use this hierarchical boundary structure. By identifying these diverse approximations as a common geometric operation, MBAM is a step toward unifying and automating this process. Indeed, one of the challenges to constructing reduced representations in complex systems is that one cannot easily identify a priori which approximations, among the many choices, one should make. Constructing approximate models therefore requires expert guidance built on years or decades of hard-won intuition. MBAM is a promising tool for identifying these approximations automatically and in a reproducible, mathematically rigorous way. Application to complex models may reveal previously unknown classes of approximations. Our hope is that MBAM may motivate useful approximations that in turn reveal which mechanisms govern complex nonlinear phenomena.
\textbf{Acknowledgments:} The author thanks Peng Qiu for helpful discussions and Gus Hart for providing the $\Pi$ matrix in Figure~\ref{fig:fccgeodesics}.
|
\section{Introduction}
\label{sec:introduction}
In this paper we will use the Euler-Maclaurin summation formula \cite{3,5} to obtain rapidly convergent series expansions for finite sums involving Stirling series \cite{1}. Our key tool will be the so called Weniger transformation \cite{1}.\newline
\noindent For example, one of our summation formulas for the sum $\sum_{k=0}^{\lfloor x\rfloor}\sqrt{k}$, where $x\in\mathbb{R}^{+}$ is
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}\sqrt{k}&=\frac{2}{3}x^{\frac{3}{2}}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)-\sqrt{x}B_{1}(\{x\})+\sqrt{x}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}(-1)^{l}\frac{(2l-3)!!}{2^{l}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{(x+1)(x+2)\cdots(x+k)}\\
&=\frac{2}{3}x^{\frac{3}{2}}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)+\left(\frac{1}{2}-\{x\}\right)\sqrt{x}+\frac{\left(\frac{1}{4}\{x\}^{2}-\frac{1}{4}\{x\}+\frac{1}{24}\right)\sqrt{x}}{(x+1)}\\
&\quad+\frac{\left(\frac{1}{24}\{x\}^{3}+\frac{3}{16}\{x\}^{2}-\frac{11}{48}\{x\}+\frac{1}{24}\right)\sqrt{x}}{(x+1)(x+2)}+\frac{\left(\frac{1}{64}\{x\}^{4}+\frac{3}{32}\{x\}^{3}+\frac{21}{64}\{x\}^{2}-\frac{7}{16}\{x\}+\frac{53}{640}\right)\sqrt{x}}{(x+1)(x+2)(x+3)}\\
&\quad+\frac{\left(\frac{1}{128}\{x\}^{5}+\frac{19}{256}\{x\}^{4}+\frac{109}{384}\{x\}^{3}+\frac{29}{32}\{x\}^{2}-\frac{977}{768}\{x\}+\frac{79}{320}\right)\sqrt{x}}{(x+1)(x+2)(x+3)(x+4)}+\ldots,
\end{split}
\end{displaymath}
where the $B_{l}(x)$'s are the Bernoulli polynomials and $S_{k}^{(1)}(l)$ denotes the Stirling numbers of the first kind.
\vspace{0.4cm}
\noindent Most of the other formulas in this article have a similar shape.\newline
\noindent Setting in the above formula for $\sum_{k=0}^{\lfloor x \rfloor}\sqrt{k}$ the variable $x:=n\in\mathbb{N}$, we obtain
\begin{displaymath}
\begin{split}
\sum_{k=0}^{n}\sqrt{k}&=\frac{2}{3}n^{\frac{3}{2}}+\frac{1}{2}\sqrt{n}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)+\sqrt{n}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}(-1)^{l}\frac{(2l-3)!!}{2^{l}(l+1)!}B_{l+1}S_{k}^{(1)}(l)}{(n+1)(n+2)\cdots(n+k)}\\
&=\frac{2}{3}n^{\frac{3}{2}}+\frac{1}{2}\sqrt{n}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)+\frac{\sqrt{n}}{24(n+1)}+\frac{\sqrt{n}}{24(n+1)(n+2)}+\frac{53\sqrt{n}}{640(n+1)(n+2)(n+3)}\\
&\quad+\frac{79\sqrt{n}}{320(n+1)(n+2)(n+3)(n+4)}+\ldots,
\end{split}
\end{displaymath}
\noindent which is the corresponding formula given in our previous paper \cite{8}.
\section{Definitions}
\label{sec:Definitions}
\noindent As usual, we denote the floor of $x$ by $\lfloor x\rfloor$ and the fractional part of $x$ by $\{x\}$.
\begin{definition}(Pochhammer symbol)\cite{1}\newline
\noindent We define the \emph{Pochhammer symbol} (or rising factorial function) $(x)_{k}$ by
\begin{displaymath}
(x)_{k}:=x(x+1)(x+2)(x+3)\cdots(x+k-1)=\frac{\Gamma(x+k)}{\Gamma(x)},
\end{displaymath}
where $\Gamma(x)$ is the gamma function defined by
\begin{displaymath}
\Gamma(x):=\int_{0}^{\infty}e^{-t}t^{x-1}dt.
\end{displaymath}
\end{definition}
\begin{definition}(Stirling numbers of the first kind)\cite{1}\newline
\noindent Let $k,l\in\mathbb{N}_{0}$ be two non-negative integers such that $k\geq l\geq0$. We define the \emph{Stirling numbers of the first kind} $S_{k}^{(1)}(l)$ as the connecting coefficients in the identity
\begin{displaymath}
(x)_{k}=(-1)^{k}\sum_{l=0}^{k}(-1)^{l}S_{k}^{(1)}(l)x^{l},
\end{displaymath}
where $(x)_{k}$ is the rising factorial function.
\end{definition}
\begin{definition}(Bernoulli numbers)\cite{2}\newline
\noindent We define the $k$-th \emph{Bernoulli number} $B_{k}$ as the $k$-th coefficient in the generating function relation
\begin{displaymath}
\frac{x}{e^{x}-1}=\sum_{k=0}^{\infty}\frac{B_{k}}{k!}x^{k}\;\;\forall x\in\mathbb{C}\;\text{with $|x|<2\pi$}.
\end{displaymath}
\end{definition}
\begin{definition}(Euler numbers)\cite{3}\newline
\noindent We define the sequence of \emph{Euler numbers} $\left\{E_{k}\right\}_{k=0}^{\infty}$ by the generating function identity
\begin{displaymath}
\frac{2e^{x}}{e^{2x}+1}=\sum_{k=0}^{\infty}\frac{E_{k}}{k!}x^{k}.
\end{displaymath}
\end{definition}
\begin{definition}(Bernoulli polynomials)\cite{2,3}\newline
\noindent We define for $n\in\mathbb{N}_{0}$ the $n$-th \emph{Bernoulli polynomial} $B_{n}(x)$ via the following exponential generating function as
\begin{displaymath}
\frac{te^{xt}}{e^{t}-1}=\sum_{n=0}^{\infty}\frac{B_{n}(x)}{n!}t^{n}\;\;\forall t\in\mathbb{C}\;\text{with $|t|<2\pi$}.
\end{displaymath}
\noindent Using this expression, we see that the value of the $n$-th Bernoulli polynomial $B_{n}(x)$ at the point $x=0$ is
\begin{displaymath}
B_{n}(0)=B_{n},
\end{displaymath}
which is the $n$-th Bernoulli number.
\end{definition}
\begin{definition}(Euler polynomials)\cite{3}\newline
\noindent We define for $n\in\mathbb{N}_{0}$ the $n$-th \emph{Euler polynomial} $E_{n}(x)$ via the following exponential generating function as
\begin{displaymath}
\frac{2e^{xt}}{e^{t}+1}=\sum_{n=0}^{\infty}\frac{E_{n}(x)}{n!}t^{n}.
\end{displaymath}
\noindent Moreover, we have that \cite{4}
\begin{displaymath}
\begin{split}
E_{n}(0)&=-2(2^{n+1}-1)\frac{B_{n+1}}{n+1}\;\;\forall n\in\mathbb{N}_{0}.
\end{split}
\end{displaymath}
\end{definition}
\section{Extended Summation Formulas involving\\
Stirling Series}
\label{sec:Extended Summation Formulas involving Stirling Series}
In this section we will prove our summation formulas for various finite sums of the form $\sum_{k=1}^{\lfloor x\rfloor}f(k)$.
For this, we need the following
\begin{lemma}(Extended Euler-Maclaurin summation formula)\cite{3,5}\newline
\noindent Let $f$ be an analytic function. Then for all $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}f(k)&=\int_{1}^{x}f(t)dt+\sum_{k=1}^{m}(-1)^{k}\frac{B_{k}(\{x\})}{k!}f^{(k-1)}(x)-\sum_{k=1}^{m}\frac{B_{k}}{k!}f^{(k-1)}(1)\\
&\quad+\frac{(-1)^{m+1}}{m!}\int_{1}^{x}B_{m}(\{t\})f^{(m)}(t)dt,
\end{split}
\end{displaymath}
where $B_{m}(x)$ is the $m$-th Bernoulli polynomial and $\{x\}$ denotes the fractional part of $x$. Therefore, for many functions $f$ we have the asymptotic expansion
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}f(k)\sim\int_{1}^{x}f(t)dt+C+\sum_{k=1}^{\infty}(-1)^{k}\frac{B_{k}(\{x\})}{k!}f^{(k-1)}(x)\;\;\text{as $x\rightarrow\infty$},
\end{split}
\end{displaymath}
for some constant $C\in\mathbb{C}$
\end{lemma}
\noindent and
\begin{lemma}(Extended Boole summation formula)\cite{3}\newline
\noindent Let $f$ be an analytic function. Then for all $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}(-1)^{k+1}f(k)&=\frac{(-1)^{x-\{x\}}}{2}\sum_{k=0}^{m}(-1)^{k+1}\frac{E_{k}(\{x\})}{k!}f^{(k)}(x)-\sum_{k=0}^{m}\frac{(2^{k+1}-1)B_{k+1}}{(k+1)!}f^{(k)}(1)\\
&\quad+\frac{(-1)^{m}}{2m!}\int_{1}^{x}(-1)^{t-\{t\}}E_{m}(\{t\})f^{(m+1)}(t)dt,
\end{split}
\end{displaymath}
where $E_{m}(x)$ is the $m$-th Euler polynomial and $\{x\}$ denotes the fractional part of $x$.\newline
\noindent Therefore, for many functions $f$ we have the asymptotic expansion
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}(-1)^{k+1}f(k)\sim C+\frac{(-1)^{x-\{x\}}}{2}\sum_{k=0}^{\infty}(-1)^{k+1}\frac{E_{k}(\{x\})}{k!}f^{(k)}(x)\;\;\text{as $x\rightarrow\infty$},
\end{split}
\end{displaymath}
for some constant $C\in\mathbb{C}$.
\end{lemma}
\noindent From the above lemma we get by setting $x:=n\in\mathbb{N}$ the following
\begin{corollary}(Boole Summation Formula)\cite{3}\newline
\noindent Let $f$ be an analytic function. Then for all $n\in\mathbb{N}$, we have
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}(-1)^{k+1}f(k)&=(-1)^{n}\sum_{k=0}^{m}(-1)^{k}\frac{(2^{k+1}-1)B_{k+1}}{(k+1)!}f^{(k)}(n)-\sum_{k=0}^{m}\frac{(2^{k+1}-1)B_{k+1}}{(k+1)!}f^{(k)}(1)\\
&\quad+\frac{(-1)^{m}}{2m!}\int_{1}^{x}(-1)^{t-\{t\}}E_{m}(\{t\})f^{(m+1)}(t)dt,
\end{split}
\end{displaymath}
\noindent and the following asymptotic expansion
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}(-1)^{k+1}f(k)\sim C+(-1)^{n}\sum_{k=0}^{\infty}(-1)^{k}\frac{(2^{k+1}-1)B_{k+1}}{(k+1)!}f^{(k)}(n)\;\;\text{as $n\rightarrow\infty$},
\end{split}
\end{displaymath}
for some constant $C\in\mathbb{C}$.
\end{corollary}
\noindent As well as the next key result found by J. Weniger:
\begin{lemma}(Generalized Weniger transformation)\cite{1}\newline
\noindent For every inverse power series $\sum_{k=1}^{\infty}\frac{a_{k}(x)}{x^{k+1}}$, where $a_{k}(x)$ is any function in $x$, the following transformation formula holds
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\infty}\frac{a_{k}(x)}{x^{k+1}}&=\sum_{k=1}^{\infty}\frac{(-1)^{k}}{(x)_{k+1}}\sum_{l=1}^{k}(-1)^{l}S_{k}^{(1)}(l)a_{l}(x)\\
&=\sum_{k=1}^{\infty}\frac{(-1)^{k}\sum_{l=1}^{k}(-1)^{l}S_{k}^{(1)}(l)a_{l}(x)}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}
\end{lemma}
\noindent Now, we prove as an example the following
\begin{theorem}(Extended summation formulas for the harmonic series)\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}&=\log(x)+\gamma-\frac{B_{1}(\{x\})}{x}+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{l+1}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}
\noindent We also have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}&=\log(x)+\gamma+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{l}S_{k}^{(1)}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}
\end{theorem}
\begin{proof}
Applying the extended Euler-Maclaurin summation formula to the function $f(x):=\frac{1}{x}$, we get that
\begin{displaymath}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}\sim\log(x)+\gamma-\sum_{k=1}^{\infty}\frac{B_{k}(\{x\})}{kx^{k}}.
\end{displaymath}
Applying now the generalized Weniger transformation to the equivalent series
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}&\sim\log(x)+\gamma-\frac{B_{1}(\{x\})}{x}-\sum_{k=2}^{\infty}\frac{B_{k}(\{x\})}{kx^{k}}\\
&\sim\log(x)+\gamma-\frac{B_{1}(\{x\})}{x}-\sum_{k=1}^{\infty}\frac{B_{k+1}(\{x\})}{(k+1)x^{k+1}},
\end{split}
\end{displaymath}
we get the first claimed formula.
To obtain the second expression, we apply the generalized Weniger transformation to the identity
\begin{displaymath}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}\sim\log(x)+\gamma-x\sum_{k=1}^{\infty}\frac{B_{k}(\{x\})}{kx^{k+1}}.
\end{displaymath}
\end{proof}
\noindent At this point, we want to give an overview on summation formulas obtained with this method:
\begin{itemize}
\item[1.)]{{\bf Extended summation formulas for the harmonic series:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}&=\log(x)+\gamma-\frac{B_{1}(\{x\})}{x}+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{l+1}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}&=\log(x)+\gamma+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{l}S_{k}^{(1)}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[2.)]{{\bf Extended summation formulas for the partial sums of $\zeta(2)$:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k^{2}}&=\zeta(2)-\frac{1}{x}-\frac{B_{1}(\{x\})}{x^{2}}+\frac{1}{x}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}(-1)^{l}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k^{2}}&=\zeta(2)-\frac{1}{x}+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}(-1)^{l}S_{k}^{(1)}(l)B_{l}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[3.)]{{\bf Extended summation formulas for the partial sums of $\zeta(3)$:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k^{3}}&=\zeta(3)-\frac{1}{2x^{2}}-\frac{B_{1}(\{x\})}{x^{3}}+\frac{1}{2x^{2}}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}(-1)^{l}(l+2)S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k^{3}}&=\zeta(3)-\frac{1}{2x^{2}}+\frac{1}{2}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}(-1)^{l}(l+1)S_{k}^{(1)}(l)B_{l}(\{x\})}{x^{2}(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[4.)]{{\bf Extended summation formulas for the sum of the square roots:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}\sqrt{k}&=\frac{2}{3}x^{\frac{3}{2}}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)-\sqrt{x}B_{1}(\{x\})+\sqrt{x}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-3)!!}{2^{l}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{(x+1)(x+2)\cdots(x+k)},
\end{split}
\end{displaymath}
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}\sqrt{k}&=\frac{2}{3}x^{\frac{3}{2}}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)-\sqrt{x}B_{1}(\{x\})+\frac{B_{2}(\{x\})}{4\sqrt{x}}+\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-1)!!}{2^{l+1}(l+2)!}S_{k}^{(1)}(l)B_{l+2}(\{x\})}{\sqrt{x}(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}\sqrt{k}&=\frac{2}{3}x^{\frac{3}{2}}-\frac{1}{4\pi}\zeta\left(\frac{3}{2}\right)+x\sqrt{x}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-5)!!}{2^{l-1}l!}S_{k}^{(1)}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[5.)]{{\bf Extended summation formulas for the partial sums of $\zeta(-3/2)$:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k\sqrt{k}&=\frac{2}{5}x^{\frac{5}{2}}-\frac{3}{16\pi^{2}}\zeta\left(\frac{5}{2}\right)-x^{\frac{3}{2}}B_{1}(\{x\})+\frac{3}{2}x^{\frac{3}{2}}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-5)!!}{2^{l-1}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{(x+1)(x+2)\cdots(x+k)},
\end{split}
\end{displaymath}
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k\sqrt{k}&=\frac{2}{5}x^{\frac{5}{2}}-\frac{3}{16\pi^{2}}\zeta\left(\frac{5}{2}\right)-x^{\frac{3}{2}}B_{1}(\{x\})+\frac{3}{4}\sqrt{x}B_{2}(\{x\})-\frac{B_{3}(\{x\})}{4\sqrt{x}}\\
&\quad+\frac{3}{2}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-1)!!}{2^{l+1}(l+3)!}S_{k}^{(1)}(l)B_{l+3}(\{x\})}{\sqrt{x}(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k\sqrt{k}&=\frac{2}{5}x^{\frac{5}{2}}-\frac{3}{16\pi^{2}}\zeta\left(\frac{5}{2}\right)+\frac{3}{2}x^{\frac{5}{2}}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-7)!!}{2^{l-2}l!}S_{k}^{(1)}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}
}
\item[6.)]{{\bf Extended summation formulas for the partial sums of $\zeta(-5/2)$:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k^{2}\sqrt{k}&=\frac{2}{7}x^{\frac{7}{2}}+\frac{15}{64\pi^{3}}\zeta\left(\frac{7}{2}\right)-x^{\frac{5}{2}}B_{1}(\{x\})+\frac{15}{4}x^{\frac{5}{2}}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-7)!!}{2^{l-2}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{(x+1)(x+2)\cdots(x+k)},
\end{split}
\end{displaymath}
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k^{2}\sqrt{k}&=\frac{2}{7}x^{\frac{7}{2}}+\frac{15}{64\pi^{3}}\zeta\left(\frac{7}{2}\right)-x^{\frac{5}{2}}B_{1}(\{x\})+\frac{5}{4}x^{\frac{3}{2}}B_{2}(\{x\})-\frac{5}{8}\sqrt{x}B_{3}(\{x\})+\frac{5B_{4}(\{x\})}{64\sqrt{x}}\\
&\quad+\frac{15}{4}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-1)!!}{2^{l+1}(l+4)!}S_{k}^{(1)}(l)B_{l+4}(\{x\})}{\sqrt{x}(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k^{2}\sqrt{k}&=\frac{2}{7}x^{\frac{7}{2}}+\frac{15}{64\pi^{3}}\zeta\left(\frac{7}{2}\right)+\frac{15}{4}x^{\frac{7}{2}}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-9)!!}{2^{l-3}l!}S_{k}^{(1)}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[7.)]{{\bf Extended summation formulas for the sum of the inverse square roots:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{\sqrt{k}}
&=2\sqrt{x}+\zeta\left(\frac{1}{2}\right)-\frac{B_{1}(\{x\})}{\sqrt{x}}+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-1)!!}{2^{l}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{\sqrt{x}(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{\sqrt{k}}
&=2\sqrt{x}+\zeta\left(\frac{1}{2}\right)+\sqrt{x}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-3)!!}{2^{l-1}l!}S_{k}^{(1)}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[8.)]{{\bf Extended summation formulas for the partial sums of $\zeta(3/2)$:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k\sqrt{k}}
&=\zeta\left(\frac{3}{2}\right)-\frac{2}{\sqrt{x}}-\frac{B_{1}(\{x\})}{x\sqrt{x}}+\frac{2}{\sqrt{x}}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l+1)!!}{2^{l+1}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k\sqrt{k}}
&=\zeta\left(\frac{3}{2}\right)-\frac{2}{\sqrt{x}}+2\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l-1)!!}{2^{l}l!}S_{k}^{(1)}(l)B_{l}(\{x\})}{\sqrt{x}(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[9.)]{{\bf Extended summation formulas for the partial sums of $\zeta(5/2)$:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k^{2}\sqrt{k}}&=\zeta\left(\frac{5}{2}\right)-\frac{2}{3x^{\frac{3}{2}}}-\frac{B_{1}(\{x\})}{x^{2}\sqrt{x}}+\frac{4}{3x\sqrt{x}}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l+3)!!}{2^{l+2}(l+1)!}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k^{2}\sqrt{k}}&=\zeta\left(\frac{5}{2}\right)-\frac{2}{3x^{\frac{3}{2}}}+\frac{4}{3\sqrt{x}}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}(2l+1)!!}{2^{l+1}l!}S_{k}^{(1)}(l)B_{l}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[10.)]{{\bf Extended Generalized Faulhaber Formulas:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$ and for every complex number $m\in\mathbb{C}\setminus\{-1\}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}k^{m}=\frac{1}{m+1}x^{m+1}+\zeta\left(-m\right)+\frac{x^{m+1}}{m+1}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}{m+1\choose l}S^{(1)}_{k}(l)B_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
\noindent and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}k^{m}=\frac{1}{m+1}x^{m+1}+\zeta\left(-m\right)-x^{m}B_{1}(\{x\})+\frac{x^{m}}{m+1}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}{m+1\choose l+1}S^{(1)}_{k}(l)B_{l+1}(\{x\})}{(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
\noindent and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}k^{m}&=\frac{1}{m+1}x^{m+1}+\zeta\left(-m\right)+\frac{1}{m+1}\sum_{k=1}^{\lfloor m+1\rfloor}(-1)^{k}{m+1\choose k}B_{k}(\{x\})x^{m-k+1}\\
&\quad+(-1)^{\lceil m+1\rceil}\frac{x^{m-\lceil m+1\rceil+2}}{m+1}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}{m+1\choose l+\lceil m+1\rceil-1}S^{(1)}_{k}(l)B_{l+\lceil m+1\rceil-1}(\{x\})}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}
}
\item[11.)]{{\bf Generalized Faulhaber Formulas:}\newline
\noindent For every complex number $m\in\mathbb{C}\setminus\{-1\}$ and every natural number $n\in\mathbb{N}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}k^{m}=\frac{1}{m+1}n^{m+1}+\zeta\left(-m\right)+\frac{n^{m+1}}{m+1}\sum_{k=1}^{\infty}\frac{(-1)^{k}\sum_{l=1}^{k}{m+1\choose l}B_{l}S^{(1)}_{k}(l)}{(n+1)(n+2)\cdots(n+k)}
\end{split}
\end{displaymath}
\noindent and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}k^{m}=\frac{1}{m+1}n^{m+1}+\zeta\left(-m\right)+\frac{1}{2}n^{m}+\frac{n^{m}}{m+1}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}{m+1\choose l+1}B_{l+1}S^{(1)}_{k}(l)}{(n+1)(n+2)\cdots(n+k)}
\end{split}
\end{displaymath}
\noindent and
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}k^{m}&=\frac{1}{m+1}n^{m+1}+\zeta\left(-m\right)+\frac{1}{m+1}\sum_{k=1}^{\lfloor m+1\rfloor}(-1)^{k}{m+1\choose k}B_{k}n^{m-k+1}\\
&\quad+(-1)^{\lceil m+1\rceil}\frac{n^{m-\lceil m+1\rceil+2}}{m+1}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}{m+1\choose l+\lceil m+1\rceil-1}B_{l+\lceil m+1\rceil-1}S^{(1)}_{k}(l)}{(n+1)(n+2)\cdots(n+k)}.
\end{split}
\end{displaymath}
}
\item[12.)]{{\bf Extended convergent versions of Stirling's formula:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\log(k)&=x\log(x)-x+\frac{1}{2}\log\left(2\pi\right)-\log(x)B_{1}(\{x\})+\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{l(l+1)}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{(x+1)(x+2)\cdots(x+k)},
\end{split}
\end{displaymath}
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\log(k)&=x\log(x)-x+\frac{1}{2}\log\left(2\pi\right)-\log(x)B_{1}(\{x\})+\frac{B_{2}(\{x\})}{2x}\\
&\quad+\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{(l+1)(l+2)}S_{k}^{(1)}(l)B_{l+2}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\log(k)&=x\log(x)-x+\frac{1}{2}\log\left(2\pi\right)-\log(x)B_{1}(\{x\})+x\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=2}^{k}\frac{(-1)^{l}}{l(l-1)}S_{k}^{(1)}(l)B_{l}(\{x\})
}{(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[13.)]{{\bf Extended first logarithmic summation formulas:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k\log(k)&=\frac{1}{2}x^{2}\log(x)-\frac{1}{4}x^{2}+\frac{1}{12}-\zeta^{'}(-1)-x\log(x)B_{1}(\{x\})+\frac{1}{2}\log(x)B_{2}(\{x\})\\
&\quad+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{l(l+1)(l+2)}S_{k}^{(1)}(l)B_{l+2}(\{x\})}{(x+1)(x+2)\cdots(x+k)}
\end{split}
\end{displaymath}
and
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}k\log(k)&=\frac{1}{2}x^{2}\log(x)-\frac{1}{4}x^{2}+\frac{1}{12}-\zeta^{'}(-1)-x\log(x)B_{1}(\{x\})+\frac{1}{2}\log(x)B_{2}(\{x\})-\frac{B_{3}(\{x\})}{6x}\\
&\quad+\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{(l+1)(l+2)(l+3)}S_{k}^{(1)}(l)B_{l+3}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[14.)]{{\bf Extended second logarithmic summation formula:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{\log(k)}{k}&=\frac{1}{2}\log(x)^{2}+\gamma_{1}-\frac{\log(x)}{x}B_{1}(\{x\})+\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{(l+1)!}S_{l+1}^{(1)}(2)S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+1)\cdots(x+k)}\\
&\quad+\log(x)\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{1}{l+1}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[15.)]{{\bf Extended third logarithmic summation formula:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{\log(k)}{k^{2}}&=-\zeta^{'}(2)-\frac{\log(x)}{x}-\frac{1}{x}-\frac{\log(x)}{x^{2}}B_{1}(\{x\})\\
&\quad+\frac{1}{x}\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\sum_{l=1}^{k}\frac{1}{l+1}\left(\sum_{m=0}^{l-1}\frac{m+1}{l-m}\right)S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}\\
&\quad+\frac{\log(x)}{x}\sum_{k=1}^{\infty}\frac{(-1)^{k}\sum_{l=1}^{k}S_{k}^{(1)}(l)B_{l+1}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[16.)]{{\bf Extended fourth logarithmic summation formula:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\log(k)^2&=x\log(x)^{2}-2x\log(x)+2x+\frac{\gamma^{2}}{2}-\frac{\pi^{2}}{24}-\frac{\log(2)^{2}}{2}-\log(2)\log(\pi)-\frac{\log(\pi)^{2}}{2}+\gamma_{1}\\
&\quad-\log(x)^{2}B_{1}(\{x\})+\frac{\log(x)}{x}B_{2}(\{x\})+2\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{(-1)^{l}}{(l+2)!}S_{l+1}^{(1)}(2)S_{k}^{(1)}(l)B_{l+2}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}\\
&\quad+2\log(x)\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{1}{(l+1)(l+2)}S_{k}^{(1)}(l)B_{l+2}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}}
\item[17.)]{{\bf Extended summation formula for partial sums of the Gregory-Leibniz series:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}\frac{(-1)^{k}}{2k+1}&=\frac{\pi}{4}+\frac{(-1)^{x-\{x\}}}{2}\sum_{k=0}^{\infty}\frac{(-1)^{k}\sum_{l=0}^{k}(-1)^{l}2^{l}S^{(1)}_{k}(l)E_{l}(\{x\})}{(2x+1)(2x+2)(2x+3)\cdots(2x+k+1)}.
\end{split}
\end{displaymath}
\noindent Setting $x:=n\in\mathbb{N}$, we get that
\begin{displaymath}
\begin{split}
\sum_{k=0}^{n}\frac{(-1)^{k}}{2k+1}&=\frac{\pi}{4}+(-1)^{n+1}\sum_{k=0}^{\infty}\frac{(-1)^{k}\sum_{l=0}^{k}\frac{(-1)^{l}}{l+1}2^{l}(2^{l+1}-1)S^{(1)}_{k}(l)B_{l+1}}{(2n+1)(2n+2)(2n+3)\cdots(2n+k+1)}.
\end{split}
\end{displaymath}}
\item[18.)]{{\bf Extended formula for the partial sums of the alternating harmonic series:}\newline
\noindent For every positive real number $x\in\mathbb{R}^{+}$, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{(-1)^{k+1}}{k}&=\log(2)+\frac{(-1)^{x-\{x\}}}{2}\sum_{k=0}^{\infty}(-1)^{k+1}\frac{\sum_{l=0}^{k}(-1)^{l}S^{(1)}_{k}(l)E_{l}(\{x\})}{x(x+1)(x+2)\cdots(x+k)}.
\end{split}
\end{displaymath}
\noindent Setting $x:=n\in\mathbb{N}$, we get
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}\frac{(-1)^{k+1}}{k}&=\log(2)+(-1)^{n}\sum_{k=0}^{\infty}(-1)^{k}\frac{\sum_{l=0}^{k}\frac{(-1)^{l}}{l+1}(2^{l+1}-1)S^{(1)}_{k}(l)B_{l+1}}{n(n+1)(n+2)\cdots(n+k)}.
\end{split}
\end{displaymath}}
\end{itemize}
\section{Other Extended Summation Formulas for Finite Sums}
\label{sec:Other Extended Summation Formulas for Finite Sums}
\noindent In this section, we denote by $\eta(s):=\sum_{k=1}^{\infty}\frac{(-1)^{k+1}}{k^{s}}$ the Dirichlet eta function.
\begin{itemize}
\item[1.)]{{\bf The extended alternating Faulhaber formula:}\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}(-1)^{k+1}k^{m}&=\eta(-m)+\frac{(-1)^{x-\{x\}}}{2(m+1)}\sum_{k=0}^{m}(-1)^{k+1}(k+1){m+1\choose k+1}E_{k}(\{x\})x^{m-k}\;\;\forall m\in\mathbb{N}_{0}.
\end{split}
\end{displaymath}
\noindent Setting $x:=n\in\mathbb{N}$, we get
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}(-1)^{k+1}k^{m}&=\eta(-m)+\frac{(-1)^{n}}{m+1}\sum_{k=0}^{m}(-1)^{k}(2^{k+1}-1){m+1\choose k+1}B_{k+1}n^{m-k}\;\;\forall m\in\mathbb{N}_{0}.
\end{split}
\end{displaymath}}
\item[2.)]{{\bf The extended generalized alternating Faulhaber formula:}
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}(-1)^{k+1}k^{m}&=\eta(-m)+\frac{(-1)^{x-\{x\}}x^{m}}{2(m+1)}\sum_{k=0}^{\infty}(-1)^{k+1}\frac{\sum_{l=0}^{k}(l+1){m+1\choose l+1}S^{(1)}_{k}(l)E_{l}(\{x\})}{(x+1)(x+2)\cdots(x+k)}\;\;\forall m\in\mathbb{C}.
\end{split}
\end{displaymath}
\noindent Setting $x:=n\in\mathbb{N}$, we get
\begin{displaymath}
\begin{split}
\sum_{k=1}^{n}(-1)^{k+1}k^{m}&=\eta(-m)+\frac{(-1)^{n}n^{m}}{(m+1)}\sum_{k=0}^{\infty}(-1)^{k}\frac{\sum_{l=0}^{k}{m+1\choose l+1}(2^{l+1}-1)S^{(1)}_{k}(l)B_{l+1}}{(n+1)(n+2)\cdots(n+k)}\;\;\forall m\in\mathbb{C}.
\end{split}
\end{displaymath}}
\item[3.)]{{\bf The Geometric Summation Formula:}
\begin{displaymath}
\sum_{k=0}^{\lfloor x\rfloor}a^{k}=\frac{a^{x}}{\log(a)}+\frac{1}{1-a}+a^{x}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{\log(a)^{l-1}}{l!}S^{(1)}_{k}(l)B_{l}(\{x\})x^{l}
}{(x+1)(x+2)\cdots(x+k)}\;\;\forall a\neq1.
\end{displaymath}
\noindent Setting $x:=n\in\mathbb{N}$, we get
\begin{displaymath}
\sum_{k=0}^{n}a^{k}=\frac{a^{n}}{\log(a)}+\frac{1}{1-a}+a^{n}\sum_{k=1}^{\infty}(-1)^{k}\frac{\sum_{l=1}^{k}\frac{\log(a)^{l-1}}{l!}S^{(1)}_{k}(l)B_{l}n^{l}
}{(n+1)(n+2)\cdots(n+k)}\;\;\forall a\neq1.
\end{displaymath}}
\item[4.)]{{\bf The alternating Geometric Summation Formula:}
\begin{displaymath}
\sum_{k=0}^{\lfloor x\rfloor}(-1)^{k}a^{k}=\frac{1}{1+a}+\frac{(-1)^{x-\{x\}}}{2}a^{x}\sum_{k=0}^{\infty}(-1)^{k}\frac{\sum_{l=0}^{k}\frac{\log(a)^{l}}{l!}S^{(1)}_{k}(l)E_{l}(\{x\})x^{l}}{(x+1)(x+2)\cdots(x+k)}
\;\;\forall a\neq-1.
\end{displaymath}
\noindent Setting $x:=n\in\mathbb{N}$, we get
\begin{displaymath}
\sum_{k=0}^{n}(-1)^{k}a^{k}=\frac{1}{1+a}+(-1)^{n+1}a^{n}\sum_{k=0}^{\infty}(-1)^{k}\frac{\sum_{l=0}^{k}\frac{\log(a)^{l}}{(l+1)!}(2^{l+1}-1)S^{(1)}_{k}(l)B_{l+1}n^{l}}{(n+1)(n+2)\cdots(n+k)}
\;\;\forall a\neq-1.
\end{displaymath}}
\item[5.)]{{\bf The Euler-Maclaurin Geometric Summation Formula:}
\begin{displaymath}
\begin{split}
&\sum_{k=0}^{\lfloor x\rfloor}a^{k}=\frac{a^{x}}{\log(a)}+\frac{1}{1-a}+a^{x}\sum_{k=1}^{\infty}(-1)^{k}\frac{\log(a)^{k-1}}{k!}B_{k}(\{x\})\;\;\text{for}\;\;\frac{1}{e^{2\pi}}<a\neq1<e^{2\pi}.
\end{split}
\end{displaymath}}
\item[6.)]{{\bf The Euler-Maclaurin alternating Geometric Summation Formula:}
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}(-1)^{k}a^{k}&=\frac{1}{1+a}+(-1)^{x-\{x\}}\frac{a^{x}(a-1)}{a+1}\sum_{k=0}^{\infty}(-1)^{k}\frac{\log(a)^{k-1}}{k!}B_{k}(\{x\})\;\;\text{for}\;\;\frac{1}{e^{2\pi}}<a<e^{2\pi}.
\end{split}
\end{displaymath}}
\item[7.)]{{\bf The Exponential Geometric Summation Formula:}\newline
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}e^{k}&=e^{x}+\frac{1}{1-e}+e^{x}\sum_{k=1}^{\infty}(-1)^{k}\frac{B_{k}(\{x\})}{k!}\;\;\forall x\in\mathbb{R}^{+}.
\end{split}
\end{displaymath}}
\item[8.)]{{\bf The Self-Counting Summation Formula:}\newline
\noindent Let $\{a_{k}\}_{k=1}^{\infty}:=\{1,2,2,3,3,3,4,4,4,4,...\}$ be the self-counting sequence \cite{6,7} defined by
\begin{displaymath}
\begin{split}
a_{k}&:=\left\lfloor\frac{1}{2}+\sqrt{2k}\right\rfloor.
\end{split}
\end{displaymath}
\noindent Then, we have that
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}a_{k}&=
\frac{x\sqrt{8x+1}}{3}-\frac{5\sqrt{8x+1}}{24}-\frac{\sqrt{8x+1}}{2}B_{1}(\{x\})+B_{1}\left(\left\{\frac{\sqrt{8x+1}}{2}-\frac{1}{2}\right\}\right)B_{1}(\{x\})\\
&\quad+\frac{1}{2}B_{1}\left(\left\{\frac{\sqrt{8x+1}}{2}-\frac{1}{2}\right\}\right)-\frac{\sqrt{8x+1}}{4}B_{2}\left(\left\{\frac{\sqrt{8x+1}}{2}-\frac{1}{2}\right\}\right)\\
&\quad+\frac{1}{6}B_{3}\left(\left\{\frac{\sqrt{8x+1}}{2}-\frac{1}{2}\right\}\right).
\end{split}
\end{displaymath}}
\item[9.)]{{\bf Slowly convergent summation formula for the sum of the square roots:}
\begin{displaymath}
\begin{split}
\sum_{k=0}^{\lfloor x\rfloor}\sqrt{k}&=\frac{2}{3}x^{\frac{3}{2}}-\sqrt{x}B_{1}(\{x\})-\frac{1}{2\pi}\sum_{k=1}^{\infty}\frac{\text{FresnelS}\left(2\sqrt{k}\sqrt{x}\right)}{k^{\frac{3}{2}}}.
\end{split}
\end{displaymath}}
\item[10.)]{{\bf Slowly convergent summation formula for the harmonic series:}
\begin{displaymath}
\begin{split}
\sum_{k=1}^{\lfloor x\rfloor}\frac{1}{k}&=\log(x)+\gamma+2\sum_{k=1}^{\infty}\text{CosIntegral}(2\pi kx).
\end{split}
\end{displaymath}}
\end{itemize}
\section{Conclusion}
\label{sec:Conclusion}
\noindent This paper presents an overview on some rapidly convergent summation formulas obtained by applying the Weniger transformation \cite{1}.
|
\section{Introduction}
\input{sections/introduction}
\section{Background}
\input{sections/background}
\input{sections/problem_definition_and_algorithm}
\section{Experimental Results}
\input{sections/experimental_evaluation}
\vspace{-0.5cm}
\section{Conclusion}
\input{sections/conclusion}
\section{Acknowledgment}
This research was supported in part by the DARPA DEFT program under AFRL grant FA8750-13-2-0026 and by MURI ARO grant W911NF-08-1-0242.
\section{Overview of the Tasks}
In this section we give a short overview of each of the three tasks considered in this paper to demonstrate successful deployment of our algorithm.
\subsection{Cold Start Slot Filling}
The goal of CSSF is to collect information (fills) about specific attributes (slots) for a set of entities (queries) from a given corpus. The queries entities can be a person (PER), organization (ORG) or geo-political entity (GPE). The slots are fixed and the 2015 task also included the inverse of each slot, for example the slot org:subsidiaries and its inverse org:parents. Some slots (like per:age) are
{\it single-valued} while others (like per:children) are {\it list-valued} i.e., they can take multiple slot fillers.
The input for CSSF is a set of queries and the corpus in which to look for information. The queries are provided in an XML format that includes an ID for the query, the name of the entity, and the type of entity (PER, ORG or GPE). The corpus consists of documents in XML format from discussion forums, newswire and the Internet, each identified by a unique ID. The output is a set of slot fills for each query. Along with the slot-fills, systems must also provide its {\it provenance} in the corpus in the form {\it docid:startoffset-endoffset}, where {\it docid} specifies a source document and the offsets demarcate the text in this document containing the
extracted filler. Systems also provide a confidence score to indicate their certainty in the extracted information.
\subsection{Tri-lingual Entity Discovery and Linking}
The goal of TEDL is to discover all entity mentions in a corpus with English, Spanish and Chinese documents. The entities can be a person (PER), organization (ORG), geo-political entity (GPE), facility (FAC), or location (LOC). The FAC and LOC entity types were newly introduced in 2015. The extracted mentions are then linked to an existing English KB entity using its ID. If there is no KB entry for an entity, systems are expected to cluster all the mentions for that entity using a NIL ID.
The input is a corpus of documents in the three languages and an English KB (FreeBase) of entities, each with a name, ID, type, and several relation tuples that allow systems to disambiguate entities. The output is a set of extracted mentions, each with a string, its provenance in the corpus, and a corresponding KB ID if the system could successfully link the mention, or else a mention cluster with a NIL ID. Systems can also provide a confidence score for each mention.
\subsection{Object Detection for Images}
The goal of the object detection task is to detect all instances of object categories (out of the $200$ predefined categories) present in the image and localize them by providing coordinates of the axis-aligned bounding boxes for each instance. The ImageNet dataset is organized according to the WordNet hierarchy and thus the object categories are WordNet {\it synsets}.
The object detection corpus is divided into training, validation and test sets. The training set consists of approximately $450$K images including both positive and negative instances, annotated with bounding boxes; the validation set consists of around $20$K images also annotated for all object categories and the test set has $50$K images. The output for the task is the image ID, the object category ($1$-$200$), a confidence score and the coordinates of the bounding box. In case of multiple instances in the same image, each instance is mentioned on a separate line.
\section{Methodology}
This section describes our approach to stacking multiple systems using their confidence scores and other auxiliary features. Figure~\ref{fig:sup} shows an overview of our system which trains a final meta-classifier for combining multiple systems. The auxiliary features depend on the task into consideration as described in the Section~\ref{sec:auxiliary}.
\subsection{Stacking}
Stacking is a popular ensembling methodology in machine learning \cite{wolpert:nn92} and has been very successful in many applications including the top performing systems in the Netflix competition
\cite{sill:2009}. The idea is to employ multiple learners and combine
their predictions by training a ``meta-classifier'' to weigh and
combine multiple models using their confidence scores as features. By
training on a set of supervised data that is disjoint from that used
to train the individual models, it learns how to combine
their results into an improved ensemble model that performs better than each individual component system. In order to successfully use stacking, the output must be represented as a {\it key-value} pair. The meta-classifier makes a binary decision for each distinct output pair. Thus before deploying the algorithm on a task, it is crucial to identify the {\it key} in the task output which serves as a unique handle for ensembling systems as well as the {\it values} which are results for a key provided by systems. Note that there is only one instance of a key in the output while there could be multiple values for a key from component systems. The output of the ensembling system is similar to the output of an individual system, but it productively aggregates results from different systems. In a final {\it post-processing} step, the outputs that get classified as ``correct" by the classifier are kept while the others are removed from the output.
The first step towards using the stacker is to represent the output as {\it key-value} pair. For the CSSF task, the {\it key} for ensembling multiple systems is a query along with a slot type, for example, per:age of ``Barack Obama'' and the {\it value} is a computed {\it slot fill}. For list-valued slot types such as org:subsidiaries, the key instance is repeated in the output for each value. For the TEDL task, we define the key to be the {\it KB (or NIL) ID} and the value to be a {\it mention}, that is a specific reference to an entity in the text. For the ImageNet object detection task, we represent the image ID as the {\it key} for ensembling and the {\it value} is a detected object category. The next step is to represent the output pair instances consistently. For a particular {\it key-value} pair if a system produced it then it also provides a confidence score else we use a confidence score of zero i.e. the output instance is incorrect according to that system. The output is now ready to be fed into the stacker as shown in Figure~\ref{fig:sup}. Along with confidence scores, we also feed in auxiliary features, described in the next section, for each task that enable the classifier to discriminate across component systems effectively and thus make better decisions.
\subsection{Auxiliary Features}
\label{sec:auxiliary}
As discussed earlier, systems must provide evidence in the form of provenance for each generated output pair. We use them as part of our auxiliary features that go into the stacker along with the confidence scores as shown in top part of Figure~\ref{fig:sup}. The provenance indicates origin or the source for the generated output instance and thus depends on the task under considerations. For the CSSF task, if a system successfully extracts a relation then it must provide a slot filler provenance indicating the location of the extracted slot fill in the corpus. Provenance is more explicit for the two KBP tasks than the object detection task. For the KBP tasks, it serves as output justification in the form of text. On the other hand, for the detection task, output justification is in the form of bounding boxes and thus serves the same purpose as provenance. For the TEDL task, if a system successfully links a mention to a KB ID then it must provide the mention provenance indicating the origin of the mention in the corpus. For both the CSSF and TEDL tasks, the provenance is in the form of \emph{docid} and \emph{startoffset-endoffset} that gives information about the document in the corpus and offset in the document. On the other hand, for the ImageNet object detection task, if a system successfully detects a target category then it must provide the object bounding box localizing the object in the image. The bounding box is in the form of $\langle x_{min}, y_{min}, x_{max}, y_{max}\rangle$. The bounding box for object detection is similar to provenance for the KBP tasks and can thus be used as auxiliary features for stacking.
The idea behind using provenance as auxiliary features is that a output is more reliable if not just multiple systems produce it but also agree on the source/provenance of the decision. In order to enable the stacker to leverage the auxiliary features for discriminating among systems, we develop features that measure provenance similarity {\it across} systems. The Jaccard similarity
coefficient is one such statistical measure of similarity between sets and is thus useful in measuring the degree of overlap between the provenance provided by systems. For the CSSF and TEDL tasks, the provenance offsets(PO) are used to capture similarity as follows. For a given {\it key}, if $N$ systems that generate a {\it value} have the same \emph{docid} for their document provenance, then the provenance offset (PO) score is calculated as the intersection of offset strings divided by its union. Thus systems that generate a {\it value} from
\emph{different} documents for the same {\it key} have zero
overlap among offsets.
\[PO(x)=\frac{1}{|N|}\times\sum_{i\in N,i\neq x}^{}\frac{|\textsf{substring(i)}\cap \textsf{substring(x)}|}{|\textsf{substring(i)}\cup \textsf{substring(x)}|}\]
For the object detection task, the Jaccard coefficient is used to measure the overlap between bounding boxes across systems. For a given image ID, if $N$ systems detect the same object instance, then the bounding box overlap (BBO) score is calculated as the the intersection of the areas of bounding boxes, divided by their union:
\[BBO(x)=\frac{1}{|N|}\times\sum_{i\in N,i\neq x}^{}\frac{|\textsf{Area(i)}\cap \textsf{Area(x)}|}{|\textsf{Area(i)}\cup \textsf{Area(x)}|}\]
We note that for the CSSF task, two systems are said to have extracted the same slot fill for a {\it key} if the fills are exactly same, however for the TEDL task, two systems are said to have linked the same mention for a {\it key} if the mentions overlap to any extent and finally for the ImageNet task, two systems are said to have detected the same object instance for a {\it key} if the Intersection Over Union (IOU) of the areas of their bounding boxes is greater than $0.5$. If the output {\it values} don't meet this criteria for a given {\it key}, then they are considered to be two different values for the same key.
For the two KBP tasks, we also use the \emph{docid}
information as auxiliary features. For a given {\it key}, if $N$ systems provide a {\it value} and a
maximum of $n$ of those systems give the same \emph{docid} in their
provenance, then the document provenance score for those $n$ systems is $n/N$. Similarly, other systems are given lower
scores based on the fraction of systems whose provenance document
agrees with theirs. Since this provenance score is weighted by the
number of systems that refer to the same provenance, it measures the
reliability of a {\it value} based on the document from where it originated. We note that the use of provenance
as features does not require access to the large corpus
of documents or image and is thus very computationally inexpensive.
Additional auxiliary features that we use are the slot type (e.g. per:age) for the CSSF task, and for the TEDL, we use the entity type and for the ImageNet task, we use the object category as an additional feature. For the CSSF task, features related to the provenance of the fill, as discussed above have also been used in \cite{viswanathan:acl15}. However, our novel auxiliary features for both the KBP tasks, that require access to the source corpus, further boosts our performance and beats their best ensemble. The 2015 CSSF task had a much smaller corpus of shorter documents compared to the previous year's slot-filling corpus \cite{ellis:tac15,surdeanu:tac14}. Thus, the provenance feature of \cite{viswanathan:acl15} did not sufficiently capture the reliability of a slot fill based on where it was extracted. Our new auxiliary feature measures the similarity between the {\it key} document and the {\it value} document. For the CSSF task, the {\it key} is the query entity along with slot type and thus the {\it key} document is the query document provided to participants to disambiguate query entities that could potentially have the same name but refer to different entities. For the TEDL task, the {\it key} is the KB ID of an entity and thus the {\it key} document is the pseudo-document made up of that entity's KB description as well as several relations involving the entity that exist in the KB. The document that the CSSF and TEDL systems provide as provenance is the {\it value} document for both the tasks. So the auxiliary feature uses cosine similarity to compare the {\it key} and {\it value} documents represented as standard TF-IDF weighted vectors. For the ImageNet task, we use the object class label as an additional feature. Some systems do well only on a set of categories such as deformable objects. Using the class label enables the stacker to learn to discriminate across such systems.
\begin{table*}[!ht]
\centering
\begin{tabular}{| c | c | c | c | }
\hline
\rule{0pt}{2.5ex}
\textbf{Methodology} & \textbf{Precision} & \textbf{Recall} & \textbf{F1} \\
\hline
\hline
\rule{0pt}{2ex}
Stacking with auxiliary features&0.4656&\textbf{0.3312}&\textbf{0.3871}\\
Stacking approach described in \citet{viswanathan:acl15}& \textbf{0.5084}&0.2855&0.3657\\
Top ranked CSSF system in 2015 \citet{angeli:tac15}& 0.3989&0.3058&0.3462\\
Oracle Voting baseline (3 or more systems must agree)&0.4384&0.2720&0.3357\\
\hline
\end{tabular}
\caption{Results on 2015 Cold Start Slot Filling (CSSF) task using the official NIST scorer}
\label{table:results1}
\end{table*}
\subsection{Post-processing}
Once we obtain the decisions on each of the key-value pairs from the stacker, we perform some final post-processing so that the output from the stacker is as though it is generated by a single system. For CSSF, this is straight forward. Each list-valued slot fill that is classified as correct is included in the final output. For single-valued slot fills, if they are multiple fills that were classified correctly for the same query and slot type, we include the fill with the highest meta-classifier confidence. For TEDL, for each entity mention link that is classified as correct, if the link is a KB cluster ID then we include it in the final output, but if the link is a NIL cluster ID then we keep it aside until all mention links are processed. Thereafter, we resolve the NIL IDs across systems since NIL ID's for each system are unique. We merge NIL clusters across systems into one if there is at least one common entity mention among them. Finally, we give a new NIL ID for these newly merged clusters.
For the ImageNet object detection task, for each object instance that is classified as correct by the stacker is included in the final output and the bounding box for that output instance is calculated as follows. If multiple systems successfully detected an object instance, then we sum the overlapping areas between a system's bounding box and every other system's that also detected the exact same instance and we do this for every such system. The bounding box produced by the system that has maximum overlapping area is included in the final output. Note that in case of two systems, this method is redundant and we include the bounding box produced by the systems with a higher confidence score.
\begin{table*}[!ht]
\centering
\begin{tabular}{| c | c | c | c | }
\hline
\rule{0pt}{2.5ex}
\textbf{Methodology} & \textbf{Precision} & \textbf{Recall} & \textbf{F1} \\
\hline
\hline
\rule{0pt}{2ex}
Stacking with auxiliary features&0.803&\textbf{0.525}&\textbf{0.635} \\
Stacking approach described in \citet{viswanathan:acl15}& \textbf{0.814}&0.508 &0.625\\
Top ranked TEDL system in 2015 \citet{sil:tac15}& 0.693&0.547 &0.611\\
Oracle Voting baseline (4 or more systems must agree)&0.514 &0.601 &0.554\\ \hline
\end{tabular}
\caption{Results on 2015 Tri-lingual Entity Discovery and Linking (TEDL) task using the official NIST scorer and the CEAFm metric}
\label{table:results2}
\end{table*}
\begin{table*}[!ht]
\centering
\begin{tabular}[h]{| c | c | c | }
\hline
\rule{0pt}{2.5ex}
\textbf{Methodology} & \textbf{Median AP} & \textbf{Mean AP} \\
\hline
\hline
\rule{0pt}{2ex}
Stacking with auxiliary features&\textbf{0.526}&\textbf{0.546} \\
Best standalone system (VGG + selective search)\citet{renNIPS15fasterrcnn}&0.450&0.454\\
Oracle Voting baseline (1 or more systems must agree) &0.367&0.353\\
\hline
\end{tabular}
\caption{Results on 2015 ImageNet object detection task using the official ImageNet scorer.}
\label{table:results3}
\end{table*}
|
\section{Introduction}
The discovery of iron based superconductors in 2008 has stimulated much discussion\cite{GRStewart}. Their rich phase diagrams have suggested various bosonic fluctuations, such as spin, orbital and charge, possibly acting as a glue between electrons. Many theories have been proposed to explain the superconducting mechanism in terms of these fluctuations. However, which fluctuation plays the most important role in superconductivity (SC) remains in dispute.
In some theories, Fermi surface (FS) nesting plays a crucial role. For example, spin fluctuation is enhanced when the condition of nesting between the hole and electron FSs is good, which induces unconventional SC\cite{I.I.Mazin, K.Kuroki1, A.V.Chubukov}. In the case of \BaFeAsP\ (Ba122P), the experimental results of inelastic neutron scattering and nuclear magnetic resonance (NMR) studies revealed that spin fluctuation was clearly enhanced in the optimally doped $x$-region\cite{M.Ishikado, Y.Nakai}. Furthermore, the study of angle-resolved photoemission spectroscopy (ARPES)\cite{T.Yoshida} demonstrated that the observed FSs fulfill the nesting condition between the electron and hole FSs, which was consistent with the spin fluctuation theory in Ba122P\cite{K.Kuroki2, K.Suzuki}. From the crystallographic viewpoint, the appearance of SC would be ascribed to the optimization of the pnictogen ($Pn$) height from the Fe layer (\hpn )\cite{Y.Mizuguchi} or the $Pn$-Fe-$Pn$ bond angle (\al )\cite{C.H.Lee}, which was also explained by nesting-based theories\cite{K.Kuroki2, T.Saito}.
On the other hand, according to the nesting scenario, the distance between neighboring Fe layers, which must be correlated with interlayer hybridization and thus with anisotropy, should have a distinct influence on antiferromagnetism (AFM) and SC. Nevertheless, both N$\acute{\text{e}}$el temperature (\Tn ) and superconducting transition temperature (\Tc ) are not higher in Ba122P than in Sr122P, although the ratio of $a$- and $c$-axes lattice constants, $c/a$, which is an index of structural anisotropy, is larger in Ba122P than in Sr122P. These behaviors are inconsistent with the nesting-based model\cite{J.M.Allred, T. Kobayashi}. The change in the FS with structural anisotropy has been confirmed in a recent ARPES study\cite{H.Suzuki}.
According to that study, the \dz\ FS at the Brillouin zone center was warped more strongly in Sr122P than in Ba122P, reflecting the difference in the structural anisotropy. This result threw doubt on the role of nesting in the stabilization of AFM and SC.
Thus, the importance of FS nesting for SC and AFM remains unclear.
In this work, we extended the previous study of Ba122P and Sr122P to a more systematic study by synthesizing single crystals of $A$Fe$_2$(As$_{1-x}$P$_x$)$_2$ ($A$=Ba$_{0.5}$Sr$_{0.5}$, Sr$_{1-y}$Ca$_y$ and Eu), where not only the local structure of \FePn\ tetrahedra but also $c/a$ can be controlled by As/P substitution and a change in the $A$ site ions. Their phase diagrams and crystallographic structures have been precisely investigated in order to clarify the relationships among the local structure around the Fe site, structural anisotropy $c/a$, and stabilities of SC and AFM.
Single crystals of $A$122P ($A$=Ba$_{0.5}$Sr$_{0.5}$, Sr$_{0.92}$Ca$_{0.08}$, Sr$_{0.84}$Ca$_{0.16}$, and Eu) were grown by the self-flux method. The starting materials were Ba, Sr, Ca, Eu (flakes), FeAs, and FeP (powders). They were mixed stoichiometrically and put into alumina crucibles sealed in quartz tubes with Ar gas of 0.2 bar. They were heated to 1300 \doc , held for 12 h, and then slowly cooled to 1050 \doc\ at a rate of 2 \doc /h. Plate-like single crystals with a typical size of 1$\times 1\times $0.05 $\mathrm{mm^3}$ were obtained. As-grown crystals were annealed by the procedure described in Ref. \citen{T.Kobayashi} to remove defects or distortion within crystals. Single crystals with $x\geq $0.24 for $\mathrm{Sr}_{0.84}\mathrm{Ca}_{0.16}$122P could not been synthesized because of the solubility limit, as in the case of high P content for Ca122P\cite{S.Kasahara}.
\Tn\ and \Tc\ were determined by the electrical resistivity and magnetic susceptibility measurements. The lattice constants of the $a$- and $c$-axes were estimated by single-crystal X-ray diffraction analysis. The composition of single crystals was determined by energy dispersive X-ray spectroscopy (EDX). In order to discuss the effects of structural change on the physical parameters, the local structure of \FePn\ tetrahedra of optimally doped crystals were precisely determined at room temperature by synchrotron X-ray (15 keV) diffraction analysis at BL-8A/8B of the Photon Factory, KEK, Japan. The atomic positions were estimated by the least-squares method using Rigaku CRYSTAL STRUCTURE.
\begin{figure}[htpb]
\centering
\includegraphics[width=70mm,origin=c,keepaspectratio,clip]{17193Fig1.eps}
\caption{(Color online) (a), (b) P-doping ($x$) dependences of the lattice constants ($a$ and $c$), and (c) the ratio of lattice constants ($c/a$) defined as structural anisotropy of \AFeAsP\ ($A$=Ba\cite{M.Rotter}, Ba$_{0.5}$Sr$_{0.5}$, Sr\cite{T.Kobayashi}, Sr$_{1-y}$Ca$_y$ and Eu).}
\label{fig1}
\end{figure}
The P content ($x$) dependences of lattice constants ($a$ and $c$) and their ratio ($c/a$) are shown in Fig. \ref{fig1}. The results for Ba122P and Sr122P in previous reports are also plotted in the same figure\cite{M.Rotter, T.Kobayashi}. Both lattice constants $a$ and $c$ decrease linearly with $x$ for all systems, which indicates that P is successfully substituted.
When the $A$ site atom is changed from the large Ba ion to the small Eu ion via the intermediate Sr ion, the $c$-axis decreases more strongly than the $a$-axis. As a result, the anisotropy ratio $c/a$ decreases. In addition, the slope of $c/a$ becomes steeper with the change in the $A$ site atom from Ba to Eu.
\begin{figure*}[htb]
\centering
\includegraphics[width=120mm,origin=c,keepaspectratio,clip]{17193Fig2.eps}
\caption{(Color online) Temperature dependences of resistivity scaled by the value at room temperature (RT) for (a) \BaSrFeAsP , (b) \Caa , (c) \Cab\ and (d) \EuFeAsP\ with various P contents ($x$).}
\label{fig2}
\end{figure*}
Figure \ref{fig2} shows the temperature dependences of resistivity normalized at room temperature for (a) \BaSr 122P , (b) \Cae 122P , (c) \Caf 122P and (d) Eu122P. A kink, indicating antiferromagnetic and structural transitions, was observed in the low-$x$ region for all the systems. The \Tn\ values for \BaSrFeAs\ and Eu122P are consistent with those in previous studies\cite{K.Kirshenbaum, H.S.Jeevan}.
In a certain composition range, SC was observed, and \Tc\ reached maximum values of 30, 32, 30, and 27 K at \xc\ (=0.30, 0.25, 0.24, and 0.18) for $A$=\BaSr , \Cae , \Caf , and Eu, respectively. It is noted that, near \xc , the resistivity shows an almost $T$-linear dependence for all systems, indicating the enhancement of two-dimensional antiferromagnetic fluctuation. For \BaSr 122P, the AFM phase seems to remain at \xc . However, a $T$-linear resistivity was observed at $x$=0.35, where \Tc\ ($\sim $30K) is nearly the same as that at \xc .
Figure \ref{fig3} shows a summary of \Tn\ and \Tc\ for all $A$122P ($A$=Ba, \BaSr , Sr, \Cae , \Caf , Eu, and Ca), where the $x-T$ phase diagrams are arranged in order of structural anisotropy from Ba to Ca, and the data for $A$=Ba, Sr, and Ca are taken from previous studies\cite{M.Nakajima, T.Kobayashi, S.Kasahara}. Magnetic transition related to \Eu\ is omitted in Fig. \ref{fig3}. Upon substituting Ca for Sr, AFM was more rapidly suppressed by P doping, and \xc\ shifted from 0.35 ($A$=Sr) to 0.24 ($A$=\Caf ) through 0.25 ($A$=\Cae ). This behavior of \xc\ can be understood in terms of the change in the structural anisotropy (see Fig. 1.). When Ca is substituted, the anisotropy of the crystal structure and the resultant FS anisotropy rapidly decrease with increasing P content. Hence, the nesting condition between FSs becomes worse and \Tn\ vanishes at a lower P content, $x$. (Namely, \xc\ decreases.) The same tendency is retained in Eu122P. In Eu122P, $c/a$ is smaller than that in $\mathrm{Sr}_{1-y}\mathrm{Ca}_{y}$122P systems, and P doping more rapidly suppresses $c/a$, i.e., the Eu122P system has the more isotropic crystal structure. As a result, $T_N$ disappears at a lower P content ($x \sim$0.18).
On the other hand, we should consider the contribution of structural transition (orbital order) or nematicity to the rapid decrease in \Tn. Because the magnetic order, orbital order, and nematicity are intertwined with each other in iron-based superconductors\cite{R.M.Fernandes1}, it is possible that orbital order or nematicity is suppressed by structural change (not through the change of FS nesting condition) and causes the suppression of magnetic order\cite{M.Yoshizawa, H.Kontani, R.M.Fernandes2}. However, there has been no report on the relationship between structural anisotropy and orbital order or nematicity, while our results are consistent with the nesting scenario and the resluts of ARPES study\cite{H.Suzuki}. Therefore, in this work, we consider that the rapid change of nesting condition is the main cause of the rapid suppression of magnetic order.
Here, we note that \Tn\ values for Ba122 and Sr122 do not follow this correlation between structural anisotropy and the stability of AFM. A previous study of \AFeAs\ ($A$=Ba$_{1-y}$Sr$_y$, Sr$_{1-y}$Ca$_y$)\cite{K.Kirshenbaum} also revealed that \Tn\ shows a nonmonotonic behavior from $A$=Ba to Sr$_{0.3}$Ca$_{0.7}$, although the lattice constants ($a$ and $c$) and the As-Fe-As bond angle \al\ decrease monotonically. So far, it is unclear what determines the \Tn\ values in this system.
\begin{figure}[htb]
\centering
\includegraphics[width=70mm,origin=c,keepaspectratio,clip]{17193Fig3.eps}
\caption{(Color online) Electronic phase diagram for \AFeAsP\ ($A$=Ba\cite{M.Nakajima}, \BaSr , Sr\cite{T.Kobayashi}, \Cae , \Caf , Eu, and Ca\cite{S.Kasahara}).}
\label{fig3}
\end{figure}
In contrast to the relationship between $c/a$ and \xc, no correlation can be seen between \Tc\ and structural anisotropy. As is shown in Fig. \ref{fig3}, \Tc\ reaches a maximum value at \xc. Surprisingly, the maximum \Tc\ was about 30 K in all the systems, although the structural anisotropy monotonically decreases from $A$=Ba to $A$=Eu and \xc\ changes accordingly. Therefore, we conclude that structural anisotropy, and thus electronic anisotropy, have little effect on \Tc\ in the 122 system. This is not intuitively reconciled with the nesting scenario because the FS topology and nesting condition are affected by the structural anisotropy depending on the $A$ site atom. It is possible that the FSs sensitive to structural anisotropy, such as \dz\ FS, do not contribute to SC, as was proposed in a previous study\cite{H.Suzuki}. However, even in this case, there remains the question why a slight but finite change in the FS topology with \dxy\ and \dyzzx\ character does not alter \Tc.
Table \ref{table1} shows the crystallographic parameters of optimally doped $A$122P ($A$=\BaSr , \Cae , \Caf , and Eu), determined by the least-squares refinement of X-ray diffraction profiles for single crystals. The determined lattice constants and anisotropy ratio are in good agreement with the results in Fig. \ref{fig1}. As reference, the local structural parameters of \FePn\ tetrahedra of optimally doped crystals for $A$122P ($A$=Ba and Sr) are given in the caption of Table \ref{table1}\cite{S.Jiang, T.Kobayashi2}. It is surprising that the \hpn\ ($\sim $1.32 \AA) and/or \al\ ($\sim $112$^{\circ}$) values at \xc\ are nearly identical in all the systems, except for $A$=Eu, although \xc\ significantly varies upon changing the $A$ ion. This indicates that \hpn\ and \al\ are crucial parameters for SC, as pointed out in previous studies\cite{Y.Mizuguchi, C.H.Lee}.
Moreover, the fact that \hpn\ (or \al) is the same at \xc\ implies that AFM is completely suppressed when \hpn\ (or \al) is this value, irrespective of the anisotropy ratio. This highlights that these local structural parameters are also important in explaining the mechanism of AFM suppression. P substitution plays a role in tuning these local structural parameters in iron pnictides, as was pointed out in Ref. \citen{M.Rotter}.
\begin{table*}[htb]
\caption{Crystallographic parameters at room temperature for \AFeAsP\ ($A$=\BaSr , \Cae , \Caf , and Eu) determined by the least-squares refinement of the single crystal X-ray diffraction profile. The reliabilities are $R_1~(I>2.00\sigma(I))=6.05\%, 5.61\%, 9.42\%, 6.07\%$ and $wR_2 (Rw)~(I>2.00\sigma(I))=8.91\%, 9.77\%, 10.73\%, 9.54\%\ $for $A$=Ba$_{0.5}$Sr$_{0.5}$, Sr$_{0.92}$Ca$_{0.08}$, Sr$_{0.84}$Ca$_{0.16}$, and Eu, respectively. \hpn\ and bond angle of $Pn$-Fe-$Pn$ (\al ) of $\mathrm{BaFe_{2}}(\mathrm{As}_{0.68}\mathrm{P}_{0.32}\mathrm{)_{2}}$ and $\mathrm{SrFe_{2}}(\mathrm{As}_{0.65}\mathrm{P}_{0.35}\mathrm{)_{2}}$ by the previous studies\cite{S.Jiang, T.Kobayashi2} are 1.319(3)\AA , 112.2(1)$^\circ$ and 1.325(1)\AA\, 111.566(17)$^\circ$, respectively.}
\label{table1}
\begin{center}
\begingroup
\scalefont{.65}
\begin{tabular}{lcccc}
\hline
Compound & $\mathrm{Ba}_{0.5}\mathrm{Sr}_{0.5}\mathrm{Fe_{2}}(\mathrm{As}_{0.70}\mathrm{P}_{0.30}\mathrm{)_{2}}$ & $\mathrm{Sr}_{0.92}\mathrm{Ca}_{0.08}\mathrm{Fe_{2}}(\mathrm{As}_{0.75}\mathrm{P}_{0.25}\mathrm{)_{2}}$ & $\mathrm{Sr}_{0.84}\mathrm{Ca}_{0.16}\mathrm{Fe_{2}}(\mathrm{As}_{0.76}\mathrm{P}_{0.24}\mathrm{)_{2}}$ & $\mathrm{EuFe_{2}}(\mathrm{As}_{0.82}\mathrm{P}_{0.18}\mathrm{)_{2}}$ \\
\hline
Space group & $I4/mmm$ & $I4/mmm$ & $I4/mmm$ & $I4/mmm$ \\
$a$ (\AA) &3.9099(10)& 3.8983(2) & 3.8970(8) &3.9039(3) \\
$c$ (\AA) &12.482(3)& 12.0723(7) & 12.022(2) &11.967(1) \\
$c/a$ &3.192(1) &3.0968(2) &3.08970(8) &3.0654(3) \\
$A$& (0, 0, 0)&(0, 0, 0) & (0, 0, 0) & (0, 0, 0) \\
Fe & (1/2, 0, 1/4) & (1/2, 0, 1/4) & (1/2, 0, 1/4) & (1/2, 0, 1/4) \\
As/P & (0, 0, z) & (0, 0, z) & (0, 0, z) & (0, 0, z)\\
& z=0.35571(4) & z=0.35978(3) & z=0.35988(8) & z=0.36212(8) \\
$h_{Pn}$(\AA)& 1.3195(8) & 1.3253(4) & 1.321(1) & 1.342(1) \\
As-Fe-As(deg.)& 108.238(6)$\times$4& 108.430(6)$\times$4 & 108.354(13)$\times$4 & 108.717(14)$\times$4\\
& 111.967(14)$\times$2& 111.574(11)$\times$2 & 111.73(2)$\times$2 & 110.99(3)$\times$2\\
Number of \\reflections & 267 & 202 & 202 & 143 \\
(I$>$2.00$\sigma$(I))& & & \\
Good of fitness& 1.067 &1.111& 1.124 & 1.216 \\
\hline
\end{tabular}
\endgroup
\end{center}
\end{table*}
Note that \hpn\ and \al\ at \xc\ in the Eu122P system are very different from those in other systems. These longer \hpn\ and smaller \al\ can be understood by considering the Ruderman-Kitetl-Kasuya-Yoshida (RKKY) interaction between interlayer Eu$^{2+} $moments proposed in several studies\cite{Y.Xiao, Y.Tokiwa, C.Feng}. The results of a previous neutron diffraction study suggested that the Eu 4f electrons participate in the Eu-(As/P) bonding through the RKKY interaction\cite{Y.Xiao}. This enhanced Eu-(As/P) bonding could result in the longer \hpn\ and smaller \al\ of Eu122P.
According to the \hpn\ or \al\ vs \Tc\ plots reported by Mizuguchi \etal\ \cite{Y.Mizuguchi} and Lee \etal\ \cite{C.H.Lee}, \Tc\ of the Eu122P system is expected to be higher than the observed value (=27 K). The \Tc\ suppression can be explained in terms of the magnetic moment of \Eu . In the annealed crystals of the Eu122P system, the magnetic order of Eu$^{2+}$ 4$f$ moments occurs at around \Tn =17--24 K, which was observed in the present crystals and also reported previously\cite{H.S.Jeevan}.
According to a previous report\cite{S.Zapf}, for Eu122P, the magnetic moments of the \Eu\ ions aligned along the $a$ direction antiferromagnetically are canted, yielding a ferromagnetic contribution along the $c$-direction.
This must cause magnetic pair breaking.
One of the concerns is the effect of structural instability on \Tc. In the case of Ca122P, owing to the very small $c$-axis length, the interlayer As-As hybridization is strong\cite{T.Yildirim} and at low temperatures the system enters into a collapsed-tetragonal (cT) phase where no bulk SC appears\cite{S.Kasahara}.
However, the interlayer As-As bond lengths of the optimally doped crystals for \Cae 122P, \Caf 122P, and Eu122P are 3.3856(9), 3.369(2), and 3.300(2), respectively. These values are close to that for Sr122P [=3.396(1)] but far from the critical value ($\sim 3.0$ \AA\cite{J.R.Jeffries}) where the system enters into the cT phase at low temperatures. Therefore, this interlayer As-As hybridization effect on \Tc\ can be ignored.
In summary, we succeeded in systematically studying the $c/a$ anisotropy effect by growing $A$122P ($A$=\BaSr , \Cae , \Caf , and Eu) crystals. For the smaller $A$ ion, \Tn\ decreases with P doping more rapidly and thus AFM disappears at a smaller $x$, which is qualitatively consistent with the nesting-based theoretical model. On the other hand, regarding SC, in all the systems, the maximum \Tc\ is approximately 30 K, although the structural anisotropy monotonically decreases from $A$=Ba to $A$=Eu. The precise X-ray diffraction analysis has revealed that the pnictogen height \hpn\ and/or As-Fe-As bond angle \al\ show the universal values at \xc\ where AFM disappears and \Tc\ shows a maximum. The fact that \Tc\ is sensitive not to the structural anisotropy but to \hpn\ and \al\ casts doubt on a simple nesting mechanism for SC.
\begin{acknowledgments}
The authors are grateful to K. Kuroki, H. Usui, and M. Nakajima
for fruitful discussions. The x-ray diffraction experiment has been
carried out under the approval of the Photon Factory Program Advisory
Committee (Proposals No. 2012S2-005). T. A. and T. K. acknowledge the Grant-in-Aid for JSPS Fellows.
The present work was supported by the JST project in Japan (TRIP and IRON-SEA).
\end{acknowledgments}
|
\section{Introduction}
Emission of broad-line region (BLR) of active galactic nuclei (AGNs) is explained by models which propose continuous steady flows or presence a very large number of clouds which exhibit some bulk motion \citep[e.g.,][]{oster,netzerbook,Rees,Rees89}. In the early models for the BLRs, the clouds are moving inward or outward through a static or slowly moving background gas \citep[e.g.,][]{mathews74,mat75,mat79}. Subsequent studies showed that the clouds may exhibit orbital motion under the strong gravitational potential of a central object. \cite{kwan82} constructed a kinematic model in which the BLR clouds orbit the central object in nearly parabolic orbits and this model has been extended by \cite{carroll85} to include the effects of a finite infalling cloud number and size. There are considerable uncertainties about formation of the BLR clouds and their dynamical stability \citep[e.g.,][]{mat86,mat89}, however, many authors have successfully produced emission of BLR systems based on the discrete cloud concept \citep[e.g.,][]{cap79,cap80,cap81,frome,netzer2010}.
Most of these models assume that the BLR clouds are pressure-confined, though a few authors argue that the clouds are transient rather than stable long-lived objects. Orbital motion of the BLR clouds is a rich source of information to estimate the mass of the central black hole. While early models are considering gravitational force of the central black hole as a dominant force, it has been argued that BLR clouds are also subject to a force due to the intense radiation of a central source (e.g., an accretion disc) and the mass of the central black hole is underestimated if radiation pressure is neglected \citep[e.g.,][]{marconi,namekata}. Moreover, it has been suggested that the central radiation is nonisotropic \citep{Liu11} and the orbits of BLR clouds are significantly modified when this feature of radiation is considered \citep{plewa,khajenabi15}.
Clouds embedded in a hot gaseous medium have also been discovered near to the Galactic center \citep[][]{Gill}. These low-mass gas clouds, known as G1 and G2, are moving on highly eccentric orbits through gaseous medium around a central black hole. Using orbital motions of these clouds, one can probe the accretion flow feeding Sgr ${\rm A}^{\star}$ \citep{McCourt2015,Pfuhl2015}. There are considerable uncertainties about the true nature of intercloud medium and physical mechanisms that may lead to the formation of these clouds. Although various processes have been proposed for the formation of G1 and G2 or BLR clouds, we don't yet know for sure if these clouds are formed as a result of such mechanisms. The intercloud medium of G1 and G2 is described, however, successfully using a kind of accretion flow which is known as Advection-Dominated Accretion Flow (ADAF; Narayan \& Yi 1994). It is a good motivation to assume that BLR clouds are also moving through this type of accretion flows \citep{krause11}. Nevertheless, most of the previous semi-analytical studies of BLR clouds' dynamics prescribe pressure profile of the intercloud medium as a simple power-law function of the radial distance \citep[e.g.,][]{netzer2010,krause11,krause12,plewa,khajenabi15}.
Since each BLR cloud is assumed to be in a pressure-confined state, its radius is determined by a balance between the interior pressure of a cloud and its ambient pressure. Then, orbital motion of a BLR cloud is treated like a classical two-body problem where a cloud with a fixed mass is subject to the central gravity and a force due to the radiation. Most of the previous analytical studies of BLR cloud's dynamics actually follow this approach.
\cite{netzer2010} studied orbital motion of pressure-confined BLR clouds in AGNs considering the combined influence of the central gravity and the radiation pressure. A modified estimate for the mass of the central black hole is presented according to their orbital analysis. \cite{krause11} addressed stability of the orbits using analytical calculations for both isotropic and anisotropic light sources and found that stable orbits may exist under certain circumstances. Although it is unlikely to obtain analytical solutions for the orbital motion of BLR clouds under general conditions, an interesting analytical solution for the orbit of BLR clouds with a fixed column density has been obtained by \cite{plewa}. In all these works, the intercloud is a simple power-law prescription not based on a physically supported model. Moreover, variations of the intercloud's pressure profile with the polar angle has been neglected. These issues motivated \cite{khajenabi15} to study orbital motion of BLR clouds through an ADAF atmosphere where its pressure profile is based on a two-dimensional self-similar analytical solution for ADAFs \citep{shadmehri14}. Under these conditions it was shown that stability of the orbits implies that the ensemble of clouds tends to have a disc like configuration.
None of the above studies about orbital motion of BLR clouds considered interaction of the clouds with the surrounding gas via a drag force. As for the G1 and G2 clouds, recent studies show that the drag force has a vital role in the orbits of these clouds \citep[e.g.,][]{McCourt2015,Pfuhl2015}. Just recently, \cite{shadmehri15} studied orbits of BLR clouds subject to a drag force proportional to the velocity. For a particular set of the input parameters, \cite{shadmehri15} presented an analytical solution for the orbits of the clouds which reduces to the analytical solution of \cite{plewa} in the absence of the drag force. In the presence of the drag force, irrespective of the input parameters, orbit of a BLR cloud would decay in a way that it will eventually fall onto the central region. According to the arguments of \cite{shadmehri15} if the time that takes a cloud to reach from its initial position to the central part, or time-of-flight, becomes less than the lifetime of the whole system, then BLR clouds are transient structures rather than long-lived objects so that mechanisms for continually forming BLR clouds are needed. In other words, drag force implies a physical constraint for analyzing orbits of the clouds. \cite{shadmehri15} found that time-of-flight of a BLR cloud is proportional to the inverse of the dimensionless drag coefficient and using this relation he showed that time-of-flight is indeed shorter than the lifetime of the whole system for a wide range of the input parameters. This interesting finding implies existence of mechanisms for continuously forming these clouds.
However, there are caveats regarding to the analysis of \cite{shadmehri15}. First of all, in his study the drag force is proportional to the velocity which is valid as long as the intercloud is laminar. Although he argues that Reynolds number is less than one which confirms the adopted drag force, for some other input parameters one can easily show that Reynolds number could be much larger than one. Introducing Reynolds number as ${\rm Re}=\rho v L /\mu$, it can be rewritten as ${\rm Re}=2(v/\bar{u})(L/\lambda)$, where $\rho$, $v$, $L$, $\mu$, $\bar{u}$ and $\lambda$ are the density of gas, the mean velocity of the cloud relative to the gas, characteristic length, dynamic viscosity, the average molecular speed and the mean free path, respectively. If the number density of the intercloud gas is $10^4$ cm$^{-3}$ and its average temperature is $10^8$ K, then we have $\bar{u}\approx 2\times 10^6$ m s$^{-1}$ and $\lambda \approx 10^{10}$ m. The Keplerian velocity at the radial distance 1 pc from the central mass $10^6$ solar mass is around $6.5\times 10^4$ m s$^{-1}$. If we adopt velocity of a cloud approximately equal to this Keplerian velocity and for a typical length $10^{13}$ cm, the Reynolds number becomes around 0.65. Obviously, if the typical length is taken larger, say $10^{15}$ cm, then we have ${\rm Re} = 65$. Also, for a more massive central mass, the Reynolds number is larger than unity. It means that the intercloud medium is turbulent and the drag force should be taken in proportion to the velocity square. In the present work, we plan to study orbits of BLR clouds with a quadratic drag force. At variance with previous work, nevertheless, the intercloud is prescribed using the standard ADAF solutions. Under these circumstances, we calculate time-of-flight of the clouds to see if the main finding of \cite{shadmehri15} is still valid when the drag force is a quadratic function of the velocity. Moreover, in most previous studies, the background gas is assumed to be in a static configuration. We also consider rotation of the medium which a cloud moves through it. In next section, we present basic assumptions and the orbital equations. In section 3, time-of-flight is calculated numerically. We then conclude with our main findings in section 4.
\section{General Formulation}
\subsection{Basic Assumptions}
We study orbital motion of a BLR cloud with mass $m$ subject to three main forces, i.e. gravitational force of a central black hole with mass $M$, a non-isotropic force due to the radiation of a central accretion disc \citep{Liu11}, and a drag force in the opposite direction of the BLR orbital motion. Under these circumstances, direction of cloud's angular momentum is conserved, though its magnitude gradually decreases because of the resistive force. Therefore, motion of a BLR cloud will be in a plane where its inclination $i$ is fixed by the initial angular momentum and it is an input parameter in our model. A system of coordinates $(x,y,z)$ is constructed so that the central black hole locates at its origin and the central radiating thin accretion is at $z=0$ plane (Figure 1). Thus, location of a cloud in its orbit with inclination angle $i$ with respect to the $x-y$ plane is uniquely determined by its distance $r$ from the origin and the polar angle $\theta$. The cloud orbit intersects the $x - y$ (disk)plane at the ascending node $A$ so that we define the angle $\angle AOC=\psi$. Having the inclination angle $i$, position of a cloud is determined by $r$ and $\psi$.
\begin{figure}\label{fig:f1}
\includegraphics[scale=0.45]{f1.eps}
\caption{The central black hole locates at origin $O$ and position of the cloud C is determined by the radial distance $r$ and the polar angle $\theta$ and the inclination angle of the orbital plane. Since direction of the angular momentum is conserved, the orientation of the orbital plane is fixed and its intersection with the $x-y$ plane is denoted by OA. Thus, position of the cloud C is equivalently determined by the radial distance $r$ and the angle $\angle AOC=\psi$.}
\end{figure}
Geometrical shape of a BLR cloud is assumed to be spherical, for simplicity. Moreover, the clouds are considered to be optically thick. But since the clouds are pressure-confined by definition, physical properties of the ambient gaseous medium like its pressure profile determines how the radius of a cloud varies depending on its position in orbital motion. There are considerable uncertainties about the true nature of intercloud medium. In other words, irrespective of the confinement mechanisms, internal pressure of a cloud is in balance with the ambient pressure.
Since each cloud is in pressure equilibrium with the hot background gas, its radius becomes $R_{\rm cl}\ \propto P_{\rm gas}^{-1/3}$ where $P_{\rm gas}$ is
the intercloud gas pressure. On the other hand, according to the standard similarity solutions for
ADAFs \citep{narayan94}, the pressure distribution is proportional to a power-law function
of the radial distance as $P_{\rm gas} \propto r^{-s}$ where $s$ is 5/2.
Therefore, we can rewrite radius of a single cloud as a function of its location, i.e.
\begin{equation}
R_{\rm cl} = R_{\rm cl0} \left ( \frac{r}{r_0}\right )^{5/6},
\end{equation}
where $r_0$ is the initial radial distance of the cloud, and $R_{\rm cl0}$ is the radius of the cloud at $r_0$. We can calculate the column density $N_{\rm cl} = 3m/2\mu_{\rm m} m_p A $, where $\mu_{\rm m}$ is the mean molecular weight and $A$ is cross section of a cloud. Having the above relations for $R_{\rm cl}$ and $P_{\rm gas}$, the column density of a pressure-confined cloud becomes
$N_{\rm cl} \propto P_{\rm gas}^{2/3}$ or $N_{\rm cl}=N_0 (r/r_0 )^{-5/3}$
where $N_0 = 3m/(2\mu_{\rm m} m_p \pi R_{\rm cl0}^2)$ is a constant column density. Also, the density of gas in the standard ADAF similarity solution is written as a power-law function of the radial distance, i.e.
\begin{equation}\label{eq:adaf-density}
\rho = \rho_0 \left( \frac{r}{r_0} \right)^{-3/2} ,
\end{equation}
where $\rho_0$ is the mass density of the intercloud gas at radius $r_0$. We adopt the outer radius of the system as $r_0$ with a value between 0.01 pc to 1 pc and the number density is $n_0 \sim 10^4 $ cm$^{-3}$ according to the observations \citep[e.g.,][]{Rees89,netzerbook,marconi,plewa}.
The radial velocity $v_{r}$ and the rotational velocity $v_{\varphi}$ of an ADAF are also power-law function of the radial distance. Similarity solutions of \cite{narayan94} are written $v_{r}=-v_{0r} v_{K}$ and $v_{\varphi}=v_{0\varphi} v_{K}$, where $v_{K}=\sqrt{GM/r}$ is Keplerian velocity and the coefficients $v_{0r}$ and $v_{0\varphi}$ are obtained as,
\begin{equation}\label{eq:vr}
v_{0r}=(5+2\varepsilon ')\frac{g(\alpha,\varepsilon ')}{3\alpha},
\end{equation}
and
\begin{equation}\label{eq:vphi}
v_{0\varphi}=\sqrt{\frac{2\varepsilon '(5+2\varepsilon')}{9\alpha^2}g(\alpha,\varepsilon ')} .
\end{equation}
Here, we have
\begin{equation}\label{eq:ga}
g(\alpha,\varepsilon ')=\sqrt{1+\frac{18\alpha^{2}}{(5+2\varepsilon ')^2}}-1 ,
\end{equation}
where $\alpha$ is the standard Shakura-Sunyaev viscosity parameter for modeling ADAF's turbulence. Moreover, the parameter $ \varepsilon '$ is written as $ \varepsilon '=\varepsilon/f$ where
$\varepsilon = (5/3-\gamma)/(\gamma-1)$, and $\gamma$ is the ratio of specific heats and the parameter $f$ measures the amount of the advected energy. For example, in a fully advective flow with $\alpha =0.1$ and $\epsilon ' =1$, we obtain $v_{0r} \simeq 0.04$ and $v_{0\varphi} \simeq 0.53$.
\subsection{Equations of Motion in a Static Atmosphere}
We can now write equations of motion of a cloud which is under the influence of the three main forces: the gravitational force of the central mass, ${\bf F}_{\rm grav}$, a force due to non-isotropic radiation of the central accretion disc \citep{Liu11}, ${\bf F}_{\rm rad}$, and a drag force proportional to the velocity square in the opposite direction of the cloud's motion, ${\bf F}_{\rm drag}$. These forces can be written as
\begin{equation}
{\bf F}_{\rm grav} =- \frac{GMm}{r^2} {\bf e}_{\rm r},
\end{equation}
\begin{equation}
{\bf F}_{\rm rad} = \frac{A}{c} \frac{L_{\rm a}}{2\pi r^2} |\cos \theta | {\bf e}_{\rm r},
\end{equation}
\begin{equation}
{\bf F}_{\rm drag} = -\frac{1}{2} \rho C_D A |{\bf v}| {\bf v},
\end{equation}
where $A$ is the cross sectional area of a cloud and $L_{\rm a}$ is the luminosity of the central source. In the above equation for the drag force, $C_{D}$ is the drag coefficient which depends on the shape of the cloud and even Reynolds number \citep[e.g.,][]{fluid}. For a sphere, value of $C_D$ may vary from large values for laminar flow to 0.47 for turbulent flow \citep{fluid}.
It is more convenient to re-write the force due to the radiation in terms of the column density $N_{\rm cl}$ and the Eddington ratio $l= L_{\rm a}/ L_{\rm edd}$, where $L_{\rm edd} = 4\pi GM m_p c /\sigma_{\rm T}$ is the Eddington luminosity. Here, $\sigma_{\rm T}$ is the Thompson cross-section. Thus, the force due to the radiation becomes
\begin{equation}
{\bf F}_{\rm rad} = \frac{GMm}{r^2} \frac{3l}{\mu_{\rm m} N_{\rm cl} \sigma_{\rm T}} |\cos\theta | {\bf e}_{\rm r},
\end{equation}
or
\begin{equation}
{\bf F}_{\rm rad} = \frac{GMm}{r^2} k |\sin \psi | {\bf e}_{\rm r},
\end{equation}
where the dimensionless parameter $k$ is defined as
\begin{equation}
k = \frac{3l}{\mu_{\rm m} N_{\rm cl} \sigma_{\rm T}} \sin (i) .
\end{equation}
Substituting radial dependence of the column density, the parameter $k$ becomes $k=k_0 (r/ r_0 )^{5/3}$ where $k_0 = 3l \sin (i) /{\mu_{\rm m}} N_0 \sigma_{\rm T}$.
Most of previous authors assume that the intercloud is static which means the background gas does not move. Thus, the relative velocity of a cloud with respect to its ambient medium is cloud's velocity itself. We first consider this simplified situation. Thus, equations of the orbital motion are written as
\begin{equation}\label{eq:main11}
\ddot{r} - r\dot{\psi}^{2} = \frac{GM}{r^2} \left ( k |\sin \psi | - 1 \right ) - \frac{\rho C_{D} A}{2}
\dot{r} \sqrt{\dot {r}^2+r^2 \dot{\psi}^2} ,
\end{equation}
\begin{equation}\label{eq:main22}
r \ddot{\psi} + 2 \dot{r} \dot{\psi} = - \frac{\rho C_{D} A}{2} r \dot{\psi} \sqrt{\dot {r}^2+r^2 \dot{\psi}^2} ,
\end{equation}
where $\dot{r}=dr/dt$, $\ddot{r}=d^2 r/dt^2$ and $\dot{\psi}=d\psi /dt$. Note that the temperature of a cloud during its orbital motion is almost constant.
The above orbital equations (\ref{eq:main11}) and (\ref{eq:main22}) are now written in the non-dimensional forms which are more convenient for the numerical integration. Thus, we use the initial radial distance $r_0$ as a reference length scale. Then, Keplerian velocity at this radial distance is written as $v_{\rm K}(r_0 ) = \sqrt{GM/ r_0 }$ and our unit time becomes $t_0 = r_0 / v_{\rm K}(r_0 )$. We now change the variables as $r = r_0 \tilde{r}$ and $t = t_0 \tau $. Thus, equations (\ref{eq:main11}) and (\ref{eq:main22}) become
\begin{equation}\label{eq:main1}
\ddot{\tilde{r}} -\tilde{r} \dot{\psi}^2 = \frac{1}{\tilde{r}^2}\left(k_0 \tilde{r}^{5/3}|\sin\psi | - 1 \right )-\beta \dot{\tilde{r}} \tilde{r}^{\frac{1}{6}}\sqrt{\dot{\tilde{r}}^2+\tilde{r}^2\dot{\psi}^2},
\end{equation}
and
\begin{equation}\label{eq:main2}
\tilde{r}\ddot{\psi} + 2\dot{\tilde{r}}\dot{\psi} = -\beta \dot{\psi}\tilde{r}^{\frac{7}{6}}\sqrt{\dot{\tilde{r}}^2+\tilde{r}^2\dot{\psi}^2},
\end{equation}
where $\dot{\tilde{r}}=d\tilde{r}/d\tau $, $\ddot{\tilde{r}}=d^2 \tilde{r}/d\tau^2 $ and $\dot{\psi}=d\psi /d\tau $. The dimensionless drag coefficient is denoted by $\beta$, i.e.
\begin{equation}\label{eq:beta}
\beta = \frac{3}{8}C_{D}\left ( \frac{r_{0}}{R_{\rm cl0}} \right ) \left (\frac{\rho_0}{\rho_{\rm cl0}} \right ).
\end{equation}
In writing the above equations, we assume that the mass of cloud is conserved during its orbital motion. We note that radius of a cloud and its density at the distance $r_0$ are denoted by $R_{\rm cl0}$ and $\rho_{\rm cl0}$, respectively. Equations (\ref{eq:main1}) and (\ref{eq:main2}) are our main equations for determining orbit of a cloud in a static atmosphere when the drag force is proportional to the velocity square. In section 3, we solve these equations to analyze orbits of a cloud.
\subsection{Equations of Motion in a Rotating Atmosphere}
We now consider a more realistic situation where the ambient gas is rotating and has radial velocity according to equations (\ref{eq:vr}) and (\ref{eq:vphi}). In writing the drag force, then, the relative velocity is considered. Thus, orbital equations become
\begin{displaymath}
\ddot{r} - r\dot{\psi}^{2} = \frac{GM}{r^2} \left ( k |\sin \psi | - 1 \right ) - \frac{\rho C_{D} A}{2}\times
\end{displaymath}
\begin{equation}\label{eq:main11b}
(\dot{r}+v_{0r}v_{K}) \sqrt{(\dot {r}+v_{0r}v_{K})^2+(r \dot{\psi}-v_{0\varphi}v_{K}})^2,
\end{equation}
\begin{displaymath}
r \ddot{\psi} + 2 \dot{r} \dot{\psi} = - \frac{\rho C_{D} A}{2}\times
\end{displaymath}
\begin{equation}\label{eq:main22b}
(r \dot{\psi}-v_{0\varphi}v_{K} )\sqrt{(\dot {r}+v_{0r}v_{k})^2+(r \dot{\psi}-v_{0\varphi}v_{K}})^2 .
\end{equation}
\begin{figure}\label{fig:f2}
\includegraphics[scale=0.45]{f2.eps}
\caption{Orbit of a BLR cloud in the plane of motion with inclination angles $i=\pi /18$ (top) and $i=\pi /3$ (bottom) for $\beta =0.01$. The other input parameters are $\alpha =0.1$, $\epsilon' =1$, ${\mu_{\rm m}} N_0 \sigma_{\rm T} =3/2$ and $l =0.1$. Initial conditions are $\tilde{r}(\tau =0 )=1$, $\dot{\tilde{r}} (\tau =0)=0$, $\psi (\tau =0) =0$ and $\dot{\psi} (\tau =0)=1$.}
\end{figure}
\begin{figure}\label{fig:f3}
\includegraphics[scale=0.45]{f3.eps}
\caption{Same as Figure 2, but with inclination angles $i=\pi /18$ (solid) and $i=\pi /3$ (dashed) and a larger dimensionless drag coefficient, i.e. $\beta =0.1$. }
\end{figure}
\begin{figure}\label{fig:f4}
\includegraphics[scale=0.45]{f4.eps}
\caption{Same as Figure 2, but with inclination angles $i=\pi /18$ (solid) and $i=\pi /3$ (dashed) and a larger dimensionless drag coefficient, i.e. $\beta =0.5$.}
\end{figure}
\begin{figure}\label{fig:f5}
\includegraphics[scale=0.45]{f5.eps}
\caption{ Orbit of a BLR cloud in the plane of motion for cases with different values of the parameter $l$. Here, we have $\beta =1$ and $i=\pi /3$ and the rest of the input parameters and the initial conditions are similar to figure 2.}
\end{figure}
\begin{figure}\label{fig:f6}
\includegraphics[scale=0.45]{f6.eps}
\caption{ Ratio of $\left [ \tau_f (i) - \tau_f (i=0) \right ] /\tau_f (i=0)$ as a function of the parameter $\beta$ when the intercloud is static (solid) or rotating (dashed) for different inclination angles. Input parameters are $\alpha =0.1$, $\epsilon' =1$, ${\mu_{\rm m}} N_0 \sigma_{\rm T} =3/2$ and $l =0.25$. The initial conditions are $\tilde{r}(\tau =0 )=1$, $\dot{\tilde{r}} (\tau =0)=0$, $\psi (\tau =0) =0$ and $\dot{\psi} (\tau =0)=1$. Each curve is labeled with corresponding value of $i$. When the intercloud is static and the inclination angle is $i=\pi /6$ (solid), there is an increase in the time-of-flight over the case when the inclination angle is zero (i.e., radiation forces do not operate) as the dimensionless drag coefficient decreases. This enhancement in the time-of-flight is more significant for the smaller values of the dimensionless drag coefficient. Exactly the same behavior is found when the intercloud is rotating and the inclination angle is $i=\pi /6$, but its difference with the shown case with static atmosphere (solid) is negligible and for this reason, this particular case has not been shown here. }
\end{figure}
Again, it is more convenient to use non-dimensional equations instead of the above orbital
equations. So, we transform equations (\ref{eq:main11b}) and (\ref{eq:main22b}) to the following non-dimensional forms:
\begin{displaymath}
\ddot{\tilde{r}} - \tilde{r}\dot{\psi}^{2} = \frac{1}{\tilde{r} ^2}\left(k_0 \tilde{r}^{5/3}|\sin\psi | - 1 \right) - \beta\times
\end{displaymath}
\begin{equation}\label{eq:main111}
(\dot{\tilde{r}}+v_{0r}\tilde{r}^{-1/2}) \tilde{r}^{\frac{1}{6}}\sqrt{(\dot {\tilde{r}}+v_{0r}\tilde{r}^{-1/2})^2+(\tilde{r} \dot{\psi}-v_{0\varphi}\tilde{r}^{-1/2})^2},
\end{equation}
\begin{displaymath}
\tilde{r} \ddot{\psi} + 2 \dot{\tilde{r}} \dot{\psi} = -\beta\times
\end{displaymath}
\begin{equation}\label{eq:main222}
( \dot{\psi}-v_{0\varphi}\tilde{r}^{-3/2}) \tilde{r}^{\frac{7}{6}} \sqrt{(\dot {\tilde{r}}+v_{0r}\tilde{r}^{-1/2})^2+(\dot{\tilde{r}} -v_{0\varphi}\tilde{r}^{-1/2})^2},
\end{equation}
where dimensionless parameter $\beta$ is defined in equation(\ref{eq:beta}). The above equations are solved subject to the appropriate initial conditions in the next section.
\begin{table}
\caption{ Our input parameter for describing a BLR cloud and its ambient gas and the resulting dimensionless drag coefficient. Note that each row corresponds to a cloud with a certain mass.}
\begin{tabular}{|l|l|l|l|l| }
\hline
$r_0$ & $n_0$ & $R_{\rm cl0}$ & $n_{\rm cl0}$ & $\beta / C_D $ \\ \hline
1 pc & $10^4$ cm$^{-3}$ & $10^{14}$ cm & $10^{10}$ cm & $1.15\times 10^{-2}$ \\ \hline
1 pc & $10^4$ cm$^{-3}$ & $10^{12}$ cm & $10^{10}$ cm & $1.15$ \\ \hline
0.01 pc & $10^4$ cm$^{-3}$ & $10^{14}$ cm & $10^{10}$ cm & $1.15\times 10^{-4}$ \\ \hline
0.01 pc & $10^4$ cm$^{-3}$ & $10^{12}$ cm & $10^{10}$ cm & $1.15\times 10^{-2}$ \\ \hline
\end{tabular}
\end{table}
\begin{table}
\caption{ Same as Table 1, but each row corresponds to a BLR cloud with a mass equal to $10^{-8}$ solar mass.}
\begin{tabular}{|l|l|l|l|l| }
\hline
$r_0$ & $n_0$ & $R_{\rm cl0}$ & $n_{\rm cl0}$ & $\beta / C_D $ \\ \hline
1 pc & $10^4$ cm$^{-3}$ & $10^{14}$ cm & $4.6\times 10^{6}$ cm & $25$ \\ \hline
1 pc & $10^4$ cm$^{-3}$ & $10^{12}$ cm & $4.6\times 10^{12}$ cm & $2.5\times 10^{-3}$ \\ \hline
0.01 pc & $10^4$ cm$^{-3}$ & $10^{14}$ cm & $10^{10}$ cm & $0.25$ \\ \hline
0.01 pc & $10^4$ cm$^{-3}$ & $10^{12}$ cm & $10^{10}$ cm & $2.5\times 10^{-5}$ \\ \hline
\end{tabular}
\end{table}
\section{ANALYSIS}
We now examine orbits of the BLR clouds in the plane of motion by solving the orbital equations. The background gas is rotating according to ADAF solutions. Describing the results is easier if the same initial conditions are used for all the considered cases. Thus, we assume that a cloud starts its journey from the initial location $\tilde{r}=1$ (note that all variables are dimensionless). The rest of the initial conditions are $\dot{\tilde{r}}(\tau =0)=0$, $\psi (\tau =0)=0$ and $\dot{\psi}(\tau =0)=1$. We found that shape of the orbits is qualitatively similar to when a linear drag is used, i.e. orbit of a BLR cloud decays due to the resistive nature of the drag force. But we can calculate the time-scale of this orbital decay when the drag force is quadratic. In doing so, time-of-flight $\tau_f$ is defined as the time needed for traveling of a cloud from its initial location to the center. In order to determine the orbital shape of a BLR cloud and its time-of-flight, we have to adopt the input parameters consistent with the observational data. According to the observations, we have $r_0 \sim 0.01-1 $ pc, $R_{\rm cl0} \sim 10^{14}$ cm, $n_0 \sim 10^4 $ cm$^{-3}$ and $n_{\rm cl0} \sim 10^{10}$ cm$^{-3}$ \citep[e.g.,][]{Rees89,netzerbook,marconi,plewa}. Tables 1 and 2 summarize our input parameters; however, in Table 2 the mass of a BLR cloud is assumed to be $10^{-8}$ ${\rm M}_{\odot}$.
Figure 2 displays orbital shape of a BLR cloud in the plane of motion (i.e., $XY$-plane where $X$-axis is along OA in Figure 1) with inclination angles $i=\pi /18$ (top) and $i=\pi /3$ (bottom) for dimensionless drag coefficient $\beta =0.01$. Input parameters are $\alpha =0.1$, $\epsilon' =1$ and $k_0 =0.2\sin i$. Also, the initial conditions are $\tilde{r}(\tau =0 )=1$, $\dot{\tilde{r}} (\tau =0)=0$, $\psi (\tau =0) =0$ and $\dot{\psi} (\tau =0)=1$. Radial distance of a cloud gradually decreases because of considering the drag force. The non-isotropic nature of the central radiation becomes more significant with increasing the inclination angle $i$. In Figures 3 and 4 orbital motion of clouds with the same initial and input parameters are explored but with larger values for the drag coefficient.
In our model, the effect of the radiation force on the orbit of a cloud appears through the dimensionless parameter $k_0$ which is directly proportional to the Eddington ratio $l$ and $\sin i$. Thus, radiation force operates more significantly in cases with a high inclination angle or a large Eddington ratio. In order to have bound orbits, the radiation force can not be arbitrary large and for a given set of the input parameters, however, there is always a maximum critical value of parameter $k_0$ so that beyond this value the gravitational force is not able to keep a cloud in a bound orbit. In Figures 2-4 the orbits are shown for two values of inclination. Since radiation force pushes a cloud toward larger radii, one can expect cases with a larger inclination angle exhibit wider orbits in comparison to a case with a smaller inclination angle. This speculation has been confirmed in Figures 2-4. The effect of the Eddington ratio on the shape of orbits are explored in Figure 5 for different values of the Eddington ratio $l$. Here, we have $\beta =1$ and $i=\pi /3$ and the remaining input parameters are similar to Figure 2. Having all the parameters fixed, we found that the orbits are no longer bound once the Eddington ratio exceeds a value around 0.3. Nevertheless, the shape of orbits is not modified significantly so long as the ratio $l$ is roughly less than 0.1. Corresponding to the cases with $l=0.01, 0.1$ and 0.3, the dimensionless time-of-flight is found 7.47, 7.78 and 12.75, respectively.
We explored various cases with different sets of the input parameters and the corresponding dimensionless time-of-flight $\tau_f$ is obtained. Interestingly, we found that time-of-flight is in proportion to the inverse of the dimensionless drag coefficient $\beta$ so the constant of the proportionality depends on the input parameters. For a static intercloud gas, we found that $\tau_f (i=0) \approx 2.13 / \beta^{0.89}$, $\tau_f (i=\pi/6) \approx 2.20 / \beta^{0.90}$ and $\tau_f (i=\pi/3) \approx 2.52 / \beta^{1.01}$. When the intercloud is rotating, the time-of-flight is obtained as $\tau_f (i=0) \approx 7.38 / \beta^{0.97}$, $\tau_f (i=\pi/6) \approx 8.00 / \beta^{0.96}$ and $\tau_f (i=\pi/3) \approx 9.54 / \beta^{1.20}$. Except for the cases with zero inclination angle where the radiation force does not operate, however, in other cases the above fitted time-of-flight functions are not valid for the whole range of the dimensionless drag coefficient $\beta$. For a static intercloud gas, a BLR cloud will not be in bound orbit once the parameter $\beta$ drops to values less than 0.005 for $i=\pi/6$ and 0.1 for $i=\pi/3$. In a rotating intercloud gas, these critical values are larger so that we do not observe bound orbits when $\beta$ is less than 0.02 for $i=\pi/6$ and 0.43 for $i=\pi/3$. Figure 6 shows the ratio $\left [ \tau_f (i) - \tau_f (i=0) \right ] /\tau_f (i=0)$ as a function of the parameter $\beta$ when the intercloud is static (solid) or rotating (dashed) for different inclination angles.
Thus, we can write $\tau_f \simeq \tau_0 /\beta$ where $\tau_0$ depends on the input parameters. Although the constant of proportionality $\tau_0$ depends on the input parameters, we found that its variations with the input parameters does not affect significantly the main conclusion in our subsequent discussions. Time-of-flight for a linear drag is also proportional to inverse of the dimensionless drag coefficient, though definition of this coefficient is different from ours (see Eq.(8) in \cite{shadmehri15}). One can easily confirm our approximate relation for the time-of-flight using dimensional analysis. A cloud loses its kinetic energy $1/2 mv^2$ due to the dissipative nature of the drag force with a rate equal to $F_{d}v$, where $v$ is the velocity of the cloud and $m$ is its mass and $F_d$ is the drag force. Thus, time-of-flight can be written as $(1/2 mv^2)/(1/2 \rho C_D \pi R_{\rm cl}^2 v^3)$ which implies the dimensionless time-of-flight to be proportional to $\beta^{-1}$, i.e. $\tau_f \propto \beta^{-1}$.
For a cloudy BLR system around a black hole with mass $10^8$ M$_{\odot}$, our time unit becomes $t_0 \simeq 1.5$ yr if we set $r_0 = 0.01$ pc. Tables 1 and 2 show that the dimensionless drag coefficient varies from $2.5\times 10^{-5} C_D$ to $25 C_D$ depending on the background gas density and the properties of a cloud such as its density and radius. Obviously, the longest cloud flight times occur when the parameter $\beta$ is as small as permissible and the radiation force is as large as it can be. The effect of the radiation force does not appear for clouds with zero inclination angle and considering the above fitted functions for the time-of-flight, this time-scale will be between $\tau_f (\beta =10^{-5}) \simeq 3\times 10^6$ yr and $\tau_f (\beta =10) \simeq 0.2$ yr for a static intercloud gas. These estimates are modified in a rotating background gas as $\tau_f (\beta =10^{-5}) \simeq 10^7$ yr and $\tau_f (\beta =10) \simeq 0.7$ yr. For clouds with non-zero inclination angles, however, radiation pressure force increases $\tau_f$ as we confirmed in Figures 5 and 6. But in these cases, there is always a lower limit for $\beta$ so that for the drag coefficient less than this critical value clouds will be pushed outward due to the strong radiation force. When the background gas is static, for example, the explored cases in Figure 6 show that the critical value of $\beta$ is $0.005$ and $0.1$ for inclination angles $\pi /6$ and $\pi /3$, respectively. Then, the time-of-flight becomes $\tau_f (\beta =0.005) \simeq 390$ yr and $\tau_f (\beta =0.1) \simeq 39$ yr which are considerably shorter than the estimated $\tau_f$ for the clouds with zero inclination angle. Critical value of $\beta$ is larger in a system with a rotating intercloud gas and the corresponding time-of-flight is found as $\tau_f (\beta =0.02, i=\pi/6) \simeq 500$ yr and $\tau_f (\beta =0.43, i=\pi/3) \simeq 40$ yr.
Using this approximate relation for the time-of-flight, we can discuss about nature of BLR clouds by comparing it with the lifetime of the whole system $\tau_{\rm life}$. If $\tau_{f}$ becomes shorter than $\tau_{\rm life}$, all clouds will fall onto the central object and the system will be depleted of clouds unless replenishment mechanisms operate to generate new clouds. Observational evidences show that BLRs are clumpy \citep[e.g.,][]{Rees89}, though we do not know if they are continuously forming or long-lived objects. But if $\tau_f < \tau_{\rm life}$, then existence of mechanisms for generating new clouds are needed. The next step is to obtain a lower limit for the lifetime of the whole system. One can argue so long as a gas reservoir which is known as intercloud gas exists, these BLR clouds may form and move in their orbits. Thus, we can introduce the accretion time-scale as a lower limit for the lifetime of the whole system, i.e. $\tau_{\rm life} = M/\dot{M}$ where $\dot{M}$ is the accretion rate. Despite of uncertainty about the geometry and the nature of the accretion in these system, an approximate relation between the Eddington ratio and the accretion rate can be written as $l\simeq \dot{M}/\dot{M}_{\rm Edd} $, where $\dot{M}_{\rm Edd}$ is the Eddington accretion rate \citep[see p. 40,][]{netzerbook}. Thus, one can obtain $\tau_{\rm life} \simeq 4\times 10^8 \frac{\eta}{l}$ yr, where $\eta \sim 0.1$ is the mass-to-luminosity conversion efficiency \citep[][]{netzerbook}. Evidently, this estimated lifetime is much longer than the time-of-flight of BLR clouds except for clouds with zero inclination angle which may have $\tau_f \sim \tau_{\rm life}$ under very restrictive circumstances. This implies that mechanisms for continuous formation of BLR clouds should operate even when the drag force is quadratic. In the absence of such cloud creation mechanisms, however, it seems only clouds with orbital plane near to equatorial plane may survive and the rest of the clouds will fall onto central black hole very quickly and thereby, a disc like configuration for the geometry of spatial distribution of BLR clouds is expected \citep[also see,][]{khajenabi15}.
Although our model is based on the existence of an ensemble of discrete independent clouds in BLRs, more recent evidence may suggest that the system is not clumpy as has been studied by \cite{arav} in their attempt to find direct signature of discrete clouds in BLR of the Seyfert galaxy NGC 4151. They argued that BLRs are not made of independent clouds. Our analysis shows that even if BLR clouds do exist, they can not be long-lived due to the effect of the drag force. In the absence of a clear physical understanding of possible mechanisms for continuous formation of BLR clouds, however, it seems the system should be depleted of clouds which is consistent with the recent observations \citep{arav}.
We note that our analysis is based on an assumption which states that the clouds are in pressure equilibrium with their ambient medium. This constraint should not lead to unphysical values for the ratio of density of cloud to the intercloud density, i.e. $\rho_{\rm cl}/\rho$, which scales as the ratio of inter-cloud medium to cloud temperature. Since mass of cloud $m$ is conserved during its journey, we obtain $\rho_{\rm cl}=\rho_{\rm cl0} (r/r_0 )^{-5/2}$ where $\rho_{\rm cl0} = 3m/(4\pi R_{\rm cl0}^3)\sim 10^{10}$ cm$^{-3}$. Having equation (\ref{eq:adaf-density}) for the intercloud density, we obtain $\rho_{\rm cl}/\rho = \rho_{\rm cl0}/\rho_0 (r/r_0 )^{-1}$ or $\rho_{\rm cl}/\rho \sim 10^6 (r/r_0 )^{-1}$ which means the ratio of the densities can not be arbitrary large so long as a cloud is not very close to the central parts.
\section{Conclusion}
Our goal is to analyze orbits of BLR cloud subject to a drag force proportional to the velocity square. We calculated time-of-flight of the clouds for different initial conditions including motion through a static and rotating atmospheres. In all cases, however, we found that a system is generally older than typical time-scale of spiraling a cloud onto the central region. It means without mechanisms for continuous creation of clouds, a typical BLR system will eventually depleted of clouds. But this feature is not supported by observations. Thus, we can conclude BLR clouds are constantly forming.
\acknowledgments
I am very grateful to referee for his/her constructive report which greatly improved the quality of this paper.
\bibliographystyle{apj}
|
\section*{Supplementary material for: Suppressed magnetic circular dichroism and valley-selective magneto-absorption due to the effective mass anisotropy in bismuth}
\subsection*{Pieter J. de Visser, Julien Levallois, Micha\"{e}l K. Tran, Jean-Marie Poumirol, Ievgeniia O. Nedoliuk, J\'{e}r\'{e}mie Teyssier, Ctirad Uher, Dirk van der Marel, Alexey B. Kuzmenko}
\maketitle
\section{Measurement of reflectivity and Kerr angle}
\begin{figure}[h]
\includegraphics[width=0.99\columnwidth]{FigSup1.pdf}
\caption{\label{figS:reflkerr} Reflectivity (a) and Kerr angle (b) measured as a function of frequency for the magnetic fields indicated in the legend. The Kerr angle curves are shifted by 0.015 for clarity. The small spike that shows up at 193 cm$^{-1}$ for some fields in the Kerr angle is due to microphonic noise.}
\end{figure}
The reflectivity and Kerr angle spectra were measured in a single measurement run using two polarisers, one before and one after the sample. For the lowest energies (2-12 meV) we have used a mercury source and a Si beamsplitter and for the higher energies (12-87 meV) a Globar source with a Ge-coated mylar beamsplitter. Both ranges were measured with He cooled bolometers with a resolution of 1 cm$^{-1}$ and 2 cm$^{-1}$ respectively, using a Bruker V70 spectrometer. The absolute reflectivity, $R(\omega,B)$, was measured with parallel polarisers, using a double reference method: $R(\omega,B) = (I_s(\omega,B)/I_{r,s}(\omega,B))/(I_g(\omega,B)/I_{r,g}(\omega,B))$. At every field, the intensity spectrum of the sample, $I_s(\omega,B)$, and a spectrum on a gold reference mirror $I_{r,s}(\omega,B)$ were taken, to compensate for temporal drifts and possible small magnetic field dependences of the alignment. The same set of measurements was performed after evaporating a gold layer on the sample, giving $I_g(\omega,B)$ on the gold-coated sample and $I_{r,g}(\omega,B)$ on the reference mirror.
The Kerr angle was measured using the so-called fast protocol, introduced and explained in detail in Ref. \onlinecite{jlevallois2015}, valid for small Kerr angles. It relies on measuring the ratio of the reflected intensity at positive and negative fields $\rho(\omega,B) = I(\omega,+B)/I(\omega,-B)$ at analyser angles of +45 and -45 degrees for both the sample and the mirror reference. The Kerr angle is then given by
\begin{equation}
\theta_K(\omega,B) \cong \frac{1}{8}\left(\frac{\rho_{+45}(\omega,B)}{\rho_{+45,r}(\omega,B)} - \frac{\rho_{-45}(\omega,B)}{\rho_{-45,r}(\omega,B)} \right).
\label{eq:Kerrmeas}
\end{equation}
Note that the evaporated gold reference is not needed for $\theta_K$.
Figure \ref{figS:reflkerr} shows the measured reflectivity and Kerr angle as a function of wavenumber for different magnetic fields. These experimental results are used to calculate the optical conductivity for left and right circularly polarised light as shown in Figure 2 of the main text.
\section{Derivation of the equation (5) from the main text}
Although the equation (5) is very general, we shall explicitly consider the two cases of parabolic and Dirac bands corresponding to the hole and electron Fermi pockets respectively. The hole Hamiltonian in zero magnetic field is well approximated by
\begin{equation}\label{Ham-h}
H_{h}=E_0+\Delta-\frac{\hbar^2(k_{x}^2+k_{y}^2)}{2M_{c}}-\frac{\hbar^2 k_{z}^2}{2M_{z}},
\end{equation}
\noindent where $M_{z}$ is the effective mass of holes along the $z$-axis, and the definition of other parameters is given in the main text. It was established long time ago \cite{pwolff1964} that the band structure near the electron pocket is in many respects similar to the famous four-component relativistic Dirac Hamiltonian describing electrons and positrons. Although the actual band structure of electrons in bismuth is highly anisotropic, it is often substituted by a simplified model\cite{zzhu2011}:
\begin{equation}\label{Ham-e}
H_{e}=\left[
\begin{matrix}
\Delta & 0 & i\hbar v_{z}k_{z} & i\hbar v(k_{x} - i k_{y})\\
0 & \Delta & i\hbar v(k_{x} + i k_{y}) & -i\hbar v_{z}k_{z}\\
-i\hbar v_{z}k_{z} & -i\hbar v(k_{x} - i k_{y}) & -\Delta & 0\\
-i\hbar v(k_{x} + i k_{y}) & i\hbar v_{z}k_{z} & 0 & -\Delta\\
\end{matrix}\right],
\end{equation}
\noindent where $v=(\Delta/m_{c})^{1/2}$ and $v_{z}=(\Delta/m_{z})^{1/2}$ and $m_{z}$ is the effective mass of electrons along the $z$-axis. This Hamiltonian describes correctly the Landau level energies and therefore is suitable to fit the transport data , such as the Nernst effect \cite{zzhu2011}, and thermodynamic properties. However, it does not capture the strong anisotropy in the $xy$ plane and therefore cannot describe the effects reported in this article. Below we will show how to modify it properly.
For a magnetic field $B$ applied along the $z$ axis, we follow a standard procedure and perform the substitution:
\begin{equation}\label{Kohn-iso}
k_x + ik_y \rightarrow \frac{l_B}{\sqrt{2}}a^{+} \mbox{ , }
k_x - ik_y \rightarrow \frac{l_B}{\sqrt{2}}a^{-}
\end{equation}
%
\noindent where ${l_B} = (\hbar/eB)^{1/2} $ is the magnetic length. The operators $a^{+}$ and $a^{-}$ act on harmonic oscillator-like wavefunctions $\psi_{n=0,1,2..}(B,x,y)$ obeying the ladder algebra:
%
\begin{equation}\label{Ladder}
a^{+}\psi_{n=0,1..} = (n+1)^{1/2}\psi_{n+1} \mbox{ , }
a^{-}\psi_{n=1,2..} = n^{1/2}\psi_{n-1} \mbox{ , } a^{-}\psi_{0} = 0.
\end{equation}
\noindent This results in:
\begin{equation}\label{Ham-h-B}
H_{h}(B)=E_0+\Delta-\hbar\omega_c \left(a^+a^-+\frac{1}{2}\right)-\frac{\hbar^2 k_{z}^2}{2M_{z}},
\end{equation}
\noindent and
\begin{equation}\label{Ham-e-B}
H_{e}(B)=\left[
\begin{matrix}
\Delta & 0 & i\hbar v_{z}k_{z} & i \frac{\sqrt{2}}{l_{B}}\hbar va^{-}\\
0 & \Delta & i \frac{\sqrt{2}}{l_{B}}\hbar va^{+} & -i\hbar v_{z}k_{z}\\
-i\hbar v_{z}k_{z} & -i \frac{\sqrt{2}}{l_{B}}\hbar va^{-} & -\Delta & 0\\
-i \frac{\sqrt{2}}{l_{B}}\hbar va^{+} & i\hbar v_{z}k_{z} & 0 & -\Delta
\end{matrix}\right],
\end{equation}
\noindent where we omitted for simplicity the Zeeman term, which does not affect the spin-conserving transitions. Diagonalisation of $H_{h}(B)$ and $H_{e}(B)$ provides the eigenvalues $E_{h}(n,k_z)$ and $E_{e}(n,k_z)$ with the corresponding wavefunctions $\Psi_{h}(n, x, y, k_{z})$ and $\Psi_{e}(n, x, y, k_{z})$, where $n$ is a Landau-level index. For the case of $k_z$ = 0, these eigenvalues are given in Eqs. (1) and (2) of the main text (apart from the Zeeman component).
The complex magneto-optical conductivity tensor can be calculated using the Kubo formula:
\begin{equation}\label{Kubo}
s_{\beta\gamma}(\omega,B) = \frac{e^3 B}{\pi^2}\int d k_{z}\sum_{n,n'}\langle \Psi_{n}|V_{\beta}|\Psi_{n'}\rangle \langle \Psi_{n'}|V_{\gamma}^{*}|\Psi_{n}\rangle
\frac{f(E_{n})-f(E_{n'})}{E_{n'}-E_{n}} \frac{i}{\hbar\omega -E_{n'}+E_{n}+i\Gamma_{nn'}},
\end{equation}
\noindent where $\beta$ and $\gamma$ are the Cartesian indices ($x$, $y$ and $z$), $n$ and $n'$ are the Landau-level indices of the initial and final states respectively , $f(E)=\left(\exp{\frac{E-E_{F}}{k_{B}T}} + 1\right)^{-1}$ is the Fermi-Dirac distribution, $E_{F}$ is the Fermi energy, $T$ is temperature, $\Gamma_{nn'}$ is the transition broadening and $V_{\beta}=\frac{1}{\hbar}\frac{\partial H}{\partial k_{\beta}}$ is the velocity operator. Note that we use here $s$ here for the conductivity of an isotropic system instead of the conventional $\sigma$, for consistency with the main text, where we reserve $\sigma$ for the measured conductivity. For a system rotationally invariant in the $xy$ plane we obviously have: $s_{yy} = s_{xx}$ and $s_{yx} = -s_{xy}$.
Next we consider the effect of anisotropy. Although the actual band structure is quite complex, giving rise to a small tilting of the electron pockets, we shall restrict ourselves to the simplest case, where effective masses $m_{x}$ and $m_{y}$ along the $x$ and $y$ axes are different. The effective cyclotron mass then becomes $m_{c} = (m_{x}m_{y})^{1/2}$. It was conjectured \cite{byang2012} that such mass anisotropy can be formally removed ('gauged out') by a proper spatial-metric transformation:
\begin{equation}\label{metric}
\tilde{x}=\alpha^{-1/2} x \mbox{ , }
\tilde{y}=\alpha^{1/2} y,
\end{equation}
\noindent or equivalently
\begin{equation}\label{metric2}
\tilde{k}_x=\alpha^{1/2} k_x \mbox{ , }
\tilde{k}_y=\alpha^{-1/2} k_y
\end{equation}
\noindent where $\alpha=(m_{x}/m_{y})^{1/2}$. In other words, the mass anisotropy is formally equivalent to a uniaxial stretching or squeezing of the spatial metric, in analogy with basic principles of general relativity. It is important that this deformation conserves not only the cyclotron mass but also the Fermi volume, the doping and the density of states. Therefore it does not influence most transport and thermodynamic properties. The transformation (\ref{metric}) maps an anisotropic model to the isotropic one. Indeed, in this case it is sufficient to redefine the a-operators:
\begin{equation}\label{Kohn-Aniso}
\tilde{k}_x + i\tilde{k}_y \rightarrow \frac{l_B}{\sqrt{2}}\tilde{a}^{+} \mbox{ , }
\tilde{k}_x + i\tilde{k}_y \rightarrow \frac{l_B}{\sqrt{2}}\tilde{a}^{-},
\end{equation}
\noindent which yields a Hamiltonian identical to (\ref{Ham-e-B}) but expressed in terms of $\tilde{a}^{\pm}$ instead of $a^{\pm}$. Therefore the Landau-level energies and wave-functions of an anisotropic Hamiltonian are easily expressed via the ones of the isotropic one:
\begin{eqnarray}\label{EBAnis}
\tilde{E}_{n}(B, k_{z})&=&E_{n}\left(B, k_{z}\right) \\
\tilde{\Psi}_{n}(B, x, y, k_{z})&=&\Psi_{n}\left(B, \alpha^{-1/2}x, \alpha^{1/2} y, k_{z}\right).
\end{eqnarray}
Our goal is to express the optical conductivity of a deformed system, $\tilde{\sigma}(\omega)$ in terms of the one of the isotropic one $s(\omega)$. If we take into account that the Landau level energies and the Fermi energy are not affected, while the velocity matrix elements are scaled as $\tilde{V}_{x}=\alpha^{-1/2} V_{x}$ and $\tilde{V}_{y}=\alpha^{1/2} V_{y}$, then, using the Kubo formula (\ref{Kubo}), we arrive straightforwardly at the equation (5) from the main text:
\begin{equation}\label{Eq5}
\begin{pmatrix}
\tilde{\sigma}_{xx} & \tilde{\sigma}_{xy} \\
-\tilde{\sigma}_{xy} & \tilde{\sigma}_{yy}
\end{pmatrix} =
\begin{pmatrix}
\alpha^{-1}s_{xx} & s_{xy} \\
-s_{xy} & \alpha s_{xx}
\end{pmatrix}.
\end{equation}
\noindent It is worth mentioning that in deriving this result we only used the fact that the original Hamiltonian is rotationally invariant, but did not use its specific form, i.e. equations (\ref{Ham-h-B}) or (\ref{Ham-e-B}). Therefore, the formula (5) from the main text is generally valid for any uniaxially deformed isotropic band structure.
\section{Valley-selective magneto-absorption for elliptical polarisation}
{A general elliptical polarization can be described by the Jones vector (in the $xy$ plane):
\begin{equation}\label{JonesVector}
\vec{E}=\frac{E_{0}}{\sqrt{1+\eta^2}}
\begin{pmatrix}
\cos\theta+i\eta\sin\theta \\
\sin\theta-i\eta\cos\theta
\end{pmatrix},
\end{equation}
\noindent where $\eta$ is the ellipticity (the ratio between the short and the long axes of the polarisation ellipse) and $\theta$ is the azimuth angle (the angle between the long axis and the $x$-axis). Note that $-1\leq\eta\leq1$ and is defined positive/negative for clockwise/counterclockwise rotations respectively. $\eta = \pm 1$ correspond to the two circular polarisations and $\eta = 0$ refers to a linear polarisation. The normalisation is chosen such that the radiation power does not depend on the ellipticity.
We shall assume that the total electric field in a certain location inside the sample is given by Equation (\ref{JonesVector}). The power (per unit volume) absorbed by electrons in the pocket e1 is given by:
\begin{equation}\label{Power}
P_{e1}=\Re(\vec{j}_{e1}^{*}\vec{E})=\Re(\vec{E}^{*}\hat{\sigma}_{e1}\vec{E}),
\end{equation}
\noindent where $\hat{\sigma}_{e1}$ is the optical conductivity tensor and $\vec{j}_{e1}$ is the current induced in this pocket. In the remaining two pockets e2 and e3 we have accordingly $P_{e2}=\Re(\vec{E}^{*}\hat{\sigma}_{e2}\vec{E})$ and $P_{e3}=\Re(\vec{E}^{*}\hat{\sigma}_{e3}\vec{E})$, where $\hat{\sigma}_{e2}=\hat{R}\hat{\sigma}_{e1}\hat{R}^{-1}$ and $\hat{\sigma}_{e3}=\hat{R}^{-1}\hat{\sigma}_{e1}\hat{R}$ ($\hat{R}$ is the rotation matrix by 120$^{\circ}$).
For intraband n-type Landau-level transitions the selection rule dictates that $s_{-} = 0$, which means that $s_{xx} = s_{+}/2$, $s_{xy} = -is_{+}/2$ in the Equation (5) from the main text:
\begin{equation}\label{SigmaMatrix}
\hat{\sigma}_{e1}=\frac{s_{+}}{2}
\begin{pmatrix}
\alpha^{-1} & -i \\
i & \alpha
\end{pmatrix},
\end{equation}
\noindent where we assume that $\alpha>1$ (in bismuth it is 14.4).
By substituting Equations (\ref{JonesVector}) and (\ref{SigmaMatrix}) into Equation (\ref{Power}), we get:
\begin{equation}\label{PowerThetaGen}
P_{e1}(\eta,\theta) = \frac{\Re (s_{+})E_{0}^{2}}{2}\left[\frac{\alpha^{-1}+\alpha}{2}+\frac{2}{\eta^{-1}+\eta}-\frac{\alpha^{-1}-\alpha}{2}\cdot\frac{\eta^{-1}-\eta}{\eta^{-1}+\eta}\cos2\theta\right].
\end{equation}
\noindent and correspondingly $P_{e2}(\eta,\theta) = P_{e1}(\eta,\theta+60^{\circ})$ and $P_{e3}(\eta,\theta) = P_{e1}(\eta,\theta-60^{\circ})$. One can see that for a given valley the absorbed power is modulated by the azimuth angle only if $\alpha > 1$ (anisotropic mass) and $\eta \neq\pm 1$ (non-circular polarisation). At the same time, the total power $P_{tot} = P_{e1} +P_{e2} +P_{e3}$ is azimuth-independent for any light ellipticity, which is expected for an optically isotropic system, such as bismuth in the $xy$-plane.
If the ellipticity of the electromagnetic wave is set to match the ellipticity of the Fermi pocket, and the rotation direction is opposite to the cyclotron motion of the electrons ($\eta = \alpha^{-1})$ then Equation (\ref{PowerThetaGen}) reduces to
\begin{equation}\label{PowerTheta}
P_{e1}(\theta) = \frac{\Re (s_{+})E_{0}^{2}}{2}\frac{(\alpha^{-1}-\alpha)^{2}}{\alpha^{-1}+\alpha}\sin^{2}\theta=P_{e1,max}\sin^{2}\theta.
\end{equation}
\noindent $P_{e1}$ becomes zero if $\theta = 0^{\circ}$, i.e. when the polarisation ellipse is orthogonal to the pocket e1, which is depicted in Fig. 5 in the main text together with plots of Eq. \ref{PowerTheta} for each pocket. Obviously, in the pockets e2 and e3 the absorption vanishes if $\theta$ is set to 120 and 60 degrees respectively.
\section{AFM of cleaved surface}
We have cleaved the Bi crystal in liquid nitrogen. The natural cleavage plane of a Bi crystal is perpendicular to the trigonal axis. To inspect the surface in detail, which looks very flat by the eye, we performed atomic force microscopy on it. At typical topographic image is shown in Figure \ref{figS:afm}a. A line profile is shown in Fig. \ref{figS:afm}b, showing height steps of approximately 0.4, 0.8 and 1.2 nm, which correspond to 1/3 cuts of the unit cell (which is 11.85 \r{A} high). By repeating several experiments on Bi crystals we have learned that the magneto-optics measurements on a pristine cleaved surface lead to the highest quality data (sharper transitions and more pronounced plasma features in reflectivity and Kerr angle).
\begin{figure}
\includegraphics[width=0.99\textwidth]{FigureSup2.pdf}
\caption{\label{figS:afm} (a) False colour atomic force microscopy image of a cleaved Bi crystal surface. (b) A typical line profile which shows atomic height variations.}
\end{figure}
\section{Extraction of the hole spectral weight from $\sigma_-$ and its uncertainty}
To derive the circular dichroism for the electron pockets from the measured $\sigma_-$ and $\sigma_+$, we need to subtract the transition peak due to the holes from $\sigma_-$ as discussed in the main text. In there we have shown in Fig. 4b one method to estimate the hole weight, by fitting Lorentzians to the electron and hole transitions in $\sigma_-$ and subtracting the weight of the hole. Note that the goal of the Lorentzians fit is only to extract the spectral weight of the hole transition, not to perfectly describe the other spectral features. However, from a careful inspection of the measurements the peak due to the hole transition is not completely Lorentzian. From $\sigma_+$ we also observe that the main electron transition at the highest energy does not have a purely Lorentzian shape either. Both aspects lead to an uncertainty in the derived hole weight.
\begin{figure}
\includegraphics[width=0.49\textwidth]{Method2Holeat5_5T.pdf}
\caption{\label{figS:methodhole2} Alternative way of deriving the hole spectral weight from the measured $\sigma_-$ and $\sigma_+$. $\sigma_+$, scaled to the height of the electron transitions in $\sigma_-$, is subtracted from the measured $\sigma_-$ to determine the weight of the hole transition in $\sigma_-$.}
\end{figure}
An alternative way to derive the weight in the hole peak is to use the knowledge that $\sigma_+$ only represents electron transitions. We can subsequently scale $\sigma_+$ to match the visible electron transitions in $\sigma_-$. We can then calculate $\sigma_{hole} = \sigma_--\sigma_{+,scaled}$ in the region where the hole peak dominates. This procedure is shown in Fig. \ref{figS:methodhole2}. This method has the advantage that it does not assume a particular shape for the electron and hole transitions. However, the shape of $\sigma_-$ and $\sigma_+$ due to the electron transitions is not exactly the same. Therefore we need to cut the integration boundaries at a finite point, which leaves an uncertainty on the spectral weight.
For the final dichroism as shown in Fig. 4d of the main text, we have taken the average dichroism obtained by the two methods described above. The uncertainty contains two parts. One part is taken to be the difference in dichroism between the two methods of hole subtraction. For the second part we estimate that each method by itself leads to an uncertainty in dichroism of 0.02, which adds to the average of the two methods an uncertainty $0.02/\sqrt{2}$. Together this leads to the error bars in Fig. 4d of the main text. The reported average dichroism $0.13\pm 0.01$ is calculated from a weighted average over the magnetic fields.
\section{Unidentified, low energy peak in the conductivity}
In the conductivity spectra in Fig. 2 of the main text one can identify a small additional peak at low energies. It is better visible in the spectra at 5.5 T as shown in Fig. 4a of the main text. It occurs in both $\sigma_-$ and $\sigma_+$ and its energy is plotted for the fields at which it is visible in Figure \ref{figS:smallpeak}. It cannot be identified in the same hole and electron picture since its field dependence does not correspond to the field dependence of Landau level, spin and combination resonances. The resonance shows up in both $\sigma_+$ and $\sigma_-$ and in this sense has characteristics similar to the electron-pockets near the L-point. However, the field dependence does not match the band structure of the bulk. A transition was observed before at the same energies (below the cyclotron frequencies) in an infrared transmission experiment \cite{jburgiel1965}, which was attributed in that paper to a combination transition (cyclotron and spin). However this explanation used a different frequency for cyclotron and spin transitions, which is not consistent with the extended two-band model which is generally accepted to describe the Landau level transitions in Bi and which we use in this work. We emphasise that the spectral weight of this feature is very small compared to the total spectral weight of the electron resonances and therefore does not influence the analysis of the circular dichroism. It is known that Bi exhibits surface states, which could also be a source of magnetic-field dependent transitions. However this experiment does not contain sufficient information to definitively ascribe this small peak to a surface state.
\begin{figure}
\includegraphics[width=0.49\textwidth]{FigureUnidentifiedPeak.pdf}
\caption{\label{figS:smallpeak} Energy of the small unidentified peak in the spectra of $\sigma_-$ and $\sigma_+$ as a function of magnetic field. For the highest three fields the energies overlap. For comparison this figure is on the same scale as in Figure 3 of the main text.}
\end{figure}
\end{document}
|
\section{Introduction and Motivation}\label{sec:introduction}
The traditional methods for user identification systems rely on entering a username and a text password. One of the major vulnerabilities of this technique is the difficulty of remembering passwords. Therefore the users tend to pick short passwords or passwords that are easy to remember. These passwords can be easily broken. Passwords that are hard to break are often hard to remember.
The ability of humans to remember pictures far better than words \cite{picture-memory-1,picture-memory-2} has recently motivated computer security researchers' in academia \cite{gp-survey-acm,gp-survey-ieee,user-authentication-categories,signature-gp,DAS,BDAS,YAGP,web-gp-doodle,pass-go,pass-points,java-gp} and industry \cite{grid-sure,android-pattern-screen-lock,blackberry-pattern-lock,passfaces,commercial-pass-points,windows8-picture-password} to study, design and develop graphical user identification systems (GUISs). More than 10 US patents on GUISs have been issued, the first in 1996 to Greg E. Blonder \cite{gp-blonder} and the last in 2014 to Microsoft Corporation \cite{gp-microsoft-2}, but yet GUISs are not widely used.
There are three basic categories of GUISs \cite{user-authentication-categories}: Cognometrics, Locimetrics, and Drawmetrics. \emph{Cognometric} systems are based on the human cognitive abilities, such as the ability to remember and recall images. \emph{Locimetric} systems are based on locating or identifying a point in an image. \emph{Drawmetric} systems are based on reproducing an already pre-drawn image (outline drawing).
Elizabeth et al. \cite{memory-gp} presents a study about the relationship between memory and graphical passwords. The results indicate that the recognition-based (\emph{cognometric}) graphical passwords are more memorable than recall-based (\emph{drawmetric}) graphical passwords, but takes more time to login. None of the previously proposed GUISs \cite{user-authentication-categories,signature-gp,DAS,BDAS,YAGP,web-gp-doodle,pass-go,pass-points,java-gp,grid-sure,android-pattern-screen-lock,blackberry-pattern-lock,passfaces,commercial-pass-points,windows8-picture-password} combines the advantages of both \emph{cognometric} and \emph{drawmetric} systems. The systems proposed in \cite{signature-gp} and \cite{java-gp} use signature-based schemes but do not use a 2D grid technology (draw, digitize and store the signature), and hence are difficult to recall. The system proposed in \cite{DAS} uses a 2D grid technology but is not a signature-based scheme. Therefore we do not categorize these three \cite{signature-gp,DAS,java-gp} as both \emph{cognometric} and \emph{drawmetric} systems.
Handwritten signatures have long been used as a proof of authorship. In general, signatures are authentic, unforegeable, not reusable, unalterable and unrepudiatable \cite{signature-gp}. Signatures personify a person and are easier to recall than other drawings when drawn on a 2D grid. Therefore we can categorize such signatures as recall-based graphical passwords that have the same or close enough memorability as recognition-based graphical passwords.
In this paper we propose a new graphical \underline{S}ignature-based \underline{U}ser \underline{I}dentification \underline{S}ystem named \textbf{SUIS}. It is based on a 2D grid technology, that is used to draw, digitize and store the signature for user identification. SUIS is categorized as both a \emph{cognometric} and \emph{drawmetric} system. The 2D grid was first used in \emph{Draw a Secret} \cite{DAS} for user identification. Our model is different than proposed in \cite{DAS}:
\begin{itemize}
\item
We take one cell in a grid, as one pixel in the drawing. This makes it much simpler to implement the model and compare the signature in practice.
\item
For signature matching, the signature drawn has to follow the same grid cells as the signature stored, independent of the sequence. This increases the usability of the system, but decreases the password space. Increasing the usability here means, since the users do not have to follow the same sequence they can draw and remember more complex signatures. The password space can be increased by increasing the number of grid cells.
\end{itemize}
Some of the characteristics of SUIS are as follows:
\begin{enumerate}
\item
SUIS is easier and faster to compare for signature matching.
\item
SUIS takes into consideration the angular changes at a coarse level.
\item
SUIS, by using a digitization technique, provides extra security and protection on top of encryption.
\item
The signature window i.e; the grid area (number of grid cells) in SUIS can be increased to increase the password space ($2^n$ where n = number of grid cells) of the system. The size of an average grid (extended) used in our empirical study is $7 \times 7$ for which the password space is $2^{49} > 10 \times 10^{14}$. The password space for a text-based user identification system for $8$ characters (average size) password is $95^{10} > 6 \times 10^{15}$. There are $95$ possible character sets including space.
\item
SUIS is rigorous enough to be a password system, but easy enough to be usable.
\item
Precision of the signature in SUIS can be changed by decreasing the size of the grid cell of the system. For example, a signature drawn with mouse/finger needs less precision, whereas a signature drawn with pen or other such pointing devices needs more precision.
\item
SUIS is independent of the language and the device used to write/draw the signature, but is more suitable for touch-based systems (the empirical study presented in this paper was performed on a touch-based system), such as, mobile devices and laptops.
\item
SUIS is efficient and suitable to be used for online (verification is performed immediately after a password is submitted) authentication systems. Our system is not based on any machine learning model as used in \cite{java-gp}. A machine learning model is not suitable and practical to be used for identifying a user for online authentication systems. Moreover, for successful signature matching such a model needs a large set of training data, i.e; the forged samples of the signatures.
\end{enumerate}
The remainder of the paper is organized as follows. Section \ref{sec:literature-review} describes and compares research efforts that are similar to SUIS. Section \ref{sec:suis} describes SUIS in detail. We conclude in Section \ref{sec:conclusion}.
\section{Related Works}\label{sec:literature-review}
This section discusses some of the previous research efforts on GUISs that are similar to the system proposed in this paper. A thorough survey of these and other systems referenced in Section \ref{sec:introduction} can be found in \cite{gp-survey-acm,gp-survey-ieee}.
Syukri et al. \cite{signature-gp} proposed a system that uses signatures drawn with mouse for identifying a user. Through experimental evaluation they selected parameters for signature verification and matching. User verification threshold was set to 70\%. The size of the signature window was set to $1024 \times 512$ pixels. The distance threshold was set to 50 pixels. The parameters selected for signature verification were:
(1) Number of signature points.
(2) Coordinate of signature points.
(3) Signature writing time.
(4) Signature writing velocity.
(5) Signature writing acceleration.
Experiments were carried out with 21 users. The successful rate achieved was 93\%. To calculate the acceleration they used the classical physics formula $a = \frac{F}{m}$, where $F$ is the user's force to push the mouse and $m$ is the mass of the mouse. Although, the number of participants in the experiments carried out in \cite{signature-gp} were small to make any definite conclusion, the successful rate reported is very encouraging, and indicates that some of the parameters can be used for a successful signature matching.
SUIS indirectly uses the signature points as part of the grid cells. We do not use the last three parameters selected by \cite{signature-gp}, as we believe these are not the true representation of a user drawing a signature. In practice the last three parameters are more dependent than the first two parameters, on different environments, such as the mood of the user (e.g; in sickness, sadness and excitement etc), times of the day or night, etc, and hence can produce more true negatives.
\begin{figure*}[htbp]
\centering
{\includegraphics[scale=0.65]{suis-system-overview.pdf}}
\caption{High level overview of SUIS (Signature-Based User Identification System).}
\label{fig:suis-system-overview}
\end{figure*}
Jermyn et al. \cite{DAS} proposed a system called \emph{Draw a Secret} (DAS), that allows the user to draw a unique password. The password is drawn on a 2D grid. If the stored and the drawn password touches the same grid cells in the same sequence, then the user is authenticated. It becomes difficult to authenticate if the strokes of the user are too close to the grid lines. In this case either, the user is presented with the internal representation to confirm if the cells were actually touched by the drawing, or, the system does not accept a drawing that is too close to a grid line.
As passwords depend on users, the proposed system in \cite{DAS} lacks an empirical study for its usability. It has only been tested through paper prototypes. Because of lack of a suitable user study we cannot comment on its effectiveness as a GUIS.
Everitt et al. \cite{java-gp} proposed a neural network-based system using graphical signatures. The system use a hybrid approach, using both text and graphical passwords. The input devices used are mouse for graphical-based password and keyboard for text-based password. The authenticity of a user is confirmed by the typing style and the signature match. For the typing style they measure two times, one is the time between the two key presses, and the other is the time a key is held down. This idea for the key metrics is similar in concept to the one used in \cite{signature-gp} for the mouse metrics. For matching two graphical signatures they use the signature traces, and measure the change in angles and euclidean distances in the two signatures.
The experiments were carried out with 41 participants between the ages of 20 and 30. The results show that they achieved a false accept rate (FAR) of 4.4\% and a range of false reject rate (FRR) from 0.2\% -- 38.6\%. FAR is the rate at which forged samples are accepted as genuine and FRR is the rate at which genuine samples are rejected as forgeries.
The system \cite{java-gp} is based on a machine learning model and hence needs training in addition to registration and verification. The training data is usually created by asking users to provide a set of forged signature samples for other users or these forge signature samples are generated automatically. We think this is one of the major problems of using a machine learning model in such systems.
\section{SUIS (Signature-Based User Identification System)}\label{sec:suis}
Figure \ref{fig:suis-system-overview} gives an overview of SUIS proposed in this paper. SUIS has two phases. In the first phase the user registers her/his signature. This signature, after digitizing\footnote{Digitizing also includes encryption. We do not discuss it in this paper because there is already a lot of literature available on encryption.}, is stored in the database. In the second phase the signature drawn by the user, after digitizing, is verified against the stored signatures. In case of a match the user is identified as a legal user.
\subsection{The Digitization Technique}\label{sec:digitization}
We use a simple but practical digitization technique to store the signature. This makes the signature easy to store and compare, but difficult enough to make its extraction non-trivial.
In SUIS, size (number of grid cells) of the 2D grid for drawing the signature range from 25 (5 x 5) -- 49 (7 x 7), to keep the password space in the range of more than ten million ($2^{25} > 1 \times 10^{7}$) to more than a trillion ($2^{49} > 1 \times 10^{14}$). Each cell in the 2D grid is stored as part of the signature. There are two grids: the \emph{drawing grid}, that is visible to the user for drawing, and the \emph{extended grid} that includes the \emph{drawing grid} and extra cells, and is used to digitize and store the signature. If a cell in a \emph{drawing grid} contains a drawing, i.e, the pixels touched in the cell are inside the drawing area of the cell, it is stored as 1, otherwise it is stored as 0. To produce a coarser signature, we keep the drawing area in a cell smaller than the area of the cell. The extra cells in the \emph{extended grid} is used to store the value of the color selected by the user and the value of the randomly selected storing technique used. The ability of selecting a color and choosing a random value for the storing technique for the signature also increase the password space.
The user can select a color to draw the signature. To elude shoulder surfing upto an extent, a different value of the color relative to the color selected (in which the signature drawn is displayed) is selected and stored. Shoulder surfing is a technique where a person looks over someone's shoulder to get information, such as passwords, PINs, other security codes and data. Each color in SUIS is assigned a number. The value of color stored is computed as follows: $color \ stored = number \ assigned \ to \ the \ selected \ color + \ceil*{\frac{N}{t}}$, where $N$ is the number of total colors used in the system, $t$ is the randomly selected storing technique used and $N$ is $\geq$ $t$. To provide more resistance for shoulder surfing, we erase the signature as soon as it's drawn completely and the user submits it for verification.
We use a number of different techniques for storing a signature. To make the extraction of the information about a signature (grid cells drawn, value of color and storing technique) non-trivial, each time a signature is stored a different storing technique is selected randomly. This information is stored as part of the signature, and also as part of the user's profile, so that during the verification phase of the user the same storing technique is used to verify the signature. Each time, when a user login, a different technique can be used and stored in the user's profile to increase the protection of the user's signature. We also encrypt the signature before storing it. Therefore, this digitization of the signature provides extra security and protection on top of encryption.
\subsection{Signature Storing Techniques}\label{sec:storing-techniques}
A signature is stored as a 2D matrix. Each cell in the matrix contains either a value 1 or a value 0. Initially all the cells in the matrix contains a value 0. The signature drawn by the user is stored in the cells of the matrix corresponding to the \emph{drawing grid}, as explained above. The \emph{extended grid} contains twice (to take care of $\ceil*{\frac{N}{t}}$) as many extra cells as the total number of colors that can be used to draw the signature. Based on the value of the color selected, the corresponding cell out of these cells of the matrix gets the value of 1. Similarly the \emph{extended grid} contains another set of additional cells equal to the number of the storing techniques available. Based on the storing technique used, the corresponding cell out of these additional cells of the matrix gets the value of 1.
Each signature's storing technique is given a number (a value) that is stored in the signature, as explained above. We introduce \emph{simple -- complex} changes in each signature technique to make it different than the other. Some of the changes introduced are:
\begin{enumerate}
\item
Changing the numbering of the signature's matrix (from left-right to right-left and so on).
\item
Changing the location (start, middle or the end) where the value of the color is stored in the signature's matrix.
\item
Changing the location (start, middle or the end) where the value of the storing technique used is stored in the signature's matrix.
\item
Instead of storing a 1 with 0's, we could just store the value of the color and the value of the storing technique in the matrix.
\item
By splitting or merging the grid cells and storing them at different locations in the signature's matrix.
\item
By just storing the information about either 0's, 1's or both in the signature's matrix. For example storing only the location of all the 1's in the matrix.
\item
Combination of two or more of the above techniques.
\end{enumerate}
Other complex storing techniques can also be used to increase protection, such as matrix manipulations, etc, and we leave this to the reader. Using different number of storing techniques gives SUIS the ability to randomize the value of the storing technique, each time a signature is stored, that makes it non-trivial to extract the signature information.
\subsection{Example}\label{sec:example}
We explain the digitization technique described in Section \ref{sec:digitization} and how it is used for storing (using one of the randomly selected storing techniques) and matching a signature, using an example shown in Figure \ref{fig:suis-digitization}.
\begin{figure}[htbp]
\centering
{\includegraphics[scale=0.45]{suis-digitization.pdf}}
\caption{An example of digitization technique used in SUIS. The signature \emph{Babajee} (as pronounced in English language), a word used in Urdu and Hindi languages to \emph{refer an old man with respect}, is digitized.}
\label{fig:suis-digitization}
\end{figure}
A 2D grid of $8 \times 5$ is used in Figure \ref{fig:suis-digitization} for drawing, called the \emph{drawing grid}. The same 2D grid is extended to $10 \times 6$ to store the signature, called the \emph{extended grid}. Whenever the signature touches a grid cell, a value of 1 is stored in that grid cell. We store a value of 0 in the grid cell at column 3 and row 5 (shown in \emph{lightgray} color), because the number of pixels touched in the grid cell are less than a predefined threshold value for this grid.
The user, used color \emph{green} to draw the signature and the storing technique (randomly selected) used for storing the signature is numbered 1. So we store number 7 (assuming color \emph{green} = 1 and color \emph{white} = 7, 1 + $\ceil*{\frac{16}{3}} = 7$) and 1 in the \emph{extended grid}, at the end of the matrix (the last 2 columns and the last row). For storing the color number 7, we store a 1 at the corresponding cell, at column 9 and row 4, in the matrix. For this example we have assumed there are 4 signature storing techniques available in SUIS, so the \emph{extended grid} has 16 + 4 = 20 extra cells. For storing the signature's storing technique number 1, we store a 1 at the corresponding cell, at column 7 and row 6, in the matrix, and the same number is also stored as part of the user's profile.
For matching a signature, we exactly match all the extra cells added to the \emph{extended grid} of all the stored signatures with all the extra cells added to the \emph{extended grid} of the signature drawn for the verification. In case of a successful match we match the corresponding \emph{drawing grid} of the stored signature with the \emph{drawing grid} of the signature drawn for the verification, based on a predefined threshold value for the grid. If the difference is $\geq$ to the predefined threshold value the match is successful, and the user is verified as a legal user.
\section{Conclusion and Future Work}\label{sec:conclusion}
In this paper, we have proposed a new graphical signature-based user identification system, that is efficient and practical to be used in login systems. The system is rigorous enough to be a password, but easy enough to be usable. It is independent of the language used to write/draw the signature.
Currently we are developing a prototype tool in Java to implement SUIS. In future we will carry out a study with a large number of participants and evaluate the validity of SUIS using a number of metrics, such as \emph{Usability}, \emph{Deployability} and \emph{Security} described in \cite{Evaluation-Web-Authentication}. To improve the usability of the current technique, in future, we will develop different techniques, such as by integrating joining blocks/lines that will make it easier to form a shape or a drawing as a graphical signature, etc, and test the pros and cons of each such technique through a large scale empirical study.
|
\section{Introduction}
Resonance phenomena are prevalent in physics; see \textit{e.g.}\ Ref.~\cite{Prigogine96}.
Quantum resonance, in particular, has been a central issue of quantum mechanics from the early days of its development~\cite{Gamow28,Eckart30,Bethe36,Siegert39,Hulthen42a,Hulthen42b,Jost51,Vogt54,Corinaldesi56,Peierls59,Nussenzveig59,Landau77,Brandas90,Kukulin89}.
It is ever more important these days because we can closely observe and even manipulate quantum-mechanical systems such as mesoscopic systems, molecules and nuclides~\cite{Kobayashi02,Kobayashi03,Kobayashi04,Sato05,Brisker08,Morita04}:
for example, the resonant conduction in experiments of mesoscopic systems yields a variety of functions of nano-devices;
producing unstable nuclides experimentally is nothing but looking for resonant states.
It may be said, however, that such experimental studies on quantum resonance far surpass theoretical ones.
Further development of theoretical and numerical methods of analyzing quantum resonance is awaited.
In the present note, we review two theoretical methods of finding resonant states in open quantum systems, namely the approach of the Siegert boundary condition~\cite{Siegert39,Landau77,Hatano08,Nakamura07,Hatano09,Garmon09,Klaiman11} and the Feshbach formalism~\cite{Feshbach58,Feshbach62,Sadreev03,Rotter09}.
One of the purposes of the note is to show the algebraic equivalence of the two approaches.
Another point of the note is to stress that the full Hamiltonian of an open quantum system is Hermitian only in the Hilbert space, which is spanned by the bound states and the scattering continuum states.
It is generally non-Hermitian outside the Hilbert space, where the resonant states reside~\cite{Hatano08,Aguilar71,Balslev71,Simon72}.
That is why the resonant states can have complex eigenvalues of the seemingly Hermitian Hamiltonian.
The two numerical methods named above produce an explicitly non-Hermitian effective Hamiltonian in a contracted finite-dimensional state space out of the implicitly non-Hermitian full Hamiltonian in the infinite-dimensional state space.
The note is organized as follows.
First in the next section~\ref{sec2}, we introduce the simple model that we are going to use in the present note.
We review the approach of the Siegert boundary condition in Sec.~\ref{sec5}.
We then go on to the Feshbach formalism in Sec.~\ref{sec3}.
We finally present in Sec.~\ref{sec7} the statement that the full Hamiltonian is Hermitian in the Hilbert space, whereas non-Hermitian outside the Hilbert space.
\section{Model}
\label{sec2}
Let us consider hereafter the simple model in Fig.~\ref{fig1}(a) for concreteness.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{fdp-hatanofig1.eps}
\caption{(a) The tight-binding model of a T-type dot.
The sites on the infinite lead are labeled by integers $0,\pm1,\pm2,\ldots$, whereas the site on the dot is labeled by ``$\mathrm{d}$."
(b) A general open quantum system to which we can generalize the present argument.}
\label{fig1}
\end{figure}
It is sometimes called a T-type dot or a side-coupled dot~\cite{Porod92,Porod93,Shao94,Kim01,Kikoin01,Kang01,Affleck01,Simon01,Affleck07}.
We can obviously extend the following arguments to more general models such as the one in Fig.~\ref{fig1}(b)~\cite{Datta95,Sasada05,Sasada11}.
We will leave out any inter-particle interactions;
see \textit{e.g.}\ Refs.~\cite{Nishino07,Imamura09,Nishino11} for quantum-dot systems Coulombic interactions.
The present problem is strictly a one-body problem.
The whole state space is hence spanned by the state $|\mathrm{d}\rangle$ in which the particle resides on the dot site $\mathrm{d}$ and the state $|x\rangle$ in which the particle resides on the site $x$ of the lead, where $x$ runs from $-\infty$ to $+\infty$.
The resolution of unity is given by
\begin{align}
|\mathrm{d}\rangle\langle\mathrm{d}|+\sum_{x=-\infty}^\infty|x\rangle\langle x|=1.
\end{align}
The Hamiltonian of the T-type dot is expressed in the form
\begin{align}\label{eq10}
H&=-t\sum_{x=-\infty}^\infty
\left(|x+1\rangle\langle x|+|x\rangle\langle x+1|\right)
-t_1\left(|\mathrm{d}\rangle\langle 0|+|0\rangle\langle\mathrm{d}|\right)
+\varepsilon_\mathrm{d}|\mathrm{d}\rangle\langle\mathrm{d}|,
\end{align}
where $t$ is the transfer amplitude between neighboring sites on the lead, $t_1$ is the one between the site $x=0$ and the site $\mathrm{d}$, and $\varepsilon_\mathrm{d}$ is the potential at the dot site $\mathrm{d}$.
The issue here is to find resonance poles of the S matrix of this open quantum system.
In the early days of quantum mechanics, it was noticed that we can find the resonance pole as an eigenstate of the time-independent Schr\"{o}dinger equation under the Siegert boundary condition~\cite{Siegert39}.
Later on, Feshbach~\cite{Feshbach58,Feshbach62} as well as other researchers~\cite{Petrosky91} established the Feshbach formalism, which algebraically produces resonance poles.
In the present paper we will review the two methods respectively in Secs.~\ref{sec5} and~\ref{sec3}, in the course of which we will show their equivalence.
\section{Siegert boundary condition}
\label{sec5}
We first review the approach of the Siegert boundary condition, the condition that the resonant state has only out-going waves away from the scatterer.
Let us first explain how this condition comes out.
In solving the scattering problem in one dimension, we typically assume the incoming wave, the reflection wave and the transmission wave in the form
\begin{align}\label{eq30-1}
\Phi(x)=
\begin{cases}
Ae^{ikx}+Be^{-ikx} & \quad\mbox{on the left of the scatterer,}\\
Ce^{ikx} &\quad\mbox{on the right of the scatterer.}
\end{cases}
\end{align}
The reflection and transmission amplitudes, which are elements of the S matrix, are given by $B/A$ and $C/A$, respectively.
A textbook definition of the resonant state as a pole of the S matrix then implies that we should look for the zeros of the amplitude $A$ of the incoming wave.
We can do so by making the amplitude $A$ zero from the very beginning:
\begin{align}\label{eq40-1}
\Phi(x)=
\begin{cases}
Be^{-ikx} & \quad\mbox{on the left of the scatterer,}\\
Ce^{ikx} &\quad\mbox{on the right of the scatterer.}
\end{cases}
\end{align}
This is the Siegert boundary condition that the resonant state has only out-going waves away from the scatterer~\cite{Siegert39,Landau77,Hatano08,Klaiman11}.
In fact, the Siegert boundary condition not only supports the resonant states but all other possible discrete states including the bound states and the anti-resonant states, that is, all poles of the S matrix.
A bound state follows from a pure imaginary value of $k$.
A resonant state comes out for $\mathop{\mathrm{Re}}k>0$, whereas
an anti-resonant state comes out for $\mathop{\mathrm{Re}}k<0$.
Let us then solve the Schr\"{o}dinger equation
\begin{align}\label{eq50-1}
H|\Phi\rangle=E|\Phi\rangle
\end{align}
under the Siegert boundary condition~\eqref{eq40-1}, specifically for the T-type dot~\eqref{eq10}, under the condition
\begin{align}\label{eq220}
\langle x|\Phi\rangle=
\begin{cases}
Be^{-ikx}&\quad\mbox{for }x\leq-1,
\\
Ce^{ikx}&\quad\mbox{for }x\geq1.
\end{cases}
\end{align}
First, the Schr\"{o}dinger equation~\eqref{eq50-1} for the sites $x\leq -2$ and $x\geq 2$ is written in the form
\begin{align}\label{eq150}
-t\left(\langle x-1|\Phi\rangle+\langle x+1|\Phi\rangle\right)
=E\langle x|\Phi\rangle,
\end{align}
which along with the condition~\eqref{eq220} yields the dispersion relation
\begin{align}\label{eq130}
E=-t\left(e^{ik}+e^{-ik}\right)=-2t\cos k
\end{align}
of the lead.
Next, the Schr\"{o}dinger equation~\eqref{eq50-1} for the sites $x=\pm 1$ reads
\begin{align}\label{eq90-1}
-t\left(\langle-2|\Phi\rangle+\langle0|\Phi\rangle\right)&=E\langle-1|\Phi\rangle,
\\\label{eq100-1}
-t\left(\langle0|\Phi\rangle+\langle2|\Phi\rangle\right)&=E\langle1|\Phi\rangle.
\end{align}
By using the condition~\eqref{eq220} with the dispersion relation~\eqref{eq130}, we simply obtain
the continuity condition
\begin{align}
B=C=\langle0|\Phi\rangle.
\end{align}
Finally for the remaining sites $x=0$ and $\mathrm{d}$, the Schr\"{o}dinger equation~\eqref{eq50-1} reads
\begin{align}\label{eq120-1}
-t\left(\langle-1|\Phi\rangle+\langle1|\Phi\rangle\right)-t_1\langle\mathrm{d}|\Phi\rangle
&=E\langle0|\Phi\rangle,
\\\label{eq130-1}
-t_1\langle0|\Phi\rangle+\varepsilon_\mathrm{d}\langle\mathrm{d}|\Phi\rangle
&=E\langle\mathrm{d}|\Phi\rangle.
\end{align}
The point here is the fact that the Siegert condition~\eqref{eq220} yields~\cite{Sasada08}
\begin{align}\label{eq140-2}
\langle\pm1|\Phi\rangle=e^{ik}\langle0|\Phi\rangle.
\end{align}
We thereby arrive at a \textit{closed} set of equations of the form
\begin{align}
-2te^{ik}\langle0|\Phi\rangle-t_1\langle\mathrm{d}|\Phi\rangle
&=E\langle0|\Phi\rangle,
\\
-t_1\langle0|\Phi\rangle+\varepsilon_\mathrm{d}\langle\mathrm{d}|\Phi\rangle
&=E\langle\mathrm{d}|\Phi\rangle.
\end{align}
These two equations are cast into the matrix form
\begin{align}\label{eq320}
\begin{pmatrix}
-2te^{ik}&-t_1\\
-t_1&\varepsilon_\mathrm{d}
\end{pmatrix}
\begin{pmatrix}
\langle0|\Phi\rangle\\
\langle\mathrm{d}|\Phi\rangle
\end{pmatrix}
=E
\begin{pmatrix}
\langle0|\Phi\rangle\\
\langle\mathrm{d}|\Phi\rangle
\end{pmatrix}.
\end{align}
We will indeed show in Sec.~\ref{sec3} that the two-by-two matrix on the left-hand side is the non-Hermitian effective Hamiltonian given below by the Feshbach formalism:
\begin{align}\label{eq180-2}
{H_\mathrm{eff}}=\begin{pmatrix}
-2te^{ik}&-t_1\\
-t_1&\varepsilon_\mathrm{d}
\end{pmatrix}.
\end{align}
Equation~\eqref{eq320} is then cast into the form of the eigenproblem for the effective Hamiltonian:
\begin{align}\label{eq190-1}
{H_\mathrm{eff}}(P|\Phi\rangle)=E(P|\Phi\rangle),
\end{align}
where $P$ is a projection operator defined below, but for the present purpose it is enough to say
\begin{align}
P|\Phi\rangle=
\begin{pmatrix}
\langle0|\Phi\rangle\\
\langle\mathrm{d}|\Phi\rangle
\end{pmatrix}.
\end{align}
Once we solve the eigenproblem~\eqref{eq190-1}, the wave functions away from the scatterer is given by the Siegert condition~\eqref{eq220} as
\begin{align}\label{eq330}
\langle x|\Phi\rangle=e^{ik|x|}\langle 0|\Phi\rangle.
\end{align}
The fact that the effective Hamiltonian of the Feshbach formalism emerges when we solve the scattering problem of the full Hamiltonian is not restricted to the Siegert boundary condition, but is prevalent in other boundary conditions.
See Appendix~\ref{appA} below.
To summarize the approach of the Siegert boundary condition, we begin with solving the original Schr\"{o}dinger equation~\eqref{eq50-1} for the \textit{full} Hamiltonian
under the Siegert boundary condition~\eqref{eq220}.
We then end up with the Schr\"{o}dinger equation~\eqref{eq190-1} for the effective Hamiltonian, augmented by Eq.~\eqref{eq330}.
These are exactly the same as the equations that will be derived from the Feshbach formalism in Sec.~\ref{sec3}.
We thereby obtain discrete eigenvalues including resonant ones by finding the roots of
\begin{align}\label{eq260}
\det (E-{H_\mathrm{eff}})=0
\end{align}
in the complex $E$ plane.
Note here that the number of the roots can be more than two although the effective Hamiltonian~\eqref{eq210-1} is a two-by-two matrix, because the Hamiltonian itself depends on the energy $E$ through the dispersion relation~\eqref{eq130}.
Using the variable $z=e^{ik}$ with the mapping
\begin{align}
E=-t(z+z^{-1}),
\end{align}
we can indeed transform the secular equation~\eqref{eq260} into a fourth-order polynomial equation with respect to $z$,
\begin{align}
t^2z^4+t\varepsilon_\mathrm{d} z^3 +{t_1}^2z^2-t\varepsilon_\mathrm{d} z-t^2=0,
\end{align}
in the particular case of Eq.~\eqref{eq320}.
Its four solutions contain two bound states, one resonant state and one anti-resonant state in one parameter region, while two bound states and two anti-bound states in the other regions~\cite{Hatano08}.
We will show in Sec.~\ref{sec7} why the effective Hamiltonian~\eqref{eq180-2} can be non-Hermitian and why complex eigenvalues can emerge in spite of the fact that they come out of the seemingly Hermitian full Hamiltonian~\eqref{eq10}.
We will argue that the full Hamiltonian of an open quantum system is in fact non-Hermitian outside the Hilbert space and that the effective Hamiltonian~\eqref{eq180-2} manifests the non-Hermiticity.
\section{The Feshbach formalism of the effective Hamiltonian}
\label{sec3}
Let us now review the Feshbach formalism for the T-type dot~\eqref{eq10}~\cite{Sadreev03,Rotter09}, which will produce the equations exactly the same as Eqs.~\eqref{eq190-1} and~\eqref{eq330}.
We introduce the projection operators $P$ and $Q$ as follows:
\begin{align}
P&=|\mathrm{d}\rangle\langle\mathrm{d}|+|0\rangle\langle0|,
\\
Q&=1-P
=\sum_{x=1}^\infty |x\rangle\langle x|+\sum_{x=-1}^{-\infty} |x\rangle\langle x|.
\end{align}
This divides the T-type dot in Fig.~\ref{fig1}(a) into the subspaces indicated in Fig.~\ref{fig2}.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{fdp-hatanofig2.eps}
\caption{The subspaces projected by the operators $P$ and $Q$.}
\label{fig2}
\end{figure}
We would like to solve the Schr\"{o}dinger equation~\eqref{eq50-1} for the full Hamiltonian.
The Feshbach formalism tells us that instead of solving Eq.~\eqref{eq50-1} in the state space in the infinite dimensions, we can solve
\begin{align}\label{eq100}
\HeffP|\Phi\rangle=EP|\Phi\rangle
\end{align}
in the $P$ subspace in two dimensions.
We are going to show in generic cases that the Feshbach formalism gives the effective Hamiltonian ${H_\mathrm{eff}}$ in Eq.~\eqref{eq100} in the form
\begin{align}\label{eq110}
{H_\mathrm{eff}}=P HP+P HQ\frac{1}{E-Q HQ}Q HP.
\end{align}
Note that we can express the effective Hamiltonian as a matrix in the $P$ subspace.
Here is how we can obtain the reduced Schr\"{o}dinger equation~\eqref{eq100} with the effective Hamiltonian~\eqref{eq110}.
We can split the original Schr\"{o}dinger equation~\eqref{eq50-1} into the $P$ and $Q$ subspaces as
\begin{align}
P H(P+Q)|\Phi\rangle&=EP|\Phi\rangle,
\\
Q H(P+Q)|\Phi\rangle&=EQ|\Phi\rangle,
\end{align}
which are cast into the forms
\begin{align}\label{eq140-1}
P HP(P|\Phi\rangle)+P HQ(Q|\Phi\rangle)&=E(P|\Phi\rangle),
\\\label{eq150-1}
Q HP(P|\Phi\rangle)+Q HQ(Q|\Phi\rangle)&=E(Q|\Phi\rangle).
\end{align}
We first solve the second equation~\eqref{eq150-1} with respect to the state $Q|\Phi\rangle$ as
\begin{align}\label{eq160-1}
Q|\Phi\rangle=\frac{1}{E-Q HQ}Q HP(P|\Phi\rangle)
\end{align}
and substitute Eq.~\eqref{eq160-1} into the first equation~\eqref{eq140-1} to obtain
\begin{align}
P HP(P|\Phi\rangle)+P HQ\frac{1}{E-Q HQ}Q HP(P|\Phi\rangle)=E(P|\Phi\rangle).
\end{align}
We can cast this into the form~\eqref{eq100} with the effective Hamiltonian~\eqref{eq110}.
Specifically for the T-type dot~\eqref{eq10}, each term in the effective Hamiltonian~\eqref{eq110} is expressed as follows:
\begin{align}\label{eq40}
P HP&=-t_1\left(|\mathrm{d}\rangle\langle 0|+|0\rangle\langle\mathrm{d}|\right)
+\varepsilon_\mathrm{d}|\mathrm{d}\rangle\langle\mathrm{d}|,
\\
Q HP&=-t|1\rangle\langle0|-t|-1\rangle\langle0|.
\\\label{eq80}
Q HQ&=-t\sum_{x=1}^\infty
\left(|x+1\rangle\langle x|+|x\rangle\langle x+1|\right)
-t\sum_{x=-1}^{-\infty}
\left(|x-1\rangle\langle x|+|x\rangle\langle x-1|\right)
\\
P HQ&=-t|0\rangle\langle1|-t|0\rangle\langle-1|,
\end{align}
These expressions enable us to understand why the finite-dimensional effective Hamiltonian~\eqref{eq110} can represent the dynamics of the particle in the infinite-dimensional space.
The first term $P HP$ of the effective Hamiltonian indeed describes the dynamics only in the $P$ subspace of the sites d and 0.
Its second term, on the other hand, describes the following dynamics:
the particle first hops from the $P$ subspace to the $Q$ subspace owing to the partial Hamiltonian $Q HP$;
the particle moves around in the infinite-dimensional $Q$ subspace under the Green's function of the the partial Hamiltonian $Q HQ$;
the particle then hops from the $Q$ subspace back to the $P$ subspace owing to the partial Hamiltonian $P HQ$.
The fact that the system is open is thus encoded in the second term of the effective Hamiltonian~\eqref{eq110}.
This term indeed turns out to be non-Hermitian.
We can calculate the Green's function $(E-QHQ)^{-1}$ explicitly for the partial Hamiltonian~\eqref{eq80}.
We do not give the details of the calculation~\cite{Sasada11} but present the following result:
\begin{align}\label{eq180-1}
\left.\begin{array}{ll}
\mbox{for } x\geq 1\quad &\displaystyle \langle x|\frac{1}{E-QHQ}|1\rangle
\\
&\\
\mbox{for } x\leq -1\quad &\displaystyle\langle x|\frac{1}{E-QHQ}|-1\rangle
\end{array}\right\}
=-\frac{1}{t}e^{ik|x|}.
\end{align}
Here we express the right-hand side in terms of the wave number $k$ on the lead instead of the energy $E$;
we indeed obtain the dispersion relation $E=-2t\cos k$ during the calculation of Eq.~\eqref{eq180-1} as the eigenvalues of the partial Hamiltonian $QHQ$.
The wave number $k$, or more precisely the crystal wave number, is restricted to the first Brillouin zone $-\pi\leq k\leq\pi$ for the energy band $-2t\leq E\leq 2t$, where the point $k=-\pi$ is equivalent to the point $k=\pi$ because of the $2\pi$ periodicity of the Fourier space.
The variables $k$ and $E$ are related through the dispersion relation of the lead, Eq.~\eqref{eq130}.
The mapping from the energy $E$ to the wave number $k$ is actually a one-to-two correspondence.
We obtain the retarded Green's function out of Eq.~\eqref{eq180-1} by choosing the branch $0<k< \pi$ and the advanced one by choosing the branch $-\pi<k<0$.
Using the Green's function~\eqref{eq180-1}, we have
\begin{align}\label{eq380-1}
PHQ\frac{1}{E-QHQ}QHP&=
(-t)^2|0\rangle\langle1|\frac{1}{E-QHQ}|1\rangle\langle0|
\nonumber\\
&+
(-t)^2|0\rangle\langle-1|\frac{1}{E-QHQ}|-1\rangle\langle0|
\nonumber\\
&=-2te^{ik}|0\rangle\langle0|.
\end{align}
Therefore, we can express the effective Hamiltonian~\eqref{eq110} in the following form of the two-by-two matrix spanned by the states $|0\rangle$ and $|\mathrm{d}\rangle$~\cite{Sadreev03}:
\begin{align}\label{eq210-1}
{H_\mathrm{eff}}=\begin{pmatrix}
-2te^{ik} & -t_1 \\
-t_1 & \varepsilon_\mathrm{d}
\end{pmatrix}.
\end{align}
To summarize the effective-Hamiltonian approach, we solve the Schr\"{o}dinger equation of the non-Hermitian effective Hamiltonian, Eq.~\eqref{eq100}, which is equivalent to Eq.~\eqref{eq190-1}.
We can obtain the corresponding eigenfunctions as follows.
The $P$ part of the resonant wave function naturally follows from the eigenproblem~\eqref{eq100}.
We can obtain the part $Q|\Phi\rangle$ from $P|\Phi\rangle$ as
\begin{align}
Q|\Phi\rangle=\frac{1}{E-QHQ}QHP|\Phi\rangle,
\end{align}
or more specifically,
\begin{align}\label{eq210}
\langle x|\Phi\rangle&=
\begin{cases}
\displaystyle
\langle x|\frac{1}{E-QHQ}|1\rangle\langle0|\Phi\rangle & \quad\mbox{for } x\geq 1,
\\
&\\
\displaystyle
\langle x|\frac{1}{E-QHQ}|-1\rangle\langle0|\Phi\rangle & \quad\mbox{for } x\leq -1
\end{cases}
\nonumber\\
\nonumber\\
&=e^{ik|x|}\langle 0|\Phi\rangle,
\end{align}
which is the same as Eq.~\eqref{eq330}.
Therefore, the approach of the Siegert boundary condition and the Feshbach formalism are algebraically equivalent and produce the same eigenstates.
\section{Origin of the non-Hermiticity of the effective Hamiltonian}
\label{sec7}
Let us argue by examining the functional space of the eigenstates, that the full Hamiltonian $H$ is \textit{non-Hermitian outside the Hilbert space}.
The eigenstates obtained either with the Siegert boundary condition or in the Feshbach formalism are distributed in the complex $k$ plane as shown in Fig.~\ref{fig3}.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{fdp-hatanofig3.eps}
\caption{The distribution of eigenstates of the Hamiltonian~\eqref{eq10} in the complex $k$ plane.}
\label{fig3}
\end{figure}
Specifically for the T-type dot~\eqref{eq10} with $\varepsilon_\mathrm{d}=0$ for example, the solutions are
\begin{align}
\label{eqC110}
&
k^\mathrm{bl}=0+i\log p,
\qquad
E^\mathrm{bl}=-(p+q)t,
\\\label{eqC120}
&
k^\mathrm{bh}=\pi+i\log p,
\qquad
E^\mathrm{bh}=(p+q)t,
\\\label{eqC180}
&
k^\mathrm{res}=+\frac{\pi}{2}-i\log p,
\qquad
E^\mathrm{res}=-i(p-q)t,
\\\label{eqC190}
&
k^\mathrm{ar}=-\frac{\pi}{2}-i\log p,
\qquad
E^\mathrm{ar}=+i(p-q)t,
\end{align}
where
\begin{align}
p&=\sqrt{\frac{{t_1}^2+\sqrt{4t^4+{t_1}^4}}{2t^2}},
\\
q&=\sqrt{\frac{-{t_1}^2+\sqrt{4t^4+{t_1}^4}}{2t^2}}=\frac{1}{p}
\end{align}
are dimensionless real numbers.
The superscripts `bl,' `bh,' `res' and `ar' in Eqs.~\eqref{eqC110}--\eqref{eqC190} respectively denote the bound state with $\mathop{\mathrm{Re}}k=0$, the bound state with $\mathop{\mathrm{Re}}k=\pi$, the resonant state and the anti-resonant state.
The distribution for the tight-binding model~\eqref{eq10} is different from the standard distribution in two points~\cite{Sasada11}:
first, the whole $k$ plane is limited to $-\pi<\mathop{\mathrm{Re}}k\leq\pi$ because of the lattice periodicity;
second, as a consequence of the first point, bound and anti-bound states can exist, respectively, on the positive and negative imaginary parts of the line $\mathop{\mathrm{Re}}k=\pi$, which correspond to the energy eigenvalues above the upper threshold $E=+2t$~\cite{Garmon09}.
Other properties are the same as the standard distribution.
The upper and lower half $k$ planes, respectively, correspond to the first and second Riemann sheets of the complex $E$ plane.
First, the bound states $\{|\Phi_n^\textrm{b}|n=1,2,3,\ldots\rangle\}$ are distributed in the upper half $k$ plane.
Their eigenfunctions are localized around the scatterer.
When we substitute $k=i\kappa$ with $\kappa>0$, the wave function~\eqref{eq330} is of the form
\begin{align}
\langle x|\Phi\rangle=e^{-\kappa|x|}\langle 0|\Phi\rangle,
\end{align}
and is $L^2$-normalizable.
Next, the continuum scattering states $\{|\Phi_k\rangle|-\pi<k\leq\pi\}$ are on the real $k$ axis.
Their eigenfunctions are plane waves of the form
\begin{align}
\langle x|\Phi\rangle=Ae^{ikx}+B^{-ikx}
\end{align}
with real $k$, and hence box-normalizable.
Indeed, it was proved~\cite{Newton61} that the above two types of the normalizable eigenfunctions form a complete set
\begin{align}
\sum_{n}
|\Phi_n^\textrm{b}\rangle\langle\Phi_n^\textrm{b}|
+\int_{-\pi}^\pi\frac{dk}{2\pi}|\Phi_k\rangle\langle\Phi_k|=1.
\end{align}
We therefore refer to the functional space spanned by these normalizable eigenstates as the Hilbert space.
On the other hand, the resonant and anti-resonant states as well as the anti-bound states are all distributed in the lower half $k$ plane, or the second Riemann sheet.
Their wave functions are not normalizable in the usual sense, although there is a method of regularization called complex scaling~\cite{Aguilar71,Balslev71,Simon72,Homma97,Masui99,Moiseyev08}.
When we substitute $k=k_\textrm{r}+i\kappa$ with $\kappa<0$, the wave function~\eqref{eq330} is of the form
\begin{align}
\langle x|\Phi\rangle\propto e^{-\kappa|x|}=e^{+|\kappa||x|},
\end{align}
which is spatially divergent~\cite{leCouteur60,Zeldovich60,Hokkyo65,Romo68,Berggren68,Berggren70,Gyarmati71,Romo80,Berggren82,Berggren96,Madrid05,Hatano08,Hatano09}.
The point here is that the wave functions of the resonant, anti-resonant and anti-bound states are not normalizable and exist outside the Hilbert space.
(We argued~\cite{Hatano08,Hatano09} nonetheless that this spatial divergence combined with the temporal decay (due to the imaginary part of the eigenenergy) is followed by the probabilistic interpretation of the resonant wave function.)
To summarize, the full Hamiltonian~\eqref{eq10} is Hermitian in the Hilbert space, whereas non-Hermitian outside the Hilbert space.
Indeed, all eigenstates in the Hilbert space (the bound states and the scattering continuum) have real eigenvalues, whereas the eigenvalues of the states outside the Hilbert space (the resonant and anti-resonant states) are generally complex (with the exception of the anti-bound states).
From this viewpoint, the effective Hamiltonian explicitly reveals the implicit non-Hermiticity of the full Hamiltonian.
In the approach of the Siegert boundary condition, the non-Hermiticity of the effective Hamiltonian arises in Eq.~\eqref{eq140-2}, which comes from the Siegert boundary condition~\eqref{eq40-1}.
In the Feshbach formalism, the non-Hermiticity is due to the term~\eqref{eq380-1}.
In both cases, we can say that the explicit non-Hermiticity of the effective Hamiltonian originates from the openness of the system, which is also the origin of the implicit non-Hermiticity of the full Hamiltonian.
\section{Summary}
In the present note, we showed the equivalence of the approach of the Siegert boundary condition and the Feshbach formalism.
We can also show the equivalence among the standard scattering boundary condition~\eqref{eq30-1}, the Feshbach formalism, and the Lippmann-Schwinger equation, which we omitted here.
An important message here is that the Hamiltonian of an open quantum system is, although it is Hermitian in the Hilbert space, generally non-Hermitian outside the Hilbert space.
The effective Hamiltonian manifests the non-Hermiticity in a contracted state space.
|
\section{Observations of Reverberation Signatures}\label{sec:obs}
We now consider the observational evidence for X-ray reverberation, focussing first on the evidence for reflection reverberation in Active Galactic Nuclei, before discussing the evidence for disc thermal reverberation in black hole X-ray binaries.
\subsection{Active Galactic Nuclei}
\subsubsection{The discovery of the soft lag}
Reverberation lags in the X-rays were first robustly observed by \citet{fabian09} through a
500~ks {XMM-Newton} observation of Narrow-line Seyfert I galaxy,
1H0707-495. In this work, Fabian et al. found that the soft excess (a
traditionally contentious part of the energy spectrum) was composed of
relativistically broadened reflection features, namely the iron L
emission line, and possibly the oxygen line, as well.
Fig.~\ref{1h07_spec} shows the ratio of the spectrum of 1H0707-495 to
a phenomenological model, consisting of an absorbed power law plus
black body emission from the accretion disc. The broad asymmetric peak at 6.4~keV is the Fe~K
emission line, which had been observed many times before
\citep[e.g.][]{tanaka95} because it is in a relatively `simple' part
of the spectrum with little absorption, and because iron is the most
cosmically-abundant element with a high fluorescent yield. The broad line at
$\sim 0.7$~keV is identified as the iron L line (with possibly some
contribution from oxygen at $\sim 0.5$~keV). The lines in the soft
excess are observationally more difficult to detect, as there are many
emission lines at soft energies, contributions from the tail of the
disc black body emission and possible absorption effects.
The clear determination of disc reflection in the soft band motivated the search for time lags between the soft band
and the continuum-dominated band at 1--4~keV. Using the Fourier
timing techniques described in the previous section, the
frequency-dependent lag was measured between these two bands. A clear `soft
lag' of $\sim 30$~s was discovered at $\sim 10^{-3}$~Hz
(Fig.~\ref{1h07_lag}). In other words, soft band variations on the
order of 1000~s followed the corresponding hard band variations by an
average of 30~s. A follow-up study of the reverberation lags in
1H0707-495, \citep{zoghbi10} illustrated this frequency-dependent
result, by using light curves filtered to highlight variations on different time-scales, thus showing the lags in the time domain (Fig.~\ref{1h07_lc}).
Similar to Fig.~\ref{1h07_lag}, we are looking at the time delays
between the soft band (0.3--1~keV) and the continuum-dominated band
(1--4~keV). The top panel shows the light curve variations on the
time scale of 700~s (corresponding to a frequency of $\sim 1 \times
10^{-3}$~Hz). The soft band lags the hard band at this time scale
(corresponding to the negative lag in Fig.~\ref{1h07_lag}). However,
looking at variations on the order of 5000~s (frequency of $2 \times
10^{-4}$~Hz), the hard band lags the soft. These light curves show
the same effect we see in the frequency domain, but also elucidate why
it is much easier to do this type of analysis in the frequency domain, where distinct lags on different time-scales can be cleanly separated without any pre-selected filtering of the light curves.
Given the spectral results that the soft band is dominated by reflection, and the harder, 1--4~keV band by the continuum emission from the corona, the soft lag is naturally interpreted in terms of the average light-travel time delay between the compact corona and the inner accretion disc. For 1H0707-495, the average lag was measured to be $\sim 30$~s, which, knowing the mass of the black hole, corresponds to a source height distance of $< 5 r_{\mathrm{g}}$ above the accretion disc. Note that this is the average lag, and does not account for the fact that both energy bands contain contributions from both the directly observed continuum and the reflection. The hard and soft bands are taken to be direct proxies for the continuum and reflected emission, respectively, whereas in reality, the contributions of both components to each band will reduce the measured value or `dilute' the measured lag. Neither does this account for the fact that the corona is likely extended, or inclination effects (since we are actually seeing the lag caused by a {\it path-length difference} between the paths from corona-observer and corona-disc-observer). These effects will be discussed further in Sect.~\ref{sec:freqdep}.
\begin{figure}
\begin{center}
\includegraphics[width=6.5cm,angle=-90]{2laor_rat.ps}
\caption{The ratio spectrum of 1H0707-495 to a continuum model \citep{fabian09}. The broad iron K and iron L band are clearly evident in the data. The origin of the soft excess below 1~keV in this source had been debatable, but in this work was found to be dominated by relativistically broadened emission lines.}
\label{1h07_spec}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=6.5cm,angle=-90]{lagplot2.ps}
\caption{The frequency-dependent lags in 1H0707-495 between the continuum dominated hard band at 1--4~keV and the reflection dominated soft band at 0.3--1 keV.}
\label{1h07_lag}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=\textwidth]{show_lag.eps}
\caption{Illustration of the soft lag (top) and hard lag (bottom) in the time domain, by filtering the light curve on short time-scales (to probe the high frequency soft lag) and long time-scales (to probe the low-frequency hard lag). This is completely analogous to the Fourier analysis shown in Fig.~\ref{1h07_lag}. Figure from \citet{zoghbi10}.}
\label{1h07_lc}
\end{center}
\end{figure}
At frequencies below $10^{-3}$~Hz, the lag was observed to switch sign to become a `hard lag'. This hard lag had been observed previously in several NLS1 galaxies (e.g. \citealt{vaughanfn03,mchardy04,arevalo06}) and first in galactic black hole X-ray binaries \citep{miyamoto89,nowak99}. The time-scales of this lag negated the possibility that it was a Comptonisation delay, but rather suggest that the lag is caused by fluctuations in the mass accretion rate in the disc that get propagated inwards on the viscous time-scale, causing the outermost soft X-rays in the corona to respond before hard X-rays at smaller radii. See Sect.~\ref{sec:bhb} for more on the hard lag and its interpretation in black hole X-ray binaries.
After the initial discovery of reverberation lags in 1H0707-495, soft lags were discovered in a number of other NLS1 sources \citep{emma11, zog11_rej1034, demarco11, cackett13, fabian13}. \citet{demarco13} conducted a systematic search through the XMM-Newton archive for variable Seyfert galaxies with sufficiently long observations, and found significant high-frequency soft lags in 15 sources. Plotting the amplitude of the lags with their best-estimated black hole masses\footnote{Black hole masses used by \citet{demarco13,kara13c} and in Fig.~\ref{lag_mbh} were obtained from the literature, and estimated primarily using optical broad line reverberation. In a few cases masses were estimated using the scaling relation between optical continuum luminosity and broad line region radius, which can be used to estimate black hole mass when combined with optical line width (e.g. \citealt{kaspi00,grier12}), or the correlation between black hole mass and host galaxy bulge stellar velocity dispersion (e.g. \citealt{gebhardt00}).}, revealed that the amplitude of the lag scales approximately linearly with mass\footnote{The observed scaling is flatter than expected from a linear relationship, but this can be explained as a bias due to the fact that we only sample the higher frequency end of the soft lag range in the highest mass objects, which leads to systematically shorter lags than would be seen if we could sample the maximum amplitude of soft lags seen at lower frequencies \citep{demarco13}.}, and suggested that the X-ray source was compact and within a few gravitational radii of the accretion disc (top panel of Fig.~\ref{lag_mbh}). The plot has been updated with the additional soft lag measurements that have been found since publication.
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm]{lag_mbh.eps}
\caption{{\em Top: }A sample of 15 Seyfert galaxies with significant soft lag measurements. The amplitude of the lag was found to scale with black hole mass, and suggested that the X-ray emitting region was very close (within $10~r_{\mathrm{g}}$) to the central black hole. Figure adapted from \citet{demarco13}. {\em Bottom: } The current sample of iron K lags (defined as being the lag of 6--7~keV relative to 3--4~keV) for nine AGN (see Table~\ref{lagtable}), plotted with the corresponding black hole mass. The short amplitude lag found for both the iron K and soft lags suggest that they originate from the same small emitting region.}
\label{lag_mbh}
\end{center}
\end{figure}
\subsubsection{The iron K lag}
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm]{compare_ngc4151.eps}
\caption{The iron K lag in NGC~4151. The lag-energy spectrum is read from bottom to top, such that the continuum emission at 2~keV responds first, followed by the red wing of the iron line 1000~s later, and lastly the blue horn of the line 1000~s after that. The higher frequency lags ($5$--$50\times10^{-5}$~Hz) in red show emission from smaller radii, and therefore are dominated by the red wing of the line, while the blue shows the lower frequency lags ($1$--$2\times10^{-5}$~Hz), which show the line from larger radii, closer to the rest-frame energy. Figure from \citet{zoghbi12}.}
\label{ngc4151}
\end{center}
\end{figure}
Up to this point, reverberation lag studies focussed on the reflection-dominated soft band, and the continuum band at 1--4~keV because the signal-to-noise is higher at these low energies. The ultimate test of reverberation, though, is in the iron K band. The broad iron K line peaks at 6.4~keV, but is asymmetrically broadened due to special relativistic beaming and Doppler shifts from a rotating accretion disc and the gravitational redshift close to the black hole. Because the iron line has the main advantage of being in a `clean' part of the spectrum, where it is not affected by the reprocessed black body emission or by other broadened lines, it is our best indicator of the relativistic effects caused by the black hole, and therefore was a natural place to search for reverberation lags. This was first accomplished by \citet{zoghbi12} for the bright Seyfert, NGC~4151 (Fig.~\ref{ngc4151}). In this work, the continuum emission was found to respond first, followed by the red wing of the line (from the very innermost radii) and finally by the rest frame of the line (produced at farther radii). This showed not just the average time lag from all light paths from corona to disc, but actually identified the reverberation from light echoing from different parts of the accretion disc.
In addition to finding the first Fe~K lag, \citet{zoghbi12} discovered the frequency-dependence of the Fe~K lag. At low frequencies ($1$--$2\times10^{-5}$~Hz), the lag-energy spectrum shows the prominent core of the Fe~K line, peaking at close to the rest energy. However, at higher frequencies the line core is suppressed in the lag-energy spectrum and we see systematically shorter lags peaking more in the red wing of the line. A likely explanation for this behaviour is that we are seeing the first evidence of the systematically shorter length-scales which produce the red wing of the line, from closer to the black hole where gravitational and transverse Doppler shifts are strongest. At high Fourier frequencies, the larger-scale rest-frame emission should be washed out by the larger light-crossing times, and effectively removed from the lag-energy spectrum, revealing the red-wing of the line which is preserved in the lag-energy spectrum due to the short light-travel times from the compact central corona to the inner disc (see Sect.~\ref{sec:model:endeptf} for a more detailed explanation of this effect). Hints of frequency dependence of the Fe~K lag have been seen in other sources \citep{zoghbi13,kara13d}, but only NCG~4151 with its high count rate, shows a significant frequency dependence so far.
The iron K lag is a powerful tool for understanding the geometry and kinematics of the inner accretion flow, as it encodes spectral and timing information about the reflected emission (See Sect.~\ref{sec:modelling} for more on modelling lags in the iron K band). Since the initial discovery in NGC~4151, iron K lags have been found in nine sources, including the original reverberating source, 1H0707-495 \citep{kara13a}. Fig.~\ref{lagen_stack} shows five of those sources, overplotted on the same axis. As they all have different black hole masses (and therefore different Fe K lag amplitudes), the lags have been scaled such that the relative lag matches between 3--4~keV and 6--7~keV. It is clear from this figure, that the shape of the Fe~K lag is similar in all these maximally-spinning black hole systems, but the lags associated with the complex soft excess vary greatly. This indicates the importance of the Fe~K line in understanding the effect of strong gravity on the reverberation lag. Future work in understanding the soft lag is also important, and may help break degeneracies in spectral modelling of the soft excess.
\begin{figure}
\begin{center}
\includegraphics[width=9cm]{lagen_4review.eps}
\caption{The lag-energy spectra overplotted for five of the published sources with Fe~K lags. The amplitude of the lag has been scaled such that the lags between 3--4~keV and 6--7~keV match for all sources. The sources shown are: 1H0707-495 (blue), IRAS 13224-3809 (red), Ark~564 (green), Mrk~335 (cyan) and PG~1244+026 (purple). While the shape of the Fe~K lag is similar in all these sources, the lags associated with the soft excess vary greatly.}
\label{lagen_stack}
\end{center}
\end{figure}
The bottom panel of Fig.~\ref{lag_mbh} shows the amplitude of the iron K lag for all the published measurements thus far plotted against black hole mass \citep{kara13c}. Similar to the soft lags in the top panel of Fig.~\ref{lag_mbh}, the iron K lags show a linear dependence on mass, and confirm that both the soft excess in these sources and the X-rays illuminating the accretion disc producing the iron K line, originate from a small emitting region close to the central black hole. The fact that the lags in the iron K band are generally larger than the soft lags is likely due to greater dilution by the continuum in the soft band, which can be accounted for in modelling of the lag (see Sect.~\ref{sec:modelling} for a discussion of dilution). As we gain a clearer understanding of the geometry of the systems we are probing, we can use reverberation lags as an indicator of black hole mass. The lag measures the size of the region in physical units (i.e. in metres rather than in gravitational units), and so if we understand how far the source is from the accretion disc in gravitational units, we can use the lag to make a measurement of the black hole mass.
Finally, it is worth emphasising that the high-frequency iron~K lags represent a model-independent confirmation of the interpretation of broad iron K lines as signatures of relativistic reflection from a compact reflector close to the black hole: these results are independent of any models fitted to the time-averaged spectra and demonstrate the reality of broad iron K features as distinct emission components. This picture is supported by the first detections of reverberation lags with the {\it NuSTAR} hard X-ray observatory (\citealt{zoghbi14}, Kara et al. in prep.). Fig.~\ref{fig:nustarlags} shows the {\it NuSTAR} lag-energy spectrum for SWIFT~J2127.4+5654 (Kara et al. in prep.). The discovery of Fe~K reverberation in {\it NuSTAR} data confirms that the detections by {\it XMM-Newton} are not in any sense `pathological' to that mission. Furthermore, there is a clear indication of lags expected from the disc reflection continuum at even higher energies.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{2127_nustar.eps}
\caption{{\it NuSTAR} lag-energy spectrum of SWIFT~J2127.4+5654. Figure taken from Kara et al. (in prep.). Note the increase at energies above the Fe~K line, consistent with reverberation of the disc reflection continuum.}
\label{fig:nustarlags}
\end{center}
\end{figure}
\subsubsection{Confirming small scale reverberation}
\label{sec:smallscale}
\begin{figure}
\begin{center}
\includegraphics[width=10.5cm]{ark.eps}
\caption{The lag-energy spectrum of Ark~564 for low frequencies (left) and high frequencies (right), using frequency ranges highlighted in the top panel, which shows the lag-frequency dependence for 1.2--4~keV relative to 0.3--1~keV. The iron K reflection feature is found at high frequencies, while the low frequencies show a featureless lag, increasing with energy. This shows that the low-frequency lags are not due to reflection. See Sect.~\ref{sec:bhb} for discussion of the origin of the low-frequency hard lag.}
\label{ark564}
\end{center}
\end{figure}
An alternative to the small-scale disc reverberation model for the soft lags was discussed by \citet{lancemiller4051} (and see also \citealt{legg12}), who suggested that the low-frequency hard lags are the real reverberation signatures, produced by reflection from more distant circumnuclear material (possibly associated with a disc wind or other large-scale absorbing/scattering gas), arguing that the negative, soft lags seen at higher frequencies were an artefact of the high-frequency oscillations expected from a top-hat like impulse response, which are caused by a `phase-wrapping'-like effect (see also Sect.~\ref{sec:freqdep}). The physical interpretation of such an impulse response is that the lags can be understood as scattering/reflection from material covering a large solid-angle at a range of size scales, from a hundred to a couple of thousand light seconds from the central source, or by invoking a more distant reflector which must be aligned close to the line of sight, to account for the short time delays observed. However, work by \citet{zog11_1h0707} and \citet{emma11} showed that the broad frequency range of the soft lag in many sources cannot be explained by the oscillatory effects expected for simple impulse reponses (see also Sect.~\ref{sec:freqdep}). The unique line-of-sight argument is also inconsistent with the ubiquity of the soft lag in NLS1 sources \citep{demarco13}.
\citet{lancemiller10} developed the large-scale scatterer/reflector model for the lags further in the case of 1H0707-495, proposing instead that the broad frequency range of soft lags (which cannot be explained by oscillations in the impulse response Fourier transform) could be explained if the soft band scattering impulse response is narrower with a smaller centroid than that seen in hard X-rays. However, the interpretation of the low-frequency hard lags as a reverberation signature cannot explain the differences in the lag-energy spectra at low and high frequencies \citep{zog11_1h0707,kara13a}. As discussed above, the lag-energy spectrum at high frequencies has a clear signature of reflection in the iron K lag, however, as we probe lower frequencies, the lag shows a featureless increase with energy. The differences between the low and high frequency lag-energy spectra of Ark~564 can be seen in Fig.~\ref{ark564} \citep{kara13c}. The low-frequency lag-energy spectrum, with no strong spectral features, is very similar to that of the hard lags found in black hole binaries, which can be understood as intrinsic lags associated with the continuum, possibly linked to the propagation of variations in the accretion flow which modulate the coronal emission \citep{kotov01,arevalo06,uttley11}. Furthermore, low-frequency hard lags have recently been found in NGC 6814, a source that is well described by just an absorbed power law, with very little neutral reflection \citep{walton13}, suggesting that the hard lag is due to changes in the continuum, and not produced by large scale reflection as in the \citet{lancemiller10} model. All this evidence strongly supports the picture that the high-frequency lags are caused by small scale reflection, while the hard lags are some separate process, unrelated to reflection.
\subsubsection{The frontiers of reverberation}
\label{sec:frontiers}
We are now delving into a regime where we can begin to disentangle the many contributions to the reverberation lag, i.e. we can isolate different light paths, and thus map the geometry of the source and inner flow. In IRAS~13224-3809, we find that the reverberation lag is dependent on flux, which changes with the geometry of the corona \citep{kara13b}. At low-flux intervals, we see a small amplitude lag at high frequencies (Fig.~\ref{iras}) and infer a compact corona producing the primary X-ray emission. At high fluxes, the frequency of the negative lag range is lower and the amplitude of the lag is greater, suggesting an overall longer centroid lag of the impulse response (see Sect.~\ref{sec:freqdep}) and hence a longer light-travel time to the disc. We infer from this that as the flux increased, the corona expanded to fill a larger volume. Interestingly, the low-flux interval shows a very clear iron K line, while lag-structure from the high flux intervals cannot be well constrained. Flux-dependent lags have also been found in NGC 4051 \citep{alston13}, which shows that there is no `hard lag' when the X-ray source is at low fluxes.
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm]{iras.eps}
\caption{The flux-dependent lags of IRAS~13224-3809. The low-flux intervals (red) reveal shorter amplitude reverberation lags at a higher frequency, suggesting that the X-ray source is closer to the central potential and more compact at low fluxes. The high-flux intervals (green) show a larger lag on longer time-scales, consistent with a more extended emitting region. Figure taken from \citet{kara13b}.}
\label{iras}
\end{center}
\end{figure}
The comparison of the high-frequency lag-energy spectra between different objects (Fig.~\ref{lagen_stack}) shows that although iron K lags seem to be fairly common, the picture for the soft lags is more complex. One possibility is that the soft lags are produced by a variety of mechanisms, some linked to photoionised reflection, but others linked to other components, e.g. the direct continuum itself. It probably isn't surprising that in AGN with complex spectra that can be explained by strong reflection (e.g. IRAS~13224-3809, 1H0707-495), we see very similar soft lag behaviour \citep{kara13b}, suggesting that in these systems, reflection does dominate the soft lags. It is worth noting that the variability in these AGN shows a high coherence between hard and soft energies at high frequencies (e.g. see Fig.~\ref{fig:1h0707psdcoh}, right panel) as would be expected if a single process such as reverberation dominates the spectral variability. However, the AGN PG~1244+026 shows a low coherence between the soft and hard bands, suggesting that the soft excess in that source is not linked to strong photoionised reflection but may be associated with a separate, cool Comptonised component which is uncorrelated with the harder power-law and associated reflection \citep{jin13}. This possibility is supported by the failure of propagation-plus-reflection models to explain the low-frequency soft lag behaviour in this AGN \citep{gardner14}. Correspondingly, the lag-energy spectra of PG~1244+026 depend on the reference energy band chosen \citep{kara13d,alston14}, as expected when the variability is produced by distinct and uncorrelated spectral components, with their own lag-energy behaviours.
The detectability of reverberation lags is based on the intrinsic phase lag of the reflection (which we assume to be similar from source to source, if lag and variability time-scales both scale linearly with black hole mass) and three other parameters: the flux of the source, the amount of variablility, and the amount of data we have available. In Table~\ref{lagtable}, we highlight the exposure, 2--10~keV flux and 2--10~keV excess variance (measured from 10~ks long light curve segments) for the nine sources with Fe~K lags measured to date. We compare these sources with other variable AGN to illustrate the detectability of reverberation lags. In addition to PG~1244+026, IRAS~13224-3809 and SWIFT~J2127.4+5654, we compile a sample of variable AGN that are common between the \citet{gm12} sample (which provides the 2--10~keV luminosity and the XMM-Newton exposure as of the date of submission), and the \citet{ponti12} sample (which provides the 2--10~keV excess variance in 10~ks time bins). Fig.~\ref{agnsample} shows the flux, excess variance and exposure for this sample of variable AGN. The sources with significant reverberation lags (Fe~K or soft lags) are highlighted. Reverberation lags have been found in many high-flux, strongly variable AGN, and longer observations will uncover lags in a greater range of sources.
\begin{figure}
\begin{center}
\includegraphics[width=7cm]{agn_sample.eps}
\caption{The 2--10~keV flux and 2--10~keV excess variance (in 10~ks long segments) for the \citep{ponti12} sample of variable AGN (including also PG~1244+026, IRAS~13224-3809 and SWIFT~J2127.4+5654). The sources are color coded based on their exposure times with {\em XMM-Newton}. The sources with known Fe~K or soft lags have been highlighted. The detectability of the lag is based on the sources brightness and variability, and on how long it has been observed. }
\label{agnsample}
\end{center}
\end{figure}
\begin{table}
\centering
\begin{tabular}{c|c|c|c}
\hline
{\bf Source} & {\bf Exposure} & {\bf 2--10~keV flux} & {\bf $\sigma^{2}_{\mathrm{XS}}$}\\
& ks & $10^{-12}$ erg cm$^{-2}$ s$^{-1}$ & \\
\hline
NGC~4151$^{1}$ & 99 & 80& 0.00055 \\
1H0707-495$^{2}$ & 1042 & 1.3 & 0.154 \\
IRAS~13224-3809$^{3}$ & 430 & 0.66 & 0.18 \\
MCG-5-23-16$^{4}$ & 130 & 80 & 0.00208 \\
NGC 7314$^{4}$ & 120 &15.7 & 0.031 \\
Ark 564$^{5}$ & 465 & 18.7 & 0.0401 \\
Mrk 335$^{5}$ & 120 & 17.2 & 0.0102 \\
PG 1244+026$^{6}$ & 123 & 2.9 & 0.028 \\
SWIFT J2127.4+5654$^{7}$ & 238 & 29 & 0.04 \\
\hline
\end{tabular}
\caption{Parameters for Fe~K Lag detected sources. References corresponding to subscripts to the source names: 1. \citet{zoghbi12}; 2. \citet{kara13a}; 3. \citet{kara13b} 4. \citet{zoghbi13}; 5. \citet{kara13c}; 6. \citet{kara13d}; 7. \citet{marinucci14}.}
\label{lagtable}
\end{table}
Most of the discoveries of X-ray reverberation lags have taken place in supermassive black hole systems, due to the large number of photons per light crossing time (see Sect.~\ref{sec:analysis} for more discussion on this point). However, Galactic black hole binaries have been influential in the study of the low-frequency hard X-ray time lags, and will be very important for reverberation studies with future telescopes (Sect.~\ref{sec:future}). We now review the current reverberation observations in black hole X-ray binaries.
\section{Analysis Methods}\label{sec:analysis}
In this section we will review the methods used to study variability and search for time-lags in X-ray light curves. These methods are crucial for the discovery and exploitation of the lags, so we aim to provide important background as well as recipes for the measurement of different spectral-timing quantities. Those readers who are familiar with these methods or who wish to focus on the observations and modelling can skip over this section, which is mainly pedagogical. We will first consider the basic Fourier techniques at our disposal, and then describe how these techniques are put into practice, as well as some practical issues regarding the determination of errors and the sensitivity of lag measurements. Finally we will introduce the concept of the impulse response, which shows how we can model the observed energy-dependent variability properties in terms of reverberation and other mechanisms, which will be considered in more detail in Sect.~\ref{sec:modelling}.
\subsection{The toolbox: Fourier analysis techniques}
Time-series analysis can be done in the time- or frequency-domain. Optical reverberation studies of AGN use time-domain techniques to measure time-lags, specifically the cross-correlation function \citep[e.g.][]{peterson93,petersonetal04,bentz09,denney10}. More recently, stochastic modelling and fitting of light curves has been applied to more accurately model the optical reverberation data \citep{zu11,pancoast11}. The situation for X-ray time-series analysis is different however: here the focus to date has been on Fourier analysis techniques (with just a few exceptions, e.g. \citealt{maccarone00,dasgupta06,legg12}). There are several reasons for this difference. Firstly, the Fourier power spectrum is an easy way to describe the underlying structure of a stochastic variable process, e.g. dependence of variability amplitude on time-scale, with statistical errors that are close to independent between frequency bins and therefore is easily modellable (unlike time-domain techniques such as the autocorrelation function or structure function where errors are correlated between bins, e.g. see \citealt{emma10}).
Secondly, although aliasing effects make it difficult to cleanly apply Fourier techniques to irregularly-sampled and `gappy' data from long optical monitoring campaigns, (Fast) Fourier techniques lend themselves particularly well to analysis of the very large, high-time-resolution light curves used to study the very rapid variability in X-ray binaries. The same approaches are easily applicable to the contiguous light curves obtained from `long-look' observations of AGN by missions such as {\it EXOSAT} (e.g. \citealt{lawrence87}) and currently, {\it XMM-Newton} (e.g. \citealt{mchardy05}). Another important factor is that there has been a significant focus on comparisons of XRB and AGN X-ray variability, e.g. to uncover the mass-scaling of variability time-scales, as well as comparing spectral-timing properties \citep{mchardy06,arevalo06}. Thus it is useful to use a common approach to data from both kinds of object in the X-ray band, and these efforts have also stimulated the development of techniques to study irregularly-sampled AGN data in the Fourier domain \citep{uttley02,markowitz03}, or more recently, through a combination of Fourier and time-domain approaches \citep{lancemiller10,kelly11,zoghbi13_gaps}. Also important is the fact that Fourier-techniques allow for the easier interpretation of complex data by decomposing the data in terms of the variations on different time-scales.
\subsubsection{The discrete Fourier transform and power spectral density}
The X-ray light curves of AGN and XRBs are best described as stochastic, {\it noise processes}\footnote{Even quasi-periodic oscillations seen in XRBs follow the same statistics as noise \citep{vanderklis97}.}. A well-known type of noise process is observational, i.e. Poisson noise, but here the observed intrinsic variations themselves are also a type of noise and thus inherently unpredictable at some level. The extent to which we can predict the flux at one time based on the flux at another depends on the shape of the underlying power spectral density function or PSD, $\mathit{P}(f)$, which describes the average variance per unit frequency of a signal at a given temporal frequency $f$. Strongly autocorrelated signals, where variations between adjacent time bins are small and increase strongly with larger time-separation (e.g. `random walks' or Brownian noise) correspond to steep {\it red noise} PSDs ($P(f)\propto f^{\alpha}$ with $\alpha \leq -2$). Above some limiting time-scale the size of variations must reach some physical limit and the PSD below the corresponding frequency flattens, typically to {\it flicker noise} with $\alpha\simeq -1$. This characteristic bend in the PSD is detected in AGNs and BHXRBs, and occurs at a frequency corresponding to a characteristic time-scale which scales linearly with black hole mass and possibly also inversely with accretion rate (\citealt{uttley05,mchardy06} but see also \citealt{gm12} which confirms the linear mass-dependence but suggests a weaker accretion rate dependence). At even lower frequencies, in BHXRBs and also detected in the AGN Ark~564, the PSD can flatten again to become {\it white noise} ($\alpha=0$) so that the total variance becomes finite and there is no long-term memory in the light curve on those time-scales. Poisson noise is also a form of white noise.
The PSD can be estimated from the periodogram, which is the modulus-squared of the discrete Fourier transform of the light curve. The discrete Fourier transform (DFT) $X$ of a light curve $x$ consisting of fluxes measured in $N$ contiguous time bins of width $\Delta t$ is given by:
\begin{equation}
\label{eqn:dft}
X_n = \sum_{k = 0}^{N - 1} x_k \exp\left( 2\pi i n k / N \right)
\end{equation}
where $x_k$ is the $k$th value of the light curve and $X_n$ is the discrete Fourier transform at each Fourier frequency, $f_n = n/(N\Delta t)$, where $n=1,2,3,...N/2$. Thus the minimum frequency is the inverse of the duration of the observation, $T_{\rm obs}=N\Delta t$ and the maximum is the Nyquist frequency, $f_{\rm max} = 1/(2\Delta t)$.
The periodogram is simply given by:
\begin{equation}
\label{eqn:periodogram}
|X_n|^2 = X_n^{*} X_n
\end{equation}
Where the asterisk denotes complex conjugation. In practice the periodogram is further normalised to give the same units as the PSD:
\begin{equation}
\label{eqn:psd}
P_n=\frac{2\Delta t}{\langle x \rangle^{2} N} |X_n|^{2}
\end{equation}
where the value in angle brackets is the mean flux of the light curve and the normalised periodogram $P_{n}$ is thus expressed in units of {\it fractional} variance per Hz \citep{belloni90,miyamoto92}, so that this normalisation is often called the `rms-squared' normalisation. Thus, integrating the PSD with this normalisation over a given frequency range and taking the square root gives the fractional rms variability (often called $F_{\rm var}$) contributed by variations over that frequency range.
For a noise process, the observed periodogram is a random realisation of the underlying PSD, with values drawn from a highly-skewed $\chi^{2}_{2}$ distribution with mean scaled to the mean of the underlying PSD. It is thus important to note that the `noisiness' of the observed periodogram is in some sense intrinsic to the `signal', i.e. the underlying variability process. Since the underlying PSD of the process is the physically interesting quantity, we usually bin up the periodogram to obtain an estimate of the PSD (in frequency and also over periodograms measured from multiple independent light curve segments), so that the PSD in a frequency bin $\nu_{j}$ averaged over $M$ segments and $K$ frequencies per segment is given by:
\begin{equation}
\label{eqn:binnedpsd}
\bar{P}(\nu_{j}) =\frac{1}{KM} \sum_{n=i,i+K-1} \sum_{m=1,M} P_{n,m}
\end{equation}
where $\bar{P}(\nu_{j})$ is the estimate of the PSD obtained from the average of the periodogram in the bin $\nu_{j}$ (henceforth, we will refer to the measured quantity as the PSD) and $P_{n,m}$ is the value of a single sample of the periodogram measured from the $m$th segment with a frequency $f_n$ that is contained within the frequency bin $\nu_{j}$ (which contains frequencies in the range $f_i$ to $f_{i+K-1}$). Note that here we use $\nu$ to denote frequency {\it bins}, rather than the underlying frequency $f$. The standard error on the PSD $\Delta \bar{P}(\nu_{j})$ can be determined either from the standard deviation in the $KM$ samples (divided by $\sqrt{KM}$ to give the standard error on the mean) or simply from the statistical properties of the $\chi^{2}_{2}$ distribution, which imply that, for a large number of samples\footnote{It is important to bear in mind that due to the highly skewed nature of the $\chi^{2}_{2}$ distribution, errors on the PSD only approach Gaussian after binning a large number of samples ($KM>50$). An alternative approach, which converges more quickly to Gaussian-distributed errors, is to bin $\log(P_{n,m})$, which also necessitates adding a constant bias to the binned log-power, see \citet{papadakis93,vaughan05} for details.}:
\begin{equation}
\label{eqn:psderror}
\Delta \bar{P} (\nu_{j}) = \frac{\bar{P}(\nu_{j})}{\sqrt{KM}}
\end{equation}
Poisson noise leads to flattening of the PSD at high frequencies with a normalisation depending on the observed count rate, which is easily accounted for by subtracting a constant $P_{\rm noise}$ from the observed PSD. In the case where Poisson statistics applies and the fluxes are expressed in terms of count rates:
\begin{equation}
\label{eqn:noiselevel}
P_{\rm noise}=\frac{2\left(\langle x \rangle + \langle b \rangle \right)}{\langle x \rangle^{2}}
\end{equation}
where $\langle b \rangle$ is the average background count rate in the light curve (we assume that the background is already subtracted from $x$). For non-contiguous sampling, the noise level must be increased in line with the reduced Nyquist frequency \citep{markowitz03,vaughan03}. If fluxes are not given in count rates (or the data are expressed as a count rate but the statistics are not Poissonian), then the equivalent noise level can be determined using $P_{\rm noise}=\langle \Delta x^{2} \rangle / \left( \langle x \rangle^{2} f_{\rm Nyq} \right) $, where $\langle \Delta x^{2} \rangle$ is the average of the squared error-bars of the light curve, and $f_{\rm Nyq}$ is the Nyquist frequency ($f_{\rm Nyq}=f_{\rm max}=1/(2\Delta t)$ for contiguous sampling).
\begin{figure}
\centering
\includegraphics[width=0.33\textwidth]{zoghbi11psds.eps}
\includegraphics[width=0.57\textwidth]{zoghbi10coherence.eps}
\caption{{\it Left:} The Poisson-noise subtracted PSDs of the NLS1 AGN 1H0707-495 (taken from \citealt{zog11_1h0707}). {\it Right:} 1H0707-495 frequency-dependent coherence between the 0.3--1~keV and 1--4~keV bands. The solid and dashed blue lines give the median and upper and lower 95~per~cent confidence levels for the coherence obtained from simulations of correlated (unity intrinsic coherence) light curves with the same flux levels, variance and PSD shape as the data. Note the dip, suggesting a changeover between two processes. At high frequencies, the coherence is consistent with unity, as expected from simple reverberation. Figure taken from \citet{zoghbi10}.}
\label{fig:1h0707psdcoh}
\end{figure}
Some example noise-subtracted PSDs for the light curve of 1H0707-495 measured in different energy bands are shown in Fig.~\ref{fig:1h0707psdcoh} (left panel). All bands show the characteristic red-noise shape at high frequencies (although not clear from the figure, there is evidence for a break to a flatter slope at $\sim 1.4\times 10^{-4}$~Hz, see \citealt{zoghbi10}). However, the higher energies show a flatter PSD, showing that there is relatively more rapid variability at these energies. Similar energy-dependent behaviour is also seen in BHXRBs (e.g. \citealt{nowak99}).
\subsubsection{The cross spectrum and lags}
\label{sec:crossspeclags}
The Fourier cross spectrum between two light curves $x(t)$ and $y(t)$ with DFTs $X_n$ and $Y_n$ is defined to be:
\begin{equation}
\label{eqn:crossspec}
C_{XY,n}=X_n^{*} Y_n
\end{equation}
We can see how the cross-spectrum is used to derive the frequency-dependent (phase) lag between two bands by considering the complex polar representation of the Fourier transform $X_n=A_{X,n}\exp\left[i\psi_n\right]$, where $A_{X,n}$ is the absolute magnitude or amplitude of the Fourier transform and $\psi_n$ is the phase of the signal (which for a noise process is randomly distributed between $-\pi$ and $\pi$) at the frequency $f_n$. Thus, a linearly correlated light curve $y(t)$ with an additional phase-shift $\phi_n$ at frequency $f_n$ has a Fourier transform $Y_n=A_{Y,n}\exp\left[i(\psi_n+\phi_n)\right]$. Multiplying by the complex conjugate of $X_{n}$, the phase $\psi_n$ cancels and the cross-spectrum is given by:
\begin{equation}
C_{XY,n} = A_{X,n}A_{Y,n}\exp\left(i\phi_{n}\right)
\end{equation}
with the phase of the cross-spectrum giving the phase lag between the light curves, as expected.
In principle the cross-spectrum may also be normalised in the same way as the periodogram, except that instead of dividing by $\langle x \rangle^{2}$ to obtain fractional rms-squared units, we must divide by the product of light curve means, $\langle x \rangle \langle y \rangle$. Note that due to the well-known linear rms-flux correlation in AGN and XRB light curves, different results can be obtained if the lags depend on the flux level (see Sect.~\ref{sec:frontiers}) and either the cross-spectrum is normalised by the means of each light curve segment {\it before} averaging segments, or a single combined mean value for all segments is used {\it after} averaging \citep{alston13}.
In the presence of any uncorrelated signal between the two light curves (e.g. due to Poisson noise, but there may also be an {\it intrinsically incoherent} signal, perhaps due to the presence of an additional independently-varying component in one energy band but not the other), the cross-spectrum should be averaged over Fourier frequencies in a given frequency bin $\nu_{j}$, as well as over multiple light curve segments, to reduce the effects of noise:
\begin{equation}
\label{eqn:binnedcrossspec}
\bar{C}_{XY} (\nu_{j}) = \frac{1}{KM} \sum_{n=i,i+K-1} \sum_{m=1,M} C_{XY,n,m}
\end{equation}
Here we assume the same notation as in Equation~\ref{eqn:binnedpsd}. The phase of the resulting binned cross-spectrum (i.e. the {\it argument} of the complex cross-spectrum vector) $\phi(\nu_{j})$, gives the average phase lag between the two light curves in the $\nu_{j}$ frequency bin. The time lag is thus given by:
\begin{equation}
\label{eqn:timelag}
\tau(\nu_{j}) = \phi(\nu_{j})/\left(2\pi \nu_{j}\right)
\end{equation}
Conversion of phase to time-lags is often carried out for ease of interpretation of the data, but since the phase is limited to the range $-\pi$ to $\pi$, caution should be taken when interpreting any time-lags corresponding to phase lags close to these limits (e.g. so-called phase-wrapping across the phase boundaries can lead to sudden `flips' in the lag). Furthermore, for broad frequency bins, it is not always clear what value should be used for the bin frequency, $\nu_{j}$ to obtain the time-lag, e.g. should the frequency be weighted according to the power contributing from each sample frequency in the bin, or should the bin centre or unweighted average frequency be used? The more common procedure is to use the bin centre, but the question is in some sense a matter of taste and convention, because conversion of phase lags to time-lags is done mainly for the convenience of expressing the phase lag in terms of a physically useful time-scale. The choice of frequency will lead to some small biases in the observed time-lag, but this effect can be easily accounted for when modelling the lags, or by fitting models to the cross-spectrum or phase-lags directly.
Examples of lag-frequency spectra are shown for a BHXRB and AGN respectively in Fig.~\ref{fig:cygx1lagfreq} and Fig.~\ref{1h07_lag}.
\subsubsection{The coherence and errors}
\label{sec:coherr}
The {\it coherence} $\gamma^{2}$ in the frequency bin $\nu_{j}$ is defined as:
\begin{equation}
\label{eqn:coherence}
\gamma^{2}(\nu_{j}) = \frac{|\bar{C}_{XY}(\nu_{j})|^{2}-n^{2}}{\bar{P}_{X} (\nu_{j}) \bar{P}_{Y}(\nu_{j})}
\end{equation}
where the binned PSD and cross-spectrum have been normalised in the same way (see above). We stress here that it is only meaningful to measure the coherence of the binned cross-spectrum (i.e. averaged over segments and/or frequency), so that uncorrelated noise cancels during the binning, otherwise observed coherence will always be unity, by definition. The $n^{2}$ term is a bias term, which arises because the Poisson noise-level contributes to the modulus-squared of the cross-spectrum\footnote{$n^{2}=\left[(\bar{P}_{X} (\nu_{j})-P_{X,{\rm noise}})P_{Y,{\rm noise}}+(\bar{P}_{Y}(\nu_{j})-P_{Y,{\rm noise}})P_{X,{\rm noise}}+P_{X,{\rm noise}}P_{Y,{\rm noise}}\right]/KM$, where we assume that the binned PSDs are not already noise-subtracted. See \citet{vaughan_nowak97} for further details.}. Formally, the coherence indicates the fraction of variance in both bands which can be predicted via a linear transformation between the two light curves \citep{vaughan_nowak97}. A reduction in coherence can also occur due to a non-linear transformation (see the discussion in \citealt{vaughan_nowak97,nowak99}). An example plot of coherence versus frequency is shown for 1H0707-495 in Fig.~\ref{fig:1h0707psdcoh} (right panel).
Geometrically, the coherence gives an indication of the scatter on the cross-spectrum vector that is caused by the incoherent components \citep{nowak99}. It can thus be used to derive the error on the phase lag \citep{bendat10}:
\begin{equation}
\label{eqn:lagerror}
\Delta \phi(\nu_{j}) = \sqrt{\frac{1-\gamma^{2}(\nu_{j})}{2\gamma^{2}(\nu_{j})KM}}
\end{equation}
and the time-lag error is simply $\Delta \tau = \Delta \phi/(2\pi \nu_{j})$. It is easy also to work out the equivalent expression in terms of the contributions to the cross-spectrum of the signal and noise components of the Fourier transform, which we will consider later, when we discuss the sensitivity of lag measurements.
One important point to note here is that the coherence used to estimate the lag error is the {\it raw} coherence, where the PSD values in the denominator of Equation~\ref{eqn:coherence} have not had the Poisson noise level subtracted. The advantage of this approach is that the lag errors can be easily determined from the data, without any other assumptions, since the quantities used are simply the observed cross-spectrum and PSD. Problems can arise once the raw coherence becomes comparable to the $n^{2}$ bias term, since oversubtraction of the bias would lead to negative values of coherence, preventing the standard estimation of lag errors using the coherence. In that case, if bias is not subtracted then in the limiting case of negligible intrinsic variability, the raw coherence becomes $1/KM$ and the estimated error on the phase lag will saturate at $\Delta \phi= 1/\sqrt{2}$ (the real error, accounting for bias, is substantially larger than this).
If the {\it intrinsic} coherence is required for a study of the variability properties of the source, i.e. the coherence attributable to the source itself, correcting for observational noise, the noise-levels should be subtracted from the PSDs in the denominator of Equation~\ref{eqn:coherence}. Errors on the intrinsic coherence can be estimated using the approach outlined in \citet{vaughan_nowak97}, but it is important to bear in mind that the errors on intrinsic coherence are difficult to correctly determine for low variability $S/N$. This problem does not affect the lag-error estimation however, since this is based on the raw coherence.
\subsection{Practical Application}
We now consider the practical application of the Fourier methods outlined above to real data.
\subsubsection{Frequency-dependent spectral-timing products}
The cross-spectral approach outlined above can be used to measure the {\it lag-frequency spectrum}: the phase or time-lag between two broad energy bands plotted as a function of Fourier frequency, as well as the coherence, if required. This - now standard - approach was first applied to data from X-ray binaries and also used to study lower frequencies in AGN before eventually leading to the discovery of soft lags and reverberation at high frequencies \citep{fabian09}. As we have already noted, the key advantage of studying lags in the Fourier domain, rather than the time-domain (e.g. with the cross-correlation function), is that complex time-scale-dependent lag behaviour associated with multiple physical processes can easily be disentangled. Hence, this approach has become the workhorse for the discovery of soft lags in AGN, as we will see in Sect.~\ref{sec:obs}.
The practical steps to calculate the lag-frequency spectrum and other frequency-dependent products are:
\begin{enumerate}
\item Create light curves with identical time-sampling in two, separate energy-bands.
\item Split the light curves into $M$ continuous segments of equal duration, choosing the segment duration based on the lowest frequency to be sampled. Small gaps in the light curves (up to a few per cent of the light curve duration) can be interpolated over (with random errors added to match the observational errors) without significant distortion of the resulting cross-spectrum, providing that the variability on time-scales comparable or shorter than the gap size is small compared to the amplitude on longer time-scales (e.g. the gap time-scale corresponds to the steep red-noise part of the PSD). Larger gaps should be avoided using a suitable choice of segment size or a method that accounts for their effects on the data and errors \citep{zoghbi13_gaps}.
\item For each segment, obtain the PSDs in both energy bands (Equations~\ref{eqn:dft}--\ref{eqn:binnedpsd}) and the cross-spectrum (Equations~\ref{eqn:crossspec}--\ref{eqn:binnedcrossspec}), averaging them to form PSDs and the cross-spectrum binned over segments.
\item Further bin the PSD and cross-spectrum over frequency, as required. It is often convenient to bin geometrically in frequency, i.e. from frequency $f$ to frequency $Bf$ where $B$ is the selected binning factor ($B>1.0$), so that the frequency bins have the same width in log-frequency. If there are $K$ frequencies sampled in a bin, the number of samples is then equal to $K\times M$.
\item Use the phase of the binned cross-spectrum to calculate the phase-lag and/or time-lag (Equation~\ref{eqn:timelag}) for each frequency bin.
\item Calculate the {\it raw} coherence from the binned PSD and cross-spectrum (Equation~\ref{eqn:coherence}).
\item Use the coherence in each frequency bin to calculate the error bar on the lag in each frequency bin (Equation~\ref{eqn:lagerror}).
\item If desired, obtain the intrinsic coherence, following the prescription in \citet{vaughan_nowak97}.
\end{enumerate}
Note that the choice of segment size may be important if `leakage' effects - due to the finite sampling window - are a problem in the data set being considered. For example, \citet{alston13} have shown that an effect of a finite segment size, which is more pronounced for shorter segments, is that lags can leak across frequencies (the effect is related to the problem of `red-noise leak' seen in PSDs, e.g. see \citealt{uttley02} for discussion). This effect is strongest when the gradient of the phase lag-frequency spectrum is largest. The leakage effect does not qualitatively affect any of the reverberation results discovered so far, but may become important as more detailed models are applied. It can be reduced by choosing longer segments (so that the flatter part of the PSD is sampled, leading to less leakage from low frequencies), or by `end-matching' the data to take out any long-term trend \citep{alston13}, although this can lead to other biases, which should be modelled. One approach to maximising the segment length (since choosing a fixed segment length can lead to some data being ignored, when observations with different durations are combined) is to measure the Fourier transforms for the entire contiguous light curves for each observation, and combine them by binning in frequency\footnote{Since the measured Fourier frequencies depend on segment length, binning is best done by making a frequency-ordered list of frequencies and power or cross-spectral value from all segments and then binning the power/cross-spectra according to frequency.}. This approach is best used when leakage is not expected to be significant and/or segment lengths are relatively similar, since combining data with different leakage effects (due to different segment lengths) can lead to confusing results which are difficult to model.
Following the recipe presented here will produce the range of commonly-used frequency-dependent spectral-timing products: the PSD in both bands, the coherence vs. frequency and the lag vs. frequency. Multiple bands can be used to compare the frequency-dependent spectral-timing properties as a function of energy. However, new insights can be gained by using a similar approach to make energy-dependent spectral-timing products, which we outline below.
\subsubsection{The lag-energy spectrum}
Although the standard lag-frequency measurement can give valuable insights, it is especially useful to look at the dependence of the lag on energy at higher spectral resolution, to reveal which spectral components contribute to the lags, and thus the {\it causal relationship} between the different spectral components.
To plot the {\it lag-energy spectrum}, it helps to optimise the signal-to-noise in two ways. Firstly, we must select a broader frequency range to average the cross-spectrum over than the narrower bins used for lag-frequency spectra. The frequency range may be chosen to select some particularly interesting behaviour for the measurement of the lag-energy spectrum, e.g. the measurement of the spectrum corresponding to negative or 'soft' lags. However, as noted by \citet{lancemiller10}, it is important to bear in mind that impulse responses corresponding to a single physical component may have a complex effect over a wide range of frequencies, so that it is not always obvious that a given frequency selection also corresponds to the selection of distinct physical components or effects. Lag-energy spectra may therefore be treated as being suggestive of certain physical effects, but ultimately, fully self-consistent modelling of the data over a wide-range of frequencies will be the best way to test our physical understanding.
Secondly, to further increase signal-to-noise, we can choose a broad {\it reference} energy band with which to measure a cross-spectrum for each of the individual energy bins or {\it channels-of-interest} (CI). By selecting a common reference band, we can measure the lag of each energy bin relative to the same reference band, to obtain a lag-energy spectrum (see also \citealt{vaughan94} for an alternative route to the lag-energy spectrum). Provided that each energy bin shares the same intrinsic coherence with the reference band, it is easy to then interpret the relative lag {\it between} each energy bin, as the lag between the variations at those energies which are also correlated with the reference band. In principle, any choice of reference band can be made but this can potentially convey different information, if the intrinsic coherence changes between energy bins. In the simplest case where variations are intrinsically fully-coherent between energies, it is simplest to use the broadest possible reference band, i.e. across all energies with good $S/N$. Using a broad reference band has the advantage that the error associated with Poisson noise in the reference band is minimised. This effect is especially important for the study of XRBs, as we shall see.
The steps for calculation of the lag-energy spectrum are as follows:
\begin{enumerate}
\item Choose a reference energy band and make its light curve.
\item Make a light curve for each channel-of-interest using the same time-sampling as the reference band. Before making the spectral-timing products, {\it if the channel-of-interest is contained in the same energy range as the reference band and samples the same data, i.e. is not from a separate detector}, then subtract off the channel-of-interest to leave a CI-corrected reference light curve.
\item Using the CI-corrected reference light curve with the CI light curve, obtain the PSDs and cross-spectrum following steps 2-3 of the approach for making frequency-dependent spectral-timing products.
\item Average the PSDs and cross-spectra over the selected broad frequency range, and obtain lags, raw coherence and lag errors associated with the channel-of-interest. The same data may also be used to make rms and covariance spectra, outlined in the following subsection.
\item Repeat for all the channels-of-interest to make a lag-energy spectrum.
\end{enumerate}
Note that subtracting the channel-of-interest light curve from the reference band is necessary in order to subtract off the part of the Poisson noise in the CI which is correlated with itself in the reference band, thereby contaminating the cross-spectrum with a spurious zero-lag component. An equivalent procedure is to subtract the CI PSD from the cross-spectrum. Technically, doing this correction means that each CI is correlated with a slightly different reference light curve (with a slightly different average energy) and hence the relative lags are not strictly equivalent to the true relative lag between each CI. However, provided the CI bins are relatively narrow and thus contain only a small fraction of reference band photons, the effect on the lag-energy spectrum is small \citep{zog11_1h0707}.
Examples of lag-energy spectra are shown throughout Sect.~\ref{sec:obs}.
\subsubsection{The rms and covariance spectrum}
The first application of a broad reference band to obtain detailed energy-dependent spectral-timing data was for the {\it covariance spectrum}, used by \citet{wilkinson09} to uncover the variability of accretion disc emission in the hard state BHXRB GX~339-4 (see Sect.~\ref{sec:bhb}), and subsequently used in a number of other analyses of data from AGN and XRBs \citep{uttley11,middleton11,cassatella12swift,kara13a,cackett13}. The covariance spectrum\footnote{Strictly speaking, the `covariance' spectrum measures the {\it square-root} of the covariance of each channel with the reference band.} is the cross-spectral counterpart of the rms-spectrum, which measures the rms amplitude of variability as a detailed function of energy. By selecting a frequency range (with width $\Delta \nu$) over which to integrate the PSD in each energy band, a Fourier-resolved rms-spectrum can be obtained \citep{revnivtsev99,gilfanov00}. In the same way, we can use the cross-spectrum to obtain a Fourier-frequency resolved covariance spectrum, which shows the spectral shape of the components which are correlated with the reference band. Thus a careful choice of reference band can be used to identify those spectral components which vary together in a given frequency range, and those which do not. The rms and covariance spectra can be determined as follows\footnote{Note that the description of the calculation of the covariance spectrum and its errors given here should be used instead of that given in \citet{cassatella12swift}, which contains several typos. We would like to thank Simon Vaughan for bringing these errors to our attention.}:
\begin{enumerate}
\item Follow steps 1--4 for the calculation of the lag-energy spectrum described above.
\item Subtract the noise level (Equation~\ref{eqn:noiselevel}) from the CI PSD and the CI-corrected reference PSDs which have been averaged over the frequency-range of interest.
\item Multiply the CI PSD by the frequency range width $\Delta \nu$, to obtain the variance in that frequency range. If the PSD uses the fractional rms-squared normalisation and an absolute counts spectrum is desired, multiply the CI PSD by the squared-mean of the CI light curve, to obtain the variance in absolute units. Take the square root to obtain the (fractional or absolute) rms of the CI, which can then be used to make the rms spectrum.
\item Repeat the above step for the CI-corrected reference band PSD.
\item Take the amplitude of the binned bias-subtracted cross-spectrum \newline ($\sqrt{|\bar{C}_{XY} (\nu_{j})|^{2}-n^{2}}$) and multiply it by $\Delta \nu$. Then multiply by the product of reference and CI light curve means if the normalisation is to be corrected into absolute units.
\item Divide the resulting value by the (Poisson-noise subtracted) rms of the CI-corrected reference band. The final resulting value is the value of the covariance spectrum in the channel-of-interest. It has the same units as the equivalent rms spectrum.
\end{enumerate}
Mathematically then, the covariance spectrum in absolute flux units (which may then be fitted and interpreted in a similar way to a time-averaged X-ray spectrum) is given by:
\begin{equation}
\label{eqn:covspec}
Cv(\nu_{j})=\langle x \rangle \sqrt{\frac{\Delta \nu_{j} \left(|\bar{C}_{XY}(\nu_{j})|^{2}-n^{2}\right)}{\bar{P}_{Y}(\nu_{j}) - P_{Y,\rm noise}}}
\end{equation}
where $Y$ denotes the CI-corrected reference band, $\Delta \nu_{j}$ is the frequency width of the $\nu_{j}$ bin and we assume that the cross and power-spectra use the fractional rms-squared normalisation (otherwise multiplication by the $\langle x \rangle$ is not required). Note that the covariance spectrum is closely related to the coherence\footnote{The covariance spectrum can also be calculated directly from the coherence, using $Cv(\nu_{j})=\langle x \rangle \sqrt{\gamma^{2}(\nu_{j})(\bar{P}_{X}(\nu_{j}) - P_{X,\rm noise})\Delta \nu_{j}}$.}, and in the limit of unity coherence, the covariance spectrum should have the same shape as the rms spectrum. However, even in this case, the signal-to-noise of the covariance spectrum is substantially better than that of the rms spectrum, since the reference band light curve is effectively used as a `matched filter' to pick out the correlated variations in each channel-of-interest. Examples of covariance spectra measured for different frequency ranges for 1H0707-495 are shown in Fig.~\ref{fig:1h0707covspec}, and for the hard state BHXRB GX~339-4 in the insets of Fig.~\ref{fig:gx339lagvsen}.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{kara13covariance.eps}
\caption{Covariance spectra of 1H0707-495 in different frequency ranges. The covariance spectrum of the frequency range over which the high-frequency soft lags are detected is shown in black and shows a harder spectrum than is seen at lower frequencies, consistent with the energy-dependent PSD behaviour. Variable reflection features appear to be present on all time-scales, as expected if the continuum drives variable reflection (figure taken from \citealt{kara13a}).}
\label{fig:1h0707covspec}
\end{center}
\end{figure}
The errors on the rms spectrum must be calculated based on the errors expected due to Poisson noise (e.g. see \citealt{vaughan03}), since they are independent between different energy bins in the rms spectrum, whereas the errors in the rms due to intrinsic stochastic variability (i.e. the intrinsic $\chi^{2}_{2}$ scatter in the PSD) are correlated between energy bins if the coherence between bins is non-zero. In the limit where the intrinsic coherence is unity, the errors on the (absolute) rms-spectrum are given by:
\begin{equation}
\label{rmsspecerr}
\Delta \sigma_{X}(\nu_{j}) = \sqrt{\frac{2 \sigma^{2}_{X}(\nu_{j}) \sigma^{2}_{X,\rm noise} + (\sigma^{2}_{X,\rm noise})^{2}}{2KM\sigma^{2}_{X}(\nu_{j})}}
\end{equation}
where $\sigma_{X}(\nu_{j})$ is the noise-subtracted rms in the channel-of-interest $x$ and $\sigma^{2}_{X,\rm noise}$ is the absolute rms-squared value obtained from integrating under the Poisson noise level of the bands, i.e. $\sigma^{2}_{X}(\nu_{j})=(P_{X} (\nu_{j}) - P_{X,\rm noise})\langle x \rangle ^{2} \Delta \nu_{j}$, $\sigma^{2}_{X,\rm noise}=P_{X,\rm noise} \langle x \rangle ^{2} \Delta \nu_{j}$, where the PSDs are in the fractional rms-squared normalisation. See \citet{vaughan03} for discussion of the rms-spectrum and its error (determined from numerical simulations), and \citet{wilkinson11} for a formal derivation of the error.
For unity intrinsic coherence, the error on the covariance spectrum is given by:
\begin{equation}
\label{covspecerr}
\Delta Cv(\nu_{j}) = \sqrt{\frac{\left[Cv(\nu_{j})\right]^{2} \sigma^{2}_{Y,\rm noise} + \sigma^{2}_{Y}(\nu_{j})\sigma^{2}_{X,\rm noise} + \sigma^{2}_{X,\rm noise} \sigma^{2}_{Y,\rm noise}}{2KM\sigma^{2}_{Y}(\nu_{j})}}
\end{equation}
where $\sigma^{2}_{Y}(\nu_{j})$ is the noise-subtracted absolute rms-squared of the reference band and $\sigma^{2}_{Y,\rm noise}=P_{Y,\rm noise} \langle y \rangle ^{2} \Delta \nu_{j}$, where the PSDs are again in the fractional rms-squared normalisation. Note that $\left[Cv(\nu_{j})\right]^{2}$ is used instead of the absolute rms-squared of the CI ($\sigma_{X}(\nu_{j})$) in this formula, because we assume unity intrinsic coherence, in which case $\left[Cv(\nu_{j})\right]^{2}$ gives a significantly more accurate measure than the equivalent value obtained from the rms-spectrum.
Note that the errors on the covariance spectrum can be substantially smaller than those of the rms spectrum provided that $\sigma^{2}_{Y}(\nu_{j}) > \sigma^{2}_{Y,\rm noise}$ {\it and} $\sigma^{2}_{X}(\nu_{j}) < \sigma^{2}_{X,\rm noise}$. This is likely to be the case in many situations where the reference band contains substantially more photons than the channels of interest, e.g. through covering a broader energy range and/or lower X-ray energies.
\subsection{Sensitivity and signal-to-noise considerations}
\label{sec:sensitivity}
We now consider the sensitivity of lag measurements as well as the accuracy and limitations of the equation for describing the error on the lag (Equation~\ref{eqn:lagerror}). To test the efficacy of the error equation and examine the statistical uncertainty of the lag as a function of flux and number of samples measured, we carry out Monte Carlo simulations of light curves with a fairly typical PSD for an NLS1 AGN, with slope -1 breaking to -2 above a frequency of $10^{-4}$~Hz. The PSD normalisation in fractional rms-squared units is 200~Hz$^{-1}$ at the break frequency, yielding a typical fractional rms in 100~ks of a few tens of per~cent. For each flux level selected, we used the method of \citet{timmerkoenig95} to simulate light curves which were then simply shifted by 100 seconds to yield a lagging light curve (the simulated light curves were also exponentiated, to produce the appropriate log-normal flux distribution, see \citealt{uttleymchv05}). Poisson noise was then added to both light curves according to the chosen count rates: the `driving' light curve was chosen to have a count rate 30 times higher than the lagging light curve (analogous to the differences in count rate expected when measuring lag-energy spectra using a broad reference band). The lag was measured by averaging the cross-spectra over the frequency range $1$--$3\times10^{-3}$~Hz, and the raw coherence and hence lag error were estimated in the standard way. Using the same underlying light curves, we regenerated the Poisson noise on both light curves $10^{4}$ times, to examine the distribution of the observed lag and the estimated lag error.
The results of our simulations are shown in Fig.~\ref{lagerror}. Note that due to the choice of the frequency bin centre as the frequency for conversion of phase to time lag (see discussion in Sect.~\ref{sec:crossspeclags}), the observed lag is only $\sim80$~s. The figure shows the equivalent $1\sigma$ range of the distribution of observed lags\footnote{I.e. corresponding to half the separation in lag between the 15.87 and 84.13 percentile values of the distribution, which is equivalent to the standard deviation for a Gaussian distribution.} and the median analytical estimate of the error for a variety of different count rates (plotted versus the lagging light curve count rate) and also for two different regimes. Since the same PSD is used in all cases the count rates can also be multiplied by the time-scale being probed (e.g. in this case $\sim500$~s for a lag in the range $1$--$3\times10^{-3}$~Hz) to give the {\it counts per cycle} of variability, meaning that the results can be extrapolated to objects with different black hole masses but the same mass-scaled PSD. The high count rate regime (blue) uses lags calculated by averaging the cross-spectrum from 3 light curve segments, each 40960~s long (bin size is chosen to be 10~s), i.e. similar to what might be typical for AGN observations. The low count rate regime (red) uses 1000 segments, also each 40960~s long\footnote{The distributions in the low count rate regime are generated using only 300 realisations instead of $10^{4}$, due to computational speed limitations.}, thus compensating for the lower count rates. This latter situation could never be realised for AGN with realistic X-ray observatories, but is closer to the situation in X-ray binaries, where we observe many more cycles of variability, but at much lower count rate per cycle.
\begin{figure}
\begin{center}
\includegraphics[width=7cm,angle=-90]{lagsnplot.ps}
\caption{Errors on the measurement of an $\sim80$~s lag in simulated light curves, comparing the error bar obtained using Equation~\ref{eqn:lagerror} (dotted lines, note that the median error obtained from the 300 or $10^{4}$ simulated light curves was used in each case) with the equivalent $1\sigma$ error obtained from the distribution of simulated lag measurements (open and filled data points). Count rates are given for the lagging band (the 'channel-of-interest') while the driving band (the `reference band') is given a count rate which is 30 times larger. Since the same PSD is assumed for all cases, the count rate can be related to counts per cycle of variability, simply by multiplying the rate by the time-scale being probed ($\sim500$~s in this case). The blue dotted line and open squares show the results for high counts per cycle, low number of measured cycles (corresponding to AGN), while the red dotted line and filled circles show the low counts per cycle, large number of cycles case (which is more similar to BHXRBs). Note the excellent match of the simulated distribution of lags with the analytical error estimate, over a wide range of lag errors. The solid black lines show the slopes expected for lag signal-to-noise scaling linearly with count rate (low rate, many cycles case) and with the square-root of count rate (high rate, few cycles case), expected for the two signal-to-noise regimes. See text for details of the simulations.}
\label{lagerror}
\end{center}
\end{figure}
It is interesting to note that the two signal-to-noise regimes considered here show different dependences of the lag error on the flux, also highlighted by the solid black lines which show the corresponding functional forms: the high-rate (per cycle), few cycles regime shows the error scaling with $1/\sqrt(flux)$, as is familiar from conventional spectroscopy. However, the low-rate (per cycle), many cycles regime shows the error scaling with $1/flux$. To understand this behaviour, we recall that the phase-lag error depends simply on the observed `raw' coherence. If we consider the case where the intrinsic coherence is unity, it is easy to show that the raw coherence is given by:
\begin{equation}
\label{eqn:simplecoherence}
\gamma^{2}(\nu_{j}) = \frac{(P_{X}(\nu_{j}) - P_{X,\rm noise})(P_{Y}(\nu_{j}) - P_{Y,\rm noise})}{P_{X}(\nu_{j}) P_{Y}(\nu_{j})}
\end{equation}
which simplifies to:
\begin{equation}
\gamma^{2}(\nu_{j}) = \left[\left(1+\frac{P_{X,\rm noise}}{P_{X, \rm signal}}\right)\left(1+\frac{P_{Y,\rm noise}}{P_{Y, \rm signal}}\right)\right]^{-1}
\end{equation}
where we define the intrinsic signal power in the channel-of-interest as $P_{X, \rm signal}= P_{X}(\nu_{j}) - P_{X,\rm noise}$ and similarly for the reference band. Thus Equation~\ref{eqn:lagerror} can be expressed as:
\begin{equation}
\label{eqn:simplelagerror}
\Delta \phi(\nu_{j}) = \sqrt{\left(\frac{P_{X,\rm noise}}{P_{X, \rm signal}}+\frac{P_{Y,\rm noise}}{P_{Y, \rm signal}}+\frac{P_{X,\rm noise}P_{Y,\rm noise}}{P_{X, \rm signal}P_{Y, \rm signal}}\right)/2M}
\end{equation}
Therefore the error on the lag depends on the ratio of the Poisson noise level to the intrinsic variability power in the frequency range $\nu_{j}$. Moreover, the error contains terms within the square-root which are linear in this ratio, and a non-linear, squared term. When the linear terms dominate the lag error, the signal-to-noise of the lag measurements scales with the square-root of the count rate, since (in fractional rms-squared units), the Poisson noise level scales inversely with count rate.
If we assume that the reference band signal-to-noise is always larger than that of the channels-of-interest, it is easy to see that the non-linear term dominates the error when $\frac{P_{Y,\rm noise}}{P_{Y, \rm signal}}\gg 1$. This situation will occur when there are few photons per variability time-scale in the reference band, and/or the rms-amplitude of variability is small. In particular, the former condition is typically satisfied when we observe X-ray binaries at time-scales where we expect reverberation signatures to dominate the lags. In this regime, it is easy to see that the signal-to-noise of any lag measurements scales linearly with count rate, as shown from our simulations. Thus, large improvements can be gained by studying brighter sources or using X-ray detectors with larger effective area. This point will be important when we consider the future of reverberation studies, in Sect.~\ref{sec:future}. It is also interesting to note that Equation~\ref{eqn:simplelagerror} implies that in the AGN case where the linear terms inside the square-root dominate the error, the error on the lag measured over an equivalent {\it mass-scaled} frequency range (for the same exposure) is independent of black hole mass. This is because, for a source with the same mass-scaled PSD, the PSD amplitude scales increases linearly with decreasing frequency, while the number of cycles measured in a fixed exposure decreases, so that the two changes cancel out. AGN with very massive black holes will remain difficult to study however, since the frequency range where reverberation lags are expected will be shifted to time-scales longer than the exposure time (e.g. see \citealt{demarco13}).
\begin{figure}
\begin{center}
\includegraphics[width=7cm,angle=-90]{phasedisthist.ps}
\caption{Phase lag distributions for the $10^{4}$ simulations carried out for five different count rates in the high rate, few cycles regime (10, 1.3, 0.12, 0.01, $8\times10^{-4}$ count~s$^{-1}$, corresponding to successively broader distributions).}
\label{lagerrordist}
\end{center}
\end{figure}
We can use our simulations to consider the limitations of the standard lag error formula. Fig.~\ref{lagerror} shows that the error formula performs extremely well in both signal-to-noise regimes over a wide range of count rates. However, it starts to break down for large errors. This effect is linked to the fact that the phase lag is bound to lie between $-\pi$ and $\pi$ and thus the Gaussian shape of the distribution of measured lags breaks down for large errors. We demonstrate this effect in Fig.~\ref{lagerrordist}, which shows how for small lag errors, the distributions are close to Gaussian, but for large errors the lags start to wrap around in phase until at the lowest count rates the phase lag distribution is completely uniform between $-\pi$ and $\pi$. This limit corresponds to the limiting value of the lag errors, seen in Fig.~\ref{lagerror}, of 171~s (which corresponds to a phase lag of $\sim 0.34\times 2\pi$, or $1\sigma$). Of course, in this situation, the lags are completely undetermined. Note that the analytical lag error reaches this limit more quickly because of the bias term that must be subtracted from the numerator of the coherence (see Sect.~\ref{sec:coherr}). The subtracted term is the {\it expectation value} of the bias on the modulus-squared of the cross-spectrum. This bias is distributed as a $\chi^{2}_{2}$ distribution and hence is positively skewed with a median value less than the expectation value, leading to coherence which is frequently negative and so cannot be used for lag estimation (hence the error is set to the limiting value).
The simulations show that the phase lag distribution starts to become significantly non-Gaussian for phase lag errors exceeding $\Delta \phi\simeq 0.75$. This limit should apply regardless of the signal-to-noise regime or observed count rates. Thus, care should be taken when fitting lag-frequency or lag-energy spectra with errors this large, since the standard assumption of Gaussian errors will no longer apply. Ideally, the cross-spectrum should be fitted directly in this case (since its errors on the real and imaginary components will remain Gaussian), or simulations should be used to assess the uncertainty on fitted model parameters.
Finally we point out that there is some small scatter on the observed phase or time lag for the different realisations of the {\it underlying} light curves which are used as the basis of the simulations for each different count rate. This effect causes the point-to-point scatter around the expected smoother trends in Fig.~\ref{lagerror}, and the small differences in centroid of the lag distributions (even though the intrinsic lag between the light curves is the same), including for the cases where the distribution is close to Gaussian, as seen in Fig.~\ref{lagerrordist}. The effect arises because the red-noise PSD of the underlying variability is intrinsically noisy. Thus, for different realisations of the underlying light curves, power is redistributed differently between frequencies, causing a change in the average phase or time-lag measured over the chosen frequency range. This effect is small compared to the expected errors and if necessary can be mitigated by using narrower frequency ranges to measure the lags (e.g. as in a lag-frequency spectrum), or averaging over many cycles of variability. It is also important to bear in mind that, provided that the intrinsic variations in different bands are well-correlated (i.e. intrinsic coherence is close to unity), errors such as this are {\it systematic} and apply a similar fractional shift to the lags measured at all energies. Thus the shape of the lag-energy spectrum should not be changed by this effect.
\subsection{Introducing the impulse response}
\label{sec:imprespintro}
Finally, as a prelude to the discussion of more detailed models in Sect.~\ref{sec:modelling}, and also a taster for the consideration of the actual physical `meaning' of the observations presented in the next section, we briefly consider how timing behaviour can be interpreted in terms of the {\it impulse response}, which enables us to link models for variability with models for the emission, and connect these models to the observed spectral-timing properties.
\subsubsection{Basic concepts}
Consider continuous light curves measured in two bands, $x(t)$ and $y(t)$. Let us imagine that the variations in both bands are driven by the same underlying variable signal, which is described by a time-series $s(t)$. The variable signal does not necessarily emit radiation itself, e.g. it may correspond to fluctuations in mass accretion rate which themselves drive variable emission. On the other hand, the variable signal may correspond to variations of some driving continuum source (e.g. in the case of reverberation). Depending on the physics of the variability process and the emission mechanism, the emission which we see in each band may be related to the underlying driving signal by a linear impulse response, which represents the response of the emission in that band to an instantaneous flash ({\it i.e.} a delta-function impulse) in the underlying driving signal. Thus, if the impulse responses of bands $x$ and $y$ are $g$ and $h$ respectively, the signals in these bands are obtained by integrating over all time delays $\tau$:
\begin{equation}
x(t) = \int_{-\infty}^{\infty} g(\tau) s(t - \tau) \; d\tau
\label{eq:tf1}
\end{equation}
\begin{equation}
y(t) = \int_{-\infty}^{\infty} h(\tau) s(t - \tau) \; d\tau
\label{eq:tf2}
\end{equation}
I.e. the observed light curve is the convolution of the underlying driving signal time-series with the impulse response for that energy band. It is important to note that for all but the sharpest impulse responses, the effect of the convolution is not only to delay the underlying time-series, but also to smear it out on time-scales comparable to or shorter than the width of the impulse response. Thus the information about the impulse response is encoded not only in time-lags, but also in the time-scale-dependent amplitude of variability in each band. The impulse response itself depends on the physical process for variability and emission. Many emission mechanisms will incorporate delays in the observed emission. For example a delta-function `flash' of seed photons which are upscattered by thermal Comptonisation will be delayed and smeared out by the time-taken to be scattered and escape the scattering region, with escaping higher energy photons subject to longer delays since they undergo more scatterings. The propagation of signals through the accretion flow will lead to much longer delays associated with the radial (viscous) propagation time through the flow.
\subsubsection{A simple example: reverberation from a spherical shell}
\label{sec:revsphershell}
In reverberation scenarios, we have a driving continuum light curve that irradiates the accretion disc leading to reflected emission. The delay is primarily set by the light travel time between the source of the irradiating emission and the location of the reflected emission. One can build up a simple impulse response for a known geometry based on the path-length difference between the direct emission from the driving light curve and the reflected emission from each reprocessed region. At a given time $\tau$ after a delta-function flare, reprocessing from a spherical shell will be seen from any region intersecting with an isodelay surface given by
\begin{equation}
\tau = (1 + \cos\theta ) r / c
\end{equation}
where $r$ is the radius from the source of the flare and $\theta$ the angle measured from the observer's line of sight. We demonstrate this for a simple spherical shell in Fig.~\ref{fig:tophat} (left panel). From this figure, it can be seen that the path length difference between the direct and reprocessed emission will be $r(1 + \cos\theta)$, and the time delay will simply be the path length difference divided by the speed of light. The region of the shell which we see emission from at a time $\tau$ after observing a flare, is the region intersected by the isodelay surface at $\tau$. Of course, in reality we do not have delta-function flares, but continuous variability, and thus our reflected light curve will be like the driving light curve but smoothed and delayed by the range of time delays possible from different regions of the sphere.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.8\textwidth]{tophat.ps}
\caption{{\it Left:} Schematic diagram showing reprocessing by a thin spherical shell of radius $R$. The path length difference between the direct emission (dotted line) and the reprocessed emission (dashed line) is $R(1 + \cos\theta)$, and hence the time delay for a given position on the sphere is $\tau = (1 + \cos\theta ) R / c$. An isodelay surface is shown in blue. {\it Right:} The corresponding impulse response is a simple top-hat function extending from the minimum delay ($\tau = 0$) to the maximum delay at $\theta = 180^\circ$, which is $\tau = 2R/c$. }
\label{fig:tophat}
\end{figure}
The impulse response for a thin spherical shell is also shown in Fig.~\ref{fig:tophat} (right panel). Assuming that the shell reprocesses the emission equally at all places, the impulse response will simply be given by a top-hat function that extends from the minimum delay ($\tau = 0$ in this case for reprocessed emission directly along the line of sight on the near side of the sphere) to the maximum delay ($\tau = 2R/c$ which corresponds to emission along the line of sight from the far side of the sphere). The reprocessed light curve, then, will simply be the driving light curve convolved with the top-hat function following Equations~\ref{eq:tf1} and \ref{eq:tf2}.
\subsubsection{The effects on spectral-timing measurements}
\label{sec:spectimeffects}
From the convolution theorem of Fourier transforms, it is easy to see that the Fourier transform of an observed light curve is equal to the Fourier transform of the driving signal multiplied by the Fourier transform of the impulse response:
\begin{equation}
X(\nu)=S(\nu)G(\nu)
\end{equation}
\begin{equation}
Y(\nu)=S(\nu)H(\nu)
\end{equation}
It is then simple to determine the effect of the impulse response on the PSD:
\begin{equation}
|X(\nu)|^{2} = S^{*}(\nu)G^{*}(\nu)S(\nu)G(\nu) = |S(\nu)|^{2}|G(\nu)|^{2}
\end{equation}
Thus, after applying the appropriate normalisation, we find that the observed PSD is equal to the PSD of the driving signal, multiplied by the modulus-squared of the Fourier transform of the impulse response. The impulse response acts as a filter on the variabilty, modifying the shape of the PSD. Energy-dependent impulse responses will result in PSDs with energy-dependent shapes. In particular, a more extended impulse response will suppress power at high frequencies. Thus, the flattening of the high-frequency PSD which is frequently observed at higher energies may indicate a sharper impulse response at these energies \citep{kotov01}.
It is also simple to understand the meaning of the cross-spectrum and phase/time-lags in terms of the impulse response:
\begin{equation}
C(\nu)=S^{*}(\nu)G^{*}(\nu)S(\nu)H(\nu)=|S(\nu)|^{2}G^{*}(\nu)H(\nu)
\end{equation}
So that after applying the appropriate normalisation, the observed, binned cross-spectrum is equal to the driving signal PSD {\it multiplied by the cross-spectrum of the impulse response}. Thus the cross-spectrum encodes detailed information about the shape of the impulse response as a function of energy and thus the phase/time lags in each energy band.
It is important to note that the impulse responses considered here and in more detail in Sect.~~\ref{sec:modelling} are all linear, in that they represent systems which respond linearly to the driving signal. Reverberation should be well-described by a linear impulse response in cases where variations in the illuminating continuum do not significantly alter the {\it shape} of the reflected spectrum (e.g. due to significant changes in the reflector ionisation state). We will briefly discuss systems which may be modelled using non-linear impulse responses in Sect.~\ref{sec:future:models}.
\section{Future Directions}\label{sec:future}
Reverberation time lags are common in X-ray bright, variable,
AGN. They are seen at soft X-ray energies and in the Fe K band. The
presence of Fe K lags is the clearest evidence that reflection
is involved in generating the lags.
Reverberation lags have now been detected in over two
dozen AGN with timescales correlated with black hole mass. The lag
times correspond to physical lengths of 2 to $20r_{\rm g}$ which means
that at least part of the corona lies within at most $10r_{\rm g}$ of
the black hole; in many cases the implied location is much
closer. Reverberation lags are also now beginning to emerge from
BHXRBs. Modelling gives a consistent picture supporting a compact
corona above the disc of a (possibly rapidly spinning) black hole
in many of the lag-detected objects. Now that X-ray reverberation has been confirmed as a physical phenomenon, the next step is to turn the
measurements of reverberation into a tool to study the close environment of accreting black holes (and in the case of XRBs, neutron stars also). This step forward will require a combination of further data, new models and model-fitting approaches, and ultimately a new generation of large-area X-ray observatories to fully exploit this powerful new tool to get close to black holes.
\subsection{Near-term improvements in methods and data}
There is considerable scope for further advances in the study of X-ray
reverberation using current instruments, particularly {\it XMM-Newton}, which
generally gives the highest count rates and longest continuous exposures, so is optimal for the work. Longer observations of the
objects which have already shown reverberation signals will improve on
the reverberation spectra and enable changes in the corona to be
tracked \citep[as seen in IRAS 13224-3809 by ][]{kara13b} and
low-frequency propagation effects to be unravelled from high
frequency reverberation. An important advance will be to systematically detect and compare the reverberation signatures in a wide range of AGN classes, for example to determine whether there are luminosity-dependent effects or systematic differences in the structure of the corona and inner accretion flow in radio-loud and radio-quiet AGN, or between AGN with distinct spectral energy distributions (which may indicate differences in the inner accretion flow structure).
The development of techniques \citep{zoghbi13_gaps} to estimate lags from non-contiguous light curves , e.g. with gaps due to the satellite orbit, is an important development which opens up the possibility of lag measurement using a range of X-ray missions such as {\it Suzaku} and the recently launched {\it NuSTAR} mission \citep{harrison13}. The hard X-ray sensitivity of {\it NuSTAR} also allows the extension of reverberation measurements in AGN to harder X-rays, which is revealing the expected reverberation signature of the reflection continuum, around the so-called `Compton hump' (e.g. \citealt{zoghbi14}, Kara et al., in prep. and see Fig.~\ref{fig:nustarlags}).
Significant advances can also be expected in the study of reverberation in BHXRBs. As noted in Sect.~\ref{sec:sensitivity}, BHXRB spectral-timing measurements at frequencies where reverberation lags are expected are currently in the regime of few photons per characteristic time-scale, where the signal-to-noise of lag measurements scales {\it linearly} with count rate. Thus a very significant advance can be made simply by observing brighter sources: a ten times brighter source can yield the same quality data as a hundred times longer exposure time! Unfortunately however, {\it XMM-Newton} EPIC-pn is limited by instrumental pile-up effects and telemetry limitations to observing sources not much brighter than 0.1-0.2~Crab, so that the crucial higher luminosities where state transitions occur, QPOs appear and jets switch on and off, are probably out of reach for studying in detail using reverberation: these studies will probably have to wait for new missions in the near and longer-term (see Sect.~\ref{sec:future:missions}).
Another promising line of research is the study of lags in neutron stars, in particular at the frequencies of the so-called `kHz QPOs', typically ranging from 600--1000~Hz. The combination of a high rms ($\sim 10$~per~cent) concentrated into a narrow frequency range makes these timing signatures especially useful to search for time-lags, with uncertainties of only a few microseconds possible, e.g. in archival data from {\it RXTE} \citep{barret13,deavellar13}. One advantage for modelling reverberation in neutron star systems is that, at least for the kHz QPOs, it is likely that the continuum emission originates primarily from the boundary layer of accretion close to the neutron star surface \citep{gilfanov03}. Thus the driving continuum source's geometry and location is probably more tightly constrained than in accreting black holes. Reverberation mapping of accreting neutron stars could ultimately yield useful constraints on the neutron star mass and radius, and hence provide a useful tool to constrain the neutron star equation of state.
\subsection{Advances in models and model-fitting}
\label{sec:future:models}
On the theory and modelling side, only the first step has been taken
into fitting the energy dependence of Fe~K lags using the simplest
disc model consisting of an X-ray point source located some height
above the black hole. In reality the X-ray corona is almost certainly
both radially and vertically extended. Including such a geometry is
clearly the next step towards building a more realistic model.
Moreover, rather than just considering the Fe~K line, including the
lags expected from the full reflection spectrum will allow for
self-consistent modelling of the lags over a much broader energy
range. Furthermore, the lag-frequency and lag-energy spectra have
thus far generally been fitted separately. Moving to an approach
involving fitting the lags as a function of frequency and energy will
provide the best constraints from the current data. More ambitious
still would involve fitting the cross spectra directly,
self-consistently matching all the properties of the light curves in
each band. Fitting the cross-spectrum will allow us to use the valuable information contained in the {\it frequency-dependent amplitudes} of variability in each band, which can encode the
smearing effects associated with more extended impulse responses. Such methods will require a merger of current X-ray spectral-fitting tools with timing techniques such as the Fourier transform, so that the required spectral-timing products can be calculated from the energy-dependent model impulse response and also correctly folded through the appropriate detector response matrices (which need to be accounted for to properly fit energy-dependent spectral-timing data).
As is also the case with conventional spectral-fitting, models for spectral-timing data will necessarily include degeneracies, e.g. the degeneracy between black hole mass and continuum source height in the simple lamppost case (Sect.~\ref{sec:model:endeptf}). More realistic models may incur other degeneracies between their larger number of free parameters, however it is important to stress that the degeneracies encountered in fitting spectral-timing data are generally different from (and in some cases orthogonal to) the degeneracies encountered in conventional spectral fitting, e.g. the source height affects the illumination pattern on the disc and hence the detailed shape of the broad iron line, which can be constrained by fitting the line shape. Therefore there is hope that the full combination of spectral and timing information will break the model degeneracies that limit the power of model fits which use only part of the spectral and timing data. Ultimately, with good enough data, a completely model-independent measure of the impulse response of the inner disc can be determined, simply by deconvolving the driving continuum (measured from a continuum-dominated band) from the reflection-band light curve. Deconvolution is a notoriously noisy process however, so this important goal will likely remain out of reach until X-ray observatories with large collecting areas are launched (see Sect.~\ref{sec:future:missions}).
While the Fe~K reverberation lags appear to be relatively consistent between objects, it is already clear from the diversity of soft lag shapes (see Fig.~\ref{lagen_stack}) that the situation in the soft band is complex, with the possibility of multiple components contributing to the soft band, including additional soft continuum components (e.g. disc thermal and/or soft optically thin emission) as well as photoionised reflection. The modelling of continuum lag components is also important for understanding the lower-frequency lags that may be linked to propagation effects in the accretion flow and corona. In principle these lags could be studied using the same impulse-response approach as the reverberation lags (e.g. following \citealt{kotov01}), although it is not yet clear as to whether a simple linear response of these continuum components can be assumed to generate the impulse response in this case.
More generally, dealing with non-linear models for spectral variability and reverberation is an important challenge which may need to be addressed if, e.g. the reflection ionisation parameter varies with illuminating flux. Flux-correlated variations in spectral shape of the primary continuum will also lead to non-linear variations in the shape of the reflection. Also if X-ray heating of the disc is significant compared to internal (`viscous') heating, the thermal reverberation of the disc will be correlated with changes in blackbody temperature and consequent non-linear variation of the disc blackbody continuum. Non-linear impulse response functions may be investigated as an approach to modelling this behaviour: they have already been explored in the time-series and signal processing literature in other fields (e.g. see \citealt{potter00}) and these approaches could be adopted for X-ray spectral-timing analysis.
It is also important to consider how to deal with intrinsic incoherence in the light curves, because this will lead to correlated `systematic' type errors between energy-bins in spectral-timing products such as the lag spectrum. The statistical errors between adjacent energy bins will likely be overestimated in cases where intrinsic incoherence causes a drop in the observed `raw' coherence used to estimate errors in the lags. However, the broader lag versus energy shape will also be more difficult to interpret in these cases, being a composite of the lag-energy behaviour of the uncorrelated spectral components. This problem could be mitigated by careful selection of a reference energy band dominated by the illuminating continuum, which reverberation signatures are likely to be well-correlated with.
\subsection{Future X-ray observatories}
\label{sec:future:missions}
The power of X-ray reverberation measurements is that they are able to map the emitting regions close to compact objects in terms of the absolute physical scale. The scales that can be mapped are sub-micro to nanoarcseconds in angular size on the sky - smaller than can ever be accessed with any X-ray imaging technology, for decades into the future. Large detected count rates are key, especially for XRBs where we have seen that signal-to-noise scales linearly with count rate. Thus the most important requirement to access these scales is a large effective area instrument, capable of measuring large count rates, combined with good energy resolution (CCD-quality or better) to probe the variable relativistically-smeared reflection.
In the near future, two X-ray instruments will enable important advances in XRB spectral-timing. The Large Area X-ray Proportional Counter (LAXPC) on the Indian {\it ASTROSAT} mission will pick up where {\it RXTE} left off, bringing a similar effective area at the Fe~K energy, and larger area at higher energies, with good time resolution \citep{paul13}. The energy resolution of the LAXPC is too coarse for detailed energy-dependent spectral-timing studies, but it could be useful for identifying broader features. NASA's {\it Neutron star Interior Composition ExploreR}, {\it NICER}, uses X-ray concentrator optics with silicon drift detectors to obtain CCD-quality resolution and a larger soft response than {\it XMM-Newton} \citep{gendreau12}. Critically, its non-imaging concentrators and detector technology mean that {\it NICER} will not suffer from the restrictions on source flux faced by {\it XMM-Newton}, so that important advances in spectral-timing can be made for bright X-ray binaries, potentially allowing the first X-ray reverberation mapping of black holes during the state transitions.
These new missions will, during this decade, provide useful steps forward in our exploration of reverberation in X-ray binaries, but do not push to significantly larger detector effective areas than previous instruments. Significant breakthroughs which exploit the full potential of X-ray reverberation will require a step up to square metres of collecting area, which will be attained by the end of the 2020s. The {\it ATHENA} mission \citep{nandra13} will increase collecting area compared to {\it XMM-Newton} by more than an order of magnitude at soft X-rays, and by a factor of $3-4$ at Fe~K energies, enabling significantly improved X-ray spectral-timing and reverberation measurements, especially in the soft band, for both XRBs and AGN. The sensitivity of {\it ATHENA} to faint soures, especially in the soft band, will allow the reverberation signal to be discovered in many more fainter objects, so reaching to a much wider luminosity range of sources than is accessible today and opening up the study of the innermost regions of a wide variety of AGN classes.
Another promising advance, if it can be exploited, is the development of non-imaging, large-area silicon drift detectors, e.g. as suggested for use on the proposed {\it Large Observatory for X-ray Timing}, {\it LOFT} \citep{feroci12}. This technology could affordably reach collecting areas two orders of magnitude larger than {\it XMM-Newton} at Fe~K and harder energies, while maintaining CCD-like energy resolution. As with the advances made by {\it ATHENA} in the soft X-ray band, such an instrument would lead to a transformational advance in the study of XRBs and AGN via reverberation and spectral-timing, by enabling Fe~K lags to be measured with extremely high precision.
To illustrate the advances which can be made in the future, we show in Fig.~\ref{fig:sensitivitycurves} the `sensitivity curves' for lag measurements with planned or proposed new missions with CCD-class or better spectral-resolution, also including {\it XMM-Newton} (EPIC-pn) and {\it NuSTAR} for comparison. The curves assume 100~ks exposure time and typical spectral and variability properties for an AGN (top panel) and a black hole X-ray binary in the hard or intermediate state (lower panel), with the lags measured in the 1--3$\times 10^{-3}$~Hz and 50--150~Hz ranges for the AGN and BHXRB respectively. The most recent instrument response matrices were used and reference bands were chosen to be 0.5--10~keV for {\it XMM-Newton}, {\it NICER} and {\it ATHENA}\footnote{For {\it ATHENA} we use the Wide Field Imager (WFI) effective area curve for both cases, since this instrument is able to observe sources with flux up to 1~Crab with minimal signal degradation due to pile-up. However, the high-resolution spectrometer (X-ray Integral Field Unit, X-IFU) will be able to observe fainter sources such as AGN, with only slightly reduced sensitivity compared to the WFI.} and 2--20~keV for {\it LOFT} and {\it NuSTAR}. For {\it NuSTAR}, the lag measurements are also assumed to include reference photons from simultaneous {\it XMM-Newton} data (otherwise the {\it NuSTAR} performance is significantly worse). The lag uncertainties are calculated assuming that the lag-energy spectrum is binned at an energy resolution which is slightly larger than the resolution of CCD-type instruments (the lag bin-size $\Delta E=0.08\times E^{1/2}$, or $\sim 200$~eV at 6.4~keV), and double that for {\it NuSTAR} (to account for the poorer energy resolution). Further information is given in the figure caption. For comparison, the expected reverberation lag-energy spectrum is shown in black (assuming a zero-spin black hole with central source height of 4~$r_{g}$, disc thermal reverberation is also included in the BHXRB case). It should be borne in mind that the relative lags between energies in the lag-energy spectrum are most important, not the absolute scale. Thus, features will only be clear in the lag-energy spectrum if the lag uncertainty is small {\it compared to the change in lag which defines the feature}. Hence, detailed Fe~K reverberation studies will remain challenging without the very large areas of a {\it LOFT}-class mission.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.7\textwidth]{sensitivity_agn.ps}
\includegraphics[angle=270,width=0.7\textwidth]{sensitivity_xrb.ps}
\caption{{\it Top:} AGN lag-sensitivity curves (and expected lag-energy spectrum in black, with arbitrary absolute lag offset), for the 1--3$\times 10^{-3}$~Hz frequency range for 100~ks exposures with {\it XMM-Newton} EPIC-pn, {\it NuSTAR}, {\it NICER}, {\it ATHENA} and a 10~m$^{2}$ area {\it LOFT}-class mission (the larger background of the {\it LOFT} mission is included in the calculation). The curves correspond to a 2--10~keV flux of $4\times 10^{-11}$~erg~s$^{-1}$~cm$^{-2}$, with continuum photon index $\Gamma=2$ and Galactic absorbing column $N_{\rm H}=4\times 10^{20}$~cm$^{-2}$. The assumed black hole mass is $10^{6}$~M$_{\odot}$, corresponding to an NLS1. However, higher-mass black holes will show the same signal-to-noise over an equivalent mass-scaled frequency range (see Sect.~\ref{sec:sensitivity}). {\it Bottom:} BHXRB lag-sensitivity curves for the 50--150~Hz range and 100~ks exposures, assuming a 10~M$_{\odot}$ black hole with a 2--10~keV flux of 0.2~Crab (1 Crab $=2.4\times10^{-8}$~erg~s$^{-1}$~cm$^{-2}$) in the case of {\it XMM-Newton} and {\it NuSTAR} and 1~Crab for the other missions which can observe significantly brighter sources without significant pileup or deadtime effects. A typical bright hard state lag-energy spectrum is shown in black. The assumed photon index is also $\Gamma=2$ and the Galactic absorbing column is $N_{\rm H}=6\times 10^{21}$~cm$^{-2}$.}
\label{fig:sensitivitycurves}
\end{figure}
It is clear that the biggest improvements between missions can be seen in the BHXRB case, where the quality of reverberation measurements substantially overtakes that of AGN for the multi~m$^{2}$ class missions. This is because of the linear-scaling of lag signal-to-noise with count rate that is seen in the XRB regime (Sect.~\ref{sec:sensitivity}). For example, although {\it NICER} has a similar effective area to the {\it XMM-Newton} EPIC-pn (which is only able to detect disc reverberation at substantially lower frequencies where the origin is more ambiguous), it does not suffer the same flux limits due to pile-up and telemetry dropouts, so is able to observe much brighter XRBs. Thus NICER should make significant inroads into the study of disc thermal reverberation at high frequencies (e.g. to determine how disc inner radius and radial temperature profile vary with accretion state). The same bright XRBs can be observed by {\it ATHENA} and the {\it LOFT}-class mission, which have significantly larger area, leading to large improvements in sensitivity. It is clear also that {\it ATHENA} and {\it LOFT}-class missions are complementary in their energy coverage for AGN and BHXRBs, with {\it LOFT} allowing detailed study of the iron line and reflection continuum, while {\it ATHENA} covers the more complex soft excess (disc thermal emission and photoionised reflection).
\subsection{Concluding remarks}
X-ray reverberation is in some sense a phenomenon discovered ahead of its time. The original assessments of detection of X-ray reverberation signals were based on time-domain methods and studies of time-dependent spectra from the next generation of instruments (e.g. \citealt{young00}). Initial approaches using these techniques did not reveal the reverberation signatures hidden in the data \citep{reynolds00,vaughan01}, which were also convolved with other time-dependent spectral variability, e.g. continuum lags on longer time-scales. However, thanks to the combination of Fourier timing and spectral techniques, we now find these signals to be within reach. Rapid advances can and should be made in modelling the reverberation signatures, but these will require the development of new ways of understanding spectral-timing data. The requirements for improvements in data are relatively easily achievable with current developments in technology and future missions: large collecting areas coupled with moderate to good spectral resolution. Given the enormous potential already demonstrated by reverberation measurements over the past few years and the first steps into modelling these signals, it seems likely that over the next two decades spectral-timing, using combined energy-dependent timing products to fit in Fourier frequency and energy space simultaneously, will replace stand-alone spectroscopy or timing as the method of choice for studying the innermost regions around accreting compact objects. These advances will break through the model degeneracies faced by studying objects in the Fourier or spectral domain alone and allow X-ray studies of the close-environments of compact objects to reach their full potential.
\section{Introduction}
Accreting black holes illuminate their surroundings, thereby making
both near and distant gas detectable. If, as is usually the case, the
luminosity varies with time, then the response from the surrounding
gas will also vary, but after a time delay due to light crossing time.
This delay or {\em reverberation} lag ranges from milliseconds to many
hours for irradiation of the innermost accretion flow at a few
gravitational radii ($r_{\rm g}=GM/c^2$) around black holes of mass $M$
ranging from 10 to $10^9 M_{\odot}$, respectively.
A prominent feature of most unobscured Active Galactic Nuclei (AGN) is
the Broad Line Region (BLR) consisting of clouds orbiting at
thousands~km/s at distances of light-days to light-months as the
black hole mass ranges from $10^6 - 10^9 M_{\odot}$. \citet{blandmckee82} showed how the resultant reverberation of the emission
lines following changes in the central ultraviolet continuum can be
combined with models for the gas velocity and ionization state of the
broad-line clouds to map their geometry via the {\it impulse response}\footnote{In Blandford \& McKee, and some subsequent optical and X-ray reverberation mapping work (including by the authors of this review), the impulse response is also called the {\it transfer function}. However, impulse response is the formally correct signal processing term to describe the time domain response of the system to a delta-function `impulse', which is what we intend here (in signal processing terminology, the transfer function is in fact the Fourier transform of the impulse response).}, which encodes the geometry to relate the input light curve to the output reprocessed light curve. Measurement of the resulting lags combined with the line velocity widths could yield the mass of the central black hole. Such work has culminated in the measurement of black hole masses for a wide range of AGN (e.g. \citealt{grier12} and see \citealt{petersonbentz06} for a review) and is now leading to the measurement of the detailed structure of the BLR, identifying its inclination to the observer as well as whether the gas is simply orbiting, outflowing, inflowing or some combination of all three \citep{bentz10,pancoast11,pancoast13,grier13}.
Emission lines occur from the innermost accretion flows in the X-ray
band, produced by the process of X-ray reflection (see \citealt{fabian10} for a review). The term reflection here means backscattered and
fluorescent emission as well as secondary radiation generated by
radiative heating of the gas. The primary emission is usually a
power-law continuum produced by Compton-upscattering of soft disc
photons by a corona lying above the accretion disc (see Fig.~\ref{fig:disk_corona}). A
prominent emission line in the reflection spectrum is usually that of
Fe K$\alpha$ at 6.4-6.97~keV, depending on ionization state. At low
ionization this is because iron is the most abundant cosmic element
with a low Auger yield. X-ray reflection around black holes was
discussed by \citet{guilbert88}. The resulting reflection
continuum was computed by \citet{lightman88}, followed by the line
emission by \citet{georgefabian91} and \citet{matt92,matt93}. \citet{rossfabian93} showed how disc photoionisation also leads to significant reflection features at soft X-ray energies.
\begin{figure}
\begin{center}
\includegraphics[width=12cm,angle=0]{GRAVITAS_fig.eps}
\caption{Schematic diagram showing the power-law emitting corona
(orange) above the accretion disc (blue), orbiting about a central
black hole. The observer sees both the direct power-law and its
``reflection'', or back-scattered spectrum. The black hole causes
strong gravitational light bending of the innermost rays. The
reverberation signal is the time lag introduced by light-travel time differences between observed variations in the direct power-law and the corresponding changes in the reflection spectrum.}
\label{fig:disk_corona}
\end{center}
\end{figure}
\citet{fabian89} demonstrated how reflection from the inner accretion
disc around a black hole leads to the emission lines being broadened
by the Doppler effect caused by the high velocities and skewed by the
strong gravitational redshifts expected close to the black
hole. Fig.~\ref{fig:reverb_schematic} shows a more recent example using a model reflection
spectrum \citep{rossfabian05}. \citet{fabian89} also mentioned
that reverberation may be detectable within the
line wings, the broadest of which are produced at the smallest radii.
The wings should respond first followed by the line core which
originates further out. This concept was explored by \citet{stella90} and
modelled for a disc around a Schwarzschild black hole by \citet{campana95}.
1995 was also the year that the first
relativistically-broadened iron line was detected, in the AGN
MCG--6-30-15 with the {\it ASCA} satellite \citep{tanaka95}.
X-ray reverberation was modelled further for Kerr black holes by \citet{reynolds99} and by \citet{young00}, who also made predictions for the appearance of reverberation signatures with future, large-area X-ray observatories. At the time, the detection of reverberation from the inner disc was
considered to require the next generation of X-ray telescopes with
square metres of collecting area. This idea was based on the assumption that we would detect the response of the disc reflection to {\it individual} continuum flares, to directly reconstruct the impulse response. A subsequent search for reverberation using time-domain methods did not find any signals \citep{reynolds00,vaughan01}, implying that reverberation was out of reach to current instrumentation. A crucial measure for the detection of the effect is the ratio of the number of detected photons to the
light-crossing time of the gravitational radius of the source. When
considering this ratio, the typically higher brightness of
stellar-mass black holes in Galactic X-ray binaries (BHXRB) compared with AGN,
does not compensate for the $10^5$ or more fold increase in light
crossing time for the detection of reverberation lags. However, the detection of X-ray lags on significantly longer time-scales than the light-crossing time was facilitated in X-ray binaries (XRB) by the enormous number of cycles of variability (scaling inversely with black hole mass) that could be combined using time-series techniques to yield a significant detection.
\begin{figure}
\begin{center}
\includegraphics[width=10cm]{strongravity13.eps}
\caption{Relativistically-blurred reflection spectrum from an ionized disc compared with its local (unblurred) counterpart, shown as a dashed line. The reflection spectrum typically has 3 characteristic parts: a soft excess, broad iron line and a Compton hump.}
\label{fig:reverb_schematic}
\end{center}
\end{figure}
Lags from accreting stellar-mass black holes were first studied using Fourier techniques in the X-ray binary system Cyg~X-1 by \citet{miyamoto88}, although they had been observed earlier using less-powerful time-domain techniques \citep{page85}. The observed lags were {\it hard}, in that variations in hard photons lagged those in soft photons, and {\it time-scale dependent}, in that the time delay increases towards lower Fourier-frequencies (longer variability time-scales). An example of the lag-frequency dependence in Cyg~X-1 is shown in Fig.~\ref{fig:cygx1lagfreq}. Crucially, the time lags can reach 0.1~s which is much larger than expected from reverberation unless the scattering region is thousands of $r_{\rm g}$ in size. Nevertheless some interpretations of those lags did invoke enormous scattering regions and explained the spectral development in terms of Compton upscattering: harder photons scatter around in a cloud for longer \citep{kazanas97}. However, given the large low-frequency lags seen in BHXRB data obtained by the {\it Rossi X-ray Timing Explorer}, these mechanisms were considered to be unfeasible when taking into account the energetics of heating such a large corona \citep{nowak99}. To get around this difficulty \citet{reig03} and later \citet{giannios04} developed a model where the hard lags are produced by scattering at large scales in a focussed jet, which solves the heating problem, but this model suffers from other significant difficulties, not least in explaining the observed relativistically broadened reflection (see \citealt{uttley11} for a discussion).
\begin{figure}
\begin{center}
\includegraphics[width=10cm]{cyglagfreq2.ps}
\caption{Time lag (8--13~keV relative to 2--4~keV) versus frequency for a hard state observation of Cyg~X-1 obtained by {\it RXTE} in December 1996. The trend can be very roughly approximated with a power-law of slope $-0.7$, but note the clear step-like features, which correspond roughly to different Lorentzian features in the power spectrum \citep{nowak00}.}
\label{fig:cygx1lagfreq}
\end{center}
\end{figure}
Coronal upscattering models predict a log-linear dependence of time-lag versus energy, and such a dependence is {\it approximately} observed \citep{nowak99} but the detailed lag-energy dependence shows significant `wiggles' notably around the iron line \citep{kotov01}. As noted however, reverberation cannot explain the large hard lags observed at low frequencies \citep{kotov01,cassatella12}. Thus, \citet{kotov01} proposed a {\it propagation} model for the lags (later explored in detail by \citealt{arevalouttley06}), where they are interpreted in terms of the inward propagation of variations in the accretion flow through a corona which becomes hotter at smaller radii (thus harder emission is produced more centrally, leading to hard lags). Similar models where the spectrum of the emission evolves on slower time-scales than light-crossing were proposed by \citet{poutanenfabian99}, invoking the evolution of magnetic reconnection flares and \citet{misra00}, discussing waves through an extended hot accretion flow, but these models still have difficulties explaining the largest lags. In the propagation model, the delays scale with radial inflow (i.e. viscous) time-scales and hence the largest delays can be produced from relatively small radii (tens of $r_{\rm g}$ or less) where we expect coronal power-law emission to be significant. Significant support for the propagation model came from the discovery of the linear rms-flux relation in XRB (and AGN) X-ray variability \citep{uttley01}, which is easily explained by intrinsic accretion flow variability and its propagation through the flow \citep{lyubarskii97}. The observed non-linear, lognormal flux distribution in BHXRBs (and apparent non-linear variability in AGN) can also be simply explained by the propagation of variations through the flow, which multiply together as they move inwards \citep{uttleymchv05}.
While significant progress was being made in the understanding of BHXRB X-ray variability, work was also underway to study lags in AGN X-ray light curves using Fourier techniques. The first time-scale dependent hard lags at low frequencies were discovered by \citet{papadakis01}, with similar lags found in several other AGN \citep{vaughanfn03,mchardy04,arevalo06,markowitz07,arevalo08}. These AGN hard lags showed a time-scale and energy dependence consistent with those seen in BHXRBs, suggesting a similar physical origin. Prompted by the similarity with BHXRBs, \citet{mchardy07} compared the lag-frequency dependence of the AGN Ark~564 over a broad range of Fourier frequencies with that seen in BHXRBs, finding evidence for a step-change in the lags rather than a continuous power-law dependence of lag and Fourier-frequency, similar to what is seen in hard and intermediate state BHXRBs, and suggestive of different processes producing the lags on different time-scales. The high-frequency lag was short and negative, i.e. a soft lag, and was suggested to be caused by reprocessing by the disc. However, the small negative lag was detected below $3\sigma$ significance ($-11\pm4$~s) and this initial hint at the existence of soft lags was only briefly remarked upon and not subsequently followed-up.
The first robust ($>5\sigma$) detection of a short (30~s) soft, high-frequency lag was seen in a 500~ks long observation of 1H0707-495 and was identified as reverberation from the photoionised disc \citep{fabian09}. This source, a narrow-line Seyfert 1 galaxy, has prominent, bright, reflection components in soft X-rays where the count rate is high, providing good signal-to-noise to detect the reverberation signal. The lag timescale is appropriate for the light crossing time of the inner disc of a few million $M_{\odot}$ black hole. This discovery led to a flurry of detections of soft, high-frequency lags in AGN and crucially also the first Fe~K$\alpha$ lags \citep{zoghbi12}, providing a clear signature of reverberation from reflected emission. Following a detailed discussion of methods and an introduction to the impulse response in Sect.~2, we discuss these discoveries, including the evidence for reverberation in BHXRBs, in Sect.~3. Results are presented for lags both as a function of Fourier frequency and energy.
Sect. 4 explores models for the lag behaviour encoded in
the impulse response. We consider the impulse response for a source
situated above an accretion disc. Initially the source is assumed to
be point-like, then we outline the expected behaviour of extended
sources. The 2D behaviour of the impulse response in energy and time
is explained. After a brief summary in Sect.~5, we explore future directions for research on time lags due to reverberation around black holes in Sect.~6. Reverberation is a
powerful tool with which the geometry of the inner accretion flow and
its relation to the primary X-ray power source can be studied. In some
sources it is already revealing the behaviour within the innermost
regions at a few $r_{\rm g}$ of rapidly spinning black
holes. Reverberation techniques will allow us to understand the inner
workings of quasars, the most powerful persistent sources in the Universe, and in BHXRBs uncover the changes in emitting region and accretion flow structure associated with jet formation and destruction.
\section{Modelling}\label{sec:modelling}
In this section we describe the frequency and energy dependence of both simple and more realistic {\it linear} impulse responses, which is important when comparing with the observed lag properties described in Sect.~\ref{sec:obs}. We first discuss the frequency dependence expected from simple and more physically motivated impulse responses in Sects.~\ref{sec:freqdep} and \ref{sec:realfreqdep}, before going on to discuss energy dependence in Sect.~\ref{sec:endep}.
\subsection{Frequency dependence of lags: simple impulse responses}\label{sec:freqdep}
We described in Sect.~\ref{sec:analysis} the Fourier techniques to
estimate the time lag between two light curves. The results of such
analyses are time lags as a function of Fourier frequency and energy
(if multiple energy bands are used). The lag between the variability
in the two energy bands is a function of the frequency of the
underlying variability; i.e. the measured lag may be different for
variability occurring over the longest timescales (the low
frequencies) and for the most rapid variability (the high
frequencies). This may be because variability on these different
timescales and the communication of this variability from one energy
band to another is driven by different underlying physical processes,
or simply because the communication of the variation from one band to
the other occurs over a sufficiently long timescale that it is
blurring out the most rapid variability. In the case of reflection and reverberation, this communication is the passage of the radiation from the corona to the
reflector. We can understand the frequency dependence of the lags in terms of the shape of the impulse response introduced in Sect.~\ref{sec:imprespintro} and its effects on the Fourier transforms in each band, because the lags are directly determined by the cross-spectrum of the impulse response in each band (Sect.~\ref{sec:spectimeffects}). In the simple examples discussed below, we assume that the driving continuum band has a simple $\delta$-function impulse response at t=0. We then examine the effects on the lag-frequency spectrum (and PSD) of different aspects of the impulse response in the `reflection' band, before summarising these effects and considering their implications for explaining the observations.
\subsubsection{Simple delta-function: time-shift and phase-wrapping}
To understand the basic properties of lags between two light curves we first consider the lag-frequency spectrum between two light curves shifted in time by 1000~s, i.e. besides the driving continuum delta-function at t=0, the impulse response of the responding reflection band is also a delta-function, but at a delay $\tau_0 = 1000$s. The resulting lag-frequency spectrum is shown as the black curve in Fig.~\ref{fig:simplelagfreq}, left panel. The lag-frequency spectrum shows a constant lag of $\tau$ across all frequencies (variability at all frequencies is delayed by the same time between the two light curves) until the lag time $\tau$ corresponds to a half-wave shift in phase (a phase shift by angle $\pi$) of the Fourier mode of frequency $\nu=\frac{1}{2\tau_0}$. At this point, the waveform could have been shifted either backwards or forwards by half a wave and since the phase of the wave is defined to be in the range $-\pi$ to $\pi$, the phase wraps around from $\pi$ to $-\pi$ and the measured lag becomes negative (so-called `phase-wrapping' which is a type of aliasing; the effect is similar to the
`wagon-wheel effect' sometimes seen when wheels rotate in movies). If we consider the behaviour in terms of the Fourier transform, a Dirac delta-function at t=0 has a Fourier transform of unity at all frequencies. Adding a time-shift $\tau_0$ and applying the shift theorem of Fourier transforms, multiplies this constant value by $\exp(i\omega \tau_0)$, where $\omega=2\pi \nu$ is the angular frequency\footnote{Formally, a time shift multiplies the Fourier transform by $\exp(-i\omega \tau_0)$, but here and throughout we use the convention that a positive phase lag corresponds to a delay, so we correct to this convention by multiplying the imaginary part of the Fourier transform (and hence the resulting phase lags) by -1.}. In the complex plane, as frequency increases the corresponding vector rotates around the origin at a constant rate with frequency, causing successive crossings of the $\pi/-\pi$ boundary (i.e. positive-negative lag oscillations) at every interval of $\nu = \frac{n}{2\tau_0}$, where $n$ is an odd integer.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.45\textwidth]{lagfreqdelfcomp.ps}
\includegraphics[angle=270,width=0.45\textwidth]{lagfreqdfdiltau0comp.ps}
\caption{{\it Left:} The lag-frequency spectra between two light curves related by a simple time-shift of 1000~s (i.e. delta-function impulse responses), for differing amounts of dilution of the lagging `reflection' light curve by a zero-lag direct component. The relative amplitude of the reflection relative to the diluting direct continuum is $R$. Phase-wrapping starts to occur at $\nu = \frac{1}{2000} = 5\times10^{-4}$~Hz. {\it Right:} The lag-frequency spectra for a diluted ($R=1$) time-shifted light curve with shift $\tau_0=1000$~s (solid line), compared to an undiluted time-shifted light curve with $\tau_0=500$~s (dashed line). Although the amplitude of the low-frequency lag is the same, the phase-wrapping frequencies and resulting high-frequency lags are distinct.}
\label{fig:simplelagfreq}
\end{figure}
\subsubsection{Dilution}
{\it Dilution} is caused by the presence of emission from the driving continuum in the reflection band (which is expected in most physical scenarios for X-ray reverberation). One of the main effects is to reduce the amplitude of the lags, but the shape of the lag-frequency spectrum can also be significantly changed. In the time-shifted light curve case, the direct continuum component simply adds a zero-lag delta-function (representing the direct continuum) to the impulse response of the reflection band. If we define the relative amplitude of the reflected flux to the direct flux in the reflection band to be $R$, such that $R=1$ represents equal contributions of reflected and direct flux, it is easy to show that the phase lag is given by:
\begin{equation}
\phi(\omega) = \arctan\left(\frac{R\sin(\omega \tau_0)}{1+R\cos(\omega \tau_0)}\right)
\end{equation}
The resulting time lags (i.e. $\phi/\omega$) for different values of $R$ are shown in Fig.~\ref{fig:simplelagfreq}. Note that the observed lag at low frequencies is reduced, by a factor $R/(1+R)$, but that the zero-crossing point for all $R$ remains the same, so that this zero-point may be used as an indicator of the true lag of the reflected flux component (this effect is shown in Fig.~\ref{fig:simplelagfreq}, right panel). Note also that the effect of $R>1$ is to make the lag-frequency dependence convex close to the crossing points. This sharp peak gets narrower as $R$ increases and in the limit $R\rightarrow \infty$ we return to the simple delta-function case. For $R<1$ the lag-frequency dependence becomes smoother, reducing the frequency where the lags begin to drop (although the zero-crossing frequency is maintained). For $R<<1$ it is easy to show that the phase lag has a simple sinusoidal dependence on frequency.
\subsubsection{Impulse response shape and width}
In reality, the delayed component in any impulse response is unlikely to be a delta-function, since most physical models for lags produced by light-travel time predict multiple path-lengths. The lag-frequency behaviour is closely related to the delta-function case however, because an extended and lagging impulse response can be described in the time-domain by a broad function centred at zero time, convolved with a lagging delta-function. Thus the shift theorem applies and we arrive at a similar functional form for lag-frequency dependence as in the simple delta-function case. As a simple example, we first consider the case of a top-hat impulse response (see Sect.~\ref{sec:revsphershell}), with centroid lag $\tau_0$ and width $\Delta \tau$. The Fourier transform of a top-hat is a sinc function. Assuming again that $R$ is the relative amplitude of reflected to direct emission (which is modelled with a delta-function at $t=0$), it is easy to show that the phase lag is given by:
\begin{equation}
\phi(\omega) = \arctan\left(\frac{R\sin(\omega \tau_0)\mathrm{sinc}(\omega \Delta \tau/2)}{1+R\cos(\omega \tau_0)\mathrm{sinc}(\omega \Delta \tau/2)}\right)
\label{eqn:thlag}
\end{equation}
The lag-frequency dependences for $R=1$ and the same $\tau_0=1000$~s but different values of $\Delta \tau$, together with the corresponding delta-function case, is shown in Fig.~\ref{fig:tophatwidth}. The sinc function associated with the top-hat imposes an additional oscillatory structure on top of that already present due to the time-shift of the centroid. Note that in the special case where $\tau_0 = \Delta \tau/2$ the multiplying sinc function is in phase with the oscillations due to the delta-function, causing the sign of the lag to always be positive.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.45\textwidth]{thfig.ps}
\includegraphics[angle=270,width=0.45\textwidth]{lagfreqthcomp.ps}
\caption{{\it Left:} Top-hat impulse responses. All have centroid $\tau_0 = 1000$s, yet different widths of $\Delta\tau = 500$~s (red), 1000~s (green) and 2000~s (blue). All are normalized to have an area of unity. {\it Right:} The corresponding lag-frequency spectra, together with the lag-frequency spectrum for a delta-function impulse response for comparison. Dilution is included in all cases, with direct continuum flux set to be equal to the lagging component flux ($R=1$). Note how the maximum lag and the frequency where the lag first goes to zero are the same for all four impulse responses (both are set by $\tau_0$ and $R$, and only $\tau_0$, respectively).}
\label{fig:tophatwidth}
\end{figure}
We also consider a smoother impulse response for the reflection, in the form of a time-shifted Gaussian with centroid $\tau_0=1000$~s and standard deviation $\sigma=300$~s. Both types of impulse response are shown in Fig.~\ref{fig:shapedependence}, for both undiluted and diluted ($R=1$) cases. The Fourier transform of a Gaussian is also a Gaussian, which replaces the sinc function in Equation~\ref{eqn:thlag}. Since the Gaussian is real and always positive, showing no oscillatory behaviour, the lag-frequency spectrum in the undiluted case is the same as for a simple, undiluted delta-function impulse response, in contrast to the top-hat case. It is important to note that despite its smooth shape, the Gaussian impulse response shows the same sharp oscillations as seen for the other functions, simply as a result of the shifted centroid delay. Thus a sharp-edged impulse response is not a requirement for sharp oscillatory lag-frequency behaviour.
The 300~s standard deviation of the Gaussian is chosen so that the lag-frequency behaviour in the diluted case closely mimics that of the diluted top-hat impulse response with width $\Delta\tau = 1000$~s. The only significant difference between the lag-frequency spectra occurs in the high-frequency oscillations, which are heavily suppressed in the Gaussian case.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.45\textwidth]{lagfreqshapecomp.ps}
\includegraphics[angle=270,width=0.45\textwidth]{psdshapecomp.ps}
\caption{{\it Left:} Lag-frequency spectra for different impulse response shapes and dilution. The Gaussian impulse response case is shown using dashed lines. {\it Right:} The modulus-squared of the Fourier transforms of the top-hat and Gaussian response functions, which shows the `filter' which the driving signal PSD should be multiplied by to yield the output PSD from the reflection band.}
\label{fig:shapedependence}
\end{figure}
\subsubsection{Effects on the PSD}
So far we have only considered the effects of the impulse response on the lag-frequency spectrum. However, as noted in Sect.~\ref{sec:spectimeffects}, the impulse response also acts as a filter on the driving signal, multiplying its PSD by the modulus-squared of the Fourier transform of the impulse response and thus potentially affecting the shape of the observed PSD. We show the shape of this filter in Fig.~\ref{fig:shapedependence} (right panel), for the different impulse responses considered in the previous subsection. It is useful to note that a simple delta-function shows a filter that is unity at all frequencies (i.e. the observed PSD is the same as that of the driving signal, as we would expect). At low frequencies, all the considered impulse responses also show a unity filter effect, which drops at higher frequencies (the impulse responses act as {\it low pass filters}, suppressing rapid variability). However, more complex shapes are seen at higher frequencies where the undiluted top-hat and Gaussian cases simply correspond to the modulus squared of a sinc function and zero-centred Gaussian respectively. The most interesting cases are the diluted top-hat and Gaussian impulse responses, which show a prominent dip in the filter around the frequency where the lag oscillations occur. This arises as a `destructive interference' effect, since at this frequency the reflected lagging signal is $\pi$ radians out of phase with the input signal which is produced in the same band by the `diluting' delta-function, so the two signals cancel each other out. At higher frequencies the filter flattens to an approximately constant value equal to $(1+R)^{-2}$, which is due to the direct continuum component (the reflected lagging component is almost completely filtered out).
\subsubsection{Summary and comparison with observations}
We first summarise the key results from our examination of simple impulse responses:
\begin{enumerate}
\item The centroid time ($\tau_0$) of the lagging impulse response component sets the frequency where the lag drops to reach zero ($\nu=1/2\tau_0$). The centroid (together with dilution) also sets the value of the low-frequency constant time lag.
\item The dilution of the lagging (`reflected') component by the direct continuum (defined by $R$, which is the reflected flux expressed as a fraction of the direct continuum flux in the same band), causes a reduction in the low-frequency constant time-lag, by a factor $R/(1+R)$. Dilution also affects the shape of the high-frequency drop to zero lag and subsequent oscillatory behaviour, but does not change the frequency of the drop to zero lag from the value set by $\tau_0$.
\item The drop and oscillatory behaviour seen in the lags at high frequencies is a consequence of the time-shift of the lagging component and not an artefact of sharp edges in the impulse response: the same effect is also seen for smooth impulse responses. However, the shape of the impulse response does affect the more detailed shape of the lags and oscillatory behaviour close to and above the initial frequency set by $\tau_0$.
\item The impulse response shape has strong high-frequency filtering effects on the PSD and can also lead to interference `dips' in the power, e.g. at $\nu=1/2\tau_0$.
\end{enumerate}
Observationally it is important to distinguish variations in lag as a function of frequency which are associated with the time-shift and associated oscillatory effect of a single impulse response, and those which are due to a change in the lag mechanism (in fact, the latter could be modelled in terms of two different impulse responses with different shapes, centroids and widths). A key point to note is that oscillatory effects make the lag change sign across a narrow, factor~2 range in frequency, which is significantly narrower than the negative lag ranges observed in many AGN. Nor do we see any clear evidence for interference dips or oscillatory behaviour in the observed PSDs. Thus the switch from hard to soft lags at high frequencies observed in 1H0707-495 and other AGN is better explained in terms of two different lag mechanisms (i.e. propagation to reverberation), and this inference is supported by the different lag-energy spectra observed at low and high frequencies (see Sects.~\ref{sec:smallscale} and \ref{sec:tophat_en}).
It seems most likely that the complex effects of the impulse response should be looked for at even higher frequencies, where a single process such as reverberation dominates. E.g. many measurements of lag-frequency spectra show that the high-frequency lags are suppressed at even higher frequencies, potentially giving an indication of the centroid time $\tau_0$ of the reflector impulse response, which is consistent with the observed lag amplitudes and expected dilution. The associated oscillations and expected features in the PSD might therefore be expected to occur at higher frequencies than can be currently probed. Since it is the detailed high-frequency behaviour of the impulse response which contains information on its shape, it is unlikely that we can probe it in detail without better data, but we can easily measure the centroid time-shift and compare that with expectations from more physically realistic models.
Another important effect is dilution, which occurs when there is a contribution from both the direct and reflected components in the light curve in a given energy band. In practice this is always the case, since although the `soft' band may be dominated by X-rays reflected from the disc, it always contains a significant contribution from X-rays seen directly from the corona. For instance, in the X-ray spectrum of 1H\,0707-495 measured by \citet{fabian09}, up to one third of the photons in the soft `reflection' band are direct continuum photons (i.e. $R=2$ in this case). As the measured lag is the average time lag between the correlated variability in the two energy bands, the effect is to `dilute' the measured lag time. This dilution is considered by \citet{kara13a} and \citet{wilkins13} who show that the measured time lag can be reduced by up to 75 per cent once the contributions of the diluting component are considered in both the reflection- and continuum-dominated energy bands.
\subsection{Frequency dependence of lags: realistic impulse responses}
\label{sec:realfreqdep}
Next we consider the effects associated with more realistic impulse responses for reflection, generated using relativistic raytracing, before examining the impact of realistic extended coronal geometries and the first steps towards energy-dependent modelling of the lags.
\subsubsection{The impulse response for an X-ray source above an accretion disc}
\label{sec:enavgtf}
We now introduce the impulse response for an X-ray source above an accretion disc. As we are discussing an accretion disc around a black hole an appropriate impulse response must take into account the relativistic effects (including Shapiro delay) that will be present around a black hole. The time-averaged response of the disc is well explored, as we discussed in the Introduction. The time-resolved response has also been well studied \citep{campana95,reynolds99,gilfanov00, kotov01,poutanen02,nayakshin02,cassatella12,chainakun12,wilkins13,cackett13_ngc4151,emma14}.
The time-resolved response represents the flux of reflected photons received from the disc as a function of time after an initial, instantaneous flash of radiation from the primary source, and was first calculated by \citet{campana95}. These authors considered both the cases of the point source lying in the plane of the disc and a uniform spherical source extending up to the inner edge of the accretion disc, accounting for relativistic effects on the energy and flux of the reflection as a function of position in the disc, but adopting a classical approximation for the light travel time. They compute the flux seen in relativistically broadened emission lines from the disc as well as the width and centroid energy of the line as a function of time after the initial flash.
A full treatment of the illumination of an accretion disc around a black hole in general relativity was first completed by \citet{reynolds99} and \citet{young00} who model, in detail, the time and energy evolution of an emitted fluorescence line from the accretion disc due to an instantaneous flash of X-rays from the corona, in order to identify observable signatures from the reverberation of a bright flare of emission from the corona. Recently, in order to interpret X-ray reverberation measurements in terms of X-rays emitted from the corona reflecting off the accretion disc, \citet{wilkins13} conducted general relativistic ray tracing simulations of X-ray reverberation scenarios, varying the height, vertical and radial extent of the irradiating X-ray source.
We show the response from the accretion disc and corresponding lag-frequency spectrum for a point-source at a height of $10~r_{\rm g}$ above a maximally-spinning black hole in Fig.~\ref{fig:lagfreq_disc}. While there are clearly several peaks in the response, it is useful to note that, to first-order, it is approximately a top-hat function. The corresponding lag-frequency spectrum therefore looks similar to the lag-frequency spectrum of a top-hat function as we described above.
In more detail, the response of the reflection from the accretion disc to an instantaneous flash of radiation from a point source shows an initial steep rise some time after the initial flash owing to the light travel time to the disc, then gradually decaying to long times as parts of the disc further from the source receive the incoming radiation at later times.
\begin{figure}
\centering
\includegraphics[width=0.9\textwidth]{tf_h10_h20_i30_amx.ps}
\caption{{\it Left:} The energy-averaged impulse response for an X-ray source at heights 10~$r_{\rm g}$ (black) and 20~$r_{\rm g}$ (blue) above a maximally spinning black hole, with the accretion disc inclined at $i = 30^\circ$ {\it Right:} The corresponding lag-frequency spectra. Note that the sign convention means that the reverberation lags here are positive, which would be the case in the measured lag-frequency spectra if the reverberation lags are hard (e.g. Fe K lagging a medium -energy continuum band). If the reflected component is observed in soft X-rays, the model lag-frequency spectrum would be inverted, according to the convention that positive lags indicate hard variations lagging soft variations. In other words, the lags at low frequencies in this figure would correspond to the observed (negative) soft lags. The drop in lag and oscillations at high frequencies thus correspond to the high-frequency suppression of the soft lags, not the switch from hard to soft lags at lower frequencies, which we attribute to a switch between a separate mechanism (probably propagation-driven continuum lags) and reverberation.}
\label{fig:lagfreq_disc}
\end{figure}
As one might expect, relativistic effects are important when considering reverberation in the strong gravitational field around a black hole \citep{shapiro64}. The first such effect relevant to timing X-rays reverberating from the accretion disc is the delayed passage of photons through the strong gravitational field. The curvature of the spacetime means that while a local observer will measure the light travelling past them at speed $c$ at any location in the field, an outside observer, further from the black hole, will perceive the light travelling through the stronger field as having been slowed down and hence its arrival will be delayed. This is due to a combination of the space being curved, increasing the proper distance the light must travel and the outside observer's time elapsing more quickly than the time measured by an observer in the stronger gravitational field.
\citet{wilkins13} find that the delay of photons travelling through the strong gravitational field close to the black hole is significant in measurements of X-ray reverberation, particularly when the X-ray source is located at low heights. Once the delays to both the reflection and directly observed continuum have been accounted for, this Shapiro delay can cause the X-ray source to appear 0.5$r_{\rm g}$ further from the black hole than it really is if it is incorrectly assumed that the photons are travelling in a flat, Euclidean spacetime, neglecting the effects of strong gravity.
The second significant relativistic effect that influences measurements of X-ray reverberation from the accretion disc is the bending of light in the strong gravitational field around the black hole. There is a secondary sharp peak in the response from the disc. This is the re-emergence of photons reflected from the far side of the accretion disc, behind the black hole that, although classically would be blocked from our view, are lensed by the strong gravitational field around the black hole into our line of sight. Not only are these photons slowed by their passage through the strong gravitational field close to the black hole, but this part of the disc is magnified through gravitational lensing, meaning that the emission in this secondary peak is enhanced, as discussed by \citet{reynolds99}. The source of these re-emerging photons becomes apparent in a ray-traced image of the accretion disc around a black hole. To an observer on one side of the black hole, the far side of the accretion disc appears to be warped appearing above the black hole (Fig.~\ref{fig:bhdisc}). It is these photons that are delayed and redshifted, giving rise to the secondary peak in the impulse response.
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{bh_disc.eps}
\caption{Ray-traced image of the accretion disc around the black hole. To an observer on one side of the black hole, the far side of the accretion disc appears to be warped, appearing above the black hole as emitted radiation is bent around the black hole. Shading corresponds to the flux seen from each part of the disc, with a darker colour indicating that both the photon energy and arrival rate is redshifted.}
\label{fig:bhdisc}
\end{figure}
Finally, each ray that travels from the disc to the observer will be influenced by gravitational redshifts as well as Doppler shifts due to the orbital motion of the material in the disc. Considering each ray to be a stream of photons emitted at regular intervals, not only are the energies of the photons arriving along a ray shifted, but so too is their arrival rate. Reflection from the approaching side of the disc is enhanced as the emission is beamed towards the observer and reflection from the receding side of the disc is reduced, though since reflection from both sides of the disc is recorded together when computing lag-frequency spectra, this effect will cancel to first order.
There is a reduction in the response seen from the innermost parts of the accretion disc as this emission is gravitationally redshifted to lower energies and photon arrival rates and since many rays emitted from the disc close to the black hole are bent towards the black hole and lost inside the event horizon. This effect, however, is offset by the enhanced illumination of the inner parts of the accretion disc by the coronal X-ray source. Rays are focussed towards the black hole and, hence, on to the innermost parts of the disc. Both of these effects play a role in the observed impulse response of the accretion disc and are computed in detailed ray-tracing simulations, however their apparent effects on the impulse response are much less visible than the Shapiro delay and re-emergence of emission from behind the black hole. The exact shape of the lag-frequency spectrum is described in detail in \citet{cackett13_ngc4151}.
\subsubsection{Beyond point-source models: extended coronae}
While it is instructive to consider the case of a variable X-ray point source above the plane of the accretion disc as a simple model to build intuition for X-ray reverberation around black holes, \citet{wilkins12} find that the illumination pattern of the accretion disc by the coronal X-ray source (the so-called emissivity profile) inferred from the profile of the relativistically broadened iron K$\alpha$ fluorescence line suggests that the corona extends tens of gravitational radii over the inner part of the accretion disc. The corona is inferred to extend to around 35$r_{\rm g}$ over the disc in the narrow line Seyfert 1 galaxy 1H\,0707-495 \citep{wilkins11,wilkins12} and to around 10$r_{\rm g}$ in IRAS\,13224-3809 \citep{fabian13}. To this end, \citet{wilkins13} also model the lag-frequency spectra that would arise from the reverberation of X-rays originating from an extended emitting region using Monte Carlo simulations of rays arising throughout an optically thin extended region, varying both its radial extent over the plane of the accretion disc and its vertical extent perpendicular to the disc plane.
The arrival of the continuum radiation from an extended corona requires careful consideration as there are now multiple ray paths from the continuum source to the observer, so variations in the luminosity of the coronal emission that are correlated with variations in the reflection do not arrive instantaneously at the observer. Furthermore, where there is an extended emission region, it is unphysical for the whole region to vary in luminosity instantaneously, rather the change in luminosity throughout the extent of the corona must propagate causally at or below the speed of light. It is possible to construct an impulse response for the corona describing the arrival of photons at the telescope after an instantaneous flash either simultaneously across the corona or propagating causally through the corona for an instantaneous injection of energy into the corona. As for the impulse responses corresponding to the reverberation from the accretion disc, this function can be convolved with a light curve representing the underlying variability in the injected energy to produce the light curve that would be observed in the continuum emission. The lag-frequency spectrum is again computed between the light curves for the continuum and reflected X-rays.
Computing the lag-frequency spectra that would be expected for a variety of coronae, \citet{wilkins13} find that the measured lag-frequency spectrum is sensitive to the vertical height and extent of the X-ray emitting region above the plane of the accretion disc. Increasing the extent of the corona above the accretion disc, increases the average light travel time of the rays between the source and the disc, increasing the lag that is measured across all frequencies. On the other hand, the broadband lag-frequency spectrum between a reflection-dominated and continuum-dominated band is not particularly sensitive to the radial extent of the X-ray source, as shown in Fig.~\ref{fig:radialextlag}. Increasing the radial extent of the X-ray source over the disc does not increase the distance from any given part of the source to the nearest position on the disc. There is, however, a slight decrease in the reverberation lag time as the radial extent increases as more of the corona is further from the black hole meaning fewer of the reflected photons have been delayed by the strong gravitational field in the centre.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{trf_radius.eps}
\includegraphics[width=0.45\textwidth]{lagspec_radius.eps}
\caption{{\it Left:} The energy-averaged impulse responses for irradiation of the accretion disc by a radially-extended corona above the plane of the accretion disc. A $2\times 10^{6}$~M$_{\odot}$ black hole is assumed, so that $r_{\rm g}/c=10$~s. The impulse responses are given for several different radial extents, and in each case the corona is only weakly vertically extended, between 2 and 2.5$r_{\rm g}$ above the disc. {\it Right:} The corresponding lag-frequency spectra, showing the relative insensitivity to the radial extent of the X-ray source. However, notice a slight decrease in the reverberation lag time as the radial extent increases as more of the corona is further from the black hole, meaning fewer of the reflected photons have been delayed by the strong gravitational field in the centre. Note that the same caveat about sign convention applies as in Fig.~\ref{fig:lagfreq_disc}. Figures taken from \citet{wilkins13}.}
\label{fig:radialextlag}
\end{figure}
The result of \citet{kara13b} that the measured reverberation lag increased when the luminosity of IRAS\,13224-3809 increased can now be interpreted in the context of the \textit{vertical} extent of the X-ray emitting corona increasing above the plane of the disc as its luminosity increases. This is not to say that the radial extent of the X-ray source is not also increasing, merely that the dominant effect is due to the increasing vertical extent of the source. It should be noted that like-for-like, the extra lag time due to increasing the vertical extent of the source is, in itself, a greater effect than that of increasing the radial extent of the source since the vertical extent of the source acts to extend the light path to the disc, while increasing the radial extent of the source decreases the effect of the Shapiro delay close to the black hole.
\subsubsection{Towards energy-dependent models}\label{sec:towardsendep}
The lag-frequency spectrum is not just useful as a probe of the corona; measured reverberation lags are sensitive to a number of parameters of the black hole and accretion disc. \citet{cackett13_ngc4151} model general relativistic impulse responses varying a variety of parameters including the height of the illuminating X-ray point source, the mass and spin of the black hole, the inclination angle at which the accretion disc is seen to the line-of-sight and the relative flux seen in the directly observed continuum compared to that in the reflection. The effects of each of these parameters is discussed in detail in \citet{cackett13_ngc4151} and also applied to fitting the Fe~K lag-frequency spectrum for NGC 4151 (see also Sect.~\ref{sec:ngc4151comparison}). We discuss these properties in more detail in the next section when considering energy-dependent lags from an accretion disc.
A further study fitting general relativistic impulse responses to frequency-dependent lags was performed by \citet{emma14}, who fit the lags in 12 AGN using the commonly observed soft high-frequency lags signal to constrain the reflector geometry and physical parameters. They find black hole masses consistent with masses derived independently, as well as obtaining other physical parameters such as height, inclination and spin. However, the robustness of the spin measurements is unclear given the disagreement with spectral fitting results and the implied unusual coronal geometries (e.g. 1H~0707$-$495, where \citealt{emma14} deduce a coronal height smaller than the disc inner radius implied by their low spin estimate). Furthermore the effects of several assumptions on spectral shape and reflection fraction need to be further explored for their effects on the measurements. Nonetheless, it is interesting that the observed lag-frequency spectra can be shown to be consistent with expectations from inner disc reflection and yield black hole masses consistent with independent estimates, even with simplifying model assumptions.
A key consideration in modelling the frequency-dependent lags is whether the lags at high frequencies are only produced by disc reflection, or whether other components also contribute. The variety of soft lag-energy shapes (see Fig.~\ref{lagen_stack}) already hints at the latter possibility, suggesting that the Fe~K region is the cleanest part of the spectrum for modelling of frequency-dependent lags. However, with further consideration of the energy-dependent behaviour and self-consistent modelling of frequency {\it and} energy dependence, it is clear that this can be a powerful technique to uncover physical parameters of the accretion geometry and kinematics. We now consider the modelling of the energy dependence of the lags in more detail.
\subsection{Energy dependence of lags}\label{sec:endep}
The lag-frequency spectrum helps to determine the average time lag between the direct and reprocessed emission, and hence gives basic information about the geometry and the reprocessing region. Additional information can be gained if one also considers the energies at which the reflected emission occurs as velocities in the reprocessing gas will be imprinted on the emission lines formed there. In the classical picture, an emission line from an accretion disc will form a doubled-horned profile with the parts of the disc coming towards the observer being blueshifted and the parts going away from the observer being redshifted. In the case of an accretion disc around a black hole, relativistic Doppler shifts and gravitational redshifts need to be included which further broaden and skew the line profile leading to a characteristic asymmetric line profile \citep{fabian89}. Hence, the instantaneous emission line profile at a time $\tau$ after a delta-function flare will be set by the kinematics of the region where the isodelay surface intersects the reprocessing gas. We demonstrate this in Fig.~\ref{fig:isodelay} where we show the instantaneous Fe K$\alpha$ emission line profile from an accretion disc at several times after a delta-function flare from an X-ray source above the black hole. As the line profile changes over time the lags will show an energy-dependence. Before discussing energy-dependent accretion disc impulse responses in more detail we first start by building up some intuition based on simple top-hat impulse responses.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.8\textwidth]{isodelay.ps}
\caption{{\it Left:} Isodelay surfaces for a delta-function flare from an X-ray source (X) at height 10~$r_{\rm g}$ above a black hole (filled black circle). The accretion disc is inclined at 30$^\circ$ to the observer. {\it Right:} Fe~K$\alpha$ line profiles at times corresponding to the isodelay surfaces shown in the left panel.}
\label{fig:isodelay}
\end{figure}
\subsubsection{Simple energy-dependent top-hat impulse responses}\label{sec:tophat_en}
Previously we have discussed how a top-hat impulse response leads to frequency-dependent lags. Let us now consider a top-hat impulse response that has some dependence on energy. The energy-dependence can come in several forms. Firstly one can simply change the relative fractions of direct and reprocessed emission at each energy. While the lag-frequency spectrum at each energy will have the same overall shape, the dilution effects will mean that the overall lag normalization will change between energies.
The energy-dependence of such a model was considered by \citet{kara13c}, for comparing the data with models presented by \citet{lancemiller10} and \citet{legg12} which explain the low-frequency hard lags in terms of extended reflection, while the high-frequency soft lags are an artefact of the Fourier analysis. \citet{kara13c} considered a top-hat impulse response with a width of $\Delta \tau$, and centered at $t_0=1500$~s with $\Delta \tau=1000$~s and 50\% dilution. They then allowed the reflection component to become steadily more dominant with increasing energy. The resulting reflection spectrum, lag-frequency spectrum and lag-energy spectrum (at two different frequencies) are shown in Fig.~\ref{fig:kara13c_tophat}. The low-frequency lag-energy spectrum shows a steady increase in lag with energy because of the decreasing dilution with energy. A high-frequency lag-energy spectrum taken over the frequency range where the phase wraps and the lags become negative shows the opposite behavior. It is essentially a mirror image of the low-frequency lag-energy spectrum, though with a lower normalization. The mirror image behavior when comparing the low-frequency lags with the higher-frequency negative lags is even more obvious when one artificially adds a discrete feature into the reflection spectrum (see the dotted line in Fig.~\ref{fig:kara13c_tophat}). For instance, by adding a dip in the reflected emission between 6 -- 7 keV. The lag-energy spectrum at low-frequencies then shows a corresponding dip in the lags between 6 -- 7 keV, whereas the negative high-frequency lags shows the opposite -- an increase in the lag at the same energies. This is used by \citet{kara13c} as evidence that the low and high frequency lags that have been detected in AGN have different physical origins. The fact that the low-frequency lags show a monotonic increase with energy while the high-frequency lags show a discrete feature around Fe K cannot readily be explained by a single impulse response, like that considered here.
\begin{figure}
\centering
\includegraphics[width=0.95\textwidth]{kara13c_tophat.ps}
\caption{The reflection spectrum (a), the lag-frequency spectrum (b) and the low and high frequency lag-energy spectrum (c) shown in blue and red, respectively for a model given by a simple top-hat response function for two test cases: (1) where the reflection fraction increases steadily with energy (solid line), and (2) where the reflection fraction increases steadily with energy except for a demonstrative dip at 6--7~keV (dotted line). Taken from \citet{kara13c}.}
\label{fig:kara13c_tophat}
\end{figure}
Another simple energy-dependent top-hat impulse response is one where the time corresponding to the center of the top-hat, $\tau_0$, changes with energy, yet the relative contribution from the direct and reprocessed emission is the same at all energies. We will consider here a top-hat where $\tau_0$ increases linearly with energy, for instance, which might be expected if higher energy photons penetrate further into the reprocessing region. Now the lag-frequency spectrum at each energy has a different frequency dependence (see Fig.~\ref{fig:tophat_endep}), and thus the frequencies at which phase-wrapping starts and the frequency where the lags become negative will be different for each energy. The low-frequency lags will show a steady increase in lag with energy. However, in a given high-frequency band each energy will be at a different stage of phase-wrapping leading to a lag-energy spectrum which switches to negative values and then increases back toward zero.
Superficially, this behaviour is somewhat reminiscent of what is seen in the observed lag-energy spectra, with lag increasing monotonically with energy at low frequencies and a more complex shape at high frequencies, with both hard and soft X-rays lagging medium-energy X-rays. However, the energy-dependent frequency switch to negative lags is not seen in the data \citep{kara13c}, and moreover, the enhanced suppression of high-frequencies seen at higher X-ray energies would lead to steeper power spectra at higher energies, not flatter power spectra, as are observed (e.g. Fig.~\ref{fig:1h0707psdcoh}). Thus, the two-component propagation and reverberation scenario for the broadband frequency-dependence of the lags is a much better match to the data, than a single energy-dependent impulse response component. However, the intuition developed from the energy-dependent top hat impulse response can be applied to understanding the lag-energy spectra within the frequency range dominated by one of these components.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{tophat_endep.ps}
\caption{Model lag-frequency and lag-energy spectra for a top-hat impulse response with a constant width but with a centroid $\tau_0$ that increases with energy. {\it Left:} The lag-frequency spectra at 2 keV (gray), 5 keV (blue) and 8 keV (green). The vertical lines indicate the frequency ranges over width the lag-energy spectra are determined. {\it Right:} The lag-energy spectra in the two frequency ranges indicated in the left panel. The vertical dashed lines mark the energies corresponding to the lag-frequency spectra on the left.}
\label{fig:tophat_endep}
\end{figure}
When we go on to consider the more complicated impulse response for an X-ray point source above an accretion disc we will encounter both types of effect discussed here. Dilution changes as a function of energy (the relativistic Fe K line has an asymmetric shape), and furthermore both the width and centroid of the impulse response changes as a function of energy, therefore phase-wrapping occurs at different frequencies in different energy ranges. As we will demonstrate, this leads to clear features in the lag-energy spectrum that change with frequency.
\subsubsection{The energy-dependent impulse response for an X-ray source above an accretion disc} \label{sec:model:endeptf}
We introduced the energy-averaged impulse response for an X-ray source above an accretion disc in Sect.~\ref{sec:enavgtf}. Now let us consider the energy-dependence of the impulse response. Here we show the general relativistic impulse responses as calculated by \citet{reynolds99}. In these models the geometry is assumed to be a simple lamppost model where an X-ray point source is at some height $h$ above a black hole, with the accretion disc inclined at angle $i$ to the observer ($i = 0^\circ$ corresponds to face-on). In Fig.~\ref{fig:fek_gr_tf} we show four representative impulse responses for different combinations of black hole spin, height of the X-ray source and source inclination\footnote{See also \url{http://stronggravity.eu/public-outreach/animations/reverberation/} for further examples and explanatory animations by M. Dov\v{c}iak.}.
\begin{figure}
\centering
\includegraphics[width=0.475\textwidth]{plottf_h10_i30_a01.ps}
\includegraphics[width=0.475\textwidth]{plottf_h10_i30_amx.ps}
\includegraphics[width=0.475\textwidth]{plottf_h10_i60_amx.ps}
\includegraphics[width=0.475\textwidth]{plottf_h20_i30_amx.ps}
\caption{Impulse responses for an X-ray source located at a height above a black hole \citep[after][]{reynolds99,cackett13_ngc4151}. Each diagram indicates a different combination of black hole spin, $a$, height of the X-ray source, $h$, and inclination angle, $i$. In each diagram the main panel shows the impulse response, the bottom panel shows the energy-averaged response of the Fe K$\alpha$ line, while the right panel shows the time-averaged Fe K$\alpha$ line. Top left: $a = 0.1$, $h = 10~r_{\rm g}$, $i = 30^\circ$. Top right: $a = 0.998$, $h = 10~r_{\rm g}$, $i = 30^\circ$. Bottom left: $a = 0.998$, $h = 10~r_{\rm g}$, $i = 60^\circ$. Bottom right: $a = 0.998$, $h = 20~r_{\rm g}$, $i = 30^\circ$. }
\label{fig:fek_gr_tf}
\end{figure}
As we noted above, the energy-averaged response of the entire Fe K$\alpha$ line is, to first-order, approximately a top-hat function. If we then look at the energy-dependence across the Fe K$\alpha$ line we see that both the centroid and the width of the response at a given energy changes across the line. Thus, as in the simple energy-dependent top-hat examples in section~\ref{sec:tophat_en}, phase wrapping will occur at different frequencies for different energies within the line. If we look at the time-averaged Fe K$\alpha$ line profile we see the familiar shape \citep{fabian89}. This means that at each energy within the line we see a different total response, and this will behave like dilution -- the line flux relative to the direct component (which will also be present at these energies) will be strongest where the line peaks, and hence we should expect the longest delays there. As the strength of the line decreases with decreasing energy we should then also expect the lags to decrease with decreasing energy as the response becomes weaker relative to the direct component.
Given that to first-order the energy-averaged response of the Fe K$\alpha$ line looks like a top-hat, it should be no surprise that the lag-frequency spectrum for the entire line looks similar to that for a top-hat function (see left panel in Fig.~\ref{fig:fekft_lagspec}). The exact shape is discussed in detail in \citet{cackett13_ngc4151}. The energy-dependent lags are less intuitive because of phase wrapping occurring at different frequencies for different energies. In Fig.~\ref{fig:fekft_lagspec} we show the lag-energy spectrum calculated over a range of different frequencies, and now follow the discussion in \citet{cackett13_ngc4151} describing the behaviour. We start by looking at the lowest frequencies. On these longest timescales (lowest frequencies) we will see the response from the entire accretion disc, and thus the lag-energy spectrum shows long lags and a double-horned profile which comes from the outer regions of the disc (see black solid line in Fig.~\ref{fig:fekft_lagspec}). As we increase the frequency we will start to look at shorter timescales and will begin filtering out the response from the outer part of the disc (the timescale will be short compared to the range of light-travel times to the outer disc). We therefore start to see a `cut-out' region in the lag-energy spectrum where the lags at energies corresponding to the outer parts of the disc are reduced (see red dotted line). As the frequency increases further that `cut-out' region gets broader, since we are now looking at increasingly shorter timescales and hence smaller regions of the disc (blue dashed line). Eventually we get to the frequency where the lags become negative (green dash-dotted line) and we see an inverted lag profile with very small magnitude. In summary, studying the lags from an Fe~K$\alpha$ line at low frequencies we see the response from the entire disc, while as we go to higher frequencies we progressively filter out the outer parts of the disc, leaving only the response from the inner regions. Notice that at all frequencies where the lag is positive the red-wing (low energy part) of the lag profile is present -- this is the response coming from the innermost regions of the disc.
\begin{figure}
\centering
\includegraphics[width=0.9\textwidth]{fektf_lagspec.ps}
\caption{{\it Left:} Lag-frequency spectrum for the entire Fe K$\alpha$ line. The colored regions correspond to the frequency ranges used to calculate the lags as a function of energy. {\it Right:} Lag-energy spectra for four different frequency ranges: $(1 - 2)\times10^{-5}~c/r_{\rm g}$ (black, solid line), $(1 - 2)\times10^{-3}~c/r_{\rm g}$ (red, dotted line), $(4 - 8)\times10^{-3}~c/r_{\rm g}$ (blue, dashed line) and $(1.7 - 4)\times10^{-2}~c/r_{\rm g}$ (green, dot-dashed line). These are all calculated for $h = 10~r_{\rm g}$, $i = 45^\circ$ and a maximally spinning black hole. Taken from \citet{cackett13_ngc4151}.}
\label{fig:fekft_lagspec}
\end{figure}
As with the lag-frequency spectrum, the lag-energy spectrum is dependent on a number of key parameters. While for the lag-frequency spectrum the continuum source height and the black hole mass were the most important parameters, when considering the lag-energy spectrum the inclination of the accretion disc is also important in determining the energy where lags are observed. Black hole spin too has an effect on the lag-energy spectrum. The exact behaviour as each of these parameters is changed is demonstrated in detail in \citet{cackett13_ngc4151}, and we just briefly comment on each parameter here. For instance, increasing the height of the X-ray source will increase the overall path length between the direct and reprocessed emission, leading to longer lags. It will also decrease the frequency at which the lags go to zero. Increasing the black hole mass will increase the overall size-scale of the system, therefore the lags will be larger and the frequency at which the lags go to zero will be lower for a higher black hole mass. Clearly, both increasing the height and increasing the black hole mass have approximately the same effect on the lag-energy spectrum and thus these parameters are somewhat degenerate -- increasing the black hole mass requires a lower height to achieve approximately the same lag-energy profile.
The inclination does not have a large effect on the lag-frequency spectrum, however, it does have a big effect on the lag-energy spectrum. As can be seen from looking at the impulse response in Fig.~\ref{fig:fek_gr_tf}, higher inclinations get a much broader energy response from the disc. This is simply because the component of the velocity along our line of sight is larger for higher inclinations, and thus Doppler shifts increase with increasing inclination. This has the most prominent effect on the blue-wing (high energy part) of the lag-energy profile (as will be familiar to anybody who has studied time-averaged iron lines). As the inclination increases the maximum energy where we see lags increases. For instance it is at approximately 6.4 keV for nearly face-on systems while for an inclination of 60$^\circ$ it will be at a little greater than 7.5 keV.
Black hole spin also affects the lag-energy spectrum. Larger black hole spin leads to an innermost-stable circular orbit that is closer to the black hole. Therefore the surface area of the disc at short lags will increase, and thus we will see more prominent lags from the red-wing. This region is also where gravitational redshifts are strongest, and therefore extending closer to the disc will decrease the minimum energy (set by the maximum gravitational redshifts) where we see lags.
\subsubsection{Comparison with NGC~4151}\label{sec:ngc4151comparison}
Finally, we demonstrate how these models can be applied to real data. \citet{cackett13_ngc4151} fit the energy-dependent lags in NGC~4151 from \citet{zoghbi12} using the GR impulse responses of \citet{reynolds99} discussed here. Given the frequency range of the observed lags and the optical reverberation mapping mass for the black hole of $M = 4.6\times10^7~M_\odot$, \citet{cackett13_ngc4151} find the X-ray source to be located at a height of $h = 7^{+2.9}_{-2.6}~r_{\rm g}$. We show their best fit in Fig.~\ref{fig:ngc4151_bestfit}. The fact that the lags are observed to drop sharply above 6.5 keV shows that the inclination $i < 30^\circ$. Larger values of $i$ would lead to significant lags above 6.5 keV. Moreover, they found that a maximally spinning black hole was a better fit than $a= 0.1$ at the 1$\sigma$ confidence level. This first exciting step into fitting the lag-energy spectra already demonstrates the power of such methods.
\begin{figure}
\centering
\includegraphics[width=8cm]{ngc4151_bestfit.ps}
\caption{Lag-energy spectrum for NGC~4151 in the frequency range $(1 - 2)\times10^{-5}$ Hz \citep[data from][]{zoghbi12}. Best-fitting GR impulse response models for $i = 5^\circ$ (black) and $i= 20^\circ$ (blue) are shown both with $h = 7~r_{\rm g}$, $M = 4.6\times10^7~M_\odot$ and $a=0.998$. The red line shows the average of the $i = 5^\circ$ model over the same energy bins as the data. Taken from \citet{cackett13_ngc4151}.}
\label{fig:ngc4151_bestfit}
\end{figure}
\section{Summary}
The first discovery of X-ray reverberation lags, in 2009, marked the beginning of a new field in X-ray astronomy that is probing the innermost regions around compact objects in a new way. These short timescale lags measure the physical distance between the compact X-ray source and the inner accretion flow, that will help us understand how these accretion processes work in the strong-field limit. The main aims of this review are to
\begin{enumerate}
\item Put reverberation lags in the historical context of previous X-ray timing studies, specifically of the low-frequency hard lags found in BHXRBs and AGN.
\item Describe the Fourier timing techniques and the spectral-timing products most often used (i.e., the lag-frequency spectrum, the lag-energy spectrum, and covariance spectrum)
\item Present some of the most significant results, including the soft lag in 1H0707-495, the Fe~K lag in NGC~4151 and the thermal reverberation in the Galactic black hole X-ray binary, GX~339-4
\item Describe the effects of the impulse response on the lags and outline the general relativistic impulse responses that can model the lag observations and give constraints on the geometry and energetics of the accretion flow.
\end{enumerate}
This review presents the progress made in X-ray reverberation lags since 2009, and we conclude now with a view for the future.
\subsection{Black hole X-ray binaries}\label{sec:bhb}
While in recent years significant work measuring reverberation has been undertaken with AGN, X-ray time lags in accreting compact objects were first studied in neutron star and black hole X-ray binaries (e.g. \citealt{vanderKlis1987,miyamoto89,vaughan94,nowak99}). The lag-studies of the accreting stellar mass black holes revealed fairly ubiquitous hard lags in most cases. The large amplitudes of the hard lags at low frequencies are difficult to explain with models invoking light-travel time delays, either due to Compton upscattering of photons in an extended corona \citep{nowak99}, or the reverberation delay of reflection from a disc \citep{cassatella12}. The first detailed study of the lag versus energy dependence in black hole X-ray binaries was carried out on {\it RXTE} data by \citet{kotov01}, who found that although there is some evidence for structure around the energies of Fe~K emission and absorption in the data, this is not consistent with the signature expected from reflection at large radii which would be required to produce the observed lags by reverberation alone. A likely alternative is that these low-frequency hard lags are associated with the propagation of fluctuations through the accretion flow and/or corona \citep{kotov01,arevalouttley06}, as may also be the case with the hard lags seen at lower frequencies in the AGN data.
\begin{figure}
\includegraphics[width=8cm,angle=-90]{gx339lagenergycomb.ps}
\caption{{\it XMM-Newton} EPIC-pn lag-energy spectra for four temporal frequency ranges for the hard state of GX~339-4. The insets show the covariance spectra for the same energy range and frequencies, plotted as a ratio to a Galactic-absorbed power-law with photon index fixed at $\Gamma=1.55$ fitted to the 3--10~keV range, to reveal the disc blackbody emission seen as a clear soft excess. The disc photons clearly lead the power-law photons at low frequencies, but the behaviour switches at high frequencies, consistent with the switch from propagation to reverberation effects. See \citet{uttley11} for further details.}
\label{fig:gx339lagvsen}
\end{figure}
Like the AGN, the first clear evidence for X-ray reverberation in black hole X-ray binaries came from the soft X-ray band, made accessible with the high-throughput of the {\it XMM-Newton} EPIC-pn instrument in its fast-readout, timing mode. The development of the covariance spectrum allowed the study of the detailed energy dependent variability of X-ray binaries in the low-hard state, which revealed that the emission of the accretion disc in that state was significantly variable \citep{wilkinson09}. This behaviour could be explained if the disc were intrinsically variable, e.g. via fluctuations of the internal viscous heating linked to changes in accretion rate, or alternatively, if the disc variability is driven by heating of the disc by the variable X-ray power-law continuum, i.e. there is {\it thermal reverberation}. The critical test of these scenarios is to measure the time-lags between the disc and power-law emission, which was done by \citet{uttley11}. The result, shown in Fig.~\ref{fig:gx339lagvsen} is very clear. At low frequencies, below $\sim1$~Hz, the disc variations {\it lead} the power-law variations, by tenths of a second, consistent with fluctuations propagating through the disc to the inner corona on a viscous time-scale, requiring that the disc, if thin ($H/R<0.1$), extends to relatively small radii, within $\sim20$~$r_{\rm g}$ of the black hole. However, at frequencies above 1~Hz, the soft photons start to {\it lag} the power-law photons (at least at intermediate energies), with lags of a couple of ms, consistent with the thermal reverberation picture, with light-travel times out to a few tens of gravitational radii at most.
So far the best lag-measurements have been limited to a handful of sources in the hard state (since other states are too bright to study with {\it XMM-Newton} EPIC-pn in timing mode), while high-frequency soft lags have been confirmed only in one source (GX~339-4) with the longest observations \citep{uttley11}. However, the behaviour seen so far is already remarkably similar to that seen in AGN. At low frequencies we see long hard lags, with substantial evidence that intrinsic disc variability is driving the power-law variations, i.e. supporting the propagating fluctuations model for the low-frequency lags. Meanwhile, at higher frequencies, the behaviour switches to short, soft lags consistent with reverberation, albeit from the disc blackbody emission and not photoionised reflection, which is to be expected since the high inner disc temperatures in BHXRBs mean that the ion species contributing to strong photoionised reflection signatures in AGN will likely be even more highly-ionised in their stellar-mass counterparts. It is worth noting here that AGN also show evidence in their X-ray/optical correlations for intrinsic disc variability on long (months-years) time-scales and thermal reverberation on short (days-weeks) time-scales \citep{uttley03,arevalooptxray08}. In fact, the variations in the optical continuum on day-to-week timescales in AGN, which show short (sub-day) red-lags (e.g. \citealt{cackett07}), can likely only be explained by thermal reverberation of the disc in response to the rapidly-varying central EUV/X-ray source. It is interesting to note that these time-scales of optical variability are the driving time-scales for optical line reverberation from the broad-line region. Thus, even optical reverberation signatures in AGN may in some sense be driven by X-ray reverberation effects!
At this point the exact mechanism for the low-frequency hard lags is unclear, but the BHXRB data firmly indicate that they are connected to intrinsic accretion variability, as posited by \citet{kotov01} and \citet{arevalouttley06}. These accretion variations were originally thought to be variations in some sort of hot, coronal flow, which would naturally lead to hard lags within the power-law continuum, if the coronal temperature increases with decreasing radius. However, it now seems likely, from both the BHXRB lag-energy spectra and the correlated long-term optical and X-ray variability in AGN, that the cool, optically thick accretion disc itself is responsible for carrying these fluctuations. A number of possible mechanisms can connect these disc variations to the coronal power-law emission. The corona may be distributed over the disc, for example, and respond directly to accretion variations in the underlying flow. However, such a model might have difficulty in explaining the short lags and strong reflection that seem to suggest a centrally concentrated corona in many of the AGN. An alternative is that the disc variations are communicated to the corona via fluctuations of the `seed' photons from the disc, which are upscattered to produce the power-law emission. The lags are not introduced by light-travel time effects since light-travel times from disc to corona should not be much larger than the observed short reverberation lags. Instead, the lags in this scenario would be caused by the relative ($\sim$viscous-time-scale) delay between variations of seed photon luminosity (as fluctuations pass through the disc) and the intrinsic coronal heating (as the accretion fluctuations finally enter the corona) which could produce correlated changes in the observed power-law index, leading to hard lags.
Finally, we note that although clear Fe~K reverberation signatures are not yet detected in X-ray binaries, we do not expect them to be seen without better data. This is because with current instruments, X-ray binaries exist in the regime of low photon signal per characteristic time-scale (see Sect.~\ref{sec:sensitivity}). However, in this regime improvements with even brighter sources can be significant, provided that those sources are not so bright that they are unobservable with the best current detector for high-resolution lag studies, the EPIC-pn. It is at least encouraging that variable broad Fe~K emission is seen in the covariance spectra in Fig.~\ref{fig:gx339lagvsen}, so that it is probably a matter of time and good fortune before Fe~K reverberation signatures are seen in black hole X-ray binaries.
|
\section{Introduction}
\label{sec:introd}
The recent discovery of a boson with mass around 125 GeV by the ATLAS \cite{Aad:2012tfa,Aad:2013wqa,AtlasConf1,AtlasConf2}, CMS \cite{Chatrchyan:2012ufa,Chatrchyan:2013lba,CMSpas}, D\O~and CDF \cite{Aaltonen:2012qt,Aaltonen:2013kxa} collaborations is the first direct hint of the electroweak symmetry-breaking mechanism. The experimental data
confirm that it is a Higgs-like scalar with couplings compatible with the Standard Model (SM) predictions. However,
this new particle could belong to an enlarged scalar sector.
In order to give mass to fermions and gauge bosons while preserving gauge invariance, the SM assumes the presence of one SU(2) electroweak scalar doublet with a non-zero vacuum expectation value. However, no fundamental principle or symmetry forbids the presence of additional scalar doublets.
The simplest extension of the SM is the two-Higgs-doublet model (2HDM)~\cite{Gunion:1989we,Branco:2011iw}, which leads to
a richer scalar sector and very interesting phenomenological implications \cite{ilisie3,ilisie2,ilisie1,Pich:2009sp,JungTuzon,PichTuzon1,Jung:2012vu,PichDstar,PichEDP,PichBll,LHCP2013,
SDG,Barroso:2013zxa,
Grinstein:2013npa,Eberhardt:2013uba,Chen:2013rba,Craig:2013hca,Coleppa:2013dya,Shu:2013uua,
Chiang:2013ixa,Krawczyk:2013jta,SK:13,IDM1,IDM2,Belanger:2013xza,Enberg:2013ara,Dorsch:2013wja,
Dorsch:2014qja,Bhattacharyya:2014nja,Altmannshofer:2012ar,Chang:2013ona,Cheung:2013rva,Enberg:2013jba}.
Generic multi-Higgs doublet models give rise to unwanted flavour-changing neutral current (FCNC) interactions, which are found to be very suppressed experimentally. The FCNCs can be eliminated at tree level by requiring the alignment in flavour space of the Yukawa matrices \cite{Pich:2009sp}. The so-called aligned two-Higgs-doublet model (A2HDM) contains as particular cases the different versions of 2HDMs with discrete $\mathcal{Z}_2$ symmetries while at the same time introduces new sources of CP violation beyond the CKM phase.
The main feature of the 2HDM is the presence of three neutral and one charged Higgs bosons.
Finding extra neutral or charged scalar bosons would be a clear signal of an extended scalar sector.
The ATLAS \cite{Aad:2012tj,Aad:2013hla} and CMS collaborations \cite{Chatrchyan:2012vca} have performed direct searches for a charged Higgs particle. However, since no excess has been found over the SM background, this only allows us to further constrain the parameter space of the various types of 2HDMs; recent analyses within the A2HDM have been performed in \cite{ilisie3,ilisie2,ilisie1}. In their searches, both collaborations assume
that the charged Higgs is produced in a top-quark decay ($t\to H^+ b$) and that it decays dominantly into fermions; {\it i.e.}, $H^+ \to q_u \bar{q}_d, \; l^+ \nu_l$. However, all experimental bounds would be trivially evaded for a fermiophobic charged Higgs, {\it i.e.}, a charged scalar which does not couple to fermions at tree level.
In order to probe such scenario, other production channels and decay rates would have to be considered.
Although such analyses have not been yet performed by the LHC collaborations, they become more compelling as the experimental bounds on a non-fermiophobic charged Higgs are getting stronger, at least in the low mass region.
The fermiophobic scenario is a simplified model that, if it turns out to be the one preferred by Nature, would allow us to measure (or at least estimate) for the first time the parameters of the scalar potential. This is usually a rather difficult task in more generic 2HDM settings.
It is also worth mentioning that a fermiophobic charged Higgs is present in the {\it inert\/} 2HDM~\cite{Ma:2008uza,Ma:2006km}, where one of the neutral scalars is a nice candidate for dark matter
\cite{Krawczyk:2013jta,
SK:13,IDM1,IDM2,Enberg:2013ara,CMR:07,Barbieri:2006dq,LopezHonorez:2006gr,ALT:09,LHY:10,LHY:11,
DS:09,Arhrib:2013ela,Ginzburg:2010wa}. The discovery of a fermiophobic $H^\pm$ particle could be interpreted in this case as an indirect signal of the presence of dark matter.
In this work, we shall focus our analysis on the search of a light fermiophobic charged Higgs $H^\pm$, with mass in the range $M_{H^\pm}\in [M_W,M_W+M_Z]$ so that only a few relevant decay modes are kinematically open.
We will study the two most important production channels for a fermiophobic $H^\pm$: associated production with either a $W^\mp$ boson or a neutral scalar.
Due to their similarity with the SM Higgs production channels, one expects them to be experimentally accessible at LHC energies. Next-to-leading order (NLO) QCD corrections will be included for both cross sections, and the bounds on the various parameters of the model from the current LHC data \cite{ilisie3} will also be taken into account.
The main features of the A2HDM are briefly presented in section 2. Section 3 discusses the calculation of the various decay rates and production modes. Finally, in section 4 we perform a phenomenological analysis, assuming different scenarios for the scalar spectrum, and conclude in section 5 with a summary of our results. Some technical details are given in four appendices.
\section{The Aligned Two-Higgs-Doublet Model}
\label{sec:A2HDM}
The 2HDM extends the SM with a second scalar doublet of hypercharge $Y=\frac{1}{2}$.
The neutral components of the two scalar doublets acquire vacuum expectation values that are in general complex,
$\langle 0|\phi_a^{(0)}(x)|0\rangle =\frac{1}{\sqrt{2}}\, v_a\, \mathrm{e}^{i\theta_a}$ ($a=1,2$), although
only the relative phase $\theta \equiv \theta_2 - \theta_1$ is observable.
It is convenient to perform a global SU(2) transformation in the scalar space $(\phi_1,\phi_2)$,
characterized by the angle $\beta = \arctan{(v_2/v_1)}$,
and work in the so-called Higgs basis
$(\Phi_1,\Phi_2)$, where only one doublet acquires a vacuum expectation value:
\begin{equation} \label{Higgsbasis}
\Phi_1=\left[ \begin{array}{c} G^+ \\ \frac{1}{\sqrt{2}}\, (v+S_1+iG^0) \end{array} \right] \; ,
\qquad\qquad\qquad
\Phi_2 = \left[ \begin{array}{c} H^+ \\ \frac{1}{\sqrt{2}}\, (S_2+iS_3) \end{array}\right] \; ,
\end{equation}
where $G^\pm$ and $G^0$ denote the Goldstone fields.
Thus, $\Phi_1$ plays the role of the SM scalar doublet with
$v\equiv \sqrt{v_1^2+v_2^2}\simeq (\sqrt{2}\, G_F)^{-1/2} = 246~\mathrm{GeV}$.
The physical scalar spectrum contains five degrees of freedom: the two charged fields $H^\pm(x)$
and three neutral scalars $\varphi_i^0(x)=\{h(x),H(x),A(x)\}$, which are related with the $S_i$ fields
through an orthogonal transformation $\varphi^0_i(x)=\mathcal{R}_{ij} S_j(x)$.
The form of the $\mathcal{R}$ matrix is fixed by the scalar potential \cite{ilisie1}, which determines the neutral scalar mass matrix
and the corresponding mass eigenstates. A detailed discussion is given in appendix \ref{app:potential}. In general, the CP-odd component $S_3$ mixes with the CP-even fields
$S_{1,2}$ and the resulting mass eigenstates do not have a definite CP quantum number.
If the scalar potential is CP symmetric this admixture disappears; in this particular case, $A(x) = S_3(x)$
and
\bel{eq:CPC_mixing}
\left(\begin{array}{c} h\\ H\end{array}\right)\; = \;
\left[\begin{array}{cc} \cos{\tilde\alpha} & \sin{\tilde\alpha} \\ -\sin{\tilde\alpha} & \cos{\tilde\alpha}\end{array}\right]\;
\left(\begin{array}{c} S_1\\ S_2\end{array}\right) \, .
\end{equation}
Performing a phase redefinition of the neutral CP-even fields, we can fix the sign of $\sin{\ta}$. In this work we adopt the conventions\ $M_h \le M_H$\ and\
$ 0 \leq \ta \leq \pi$, so that $\sin{\ta}$ is positive.
The most generic Yukawa Lagrangian with the SM fermionic content gives rise to FCNCs because the fermionic couplings of the two scalar doublets cannot be simultaneously diagonalized in flavour space. The non-diagonal neutral couplings can be eliminated by requiring the alignment in flavour space of the Yukawa matrices~\cite{Pich:2009sp}; {\it i.e.}, the two Yukawa matrices coupling to a given type of right-handed fermions are assumed to be proportional to each other and can, therefore, be diagonalized simultaneously. The three proportionality parameters $\varsigma_f$~($f=u,d,l$) are arbitrary complex numbers and introduce new sources of CP violation.
In terms of the fermion mass-eigenstate fields, the Yukawa interactions of the A2HDM read~\cite{Pich:2009sp}
\begin{eqnarray}\label{lagrangian}
\mathcal L_Y & = & - \frac{\sqrt{2}}{v}\; H^+ \left\{ \bar{u} \left[ \varsigma_d\, V M_d \mathcal P_R - \varsigma_u\, M_u^\dagger V \mathcal P_L \right] d\, + \, \varsigma_l\, \bar{\nu} M_l \mathcal P_R l \right\}
\nonumber \\
& & -\,\frac{1}{v}\; \sum_{\varphi^0_i, f}\, y^{\varphi^0_i}_f\, \varphi^0_i \; \left[\bar{f}\, M_f \mathcal P_R f\right]
\; + \;\mathrm{h.c.} \, ,
\end{eqnarray}
where $\mathcal P_{R,L}\equiv \frac{1\pm \gamma_5}{2}$ are the right-handed and left-handed chirality projectors,
$M_f$ the diagonal fermion mass matrices
and the couplings of the neutral scalar fields are given by:
\begin{equation} \label{yukascal}
y_{d,l}^{\varphi^0_i} = \mathcal{R}_{i1} + (\mathcal{R}_{i2} + i\,\mathcal{R}_{i3})\,\varsigma_{d,l} \, ,
\qquad\qquad
y_u^{\varphi^0_i} = \mathcal{R}_{i1} + (\mathcal{R}_{i2} -i\,\mathcal{R}_{i3}) \,\varsigma_{u}^* \, .
\end{equation}
As in the SM, all scalar-fermion couplings are proportional to the corresponding fermion masses, and
the only source of flavour-changing interactions is the Cabibbo-Kobayashi-Maskawa~(CKM) quark mixing matrix $V$~\cite{Cabibbo:1963yz,Kobayashi:1973fv}.
The usual models with natural flavour conservation, based on discrete ${\cal Z}_2$ symmetries, are recovered for particular (real) values of the couplings $\varsigma_f$~\cite{Pich:2009sp}.
The full set of interactions among the gauge and scalar bosons is given in \cite{ilisie1}. The coupling of a single neutral scalar with a pair of gauge bosons takes the form ($V=W,Z$)
\begin{align}
g_{\varphi_i^0 VV} = \mathcal{R}_{i1} \; g^{\text{SM}}_{hVV}\, ,
\end{align}
which implies $g_{hVV}^2 + g_{HVV}^2 + g_{AVV}^2 = (g_{hVV}^\text{SM})^2$. Thus, the strength of the SM Higgs interaction is shared by the three 2HDM neutral bosons. In the CP-conserving limit, the CP-odd field decouples while the strength of the $h$ and $H$ interactions is governed by the corresponding $\cos\tilde\alpha$ and $\sin\tilde\alpha$ factors.
In the following analysis we are also going to need the coupling of a neutral scalar with a pair of charged Higgses. We have parametrized the corresponding interaction as:
\begin{equation}
{\cal L}_{\varphi^0 H^+H^-}\; =\; - v \;\sum_{\varphi^0_i}\, \lambda_{\varphi^0_i H^+H^-}\;\, \varphi^0_i\, H^+H^-\, .
\label{hHPHM}
\end{equation}
Explicit expressions for the reduced cubic couplings $\lambda_{\varphi^0_i H^+ H^-}$, in terms of the generic Higgs potential parameters, can be found in \cite{ilisie1}.
The phenomenological constraints on the A2HDM parameters have been studied in detail in Refs.
\cite{ilisie3,ilisie2,ilisie1,Pich:2009sp,JungTuzon,PichTuzon1,Jung:2012vu,PichDstar,PichBll,PichEDP}. For a light $H^\pm$, loop-induced processes dominated by top contributions ($\varepsilon_K$, $Z\to b\bar b$, $B^0$--$\bar B^0$ mixing) impose a tight (95\% CL) upper bound on the up-type alignment parameter: $|\varsigma_u| < 0.77 \; (1.7)$, for $M_{H^\pm} = 80$ (500) GeV. Owing to the much smaller fermion masses, the constraints on the down-type (and lepton) parameter are very weak; one imposes instead $|\varsigma_d| \le 50$ to guarantee a perturbative Yukawa coupling. In the popular type-II 2HDM
($\varsigma_u = -1/\varsigma_d = -1/\varsigma_l =\cot\beta$), the decay $\bar B\to X_s\gamma$ excludes charged Higgs masses below 380 GeV \cite{Hermann:2012fc} at 95\% CL, because the SM and charged-Higgs contributions interfere constructively. This is no longer true in the more general A2HDM framework, where one only gets a combined correlated constraint on $M_{H^\pm}$, $\varsigma_u$ and $\varsigma_d$, which allows much lighter values of the charged-scalar mass in a restricted region of the parameter space $\varsigma_u$--$\varsigma_d$ \cite{JungTuzon,PichTuzon1,Jung:2012vu}.
The symmetries of the A2HDM protect in a very efficient way the flavour-blind phases of the alignment parameters from undesirable phenomenological consequences. The experimental upper bounds on fermion electric dipole moments provide the
strongest constraints on $\mathrm{Im}(\varsigma_f)$, but ${\cal O}(1)$ contributions remain allowed at present \cite{PichEDP}.
For simplicity, in section \ref{sec:phenom}, we will restrict our analysis to the CP-conserving limit and, therefore, will consider real alignment parameters.
The LHC data require the gauge coupling of the 125~GeV boson to have a magnitude close to the SM one. Assuming that it corresponds to the lightest CP-even scalar $h$ of the CP-conserving A2HDM, the measured Higgs signal strengths imply $|\cos{\tilde\alpha}| > 0.90\; (0.80)$ at 68\% (90\%) CL \cite{ilisie3,ilisie2,ilisie1}.
Direct searches for a heavier neutral scalar ($H$) provide upper bounds on $|\sin{\tilde\alpha}|$ as a function of $M_H$, which at present result in a weaker
constraint on the mixing angle \cite{ilisie3}.
In the following we will explore the intriguing possibility that the charged scalar could be fermiophobic, {\it i.e.}, that its tree-level couplings to fermions vanish ($\varsigma_{u,d,l}=0$). All current experimental bounds are then trivially avoided,
in particular the flavour constraints \cite{JungTuzon}. The Yukawa couplings of the h(125) boson scale in this case, with respect to the SM ones, with the same factor as the gauge couplings: $y^h_f = {\cal R}_{11} = \cos{\tilde\alpha}$. The global fit to the Higgs signal strengths results in the slightly improved bound $|\cos{\tilde\alpha}| > 0.86$ at 90\% CL \cite{ilisie3}.
In the fermiophobic (and CP-conserving) limit, the CP-odd scalar $A$ has also vanishing Yukawa couplings. Therefore, it only couples via multi-Higgs interactions with an even number of $A$ bosons, or through its gauge couplings
($A W^\pm H^\mp$, $A Z h$, $A Z H$, $A^2Z^2$, $A^2W^+W^-$, $AH^\pm W^\mp \gamma$, $AH^\pm W^\mp Z$).
Thus, a light $A$ boson might be very long-lived. While this could have cosmological implications, it is not in conflict
with the relic-density constraints
\cite{Krawczyk:2013jta,SK:13,CMR:07,Barbieri:2006dq,LopezHonorez:2006gr,ALT:09,LHY:10,LHY:11,DS:09,Arhrib:2013ela,Ginzburg:2010wa}.
A more specific version of the fermiophobic scenario is provided by the {\it inert\/} 2HDM~\cite{Ma:2008uza,Ma:2006km}, which assumes a discrete ${\cal Z}_2$ symmetry in the Higgs basis such that all SM fields and $\Phi_1$ are even ($\Phi_1\to \Phi_1$) under this symmetry while the second (inert) scalar doublet is odd ($\Phi_2\to -\Phi_2$). In this restricted case, there is no mixing between the CP-even neutral scalars $h$ and $H$; {\it i.e.}, $\cos{\tilde\alpha} = 1$. The spectrum of the {\it inert\/} 2HDM is described in appendix~\ref{subsec:inert}.
\section{Decay and Production modes}
\label{sec:calc}
We are going to analyse the possibility of having a fermiophobic charged Higgs with a mass in the restricted interval $M_{H^\pm} \in [M_W, \!\ M_W+M_Z]$. In this region, the only relevant decay rates are $H^+\to W^+\gamma$ and $H^+\to W^+\varphi_i^0$.
We are mainly interested in the one-loop suppressed decay
$H^+\to W^+\gamma$, the only two-body kinematically allowed decay mode, but we need to account also for the tree-level decay into a $W^+$ boson and a neutral scalar, which cannot be both on-shell simultaneously for the whole considered kinematical region. Thus,
we shall consider three-body decays like $H^+\to W^+ f\bar{f}$ mediated by the neutral scalars $\varphi_i^0$
and $H^+\to \varphi_i^0 f_u\bar{f}_d$ mediated by a virtual $W^+$,
where $f_u\bar{f}_d$ stands for quark pairs $q_u\bar{q}_d$, or lepton-neutrino pairs $l^+\nu_l$. The loop-induced decay $H^+\to f_u\bar{f}_d$ has a strong Yukawa suppression $m_f^2/v^2$ and, therefore, it is irrelevant for this discussion.
When surpassing the $M_W+M_Z$ threshold, the one-loop decay $H^+\to W^+Z$ would enter the game and we would also be close to the top-quark production threshold. The analysis of these two extra decay modes lays beyond the goal of this paper.
\subsection{$\mathbf{H^+\to W^+\gamma}$}
The first process that we are going to analyse is $H^+(k+q)\to W^+(k)\,\gamma(q)$.
Owing to the conservation of the electromagnetic current, the
decay amplitude must adopt the form:
\begin{align}
{\cal M}\; =\;\Gamma^{\mu \nu}\,\varepsilon_\mu^*(q)\,\varepsilon_\nu^*(k)\, ,
\qquad\qquad
\Gamma^{\mu \nu}\; =\; \left( g^{\mu\nu} k \cdot q - k^\mu q^\nu\right) \, S \, + \, i \, \epsilon^{\mu\nu\alpha\beta} \, k_\alpha \, q_\beta \; \tilde{S}\, ,
\label{transverse}
\end{align}
where $S$ and $\tilde S$ are scalar form factors.
To obtain this expression,
we have considered the most general Lorentz structure for the effective $\Gamma^{\mu\nu}$ vertex, and have imposed the electromagnetic current conservation condition $q_\mu\, \Gamma^{\mu\nu} = 0$.
All terms proportional to $q^\mu$ and $k^\nu$ have been also eliminated, as they cancel when contracted with the polarization vectors of the photon and the $W$ boson. Note that, accidentally, the Ward-like identity $k_\nu \,\Gamma^{\mu\nu}=0$ also holds for (\ref{transverse}).
In the unitary gauge, the decay proceeds at one loop through
the three sets of diagrams shown in Fig.~\ref{oneLoop}: fermionic loops (set 1), scalar loops (set 2) and loops with both gauge and scalar bosons (set 3).
Each set is transverse by itself, {\it i.e.}, of the form given in (\ref{transverse}). We can then decompose the result into the three separate contributions:
$S = S_{(1)} + S_{(2)} + S_{(3)}$ and $\tilde{S}=\tilde{S}_{(1)}$ (the only contribution to the structure $\epsilon^{\mu\nu\alpha\beta} \!\ k_\alpha \!\ q_\beta$ comes from the fermionic loops).
One can further simplify the calculation of $S_{(j)}$
by only considering the terms of the transverse set $j$ that contribute to the structure $k^\mu q^\nu$.
In order to calculate these contributions, one only needs to compute diagrams 1.a and 1.b for the first set, 2.a for the second set and 3.a for the third one.
\begin{figure}[t]
\centering
\includegraphics[scale=0.44]{./Figures/diag1.pdf} \\
\includegraphics[scale=0.44]{./Figures/diag2.pdf} \\
\includegraphics[scale=0.44]{./Figures/diag3.pdf}
\caption{\it One-loop diagrams contributing to $H^+\to W^+\gamma$ in the unitary gauge.}
\label{oneLoop}
\end{figure}
\noindent We obtain the following expressions for the form factors:
\begin{align}
S_{(1)}\; = &\;\; \frac{\alpha \!\ N_C \!\ |V_{tb}|^2}{2\pi \!\ v \!\ s_{_\text{W}}}\; \int_0^1 dx \int_0^1 dy \;\, \left[ Q_t\, x + Q_b\, (1-x)\right]
\notag \\ &\;\; \times \;\; \frac{-\varsigma_u m_t^2 \, x\, (2xy -2y + 1) \, + \, \varsigma_d m_b^2 \, (1-x)(1-2xy)}{M_W^2\, x\, (x-1) + m_b^2\, (1-x)+ m_t^2\, x + (M_W^2-M_{H^\pm}^2)\, xy\, (1-x)} \; ,
\\[2.2ex]
S_{(2)} \; = &\;\; \frac{\alpha\, v}{2\pi \, s_{_\text{W}}} \; \sum_i \; \lambda_{\varphi_i^0 H^+ H^-} \!\ \big(\mathcal{R}_{i2}- i\mathcal{R}_{i3} \big) \; \int_0^1 dx \int_0^1 dy
\notag \\ & \;\; \times \;\; \frac{x^2 y\, (1-x)}{M_W^2\, x\, (x-1) + M_{\varphi_i^0}^2\, (1-x)+ M_{H^\pm}^2\, x + (M_W^2-M_{H^\pm}^2)\, xy\, (1-x)} \; ,
\\[2.2ex]
S_{(3)}\; = &\;\; \frac{\alpha}{2\pi v \, s_{_\text{W}}}\; \sum_i \; \mathcal{R}_{i1} \big(\mathcal{R}_{i2}- i\mathcal{R}_{i3} \big) \; \int_0^1 dx \int_0^1 dy \;\, x^2
\notag \\ & \;\;\times \;\;
\frac{ 2M_W^2 \, + \, \big(M_{H^\pm}^2 + M_W^2 - M_{\varphi_i^0}^2\big) \, y\, (x-1)}{M_W^2\, x^2 + M_{\varphi_i^0}^2\, (1-x)+ (M_W^2-M_{H^\pm}^2)\, xy\, (1-x)} \; ,
\\[2.2ex]
\tilde{S} \; = &\;\; \frac{\alpha\, N_C \, |V_{tb}|^2}{2\pi \, v \, s_{_\text{W}}} \int_0^1 dx \int_0^1 dy \;\,
\left[ Q_t\, x + Q_b\, (1-x)\right]
\notag \\ & \;\; \times \;\;
\frac{\varsigma_u m_t^2 \, x\, +\,\varsigma_d m_b^2\, (1-x)}{M_W^2\, x\, (x-1) + m_b^2\, (1-x)+ m_t^2\, x + (M_W^2-M_{H^\pm}^2)\, xy\, (1-x)} \; ,
\end{align}
with $s_{_\text{W}}\equiv\sin{\theta_{_{\mathrm{W}}}}$.
The calculation of $S_{(3)}$ has been also performed in the Feynman ($\xi=1$) gauge,
where additional diagrams with Goldstone bosons are present, verifying that these expressions are gauge independent. Our results are
in agreement with the recent calculation of the $H^+W^-\gamma$ effective vertex in Ref.~\cite{EffVertex}. This calculation was also done many years ago by several groups \cite{OldHWgamma1,OldHWgamma2,OldHWgamma3,OldHWgamma4} using a somewhat different notation.
The $H^+\to W^+\gamma$ decay width is easily found to be:
\begin{equation}
\Gamma(H^+\to W^+\gamma)\; =\; \frac{M_{H^\pm}^3}{32 \pi} \; \left(1-\frac{M_W^2}{M_{H^\pm}^2}\right)^3 \, \left(\, |S|^2+|\tilde{S}|^2 \,\right)\, .
\end{equation}
This one-loop decay rate is in general much smaller than the tree-level decay rates of a charged Higgs into fermions. However, it becomes relevant if the charged Higgs is fermiophobic ($\varsigma_f\to 0$). In this case, the first set of diagrams (which has only been presented for completeness) does not contribute.
\subsection{$\mathbf{H^+\to W^+\varphi_i^0}$}
\begin{figure}[t]
\centering
\includegraphics[scale=0.6]{./Figures/diag5.pdf}
\caption{\it $H^+\to W^+ f\bar{f}$ process mediated by the virtual neutral scalars $\varphi_i^0$ (left) and $H^+\to \varphi_i^0 f_u\bar{f}_d$ mediated by a virtual $W^+$ (right).}
\label{threebody}
\end{figure}
The $H^+$ decay rate to on-shell $W^+$ and $\varphi_i^0$ bosons is given by
\begin{equation}
\Gamma(H^+ \to W^+\varphi_i^0) \; = \;
\frac{\alpha}{16\, s_{_\text{W}}^2\,M_{H^\pm}^3 \, M_W^2} \;
\left(\mathcal{R}_{i2}^2 + \mathcal{R}_{i3}^2\right) \; \lambda^{3/2}( M_{\varphi_i^0}^2, M_{H^\pm}^2,M_W^2) \, ,
\end{equation}
with the usual definition of the lambda function $\lambda(x,y,z)\equiv x^2+y^2+z^2-2xy-2xz-2yz$.
The corresponding three-body decay rate to $W^+f\bar{f}$, with off-shell neutral scalars (Fig.~\ref{threebody}, left), takes the form:
\begin{align}
\Gamma(H^+ \to W^+ f \bar{f} )\; & = \;
\frac{\alpha^2\, N_C^f\, m_f^2}{128\,\pi\, s_{_{\text{W}}}^4 \, M_{H^\pm}^3 \, M_W^4}\;
\int_{4 m_f^2}^{(M_{H^\pm}-M_W)^2} ds_{23} \;\;
\lambda^{3/2}(M_{H^\pm}^2,M_W^2,s_{23})
\notag \\[1.5ex]
& \qquad\qquad \times \; \left( 1-\frac{4 m_f^2}{s_{23}} \right)^{1/2} \; \sum_{i,j} \;
\big(\mathcal{R}_{i2}-i\mathcal{R}_{i3}\big)
\big(\mathcal{R}_{j2}+i\mathcal{R}_{j3}\big) \; \mathcal{M}_{ij} \; ,
\end{align}
where $N_C^f$ stands for the number of colours of the fermion $f$, 3 for quarks and 1 for leptons, $s_{23}$ is the square of the fermion-antifermion invariant mass and
\begin{align}
\mathcal{M}_{ij} \; \equiv \; \frac{(s_{23}-2m_f^2) \; \text{Re}\big(y_f^{\varphi_i^0} y_f^{\varphi_j^0 *} \big) - 2m_f^2 \; \text{Re}\big(y_f^{\varphi_i^0} y_f^{\varphi_j^0 }\big)}{(s_{23}-M_{\varphi_i^0}^2)(s_{23}-M_{\varphi_j^0}^2)}
\; .
\end{align}
Obviously, the $b$-quark contribution will dominate because of the global factor $m_f^2$. Therefore, we will neglect the other fermionic final states.
For the decay $H^+\to \varphi_i^0 f_u\bar{f}_d$, with an of-shell $W^+$ (Fig.~\ref{threebody}, right), we are going to consider all possible final states, quarks and leptons. We exclude the top quark, since this process is well below its production threshold. Neglecting the final fermion masses, the sum over all kinematically-allowed decay modes amounts to a global factor
\begin{equation}
\Omega\; = \; \left( 3 + N_C \sum_{u_i = u,c} \,\sum_{d_j = d,s,b} |V_{u_i d_j}|^2 \right) \; = \; 9 \, ,
\end{equation}
where the unitarity of the CKM matrix has been used. The total decay width can be expressed as an integral over the fermion-antifermion invariant-mass squared:
\begin{align}
\Gamma\Bigl(H^+ \to \varphi_i^0 \sum_{f_u,f_d} f_u\bar{f}_d \Bigr) \; = \; &
\frac{\Omega}{9} \;
\frac{3 \,\alpha^2\, (\mathcal{R}_{i2}^2 + \mathcal{R}_{i3}^2 )}{64\,\pi\, s_{_{\text{W}}}^4 \, M_{H^\pm}^3}\;
\int_{0}^{(M_{H^\pm}-M_{\varphi_i^0})^2} \; ds_{23} \; \frac{\lambda^{3/2}(M_{H^\pm}^2,M_{\varphi_i^0}^2,s_{23})}{(s_{23}-M_W^2)^2} .
\end{align}
\subsection{Charged-Higgs Production}
In order to see if the fermiophobic scenario can be experimentally probed, one needs an estimation of the production cross sections for different channels. Here we
will consider two possibilities, the associated production with a neutral scalar and the associated production with a $W$ boson (Fig.~\ref{productionCH}).
The $q_u\bar{q}_d\to H^+\varphi_i^0$ production process is by far the most interesting channel, as it requires the least number of new parameters.
For initial-state massless quarks, the leading-order (LO) partonic cross section reads
\begin{equation}
\hat{\sigma}(q_u\bar{q}_d\to H^+\varphi_i^0) \; = \; \frac{g^4 \; |V_{ud}|^2}{768 \; \pi \; N_c \; \hat{s}^2} \;
\frac{(\mathcal{R}_{i2}^2+\mathcal{R}_{i3}^2)}{(\hat s - M_W^2)^2} \; \lambda^{3/2}(\hat{s},M_{H^\pm}^2,M_{\varphi_i^0}^2) \; ,
\label{drell-yan}
\end{equation}
where $\hat s$ is the partonic invariant-mass squared.
The NLO QCD corrections are available and can be expressed in a very simple form, as shown in appendix~\ref{QCDcorrections}.
\begin{figure}[t]
\centering
\includegraphics[scale=0.42]{./Figures/production.pdf}
\caption{\it LO contributions to the charged-Higgs associated production with a $W$ boson (diagrams a, b) or a neutral scalar (diagram c), in the fermiophobic scenario.}
\label{productionCH}
\end{figure}
The associated production with a $W$ boson can proceed through either $q\bar{q}$ or
$gg$ fusion. The partonic LO cross section for the $q\bar{q}$ fusion process, is given by
\begin{align}
\hat{\sigma}(q\bar{q}\to H^+ W^-)\; &=\; \frac{g^2}{128 \; \pi \, M_W^2 \, {\hat s}^2} \; \frac{m_q^2}{v^2} \; \frac{1}{N_c} \;
\lambda^{3/2}(\hat s,M_{H^\pm}^2,M_W^2) \;\left(1 -\frac{4 m_q^2}{\hat s}\right)^{-1/2}
\notag \\
& \qquad\qquad\qquad \times \; \sum_{i,j} \;
\big(\mathcal{R}_{i2}+i\mathcal{R}_{i3}\big)
\big(\mathcal{R}_{j2}-i\mathcal{R}_{j3}\big) \; \mathcal{N}_{ij} \; ,
\label{sigmaqq}
\end{align}
with the reduced amplitudes
\begin{equation}\label{eq:Nij}
\mathcal{N}_{ij} \;\equiv\;
\frac{(\hat s-2m_q^2)\; \text{Re}\big(y_q^{\varphi_i^0} y_q^{\varphi_j^0 *} \big) - 2m_q^2 \; \text{Re}\big(y_q^{\varphi_i^0} y_q^{\varphi_j^0 }\big)}{(\hat s-M_{\varphi_i^0}^2 + i M_{\varphi_i^0} \Gamma_{\varphi_i^0})\, (\hat s-M_{\varphi_j^0}^2-iM_{\varphi_j^0}\Gamma_{\varphi_j^0})} \; .
\end{equation}
We have kept the dependence on the initial quark masses, since otherwise the $q\bar q$ Yukawa coupling vanishes. This implies a strong suppression of this production mechanism by a factor $m_q^2/v^2$.
The gluon fusion mechanism dominates by far the previous one. The corresponding LO cross section at the partonic level takes the form
\begin{align}
\hat{\sigma}(gg\to H^+ W^-)\; & =\; \frac{\alpha_s^2 \; T_F^2}{4096 \; \pi^3 \; v^4} \; \lambda^{3/2}(\hat s,M_{H^\pm}^2,M_W^2)
\notag \\
& \qquad\qquad\qquad \times \; \sum_{i,j} \; (\mathcal{R}_{i2} + i \mathcal{R}_{i3})(\mathcal{R}_{j2} - i \mathcal{R}_{j3}) \; \mathcal{G}_{ij} \; ,
\label{sigmagg}
\end{align}
where $T_F=1/2$ is the $SU(3)$ colour group factor and the reduced amplitudes $\mathcal{G}_{ij}$ are given by
\begin{equation}\label{eq:Gij}
\mathcal{G}_{ij}\; \equiv\; \sum_{q q'}\;
\frac{\text{Re}\big( y_q^{\varphi_i^0} \big)\, \text{Re}\big( y_{q'}^{\varphi_j^0} \big) \,\mathcal{F}(x_q)\,\mathcal{F}(x_{q'})^*\, +\,
\text{Im}\big( y_q^{\varphi_i^0} \big)\, \text{Im} \big( y_{q'}^{\varphi_j^0} \big)\,
\mathcal{K}(x_q)\,\mathcal{K}(x_{q'})^*
}{ (\hat s-M_{\varphi_i^0}^2+ i M_{\varphi_i^0} \Gamma_{\varphi_i^0}) \, (\hat s-M_{\varphi_j^0}^2 - i M_{\varphi_j^0} \Gamma_{\varphi_j^0}) }\; ,
\end{equation}
with $x_q\equiv 4m_q^2/\hat s$.
The explicit expressions of the different loop functions are:
\begin{equation}
\mathcal{F}(x)\; =\; \frac{x}{2}\, [4+(x-1)f(x)] \, ,
\qquad\qquad\qquad
\mathcal{K}(x) \; = \; - \frac{x}{2}\, f(x) \, ,
\end{equation}
with
\begin{equation}
f(x)\; =\; \begin{cases} \; -4\arcsin^2(1/\sqrt{x})\, , \qquad\quad & x\geqslant1
\\[3pt]\; \Big[\ln\Big( \frac{1+\sqrt{1-x}}{1-\sqrt{1-x}}\Big)- i\pi \Big]^2\, , & x<1 \end{cases} \; .
\end{equation}
We have regulated the propagator poles with the term $ i M_{\varphi_i^0} \Gamma_{\varphi_i^0}$, both in Eqs.~(\ref{eq:Nij}) and (\ref{eq:Gij}),
because in our analysis one of the neutral scalars will, most likely, reach the on-shell kinematical region.
NLO QCD corrections to the gluon fusion channel are also available and will be taken into account; the details are given in appendix~\ref{QCDcorrections2}.
\section{Phenomenology}
\label{sec:phenom}
In the following phenomenological analysis, besides the fermiophobic charged-Higgs assumption ($\varsigma_f\to 0$),
we are also going to consider that the Higgs potential is CP-conserving. The consequence of this last hypothesis is that the CP-odd neutral Higgs $A$ will also be fermiophobic, as we have mentioned before in section \ref{sec:A2HDM};
moreover $\lambda_{AH^+H^-} =0$. This means that the decay $H^+\to W^+ A^* \to W^+ \bar{f}f$ does not occur and $A$ does not contribute either to
$H^+\to W^+\gamma$. The charged-Higgs production amplitudes mediated by a virtual $A$
also vanish. The CP-odd scalar can contribute to $H^\pm$ production in a direct way through the $q_u \bar{q}_d \to W^{*} \to H^+ A$ production channel or, in an indirect way, by modifying the total decay rate $\Gamma_{\varphi_i^0}$,
which regulates the pole in the CP-even scalar propagators ($\varphi_i^0=h,\, H$), through
decays like $\varphi_i^0\to AA$ or $\varphi_i^0\to AZ$. The decay $H\to Ah$ cannot occur at tree level because all cubic vertices of the scalar potential involving an odd number of $A$ fields vanish in the CP-conserving limit. The total decay width $\Gamma_{\varphi_i^0}$ is the sum of all the decay rates explicitly presented in appendix~\ref{app:decayrates}.
In our particular case, the expressions for the Yukawa couplings simplify and become equal to the reduced scalar couplings to two gauge bosons. They are given by
\begin{equation}
y^h_f = \frac{g_{hVV}^{\phantom{\mathrm{SM}}}}{g_{hVV}^{\mathrm{SM}}} = \mathcal{R}_{11} = \cos\tilde\alpha \, ,
\qquad
y^H_f = \frac{g_{HVV}^{\phantom{\mathrm{SM}}}}{g_{hVV}^{\mathrm{SM}}} = R_{21} = -\sin\tilde\alpha \, ,
\qquad
y^A_f = g_{AVV}^{\phantom{\mathrm{SM}}} = \mathcal{R}_{31} = 0 \, .
\end{equation}
Even within the restricted range of charged-Higgs masses we are interested in, $M_{H^\pm} \in [M_W,M_W+M_Z]$, the possible phenomenological signals depend on the choice of masses for the remaining scalars. In the following subsections, we will therefore consider different scenarios for the scalar spectrum. The first part of the analysis will be dedicated to the study of the various decay modes of the charged Higgs and the second part will focus on estimating the production cross sections.
\subsection{Decay rates and branching ratios}
One of the two CP-even scalars should correspond to the Higgs boson discovered at the LHC, but a broad range of masses is allowed for the other two neutral scalars. We will consider the following four scenarios, which cover the different possibilities:
\begin{enumerate}
\item $M_h = 125$~GeV \ and \ $M_{A,H} > M_W + M_Z$.
\item $M_h = 125$~GeV \ and \ $M_{A} < M_W + M_Z < M_{H}$.
\item $M_h = 125~\mathrm{GeV} < M_H < M_W + M_Z$ \ and three different options for $A$
($M_A < M_H$, \ $M_H < M_A < M_W + M_Z$ \ and \ $M_A > M_W + M_Z$).
\item $M_H = 125$~GeV, \ $M_h = 90$~GeV \ and \ $M_{A} < M_W + M_Z$.
\end{enumerate}
\subsubsection{First Scenario}
\label{sec:phenom1}
\begin{figure}[t]
\centering
\includegraphics[scale=0.53]{./Figures/BR_1_1_09.pdf} \;\;\; \includegraphics[scale=0.53]{./Figures/BR_0_0_09.pdf} \\[2ex]
\includegraphics[scale=0.53]{./Figures/BR_m1_m1_09.pdf} \;\;\; \includegraphics[scale=0.55]{./Figures/Decay_1_1_m1_m1_09.pdf}
\caption{Charged-Higgs branching ratios as functions of $M_{H^\pm} \in [M_W,M_W+M_Z]$, for $\cos\tilde\alpha = 0.9 $, $M_H \in [M_W+M_Z, \, 500\, \text{GeV}]$ and $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-} = 1$ (top-left), 0 (top-right) and -1 (bottom-left). The corresponding total decay widths are shown in the bottom-right panel $(\lambda_{h}^{\pm} \equiv \lambda_{h H^+ H^-} ,\; \lambda_{H}^\pm \equiv \lambda_{H H^+ H^-})$.}
\label{11C09}
\end{figure}
In the first scenario the mass of the lightest CP-even scalar is set to $M_h=125$ GeV. Therefore, the strong constraint on the scalar mixing angle, from the global fit to the light Higgs boson signal strengths using the LHC
data, must be used: $|\cos\tilde\alpha|> 0.9$ at 68\% CL \cite{ilisie3,ilisie2,ilisie1}.
The masses of the remaining neutral scalars are considered to be greater than $M_W + M_Z$ so that decays of a charged Higgs into an on-shell $H$ or $A$ are kinematically forbidden. In the limit $\cos\tilde\alpha \to 1$, the only surviving decay amplitude (not proportional to $\sin\tilde\alpha$) is the contribution of $H$ to the amplitude $S_{(2)}$. Thus, in this limit the branching ratio of $H^+\to W^+ \gamma$ is 100\%; all the other decay channels vanish.
If we set $\cos\tilde\alpha = 0.9$, $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-}= 1$, vary the charged Higgs mass in the region $M_{H^\pm} \in [M_W,M_W+M_Z]$ and $M_H$ from $M_W + M_Z$ up to 500 GeV, we obtain the branching ratios (top-left) and total decay width (bottom-right) shown in Fig.~\ref{11C09}. The width of the branching ratio bands reflects the variation of the input parameters in the mentioned ranges. The same consideration is valid for the following scenarios. The decay channel $H^+\to W^+ \gamma$ dominates for $M_{H^{\pm}} \lesssim M_h$. When the charged Higgs is kinematically allowed to decay into an on-shell $h$, then $H^+\to hf_u\bar{f}_d$ rapidly becomes the dominant channel as $M_{H^\pm}$ grows. The remaining $H^+\to W^+ b\bar b$ branching ratio stays at a few percent level or less for the whole allowed region. The total decay width approximately grows from $10^{-14}$ up to $10^{-8}$ GeV, in the region dominated by the radiative $H^+\to W^+ \gamma$ decay, and sizeably increases up to $10^{-5}$ GeV, once the $hf_u\bar{f}_d$ production threshold is reached.
The tree-level decay rates are significantly larger than the loop-induced one. Flipping the sign of $\cos\ta$ leads to an equivalent solution with a sign flip of the coupling $\lambda_{hH^+H^-}$. This is also valid for the next scenarios.
If, instead, we consider all the previous settings but taking this time $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-}= 0$, then the only amplitude that contributes to the $H^+\to W^+\gamma$ decay channel is $S_{(3)}$, which is suppressed by a factor $\sin{\tilde\alpha}$. As shown in Fig.~\ref{11C09} (top-right), this channel remains the dominant one up to $M_{H^\pm} \gtrsim M_h$, but with a sizeably smaller decay width
(bottom-right). The $H^+\to W^+ b\bar b$ branching ratio is also more sizeable, raising up to the 10\% level.
Let us now consider $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-} = -1$ and everything else as previously. In this particular case the amplitudes $S_{(2)}$ and $S_{(3)}$ interfere destructively and, as a consequence, the decay $H^+\to W^+b\bar{b}$ competes with $H^+\to W^+\gamma$. Thus, the $Wb\bar{b}$ decay channel can dominate in some cases. However, as soon as the charged Higgs reaches $M_{H^\pm} \gtrsim M_h$, the dominant decay mode is again $H^+\to hf_u\bar{f}_d$, as in the previous cases (Fig.~\ref{11C09}, bottom-left).
\subsubsection{Second Scenario}
\label{sec:phenom2}
In the second scenario the mass of lightest CP-even scalar is set to $M_h=125$ GeV
and $M_H > M_W+M_Z$, as in the first one, but this time we assume the CP-odd Higgs boson $A$ to have its mass below the $WZ$ threshold ($M_A < M_W+M_Z$).
The decay of the charged Higgs into an on-shell $A$ is then kinematically allowed, but into an on-shell $H$ is forbidden. The same constraint as before is considered for the scalar mixing angle. Taking the limit $\cos\tilde\alpha \to 1$, this time there are two surviving decay amplitudes, $H^+\to W^+ \gamma$ and $H^+\to A f_u \bar{f}_d$.
Let us consider $\cos\tilde\alpha=0.9$, $\lambda_{hH^+H^-}=\lambda_{HH^+H^-}=1$ and $M_A=$ 90, 130 and 150 GeV. For each value we shall vary $M_H$ from $M_W+M_Z$ up to its allowed upper bound from the oblique parameters (at 68\% CL) \cite{ilisie3,ilisie2,ilisie1}, with a maximum limit of 500 GeV. We obtain then the branching ratios and total decay widths in Fig.~\ref{11C09A90130150}. We observe that for $M_A=90$ GeV (top-left), when kinematically allowed, the decay to an on-shell $A$ boson rapidly becomes the dominant one as $M_{H^\pm}$ increases. For this configuration the $Wb\bar{b}$ channel is insignificant. When $M_A=130$ GeV (top-right), which is close to $M_h$, the decays into an on-shell $h$ or $A$ boson compete. However, the decay to $Af_u\bar{f}_d$ still dominates even if the masses are similar because of the relative suppression factor $\sin^2\tilde\alpha$ of the $hf_u\bar{f}_d$ width. As $M_A$ becomes heavier, $M_A=150$~GeV (bottom-left), the decay rate into an on-shell $A$ boson does not grow as
rapidly as in the previous cases; thus, $hf_u\bar{f}_d$ dominates over $Af_u\bar{f}_d$ in the considered region. For the last two configurations, that is $M_A=130$ and 150 GeV, the $H^+\to W^+b\bar{b}$ decay channel can also bring sizeable contributions.
The total decay width in this scenario can reach as high as $10^{-3}$ GeV, see Fig.~\ref{11C09A90130150} (bottom-right). This is approximately two orders of magnitude larger than in the previous case and it is due to the tree-level decays, as we mentioned earlier. The maximum values are reached for the smallest mass of the CP-odd scalar ($M_A=90$ GeV).
It is worth mentioning that, just as in the previous scenario, the $Wb\bar{b}$ branching ratio can be sizeably increased by decreasing the $W\gamma$ decay width through a sign flip of the $\lambda_{\varphi_i^0H^+H^-}$ couplings, creating destructive interference among the various loop contributions. The same consideration is also valid for the next scenario.
\begin{figure}[t]
\centering
\includegraphics[scale=0.53]{./Figures/BR_1_1_09_A90.pdf} \;\;\; \includegraphics[scale=0.53]{./Figures/BR_1_1_09_A130.pdf} \\[2ex]
\includegraphics[scale=0.53]{./Figures/BR_1_1_09_A150.pdf} \;\;\; \includegraphics[scale=0.55]{./Figures/Decay_1_1_09_A90_150.pdf}
\caption{\it Charged-Higgs branching ratios as functions of $M_{H^{\pm}}$, for $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-}= 1$, $\cos\tilde\alpha=0.9$ and $M_A = $ 90 (top-left), 130 (top-right) and 150 (bottom-left) GeV. $M_H$ is varied from $M_W + M_Z$ up to its permitted value by the oblique parameters. The bottom-right panel shows the corresponding total decay widths.}
\label{11C09A90130150}
\end{figure}
\subsubsection{Third Scenario}
\label{sec:phenom3}
\begin{figure}[t]
\centering
\includegraphics[scale=0.53]{./Figures/BR_1_1_09_099_hH.pdf} \;\;\; \includegraphics[scale=0.53]{./Figures/BR_1_1_09_099_hHA.pdf} \\[2ex]
\includegraphics[scale=0.53]{./Figures/BR_1_1_09_099_hAH.pdf} \;\;\; \includegraphics[scale=0.55]{./Figures/Decay_1_1_09_099_hH.pdf}
\caption{\it Charged-Higgs branching ratios as functions of $M_{H^{\pm}}$, for $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-}= 1$, $\cos\tilde\alpha\in[0.9,\, 0.99]$, $M_H = 140$ GeV, $M_A > M_W + M_Z $ (top-left); $(M_H, M_A)=(140, 150)$ GeV (top-right) and $(M_H, M_A)=(150, 140)$ GeV (bottom-left). The total decay width for the first case is also shown (bottom-right).}
\label{case3}
\end{figure}
In this scenario the mass of the lightest CP-even scalar is also set to $M_h=125$ GeV, while the heavy CP-even Higgs boson $H$ has its mass in the range $M_h < M_H < M_W+M_Z$. For the mass of the remaining CP-odd scalar we consider three different possibilities: a) $M_A > M_W+M_Z$, so that the decay into an on-shell $A$ is forbidden; b) $M_H < M_A < M_W+M_Z$, and c) $M_A < M_H < M_W + M_Z$. In the last two situations the $H^\pm$ boson could decay into any of the three neutral scalars. Again, we use the LHC constraint $|\cos\tilde\alpha|>0.9$ at 68\% CL. In the limit $\cos\tilde\alpha \to 1$, there are three possible surviving decay channels: $H^+\to W^+ \gamma$, $H^+\to H f_u \bar{f}_d$ and, when kinematically allowed, $H^+\to A f_u \bar{f}_d$.
For all three cases we set $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-}= 1$ and vary $\cos\tilde{\alpha}\in [0.9,\, 0.99]$.
In Fig.~\ref{case3} we show the $H^\pm$ branching ratios (top-left) and total decay width (bottom-right) when $M_H$ = 140 GeV and $M_A > M_W + M_Z$ (first case). To illustrate the other two possibilities, we set $(M_H, M_A)=(140, 150)$ GeV (Fig.~\ref{case3}, top-right) and $(M_H, M_A)=(150, 140)$ GeV (Fig.~\ref{case3}, bottom-left). The total $H^\pm$ decay widths for these two last configurations are very similar to the first one.
The $H^\pm$ decay into an on-shell $h$ boson has a global relative suppression factor of $\tan^2{\tilde\alpha}$ with respect to the decay into an on-shell $H$ and $\sin^2{\tilde\alpha}$ with respect to the decay into an on-shell $A$. Therefore, when $hf_u\bar{f}_d$ competes with $Hf_u\bar{f}_d$, the later one dominates as $\cos\tilde\alpha\to 0.99$ (Fig.~\ref{case3}, upper-left). When all three channels compete, the decay rate into the heaviest scalar boson grows the slowest and, therefore, brings a sub-dominant contribution to the branching ratios.
\subsubsection{Fourth Scenario}
\label{sec:phenom4}
In this last scenario we are going to set the mass of the heavy CP-even scalar to $M_H=125$ GeV; therefore, the LHC bounds
translate into $|\sin\tilde\alpha|>0.9$ at 68\% CL. The mass of the light CP-even scalar will be set to $M_h=90$ GeV. As for the CP-odd one, we will consider three possible values: $M_A$ = 150, 140 and 110 GeV.
In order to safely avoid the stringent constraints on light scalar masses from LEP \cite{Abbiendi:2002qp,Barate:2003sz}, we need to have very suppressed decay and production channels. In our particular case with $\varsigma_f=0$, CP-conserving potential, and $M_A>M_h$ (therefore the decays $h\to AA$ and $h\to AZ$ are forbidden), we have the simple relation $\Gamma_h =\cos^2\tilde\alpha \; \Gamma^{\text{SM}}_h$. Here $\Gamma_h$ is the total decay rate of the light CP-even scalar boson with $M_h<M_H=125$ GeV, and $\Gamma^{\text{SM}}_h$ the corresponding decay rate in the SM for a Higgs boson with the same mass $M_h$.
The $\cos^2{\tilde\alpha}$ suppression factor is common to all allowed $h\to f\bar f$ decay modes, and cancels out in the branching ratios. The same suppression factor appears
in the LEP production rate, so that the signal strengths, relative to the SM, are then given by
\begin{align}
\mu_{X}^h \;\equiv\; \frac{\sigma(e^+e^-\to Zh) \; \text{Br}(h\to X)}{\sigma(e^+e^-\to Zh)_{\text{SM}} \; \text{Br}(h\to X)_{\text{SM}}}\; = \;\cos^2{\tilde\alpha} \; ,
\label{SignalStrX}
\end{align}
with $X=b\bar b$ and $\tau^+\tau^-$. Thus, we have a global suppression factor
$\cos^2{\tilde\alpha}$. The LEP constraints from the $\tau^+\tau^-$ channel, which are the strongest ones in our case, can then be avoided by setting $\cos^2{\tilde\alpha}\approx 0.02$ ($\sin\tilde\alpha\approx0.99$).
The OPAL collaboration has also performed a decay-mode-independent search for a light neutral scalar and found the upper limits $\cos^2\tilde{\alpha}<0.1$ (1) for $M_h < 19$ (81) GeV \cite{Abbiendi:2002qp}, which are weaker (in our case).
It is worth mentioning that in (\ref{SignalStrX}) we have ignored the charged-Higgs contribution to the $h\to\gamma\gamma$ decay rate. If however, we choose to enhance it
through the $H^\pm$ loop contribution,
it would only further suppress the fermionic branching ratios, weakening the bound on $\sin\tilde\alpha$.
\begin{figure}[t!]
\centering
\includegraphics[scale=0.53]{./Figures/BRA150.pdf} \;\;\; \includegraphics[scale=0.53]{./Figures/BRA140.pdf} \\[2ex]
\includegraphics[scale=0.53]{./Figures/BRA110.pdf} \;\;\; \includegraphics[scale=0.55]{./Figures/GAMMA110.pdf}
\caption{\it Charged-Higgs branching ratios as functions of $M_{H^{\pm}}$, for $\sin\tilde\alpha = 0.99$ and $M_h = 90$ GeV.
The trilinear couplings are set to $\lambda_{h H^+ H^-} = \lambda_{H H^+ H^-}= 1$, with $M_A = 150 $ GeV (top-left) and $M_A = 140 $ GeV (top-right),
and $\lambda_{h H^+ H^-}, \lambda_{H H^+ H^-}\in[-5,5]$ with $M_A = 110 $ GeV (bottom-left). The total decay width (bottom-right) for the last case is also shown.}
\label{case4}
\end{figure}
With all this being said, we set $\sin\tilde\alpha=0.99$. In Fig.~\ref{case4} we plot the $H^\pm$ branching ratios for $M_A=150$ (top-left) and 140 GeV (top-right), taking $\lambda_{hH^+H^-}=\lambda_{HH^+H^-}$. In both plots we can observe that, when kinematically allowed, the tree-level $H^+\to hf_u\bar{f}_d$ decay dominates. In this case, this decay no longer has a suppression factor as its partial width is proportional to $\sin^2\tilde\alpha\sim 1$. The suppression factor appears now in the $Hf_u\bar{f}_d$ decay mode with a partial decay width proportional to $\cos^2\tilde\alpha$. This is why, when $M_A\sim M_H$, the decay into an on-shell $A$ boson dominates over the decay into an on-shell $H$.
Both $A$ and $H$ contributions are, however, very suppressed due to their heavy masses. It is also worth mentioning that a small variation of $M_A$ can produce a significant change (roughly, one
order of magnitude) in $\text{Br}(H^+ \to A f_u \bar{f}_d )$, as can be seen in Fig.~\ref{case4} (top-left and top-right).
For the last case we set $M_A$ to 110 GeV. The perturbativity bounds on neutral scalar couplings to a pair of charged Higgses, for the considered region of the charged Higgs mass, are roughly given by $|\lambda_{\varphi_i^0H^+H^-}|\leq5$ (here $\varphi_i^0=h,H$) \cite{ilisie1}. In order to see the impact of these two parameters on the $H^\pm$ branching ratios, we will vary both independently in this region. The result, shown in Fig.~\ref{case4} (bottom-left), is that $W\gamma$ and $hf_u\bar{f}_d$ compete, even after crossing the $h$ production threshold. Since $M_A$ is lighter than in the previous two cases, the $H^+\to Af_u\bar{f}_d$ branching ratio can also reach higher values. The total decay rate for this configuration is also shown in Fig.~\ref{case4} (bottom-right).
As we have seen, in the four proposed scenarios, the configuration of the $H^\pm$ branching ratios depends very sensitively on the chosen parameters. However, we can draw some important conclusions. There are only a few decay channels to be analysed and the largest decay widths are the tree-level ones, corresponding to the on-shell production of scalar bosons. Thus, the number of decay channels decreases as the number of neutral scalar bosons that are heavier than the charged Higgs ({\it i.e.}, $M_{\varphi_i^0}>M_{H^\pm}$) increases. The $W\gamma$ decay mode can bring sizeable contributions below and close to the the on-shell production threshold of a scalar boson. Short after this threshold is reached, as $M_{H^{\pm}}$ grows, the $H^+\to W^+\gamma$ branching ratio rapidly decreases. As we have shown, the $H^+\to W^+b\bar b$ decay can dominate over $H^+\to W^+\gamma$ in some cases, depending on the values of the $\lambda_{\varphi_i^0H^+H^-}$ couplings. If a fermiophobic charged Higgs is finally discovered in this
mass range, the precise values of its mass and branching ratios would provide priceless information about all other parameters. The masses of the remaining scalars would also be highly constrained by the electroweak oblique parameters. These constraints were used in our second scenario, because they put an upper bound on $M_H$; we did not mention them in the other cases, since they do not bring additional constraints. The mean lifetime of a fermiophobic charged scalar is short, ranging from $10^{-11}$ to $10^{-23}$ s, making its direct detection very compelling at the LHC.
\subsection{Production cross sections}
\label{sec:phenom5}
In order to estimate the total hadronic cross sections for the various production channels, we need to convolute the partonic cross sections with the corresponding parton distribution functions (PDFs). Here we will use the MSTW set \cite{PDFs}. Moreover, we will compute the cross sections at the NLO; {\it i.e.}, including the LO QCD corrections, for which simple analytical expressions can be obtained \cite{DJ1,DJ2}.
For the $q_u\bar{q}_d\to H^+\varphi_i^0$ associated production, the ${\cal O}(\alpha_s)$ contributions simply correspond to the QCD corrections to the Drell-Yan process $q_u\bar{q}_d\to W^{*}$, integrating over the virtuality of the W boson.
As for the $H^+W^-$ associated production, the needed QCD corrections can be easily extracted from the SM Higgs production channels $q\bar q\to h$ and
$gg\to h$. At the LHC, $gg\to H^+W^-$ production dominates over $q\bar q\to H^+W^-$. For typical LHC hadronic center-of-mass energies, {\it i.e.}, $\sqrt{s}\sim 14$ TeV, the latter only corresponds at LO to a few percent of the total $pp\to H^+W^-$ cross section, so we can safely neglect it. The detailed expressions of the hadronic cross sections and the
QCD corrections are given in appendices~\ref{QCDcorrections} and \ref{QCDcorrections2}.
In order to estimate the theoretical uncertainty of the QCD enhancement factor
$K\equiv \sigma_{NLO}/ \sigma_{LO}$, we vary the factorization ($\mu_F$) and renormalization ($\mu_R$) scales for $\sigma_{NLO}$, keeping both scales fixed at their central value $\mu_F=\mu_R=\hat{s}$ for $\sigma_{LO}$.
When one of the intermediate scalar bosons reaches its on-shell kinematical region, one needs to estimate also its total decay rate. The explicit expressions for the tree-level scalar decay rates are presented in appendix \ref{app:decayrates}.
\subsubsection{$H^+\varphi_i^0$ associated production}
Assuming the most general scalar potential, the LO partonic cross section, given in Eq.~(\ref{drell-yan}), is proportional to the combination of rotation matrix elements $R^2\equiv(\mathcal{R}_{i2}^2+\mathcal{R}_{i3}^2)$. We take away the explicit dependence on the scalar-potential parameters, plotting in Fig.~\ref{crossDY} (left) the ratio
$\sigma(pp\to H^+\varphi_i^0)/R^2$ at $\sqrt{s}=14$~TeV, as a function of $M_{H^\pm}$, for different values of $M_{\varphi_i^0}$ which can be interpreted as the mass of any of the three neutral scalars of the theory.
As expected, the cross section reaches higher values for lower scalar masses. The most interesting case is of course $M_{\varphi_i^0}$=125 GeV, which could constitute a very good detection channel, since we already know that there is one scalar with that mass. If we consider $\varphi_i^0$ to be the light CP-even scalar of the theory, the cross section is suppressed by a factor $R^2=\sin^2{\tilde\alpha}$. The measurement of this production channel can be experimentally challenging due to the small value of the cross section.
QCD corrections provide a mild enhancement of the cross section. The resulting QCD K factor is shown in Fig.~\ref{crossDY} (right), for $M_{\varphi_i^0}=125$ GeV and different choices of $\mu_R$ and $\mu_F$. Its central value is around 1.2, similarly to other cross sections of the Drell-Yan type.
\begin{figure}[t]
\centering
\includegraphics[scale=0.56]{./Figures/crossPhiW.pdf} \;\;\; \includegraphics[scale=0.54]{./Figures/KnloDy.pdf}
\caption{\it LO production cross section $\sigma(pp\to H^+\varphi^0_i)/R^2$ at $\sqrt{s}=14$~TeV (left), as function of $M_{H^\pm}$, for different values of
$M_{\varphi_i^0}$. The
QCD K factor is shown (right) for $M_{\varphi_i^0}=125$ GeV and different choices of $\mu_R$ and $\mu_F$}
\label{crossDY}
\end{figure}
\subsubsection{$H^+W^-$ associated production}
For this specific production channel we are going to consider two
alternative possibilities: we can either identify the 125 GeV boson with the
lightest CP-even scalar $h$, or with the heaviest one $H$. In the first case ($M_h=125$ GeV), the scalar $H$ can be heavy enough to reach the on-shell region and, therefore, it is necessary to regulate the propagator pole with its total decay width. In the second case ($M_H=125$ GeV), both $M_h$ and $M_H$ are below the $H^+W^-$ production threshold for the whole considered range of charged Higgs masses. Therefore, there is no need to regulate the $h$ and $H$ poles (assuming their total decay widths to be small).\\
\vskip .5cm\noindent {\bf A) $\mathbf{M_h=125}$ GeV.}
Let us first estimate the size of the $H$ decay width for three representative values of $M_H$ (150, 200 and 400 GeV) and different choices for the cubic scalar couplings.
The CP-odd mass $M_A$ will always be taken within the 68\% CL range allowed by the oblique parameters. In the following discussion, we set $\cos\tilde\alpha=0.9$ and ignore the loop-induced decays $H\to gg$ and $H\to\gamma\gamma$, which are suppressed by a $\sin^2\tilde\alpha$ factor with respect to the SM.
\begin{figure}[t]
\centering
\includegraphics[scale=0.55]{./Figures/Hwidth1.pdf} \;\;\; \includegraphics[scale=0.55]{./Figures/Hwidth2.pdf} \\[2ex]
\includegraphics[scale=0.55]{./Figures/Hwidth3.pdf} \;\;\; \includegraphics[scale=0.55]{./Figures/Hwidth4.pdf}
\caption{\it Total $H$ decay rate as a function of $M_{H^{\pm}}$ for a) $M_H = 150 $ GeV with different values of $M_A$ and $|\lambda_{HAA}|$ (top-left), b) $M_H = 200 $ GeV and $M_A > M_H-M_Z$ with different values of $\lambda^{\pm}\equiv |\lambda_{HH^+H^-}|$ (top-right), c) $M_H = 200 $ GeV and $M_A = 50$ GeV with different values of $|\lambda_{HH^+H^-}|$ and $|\lambda_{HAA}|$ (bottom-left), and d) $M_H = 400 $ GeV and $M_A = 140$ GeV with different values for the set of couplings $(|\lambda_{HAA}|,\; |\lambda_{Hhh}|, \; |\lambda_{HH^+H^-}|)$ (bottom-right).}
\label{Hdecay}
\end{figure}
For $M_H=150$ GeV, the $H$ boson does not reach the on-shell region (its mass is below the $H^+W^-$ threshold) and its total decay width is in principle not needed to regulate the propagator pole. However, $\Gamma_H$ can induce sizeable effects for
small $M_A$
and large values of the cubic coupling $\lambda_{HAA}$.
This is shown in Fig.~\ref{Hdecay} (upper-left). When $M_A>M_H/2$, the $H$ width is small because its only relevant tree-level decays are $H\to b\bar b,\, WW$ and $ZZ$. However,
extra decay channels like $H\to AA$ or $H\to AZ$ are open when one allows $A$ to be light. This possibility is exemplified in the figure, taking $M_A=50$ GeV and $\lambda_{HAA}=0$ (therefore $H\to AZ$ is the only extra channel), and also for $|\lambda_{HAA}|=0.1, \; 1$ and 5. The width $\Gamma_H$ varies roughly from around $10^{-3}$ up to 100 GeV for the considered parameter configurations.
Let us now consider $M_H=200$ GeV. If the CP-odd boson satisfies $M_A>M_H-M_Z\approx 110$ GeV, then the channels $H\to AA, \, AZ$ are closed. The open decay channels are $H\to b\bar b, \, WW, \, ZZ$ as before, plus two extra ones: $H\to H^{\pm}W^{\mp}$ (up to $M_{H^\pm} \approx 120$ GeV) and $H\to H^+H^-$ (up to $M_{H^\pm}=100$ GeV). When kinematically allowed (and if $|\lambda_{HH^+H^-}|$ is not too small), the decay into two charged scalars is the dominating channel. There is also a sizeable contribution from $H\to H^{\pm}W^{\mp}$ when this decay mode is open. The predicted values of $\Gamma_H$ are shown in Fig.~\ref{Hdecay} (upper-right) for different values of $|\lambda_{HH^+H^-}|$. If we take instead $M_A=50$ GeV, the channels $H\to AA,\, AZ$ open. The $H$ decay width is shown for this configuration in Fig.~\ref{Hdecay} (lower-left), as a function of the charged Higgs mass, taking $|\lambda_{HAA}|=0,\, 5$ and $|\lambda_{HH^+H^-}|=0,\, 5$. The total $H$ decay width obviously increases with increasing
values of $|\lambda_{HAA}|$ and $|\lambda_{HH^+H^-}|$. In the considered range of cubic couplings, $\Gamma_H$ can vary between 1 and 200 (70) GeV when $H \to H^+H^-$ is allowed (forbidden, $M_{H^\pm}>M_H/2$).
Taking a heavier mass $M_H=400$ GeV, the electroweak oblique parameters imply very stringent restrictions on $M_A$: the only value that roughly satisfies these constraints for the whole considered range of the charged Higgs mass is $M_A=140$ GeV. For this configuration, all the channels we have considered before are kinematically allowed. Besides, there is an extra one, the decay into two light CP-even scalars $H\to hh$. Thus, we have three unknown couplings $\lambda_{HAA},\; \lambda_{Hhh},$ and $\lambda_{HH^+H^-}$. The lower-right panel in Fig.~\ref{Hdecay} shows the resulting values of $\Gamma_H$, taking
$(|\lambda_{HAA}|,\; |\lambda_{Hhh}|, \;|\lambda_{HH^+H^-}|)=(0,0,0),\; (5,0,0),\; (0,5,0),\; (0,0,5),$ and $(5,5,5)$.
The total $H$ decay rate grows from around 30 GeV when the three cubic scalar couplings are zero, up to approximately 150 GeV when their values are $(5,5,5)$.
\begin{figure}[t]
\centering
\includegraphics[scale=0.55]{./Figures/sigma1.pdf} \qquad \includegraphics[scale=0.54]{./Figures/K30.pdf}
\caption{\it LO production cross section $\sigma(pp\to H^+W^-)$ at $\sqrt{s}=14$ TeV (left), as a function of $M_{H^\pm}$, for $M_h=125$ GeV, $\cos\tilde\alpha=0.9$ and different values for the pair $(M_H,\; \Gamma_H)$ in GeV. The QCD K factor is shown (right) for $(M_H, \; \Gamma_H)=(400,30)$ GeV and different choices of $\mu_R$ and $\mu_F$.}
\label{cross1}
\end{figure}
Fig.~\ref{cross1} (left) shows the predicted LO production cross sections at $\sqrt{s}=14$ TeV,
for representative values of $M_H$ and $\Gamma_H$, which cover the range of possibilities we have just discussed:
$(M_H,\; \Gamma_H)= (150,10^{-3})$, $(150,50)$, $(200,1)$, $(200,80)$, $(400,30)$, and $(400,150)$ GeV.
The cross section is very small when both CP-even scalars are off-shell. For $M_H = 150$ GeV, $\sigma(pp\to H^+W^-)$ is roughly smaller than $10^{-3}$ pb.
With $M_H=200$ GeV and a large decay width $\Gamma_H=80$ GeV, the cross section stays below $10^{-2}$ pb; however,
with a smaller width $\Gamma_H=1$ GeV, the cross section is enhanced by approximately two orders of magnitude
(three orders of magnitude with respect to the previous cases), in the region where $M_H$ is on-shell ($M_{H^\pm} \lesssim 120$ GeV).
The most interesting case is when $M_H=400$ GeV, because the cross section gets enhanced by the on-shell $H$ pole, reaching higher values around 0.1~pb.
The QCD K factor for this $H$ mass and $\Gamma_H=30$~GeV is given in Fig.~\ref{cross1} (right), and it is practically constant in the whole range of $M_{H^\pm}$; it approximately corresponds to the K factor for the production of a SM Higgs with a 400 GeV mass. Its central value is around 1.9. A very similar K factor is obtained for $\Gamma_H=150$~GeV, although with a smaller cross section.
Thus, a heavy $H$ boson would be the most favourable situation from the experimental point of view, with production cross sections between $10^{-2}$ and 1 pb at $\sqrt{s}= 14$~TeV, depending on the value of $\Gamma_H$, which are potentially measurable at the LHC. As we have seen, they are increased by a factor of $\approx$ 2 by the NLO QCD corrections. For the other configurations both CP-even scalars are off-shell and the value of the cross section decreases by a few orders of magnitude, which results pretty challenging for the LHC, if not impossible. Nonetheless, these small values could turn out to be measurable in the future if the LHC luminosity is increased by a factor of 10, as planned for its High-Luminosity option.
\vskip .5cm\noindent {\bf B) $\mathbf{M_H=125}$ GeV.}
\begin{figure}[t]
\centering
\includegraphics[scale=0.54]{./Figures/sigma2.pdf} \qquad \includegraphics[scale=0.52]{./Figures/K125.pdf}
\caption{\it LO production cross section $\sigma(pp\to H^+W^-)$ at $\sqrt{s}=14$ TeV (left), as a function of $M_{H^\pm}$, for $M_H=125$ GeV, $\sin\tilde\alpha=0.99$ and $M_h=20$, $80$, $100$ GeV. The NLO QCD K factor (right) is shown for $M_h=20$ GeV and different choices of $\mu_R$ and $\mu_F$.}
\label{cross2}
\end{figure}
In this case both CP-even neutral scalars are off-shell and their decay widths can be neglected (assuming they are small). The scalar mixing angle must be small enough to avoid the LEP constraints, thus we take $\sin\tilde\alpha=0.99$, as we have done before in the analysis of branching ratios. The mass of the light scalar will be set to $M_h=20$, $80$ and $100$ GeV. The predicted LO production cross sections at $\sqrt{s}=14$~TeV are shown in Fig.~\ref{cross2} (left). For the chosen values of $M_h$, they range in between $10^{-5}$ and $10^{-6}$ pb. These values are extremely small and lay below the experimental sensitivity attainable in the near future. This scenario is thus, the most challenging experimentally. The computed K factor, Fig~\ref{cross2} (right), has
a similar value to the one obtained in the previous scenario.
\section{Conclusions}
The recent discovery of a Higgs-like boson has confirmed the existence of a scalar sector, which so far seems compatible with the SM predictions.
As it is widely known, an enlarged scalar sector is not forbidden by the symmetries of the electroweak theory, and there exists a broad range of possibilities
satisfying all experimental constraints. The direct discovery of another scalar particle would represent a major break-through in
particle physics, opening a window into a new high-energy dynamics and providing priceless information on which type of extension,
amongst many theoretical models of the scalar sector, is preferred by Nature.
Here we have focused on a particular 2HDM scenario, characterized by a fermiophobic charged Higgs, which would have evaded all experimental searches
performed until now. It is a quite predictive case, since all Yukawa couplings are determined by the mixing among the neutral scalars.
We have assumed a CP-conserving scalar potential and have restricted our analysis to the range $M_{H^\pm}\in\left[ M_W, M_W+M_Z\right]$, so that
only a few decay modes are kinematically open. We have presented detailed formulae for the loop-induced decay $H^+ \to W^+\gamma$, which becomes very relevant
in this mass region, and for the tree-level three-body decays of the charged scalar. We have analyzed the parameter space of the model, in order to
characterize the possible values of the $H^\pm$ decay width and branching ratios, taking into account the constraints from LHC, LEP and flavour data.
The two most important production channels for a fermiophobic charged scalar have been investigated, including NLO QCD corrections:
the associated production with either a neutral scalar or a charged $W$; {\it i.e.}, $q_u\bar q_d\to H^+\varphi_i^0$ and
$gg\to H^+W^-$. The predicted cross sections are small in most of the parameter space, making the experimental search challenging,
but they become very sizeable ($\ge 10^{-3}$~pb) for large values of the mass of the heavy neutral scalar $H$. In some extreme cases, cross sections between
0.1 and 1 pb are obtained. Thus, the detection of a fermiophobic $H^\pm$ at the LHC seems plausible in the near future. The interesting features of this possible scenario should encourage specific experimental searches for such a particle in the LHC data.
\begin{appendix}
\section{Scalar Potential}
\label{app:potential}
In the Higgs basis, the most general scalar potential takes the form
\begin{eqnarray}\label{eq:potential}
V & = & \mu_1\; \Phi_1^\dagger\Phi_1\, +\, \mu_2\; \Phi_2^\dagger\Phi_2 \, +\, \left[\mu_3\; \Phi_1^\dagger\Phi_2 \, +\, \mu_3^*\; \Phi_2^\dagger\Phi_1\right]
\nonumber\\ & + & \lambda_1\, \left(\Phi_1^\dagger\Phi_1\right)^2 \, +\, \lambda_2\, \left(\Phi_2^\dagger\Phi_2\right)^2 \, +\,
\lambda_3\, \left(\Phi_1^\dagger\Phi_1\right) \left(\Phi_2^\dagger\Phi_2\right) \, +\, \lambda_4\, \left(\Phi_1^\dagger\Phi_2\right) \left(\Phi_2^\dagger\Phi_1\right)
\nonumber\\ & + & \left[ \left(\lambda_5\; \Phi_1^\dagger\Phi_2 \, +\,\lambda_6\; \Phi_1^\dagger\Phi_1 \, +\,\lambda_7\; \Phi_2^\dagger\Phi_2\right) \left(\Phi_1^\dagger\Phi_2\right)
\, +\, \mathrm{h.c.}\right]\, .
\end{eqnarray}
The hermiticity of the potential requires all parameters to be real except $\mu_3$, $\lambda_5$, $\lambda_6$ and $\lambda_7$; thus, there are 14 real parameters.
The minimization conditions
$\langle 0|\Phi_1^T(x)|0\rangle =\frac{1}{\sqrt{2}}\, (0, v)$ and $\langle 0|\Phi_2^T(x)|0\rangle =\frac{1}{\sqrt{2}}\, (0, 0)$
impose the relations
\bel{eq:minimum}
\mu_1\; =\; -\lambda_1\, v^2\, ,
\qquad\qquad\qquad
\mu_3\; =\; -\frac{1}{2}\,\lambda_6\, v^2\, .
\end{equation}
The potential can then be decomposed into a quadratic term plus cubic and quartic interactions
\bel{eq:potential2}
V\; =\; -\frac{1}{4}\,\lambda_1\, v^4\, +\, V_2 \, +\, V_3 \, +\, V_4\, .
\end{equation}
The mass terms take the form
\begin{eqnarray}\label{eq:mass_term}
V_2 & = &
M_{H^\pm}^2\, H^+ H^-\, +\, \frac{1}{2}\, \left(S_1, S_2, S_3\right)\; \mathcal{M}\; \left(\begin{array}{c} S_1\\ S_2\\ S_3\end{array}\right)
\nonumber\\[10pt] & = &
M_{H^\pm}^2\, H^+ H^-\, +\, \frac{1}{2}\, M_h^2\, h^2\, +\, \frac{1}{2}\, M_H^2\, H^2\, +\, \frac{1}{2}\, M_A^2\, A^2\, ,
\end{eqnarray}
with
\bel{eq:mplus}
M_{H^\pm}^2\; =\; \mu_2 + \frac{1}{2}\,\lambda_3\, v^2
\end{equation}
and
\bel{eq:mass_matrix}
\mathcal{M}\; =\; \left(\begin{array}{ccc}
2\lambda_1 v^2 & v^2\, \lambda_6^{\mathrm{R}} & -v^2\, \lambda_6^{\mathrm{I}}\\
v^2\, \lambda_6^{\mathrm{R}} & M_{H^\pm}^2 + v^2\left(\frac{\lambda_4}{2} + \lambda_5^{\mathrm{R}}\right)
& -v^2\, \lambda_5^{\mathrm{I}}\\
-v^2\, \lambda_6^{\mathrm{I}} & -v^2\, \lambda_5^{\mathrm{I}} & M_{H^\pm}^2 + v^2\left(\frac{\lambda_4}{2} - \lambda_5^{\mathrm{R}}\right)
\end{array}\right)\, ,
\end{equation}
where $\lambda_i^{\mathrm{R}}\equiv \mathrm{Re}(\lambda_i)$ and $\lambda_i^{\mathrm{I}}\equiv \mathrm{Im}(\lambda_i)$.
The symmetric mass matrix $\mathcal{M}$ is diagonalized by an orthogonal matrix $\mathcal{R}$, which defines the neutral mass eigenstates:
\bel{eq:mass_diagonalization}
\mathcal{M}\; =\; \mathcal{R}^T\; \mathcal{M}_D \; \mathcal{R}\, ,
\qquad\qquad\qquad
\varphi^0 \; =\; \mathcal{R}\; S \, ,
\end{equation}
where we have introduced the shorthand matrix notation
\bel{eq:mass_diagonalization}
\mathcal{M}_D \; \equiv \; \left(\begin{array}{ccc} M_h^2 & 0 & 0 \\ 0 & M_H^2 & 0 \\ 0 & 0 & M_A^2
\end{array}\right)\; \, ,
\qquad\qquad
\varphi^0 \; \equiv \; \left(\begin{array}{c} h\\ H\\ A\end{array}\right)\;
\; \, ,
\qquad\qquad S \; \equiv \; \left(\begin{array}{c} S_1\\ S_2\\ S_3\end{array}\right)\, .
\end{equation}
Since the trace remains invariant, the masses satisfy the relation
\bel{eq:mass_sum}
M_h^2 \, +\, M_H^2 \, +\, M_A^2\; =\; 2\, M_{H^\pm}^2\, +\, v^2\,\left(2\,\lambda_1 +\lambda_4\right)\, .
\end{equation}
The minimization conditions allow us to trade the parameters $\mu_1$ and $\mu_3$ by $v$ and $\lambda_6$. The freedom to rephase the field $\Phi_2$ implies,
moreover, that only the relative phases among $\lambda_5$, $\lambda_6$ and $\lambda_7$ are physical; but only two of them are independent. Therefore,
we can fully characterize the potential with 11 parameters: $v$, $\mu_2$, $|\lambda_{1,\ldots,7}|$,
$\mathrm{arg}(\lambda_5\lambda_6^*)$ and $\mathrm{arg}(\lambda_5\lambda_7^*)$. Four parameters can be determined through the physical scalar masses \cite{ilisie1}. The matrix equation
\begin{align}
(\mathcal{M} \; \mathcal{R}^T - \mathcal{R}^T \; \mathcal{M}_D) \;
= \; 0
\end{align}
relates the scalar masses and mixings. Summing the second row with $(-i)$ times the third row, one obtains the identity (imaginary parts included):
\begin{align}
v^2 \lambda_6 \mathcal{R}_{i1} + \Big[ M_{H^\pm}^2 - M_{\varphi_i^0}^2 + v^2 \Big( \frac{\lambda_4}{2} + \lambda_5 \Big) \Big]\,
(\mathcal{R}_{i2} - i \mathcal{R}_{i3}) + 2 iv^2\lambda_5 \mathcal{R}_{i3}\; =\; 0\, .
\label{2+3rows}
\end{align}
This proves in full generality that
\begin{align}
(\mathcal{R}_{i2}-i\mathcal{R}_{i3})\; \frac{M_{\varphi_i^0}^2-M_{H^\pm}^2}{v^2}\; =\; (\mathcal{R}_{i2}-i\mathcal{R}_{i3})\,
\Big( \frac{\lambda_4}{2} + \lambda_5 \Big) + 2 i \mathcal{R}_{i3} \lambda_5 + \mathcal{R}_{i1} \lambda_6
\; =\; \lambda_{H^+ G^-\varphi_i^0}\, .
\end{align}
Taking instead the first row, one gets:
\begin{align}
\big( 2 \lambda_1 v^2 - M_{\varphi_i^0}^2\big) \; \mathcal{R}_{i1} + v^2 \lambda_6^{\text{R}} \mathcal{R}_{i2} - v^2 \lambda_6^{\text{I}} \mathcal{R}_{i3}\; =\; 0\, ,
\label{firstrow}
\end{align}
which generalizes the usual relation determining $\tan\tilde\alpha$ in the CP-conserving limit ($\mathcal{R}_{13}=\mathcal{R}_{23}=0$). It also proves that the following identity holds in general
\begin{align}
\frac{M_{\varphi_i^0}^2}{v^2} \; \mathcal{R}_{i1} \; = \; 2 R_{i1} \lambda_1 + i \mathcal{R}_{i3} \lambda_6 + (\mathcal{R}_{i2}-i \mathcal{R}_{i3})\lambda_6^{\text{R}}\; =\; \lambda_{G^+G^-\varphi_i^0}\, .
\end{align}
Here, similarly to Eq.~\eqn{hHPHM}, we have parametrized the Goldstone terms of $V_3$ in the form
\begin{align}
\Big(\; v \; \lambda_{H^+ G^-\varphi_i^0} \; H^+ G^- \varphi_i^0 \; + \text{h.c.} \; \Big) \; +
\; v \; \lambda_{G^+ G^-\varphi_i^0} \; G^+ G^- \varphi_i^0 \; \; \subset \; V_3\, .
\end{align}
These identities generalize the ones from \cite{HaberCP}, that are valid only in the CP-conserving limit of the scalar potential. They turn out to be very useful if one works in $R_\xi$ gauges with a fully general potential.
Using again Eq.~(\ref{firstrow}), the orthogonality of $\mathcal{R}$ implies:
\begin{align}
\sum_i \mathcal{R}_{i1}^2\; M_{\varphi_i^0}^2\; =\; 2\lambda_1 v^2 \, ,
\qquad\;
\sum_i \mathcal{R}_{i1} \mathcal{R}_{i2} \; M_{\varphi_i^0}^2\; =\; \lambda_6^{\text{R}} v^2 \, ,
\qquad\;
\sum_i \mathcal{R}_{i1}\mathcal{R}_{i3} \; M_{\varphi_i^0}^2\; =\; - \lambda_6^{\text{I}} v^2 \, .
\label{ortogonal}
\end{align}
Eq.~(\ref{2+3rows}) gives the additional orthogonality relations.
\begin{align}
\sum_i \mathcal{R}_{i1} (\mathcal{R}_{i2}-i\mathcal{R}_{i3})\; M_{\varphi_i^0}^2 \; &= \; \lambda_6 v^2\, ,
\\
\sum_i \mathcal{R}_{i2} (\mathcal{R}_{i2}-i\mathcal{R}_{i3}) \; M_{\varphi_i^0}^2 \; &= \; M_{H^\pm}^2 + v^2 \Big( \frac{\lambda_4}{2} + \lambda_5 \Big)\, ,
\\
i\sum_i \mathcal{R}_{i3} (\mathcal{R}_{i2}-i\mathcal{R}_{i3}) \; M_{\varphi_i^0}^2 \; &= \; M_{H^\pm}^2 + v^2 \Big( \frac{\lambda_4}{2} - \lambda_5 \Big)\, .
\end{align}
The first identity reproduces in complex form the last two real equations in (\ref{ortogonal}). Separating the real and imaginary parts of the last two relations, one gets:
\begin{align}
& \sum_i \mathcal{R}_{i2}^2 \; M_{\varphi_i^0}^2 \; = \; M_{H^\pm}^2 + v^2 \Big( \frac{\lambda_4}{2} + \lambda_5^{\text{R}} \Big) \; , \\
& \sum_i \mathcal{R}_{i3}^2 \; M_{\varphi_i^0}^2 \; = \; M_{H^\pm}^2 + v^2 \Big( \frac{\lambda_4}{2} - \lambda_5^{\text{R}} \Big) \; , \\
& \sum_i \mathcal{R}_{i2} \mathcal{R}_{i3} \; M_{\varphi_i^0}^2 \; = - v^2 \lambda_5^{\text{I}} \; .
\end{align}
\subsection{Inert 2HDM}
\label{subsec:inert}
Imposing a discrete ${\cal Z}_2$ symmetry such that all SM fields remain invariant under a ${\cal Z}_2$ transformation, while
\bel{eq:Z2sym}
\Phi_1\;\to\; \Phi_1\, ,
\qquad\qquad
\Phi_2\;\to\; -\Phi_2\, ,
\end{equation}
one makes the second scalar doublet {\it inert}\/: linear interactions of $\Phi_2$ with the SM fields are odd under a ${\cal Z}_2$ transformation, and thus forbidden~\cite{Ma:2008uza,Ma:2006km}. In particular, $\Phi_2$ is fermiophobic. This inert scalar doublet can only interact with the other fields through quadratic couplings. The lightest neutral component of $\Phi_2$ is then a very good candidate for dark matter.
The ${\cal Z}_2$ symmetry implies a significant simplification of the scalar potential, because all terms with an odd number of $\Phi_2$ fields vanish:
$\mu_3=\lambda_6=\lambda_7=0$. Moreover, making an appropriate rephasing of $\Phi_2$, $\lambda_5$ can be taken real. Therefore, the neutral mass matrix \eqn{eq:mass_matrix} becomes diagonal and there is no mixing among the neutral scalars ($\mathcal{R} = I$). The neutral scalar masses are given by:
\bel{eq:InertMasses}
M^2_h\; =\; 2 \lambda_1 v^2\, ,
\qquad
M^2_H\; =\; M^2_{H^\pm} + \left(\frac{\lambda_4}{2}+\lambda_5\right)\, v^2\, ,
\qquad
M^2_A\; =\; M^2_{H^\pm} + \left(\frac{\lambda_4}{2}-\lambda_5\right)\, v^2\, .
\end{equation}
\section{Heavy neutral Higgs decay rates}
\label{app:decayrates}
In this section we are going to write down the tree-level on-shell two-body dominant decay rates of a heavy neutral Higgs. All the formulae presented here are, as in section \ref{sec:calc}, completely general (no assumptions are made on the Higgs potential and the A2HDM Yukawa structure is assumed). The decay rate of a neutral scalar to a pair of massive fermions is given by:
\begin{align}
\Gamma(\varphi_i^0 \to f \bar{f})\; =\; \frac{N_c^f \, m_f^2 \, M_{\varphi_i^0}}{8\,\pi\, v^2} \; \Big( 1-
\frac{4m_f^2}{M_{\varphi_i^0}^2} \Big)^{3/2} \; \Big[ \; \text{Re}\big( y_f^{\varphi_i^0} \big)^2 + \text{Im}\big( y_f^{\varphi_i^0} \big)^2 \; \Big(1-\frac{4m_f^2}{M_{\varphi_i^0}^2} \Big)^{-1} \; \Big]\, ,
\end{align}
where $N_c^f$ is 1 for leptons and 3 for quarks. The decay into two gauge bosons reads ($V=W,Z$)
\begin{align}
\Gamma(\varphi_i^0 \to VV)\; =\; \mathcal{R}_{i1}^2 \;\,
\frac{M_{\varphi_i^0}^3 \; \delta_V}{32 \, \pi \, v^2 }
\; \Big( 1-\frac{4M_V^2}{M_{\varphi_i^0}^2} \Big)^{1/2}
\Big( 1 - \frac{4 M_V^2}{M_{\varphi_i^0}^2} + \frac{12 M_V^4}{M_{\varphi_i^0}^4} \Big)\, ,
\end{align}
with $\delta_Z=1$ and $\delta_W=2$. Other channels that can bring important contributions are $\varphi_i^0 \to \varphi_j^0 \varphi_j^0$ and $\varphi_i^0 \to H^+H^-$. The corresponding decay widths are given by
\begin{align}
\Gamma(\varphi_i^0 \to \varphi_j^0 \varphi_j^0) &\; =\; \frac{v^2\; \lambda^2_{\varphi_i^0\varphi_j^0\varphi_j^0}}{32 \, \pi \, M_{\varphi_i^0}}\;
\Big( 1- \frac{4M_{\varphi_j^0}^2}{M_{\varphi_i^0}^2} \Big)^{1/2}\, , \\
\Gamma(\varphi_i^0 \to H^+H^-) &\; =\; \frac{v^2\; \lambda^2_{\varphi_i^0 H^+ H^-}}{16 \, \pi \, M_{\varphi_i^0}}\;
\Big( 1- \frac{4M_{H^\pm}^2}{M_{\varphi_i^0}^2} \Big)^{1/2}\, ,
\end{align}
where, for the charged Higgs interaction Lagrangian we have used the parametrization given in (\ref{hHPHM}) and we have parametrized the cubic interaction of the neutral Higgs fields as
\begin{equation}
{\cal L}_{\varphi_i^0 \varphi_j^0 \varphi_j^0 }\; =\; - \frac{v}{2} \; \lambda_{\varphi^0_i \varphi^0_j \varphi^0_j }\;\, \varphi^0_i\, \varphi^0_j\,\varphi^0_j\,\, .
\label{hHPHM2}
\end{equation}
Explicit expressions for these couplings can be found in \cite{ilisie1}. Here we didn't consider the off-shell $\varphi_i^0 \to \varphi_j^{0*} \varphi_j^{0*}$ decay mode because in addition to its kinematical suppression it also depends on the unknown parameter
$\lambda_{\varphi^0_i \varphi^0_j \varphi^0_j }$ and would not bring useful information. The last two processes that must be taken into account are
$\varphi_i^0\to \varphi_j^0 Z$ and $\varphi_i^0 \to H^+ W^-$. We have
\begin{align}
\Gamma(\varphi_i^0\to \varphi_j^0 Z) &\; =\;
\left( \mathcal{R}_{i3}\,\mathcal{R}_{j2}-\mathcal{R}_{i2}\,\mathcal{R}_{j3}\right)^2
\; \frac{1}{16\,\pi\, v^2 M_{\varphi_i^0}^3}\;
\lambda^{3/2}( M_{\varphi_i^0}^2, M_{\varphi_j^0}^2,M_Z^2)\, , \\
\Gamma(\varphi_i^0\to H^+ W^-) &\; =\; (\mathcal{R}_{i2}^2 + \mathcal{R}_{i3}^2) \;
\frac{1}{16\,\pi\, v^2 M_{\varphi_i^0}^3}\;
\lambda^{3/2}( M_{\varphi_i^0}^2, M_{H^\pm}^2,M_W^2) \, .
\end{align}
Again, the scalar couplings to gauge bosons are taken from \cite{ilisie1}.
\section{QCD corrections to $\mathbf{pp\to H^+\boldsymbol{\varphi}_i^0}$}
\label{QCDcorrections}
For the $H^+\varphi_i^0$ associated production, we write the LO hadronic cross section as
\begin{equation}
\sigma_{\text{LO}} \; =\; \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x} \; \sum_{q_u,\bar{q}_d} \; \Big[ \;
q_u(x,\mu_F) \, \bar{q}_d(\tau/x,\mu_F) + \bar{q}_d(x,\mu_F) \, q_u(\tau/x,\mu_F)
\; \Big] \; \hat{\sigma}_{\text{LO}}(\hat{s}=\tau s)\, ,
\label{crossPDFs}
\end{equation}
where we have introduced the shorthand notation $\hat{\sigma}_{\text{LO}} \equiv \hat{\sigma}(q_u\bar{q}_d\to H^+\varphi_i^0)$, for the partonic cross section given in Eq.~(\ref{drell-yan}). As usual, the partonic invariant-mass $\hat{s}$ must be expressed as a fraction of the hadronic center-of-mass energy $s$, that is $\hat{s}=\tau s$. The lower integration limit is given by $\tau_0=(M_{H^\pm}+M_{\varphi_i^0})^2/s$. The PDFs $q_i(x,\mu_F)$, for a given quark flavour `i', depend on the momentum fraction $x$ and the factorization scale $\mu_F$.
The NLO cross section, that includes first-order QCD corrections, can be cast in the simple form \cite{DJ2,DJ1}
\begin{equation}
\sigma_{\text{NLO}}\; =\; \sigma_{\text{LO}} + \Delta\sigma_{q\bar{q}} + \Delta\sigma_{qg}\, ,
\end{equation}
where $\Delta\sigma_{q\bar{q}}$ and $\Delta\sigma_{qg}$ are given by
\begin{align}
\Delta\sigma_{q\bar{q}} &\; =\; \frac{\alpha_s(\mu_R)}{\pi}\, \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x}\;\sum_{q_u,\bar{q}_d}
\; \Big[ \,
q_u(x,\mu_F) \, \bar{q}_d(\tau/x,\mu_F) + \bar{q}_d(x,\mu_F) \, q_u(\tau/x,\mu_F) \,
\Big]
\notag \\
& \qquad \times \; \int_{\tau_0/\tau}^1 \; dz \; \hat{\sigma}_{\text{LO}} (\tau s z) \; \omega_{q\bar q}(z)
\, , \\[2ex]
\Delta\sigma_{qg} &\; =\; \frac{\alpha_s(\mu_R)}{\pi}\, \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x}\;\sum_{q_u,\bar{q}_d}
\; \Big[ \,
q_u(x,\mu_F) \, g(\tau/x,\mu_F) + g(x,\mu_F) \, q_u(\tau/x,\mu_F)
\notag \\
& \qquad + \; \bar{q}_d(x,\mu_F) \, g(\tau/x,\mu_F) + g(x,\mu_F) \, \bar q_d(\tau/x,\mu_F) \, \Big] \; \int_{\tau_0/\tau}^1 \; dz \; \hat{\sigma}_{\text{LO}} (\tau s z) \; \omega_{q g}(z) \, ,
\end{align}
with $\mu_R$ the renormalization scale and
\begin{align}
\omega_{q\bar{q}}(z) \; &= \; -P_{qq}(z)\;\log\Big(\frac{\mu_F^2}{\tau s}\Big)\, +\, \frac{4}{3}\,\Big[
\Big( \frac{\pi^2}{3} -4 \Big)\,\delta(1-z) + 2\, (1+z^2)\,\Big( \frac{\log(1-z)}{1-z} \Big)_{+}
\Big] \, , \notag \\[1.5ex]
\omega_{qg}(z) \; &= \; -\frac{1}{2}\, P_{qg}(z)\,\log\Big(\frac{\mu_F^2}{(1-z)^2\tau s}\Big)\, +\, \frac{1}{8}\,\Big[
1 + 6 z - 7 z^2
\Big]\, .
\end{align}
The Altarelli-Parisi splitting functions $P_{qq}$ and $P_{qg}$ are given by
\begin{align}
P_{qq}(z)\; =\; \frac{4}{3}\,\Big[ \frac{1+z^2}{(1-z)_{+}} + \frac{3}{2}\,\delta(1-z) \Big]\, , &&&
P_{qg}(z)\; =\; \frac{1}{2}\,\Big[z^2 + (1-z)^2 \Big] \, ,
\end{align}
where $F_+$ is the `$+$' distribution defined as $F_+(x)\, =\, F(x)-\delta(1-x)\int_0^1 dx'\, F(x')$, and
\begin{equation}
\int_a^1 dz \; g(z) \; \Big( \frac{f(z)}{1-z} \Big)_+ \;\equiv\; \int_a^1 dz \; \Big( g(z)-g(1) \Big)\;\frac{f(z)}{1-z}\, -\, g(1)\int_0^a dz\; \frac{f(z)}{1-z}\, .
\end{equation}
\section{QCD corrections to $\mathbf{pp\to H^+W^-}$}
\label{QCDcorrections2}
The LO hadronic production cross section for the dominant gluon-fusion channel (in the heavy top-mass approximation) can be cast in the simple form
\begin{equation}
\sigma_{\text{LO}} \; =\; \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x} \; g(x,\mu_F) \, g(\tau/x,\mu_F) \; \hat{\sigma}_{\text{LO}}(\hat{s}=\tau s)\, ,
\label{crossPDFs2}
\end{equation}
where $\hat{\sigma}_{\text{LO}}$ stands for the partonic cross section $\hat{\sigma}(gg\to H^+ W^-)$, given in Eq.~(\ref{sigmagg}), and $\tau_0=(M_{H^\pm}+M_W)^2/s$. At the NLO, the cross section can be written as \cite{DJ2,DJ1}
\begin{equation}
\sigma_{\text{NLO}}\; =\; \sigma_{\text{LO}} + \Delta\sigma_{gg}^{\text{virt}}
+ \Delta\sigma_{gg} + \Delta\sigma_{q\bar{q}} + \Delta\sigma_{gq}\, ,
\end{equation}
where:
\begin{align}
\Delta\sigma_{gg}^{\text{virt}} &\; =\;
\frac{\alpha_s(\mu_R)}{\pi} \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x} \; g(x,\mu_F) \, g(\tau/x,\mu_F) \; \hat{\sigma}_{\text{LO}}(\tau s) \; \omega_{gg}^{\text{virt}} \, ,
\\[2ex]
\Delta\sigma_{gg} &\; =\;
\frac{\alpha_s(\mu_R)}{\pi} \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x} \; g(x,\mu_F) \, g(\tau/x,\mu_F) \; \int_{\tau_0/\tau}^1 \frac{dz}{z} \; \hat{\sigma}_{\text{LO}}(\tau s z) \; \omega_{gg}(z) \, ,
\\[2ex]
\Delta\sigma_{gq} &\; =\; \frac{\alpha_s(\mu_R)}{\pi} \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x}\; \sum_{q,\bar{q}}
\; \Big[ \;
q(x,\mu_F) \, g(\tau/x,\mu_F) + g(x,\mu_F) \, q(\tau/x,\mu_F)
\notag \\
& \qquad + \; \bar{q}(x,\mu_F) \, g(\tau/x,\mu_F) + g(x,\mu_F) \, \bar q(\tau/x,\mu_F) \; \Big] \; \int_{\tau_0/\tau}^1 \; \frac{dz}{z} \; \hat{\sigma}_{\text{LO}} (\tau s z) \; \omega_{gq}(z) \, ,
\end{align}
\begin{align}
\Delta\sigma_{q\bar{q}} &\; =\; \frac{\alpha_s(\mu_R)}{\pi} \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x}\;\sum_{q,\bar{q}}
\; \Big[ \;
q(x,\mu_F) \, \bar{q}(\tau/x,\mu_F) + \bar{q}(x,\mu_F) \, q(\tau/x,\mu_F) \; \Big]
\notag \\
& \qquad \times \; \int_{\tau_0/\tau}^1 \; \frac{dz}{z} \; \hat{\sigma}_{\text{LO}} (\tau s z) \; \frac{32}{27}(1-z)^3 \, ,
\end{align}
with the functions $\omega_{gg}^{\text{virt}}$, $\omega_{gg}$ and $\omega_{gq}$ given by
\begin{align}
\omega_{gg}^{\text{virt}} &\; =\; \pi^2 + \frac{11}{2} + \frac{33-2N_f}{6}\,\log \Big(\frac{\mu_R^2}{\tau s} \Big)
\, , \notag \\[1.5ex]
\omega_{gg} &\; =\; - z\, P_{gg}(z) \, \log \Big(\frac{\mu_F^2}{\tau s}\Big) - \frac{11}{2}\, (1-z)^3
+ 12\, \Big(\frac{\log(1-z)}{1-z}\Big)_{+} - 12\, z\, (2-z+z^2)\,\log(1-z) \, , \notag \\[1.5ex]
\omega_{gq} &\; =\; -\frac{z}{2}\, P_{gq}(z) \, \log \Big(\frac{\mu_F^2}{\tau s \, (1-z)^2}\Big) -1+2\, z-\frac{1}{3}\, z^2 \, ,
\end{align}
where $P_{gg}$ and $P_{gq}$ are the Altarelli-Parisi splitting functions
\begin{align}
P_{gg}(z) &\; =\; 6\, \Big[ \Big(\frac{1}{1-z}\Big)_{+} + \frac{1}{z} -2 + z\, (1-z) \Big] + \frac{33-2N_f}{6}\, \delta(1-z) \, ,
\notag \\[1.5ex]
P_{gq}(z) &\; =\; \frac{4}{3} \; \frac{1+(1-z)^2}{z}\, .
\end{align}
\end{appendix}
\section*{Acknowledgements}
This work has been supported in part by the Spanish
Government and ERDF funds from the EU Commission
[Grants FPA2011-23778 and CSD2007-00042
(Consolider Project CPAN)] and by Generalitat
Valenciana under Grant No. PROMETEOII/2013/007. The work of V.I. is supported by the Spanish Ministry MINECO through the FPI grant BES-2012-054676.
|
\section*{\Large\bf Supplementary Information}
\begin{center}
{\large \bf Chiral spin superfluidity and spontaneous spin Hall effect of interacting bosons}
\end{center}
\section{Details of the Bogoliubov analysis}
\label{sec:BogDetails}
In this section, we give the details of our Bogoliubov analysis. The ground state energy corrections are explicitly calculated for
the $(++)$ and $(+-)$ states. Symmetry guarantees the same result for the $(--)$ and $(-+)$ states. Fluctuations
are included as
\begin{eqnarray}
\phi_{\uparrow \textbf{r}} &=& \sqrt{\rho_\uparrow} e^{ i\textbf{K} \cdot \textbf{r}}
+ \int _\textbf{k} \phi_\uparrow(\textbf{k}) e^{i\textbf{k} \cdot \textbf{r}} \nonumber \\
\phi_{\downarrow \textbf{r}} &=& \sqrt{\rho_\downarrow} e^{ip\textbf{K} \cdot \textbf{r}}
+ \int _\textbf{k} \phi_\downarrow (p\textbf{k}) e^{ip\textbf{k} \cdot \textbf{r}},
\end{eqnarray}
with $p=$ $+$ and $-$ for chiral charge and spin states, respectively.
The effective Bogoliubov Hamiltonian controlling these fluctuations is
\begin{eqnarray}
H _{\rm eff} = \frac{1}{2} \int _\textbf{k} \Psi ^\dag (\textbf{k}) {\cal H} (\textbf{k} ) \Psi (\textbf{k}) + const,
\label{eq:SHeff}
\end{eqnarray}
where $\Psi (\textbf{k}) =
[ \phi_\uparrow (\textbf{k}_{1\uparrow} ), \phi_\uparrow ^\dag (\textbf{k}_{2\uparrow} ),
\phi_\downarrow (\textbf{k}_{1\downarrow} ), \phi_\downarrow ^\dag (\textbf{k}_{2\downarrow}) ]^T $ and
\begin{equation}
{\cal H } ( \textbf{k}) = \left [
\begin{array}{cc}
M_{\uparrow \uparrow} (\textbf{k}_{1\up}, \textbf{k}_{2\up} ) & M_{\uparrow \downarrow} (\textbf{k}) \\
M_{\downarrow \uparrow} (\textbf{k}) & M_{\downarrow \downarrow} (\textbf{k}_{1\downarrow}, \textbf{k}_{2\downarrow} )
\end{array}
\right] ,
\label{eq:Hkmat}
\end{equation}
with $(\textbf{k}_{1\up}, \textbf{k}_{2\up}, \textbf{k}_{1\downarrow}, \textbf{k}_{2\downarrow})$
defined to be $(\textbf{k}, 2\textbf{K} -\textbf{k}, \textbf{k}, 2\textbf{K} - \textbf{k})$ and
$(\textbf{k}, 2\textbf{K} - \textbf{k}, -2 \textbf{K} + \textbf{k}, -\textbf{k})$ for chiral charge and spin
states, respectively. These $M_{\sigma \sigma'}$ matrices are
\begin{eqnarray}
&& M_{\sigma \sigma} (\textbf{k}_{1\sigma}, \textbf{k}_{2 \sigma} ) \nonumber \\
&=& \left[
\begin{array}{cc}
\epsilon (\textbf{k}_{1\sigma} ) + \rho_\sigma U_{\sigma \sigma} (\textbf{K} - \textbf{k})
& \rho_{\sigma} U_{\sigma \sigma} (\textbf{k} - \textbf{K}) \\
\rho_\sigma U_{\sigma \sigma}^* (\textbf{k} - \textbf{K}), & \epsilon (\textbf{k}_{2\sigma})
+ \rho_\sigma U_{\sigma \sigma} (\textbf{K} - \textbf{k})
\end{array}
\right] \\
&& M_{\uparrow \downarrow} (\textbf{k}) \nonumber \\
& =& \left[
\begin{array}{cc}
U_{\up \downarrow} (\textbf{k} - \textbf{K}) \sqrt{\rho_\uparrow \rho_\downarrow} & U_{\up \downarrow} (\textbf{k} - \textbf{K}) \sqrt{\rho_\uparrow \rho_\downarrow} \\
U_{\up \downarrow} (\textbf{k} - \textbf{K}) \sqrt{\rho_\uparrow \rho_\downarrow} & U_{\up \downarrow} (\textbf{k} - \textbf{K}) \sqrt{\rho_\uparrow \rho_\downarrow}
\end{array}
\right] .
\end{eqnarray}
From Bogoliubov Hamiltonian in Eq.~\eqref{eq:Hkmat}, we treat the spin-mixing part $M_{\uparrow \downarrow} $
as a perturbation which is well justified in the weakly interacting limit (see Fig.~\ref{fig:Esplit}a).
We thus write $H_{\rm eff} = H_{\rm eff} ^{(0)} + H_{\rm eff} ^{(1)}$, with $H_{\rm eff} ^{(0)}$ block
diagonal in the spin space. The leading part is readily diagonalized in terms of
$\tilde{\Psi} (\textbf{k}) =
[ \tilde{ \phi}_\uparrow (\textbf{k}_{+\uparrow} ), \tilde{\phi}_\uparrow ^\dag ( \textbf{k}_{-\up } ),
\tilde{\phi}_\downarrow (\textbf{k}_{+\downarrow}), \tilde{\phi}_\downarrow ^\dag ( \textbf{k}_{-\downarrow }) ]^T $, with
\begin{eqnarray}
&& \tilde{\phi}_\sigma(\textbf{k}_{+,\sigma})
= u_\sigma (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) \phi_\sigma (\textbf{k}_{1\sigma} )
+ v_\sigma (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) \phi_\sigma ^\dag (\textbf{k}_{2\sigma} ), \nonumber \\
&& \tilde{\phi}_\sigma (\textbf{k}_{-,\sigma})
= v_\sigma (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) \phi_\sigma (\textbf{k}_{1\sigma} )
+ u_\sigma (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) \phi_\sigma ^\dag (\textbf{k}_{2\sigma} ) .
\end{eqnarray}
The coefficients are determined to be
\begin{eqnarray}
&& v_\sigma ^2 (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) = u_\sigma ^2 (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) -1 \nonumber \\
&&= \frac{1}{2} \left[ \frac{ \overline{\epsilon} (\textbf{k}_{1\sigma} , \textbf{k}_{2\sigma} ) + \rho_\sigma U_{\sigma \sigma} (\textbf{k} -\textbf{K}) }
{\varepsilon_\sigma (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma} )}
-1 \right], \nonumber
\end{eqnarray}
with
\begin{eqnarray}
\varepsilon_\sigma ^2 (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma})
&=& \overline{\epsilon} (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma} )
\left[ \overline{\epsilon} (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) + 2\rho_\sigma U_{\sigma \sigma} (\textbf{K} -\textbf{k}) \right]\nonumber \\
\overline{\epsilon} ( \textbf{k}_{1\sigma} , \textbf{k}_{2\sigma} ) &=&
\left( \epsilon(\textbf{k}_{1\sigma}) + \epsilon (\textbf{k}_{2\sigma} ) \right) /{2}.\nonumber
\end{eqnarray}
The Bogoliubov spectra are
\begin{eqnarray}
\xi_{\pm , \sigma}
= \varepsilon_\sigma (\textbf{k}_{1\sigma}, \textbf{k}_{2\sigma}) \pm \Delta \epsilon (\textbf{k}_{1\sigma} , \textbf{k}_{2\sigma})
\end{eqnarray}
with
$$
\Delta \epsilon (\textbf{k}_{1\sigma} , \textbf{k}_{2\sigma}) =(\epsilon(\textbf{k}_{1\sigma}) - \epsilon(\textbf{k}_{2\sigma} ))/{2}.
$$
Under the condition $U_{\sigma \sigma} (\textbf{k}) >0$ already assumed , we have $\xi_{\pm, \sigma} >0$,
which means the system is stable~\cite{2013_Stamper_Kurn_RMP,2008_Pethick_Smith}.
Then $H_{\rm eff}^{(0)}$ takes a diagonal form
$
H_{\rm eff} ^{(0)} = \frac{1}{2}\int_\textbf{k}
\sum_p \xi_{p ,\sigma} \tilde{\phi}_{\sigma} ^\dag (\textbf{k}_{p, \sigma}) \tilde{\phi}_{\sigma} (\textbf{k}_{p, \sigma})
+ { E} ^{(0)},
$
with ${ E} ^{(0)}$,
\begin{eqnarray}
&& \textstyle { E} ^{(0)}/N_s \nonumber \\
&=&\textstyle \frac{1}{2} \int _\textbf{k}
\left\{ -2 \bar{\epsilon} (\textbf{k}, 2\textbf{K} - \textbf{k}) +
\sum_\sigma
\varepsilon_\sigma (\textbf{k}, 2\textbf{K} - \textbf{k})
- \rho_\sigma U_{\sigma \sigma} (\textbf{K} - \textbf{k})
\right\}
\end{eqnarray}
the same for chiral charge and spin states.
Treating the spin mixing terms perturbatively, (see Sec.~\ref{sec:perttheory}) the ground state receives an energy correction
\begin{eqnarray}
\textstyle {E}^{(2)}/N_s
\textstyle & =&\textstyle -\frac{1}{2} {\rho_\up \rho_\downarrow} \int _\textbf{k}
\Gamma^2 (\textbf{k}_{1\up}, \textbf{k}_{2\up}, \textbf{k}_{1\downarrow},\textbf{k}_{2\downarrow} ) \nonumber \\
&& \times \textstyle \left\{\frac{1 }{\xi_{+\up} + \xi_{- \downarrow} }
+ \frac{1 }{\xi_{+ \downarrow} + \xi_{-\up}} \right\},
\end{eqnarray}
with
\begin{eqnarray}
&& \Gamma (\textbf{k}_{1} , \textbf{k}_{2}, \textbf{k}_{3},\textbf{k}_{4} ) \nonumber \\
&=& \textstyle | U_{\up \downarrow} ( \textbf{k} - \textbf{K}) |^2
\left(u_\up (\textbf{k}_{1}, \textbf{k}_{2} ) - v_\up (\textbf{k}_{1}, \textbf{k}_{2} ) \right)
\left( u_\downarrow (\textbf{k}_{3}, \textbf{k}_{4} ) - v_\downarrow (\textbf{k}_{3}, \textbf{k}_{4}) \right) .
\end{eqnarray}
Introducing
\begin{equation}
g(\textbf{k}) = \Gamma (\textbf{k}, 2\textbf{K} -\textbf{k}, \textbf{k}, 2\textbf{K} -\textbf{k} ),
\end{equation}
for the chiral charge case we have
\begin{eqnarray}
\textstyle { E}_{\chi_c} ^{(2)} /N_s =
\textstyle -\frac{1}{2} {\rho_\up \rho_\downarrow}
\int _\textbf{k} g^2(\textbf{k} )
\times \textstyle
\left\{ \frac{2 } {\varepsilon_\up (\textbf{k}, \textbf{Q} -\textbf{k})
+ \varepsilon_\downarrow(\textbf{k}, \textbf{Q} - \textbf{k}) }
\right\},
\label{eq:Echic}
\end{eqnarray}
with $\textbf{Q} = 2\textbf{K}$,
while for the chiral spin case we have
\begin{eqnarray}
&& \textstyle {E}_{\chi_s} ^{(2)} /N_s = -\frac{1}{2} {\rho_\up \rho_\downarrow}
\int _\textbf{k} g^2 (\textbf{k}) \nonumber \\
& \textstyle \times & \textstyle
\left\{ \frac{1}{\varepsilon_\up (\textbf{k}, \textbf{Q} -\textbf{k})
+ \varepsilon_\downarrow(-\textbf{Q} + \textbf{k}, -\textbf{k} )
+ \Delta \epsilon (\textbf{k}, \textbf{Q} - \textbf{k})
- \Delta \epsilon (-\textbf{Q} + \textbf{k}, -\textbf{k}) } \right. \nonumber \\
&& \left. \textstyle
+ \frac{1}{\varepsilon_\downarrow (-\textbf{Q} +\textbf{k}, -\textbf{k})
+ \varepsilon_\up (\textbf{k}, \textbf{Q} -\textbf{k})
+ \Delta \epsilon (-\textbf{Q} + \textbf{k}, -\textbf{k})
- \Delta \epsilon (\textbf{k}, \textbf{Q} - \textbf{k} ) } \right\}.
\label{eq:Echis}
\end{eqnarray}
We also calculate the energy correction by numerically diagonalizing the full Bogoliubov Hamiltonian (Eq.~\eqref{eq:Hkmat}), finding
excellent agreement with our analytic results when the inter-species interactions are weak (see Fig.~\ref{fig:Esplit}a).
\begin{figure}[htp]
\includegraphics[angle=0,width=0.5\linewidth]{BogSpectra.eps}
\caption{Bogoliubov spectra along the $k_x$ axis in the chiral spin superfluid state. In this plot we use $U/t = 1$ and $V/t = 0.3$. }
\label{fig:bogspectra}
\end{figure}
\section{Perturbation theory for Bogoliubov ground states}
\label{sec:perttheory}
In this section, we discuss the perturbative method to calculate the ground state energy of a Bogoliubov problem
$$
H_{\rm Bog} = \Psi^\dag {\cal H}_{\rm Bog} \Psi - {\cal H} _{\rm Bog} (2,2) - {\cal H}_{\rm Bog} (4,4),
$$
with $\Psi$ a column vector of bosonic operators
$[\phi_{\up 1}, \phi_{\up 2} ^\dag, \phi_{\downarrow 1}, \phi_{\downarrow 2} ^\dag] ^T$.
This Bogoliubov Hamiltonian is one momentum slice of Eq.~\eqref{eq:SHeff}
and the momentum ${\bf k}$ index is suppressed for brevity.
The $4\times4$ matrix ${\cal H}_{\rm Bog}$ can be rewritten as
\begin{eqnarray}
{\cal H}_{\rm Bog} = \left[
\begin{array}{cc}
M_\up &G \\
G^\dag &M_\downarrow
\end{array}
\right],
\end{eqnarray}
where the $2\times 2$ matrices can be expanded in terms of Pauli matrices,
$M_\sigma = c_{0\sigma} \mathbb{1} + c_{x\sigma} \sigma_x + c_{z \sigma} \sigma_z$, and
$G$ takes a special form $g (\mathbb{1} + \sigma_x)$. The terms $c_{0\sigma}$, $c_{x\sigma}$ and $c_{z\sigma}$ can be read off from Eq.~\eqref{eq:Hkmat}.
Here we will treat the off-diagonal part $G$ perturbatively.
The leading part is readily diagonalized as
\begin{eqnarray}
\textstyle H^{(0)} =\sum_ \sigma \textstyle \left[
\tilde{\phi}_{\sigma + }^\dag , \tilde{\phi} _{\sigma -}
\right]
D_\sigma
\left[
\begin{array}{c}
\tilde{\phi}_{\sigma + } \\
\tilde{\phi} _{\sigma -}^\dag
\end{array}
\right]
+ D_\sigma (2,2) - M_\sigma (2,2) \nonumber
\end{eqnarray}
with
\begin{eqnarray}
D_\sigma &=& \epsilon_\sigma + c_{z\sigma} \sigma_z \\
\epsilon_\sigma &=& \sqrt{c_{0\sigma}^2 - c_{x \sigma}^2} \nonumber
\end{eqnarray}
and
\begin{eqnarray}
\left[
\begin{array}{c}
\tilde{ \phi}_{\sigma +} \\
\tilde{\phi} _{\sigma -} ^\dag
\end{array}
\right] = T_\sigma
\left[
\begin{array}{c}
{ \phi}_{\sigma 1} \\
{\phi} _{\sigma 2} ^\dag
\end{array}
\right], \nonumber
\end{eqnarray}
\begin{eqnarray}
T_\sigma =
\left[
\begin{array}{cc}
u_\sigma &v_\sigma \\
v_\sigma &u_\sigma
\end{array}
\right] \nonumber
\end{eqnarray}
and
\begin{equation}
u_\sigma^2 = v_\sigma ^2 + 1 = \frac{1}{2}\left[ \frac{c_{0\sigma}}{\epsilon_\sigma} +1\right].
\end{equation}
The Bogoliubov spectra are
\begin{eqnarray}
\xi_{\sigma, \pm} = \epsilon_\sigma \pm c_{z \sigma}
\end{eqnarray}
In terms of $\tilde{\phi}$,
the perturbative part reads
\begin{eqnarray}
&& H^{(1)} =
g(u_\up - v_\up ) (u_\downarrow - v_\downarrow) \nonumber \\
&\times& \textstyle
\left\{
\tilde{ \phi}_{\up + } ^\dag \tilde{\phi}_{\downarrow +} + \tilde{\phi}_{\up -} \tilde{\phi}_{\downarrow -} ^\dag
+ \tilde{\phi} _{\up - } \tilde{\phi}_{\downarrow +} + \tilde{\phi}_{\up +} ^\dag \tilde{\phi}_{\downarrow -} ^\dag
\right\} + h.c.
\label{eq:SH1}
\end{eqnarray}
Then standard perturbation theory applies, and only the third and fourth terms in Eq.~\eqref{eq:SH1} contribute at second order.
The ground state energy is thus obtained to be
\begin{eqnarray}
&& { E}
=\sum_\sigma (\epsilon_\sigma - c_{0, \sigma} ) \\
&-& \left|g(u_\up - v_\up) (u_\downarrow -v_\downarrow) \right|^2
\times \left\{
\frac{1}{\xi_{\up +} + \xi_{\downarrow -} }
+ \frac{1}{\xi_{\up +} + \xi_{\downarrow -} } \right\}. \nonumber
\end{eqnarray}
\section{Large $U$ limit of the hexagonal lattice model}
\label{sec:largeU}
In the large $U$ limit, we can project out double occupancy, and the Gutzwiller state is
\begin{eqnarray}
|G \rangle
= \prod_\textbf{r} \left( f_{\up, \textbf{r},0} + f_{\up, \textbf{r}, 1} \phi_{\up, \textbf{r}} ^\dag \right)
\left( f_{\downarrow, \textbf{r}, 0} + f_{\downarrow, \textbf{r}+\hat{e}_1, 1} \phi_{\downarrow, \textbf{r}+\hat{e}_1} ^\dag \right)
|{\rm vac} \rangle, \nonumber
\end{eqnarray}
with a normalization condition $|f_{\sigma, \textbf{r}, 0}|^2 + |f_{\sigma, \textbf{r}, 1}|^2 =1$.
To minimize kinetic energy we take $f_{\up/\downarrow, \textbf{r}, 1} = f_{\up/\downarrow,1} e^{\pm i\textbf{K} \cdot \textbf{r}}$
and $f_{\sigma, \textbf{r}, 0} = \sqrt{1-f_{\sigma, 1}^2}$, where $f_{\sigma,1}$ is a real number. Then the energy
cost of the Gutzwiller state is
\begin{eqnarray}
&& E /N_s = \left[-\mu -3t_{\rm eff} \right] \left( f_{1\up} ^2 + f_{1\downarrow}^2 \right) \\
&+& \frac{3t_{\rm eff} + 3V/2}{2} \left( f_{1\up} ^2 + f_{1\downarrow}^2 \right) ^2
+ \frac{3t_{\rm eff} -3V/2} {2} \left( f_{1\up} ^2 - f_{1\downarrow}^2 \right) ^2 , \nonumber
\end{eqnarray}
which after minimization leads to
\begin{eqnarray}
\left\{
\begin{array}{cccc}
| f_{1\up} ^2 - f_{1\downarrow} ^2|&=&0 , &\text{if $t_{\rm eff} >V/2$} \\
| f_{1\up}^2 - f_{1\downarrow} ^2|& =& \frac{2\mu + 6t_{\rm eff} }{6 t_{\rm eff} + 3V}, &\text{otherwise}
\end{array}
\right. .
\end{eqnarray}
The transition from the unpolarized superfluid to the fully polarized state is at $V_c = 2t_{\rm eff}$ in this large $U$ limit,
where the transition is first order.
\end{document}
|
\section{Introduction}
In this work, we join and expand three threads of research in the analysis of modern finite element methods: polytope domain meshing, generalized barycentric coordinates, and families of finite-dimensional solution spaces characterized by finite element exterior calculus.
It is well-known that on simplicial meshes, standard barycentric coordinates provide a local basis for the lowest-order $H^1$-conforming scalar-valued finite element spaces, commonly called the Lagrange elements.
Further, local bases for the lowest-order vector-valued Brezzi-Douglas-Marini~\cite{BDM85}, Raviart-Thomas~\cite{RT1977}, and N{\'e}d{\'e}lec~\cite{BDDM87,N1980,N1986} finite element spaces on simplices can also be defined in a canonical fashion from an associated set of standard barycentric functions.
Here, we use generalized barycentric coordinates in an analogous fashion on meshes of convex polytopes, in dimensions 2 and 3, to construct local bases with the same global continuity and polynomial reproduction properties as their simplicial counterparts.
We have previously analyzed linear order, scalar-valued methods on polygonal meshes~\cite{GRB2011,RGB2011b} using four different types of generalized barycentric coordinates: Wachspress~\cite{W1975,W2011}, Sibson~\cite{F1990,S1980}, harmonic~\cite{C2008,JMRGS07,MKBWG2008}, and mean value~\cite{F2003,FHK2006,FKR2005}.
The analysis was extended by Gillette, Floater and Sukumar in the case of Wachspress coordinates to convex polytopes in any dimension~\cite{FGS2013}, based on work by Warren and colleagues~\cite{JSWD2005,W1996,WSHD2007}.
We have also shown how taking pairwise products of generalized barycentric coordinates can be used to construct quadratic order methods on polygons~\cite{RGB2011a}.
Applications of generalized barycentric coordinates to finite element methods have primarily focused on scalar-valued PDE problems~\cite{MP2008,RS2006,SM2006,ST2004,WBG07}.
\begin{table}
\centering
\begin{tabular}{c|c|c}
~n & k & functions\\
\hline
~2 & 0 & $\lambda_i$ \\
& 1 & $\lnl ij$\\
& & $\mathcal{W}_{ij}$ \\
& & $\textnormal{rot}~\lnl ij$ \\
& & $\textnormal{rot}~\mathcal{W}_{ij}$ \\
& 2 & $\lnldrnl ijk$ \\
& & $\mathcal{W}_{ijk}$ \\
\end{tabular}
\qquad
\qquad
\begin{tabular}{c|c|c}
~n & k & functions\\
\hline
~3 & 0 & $\lambda_i$ \\
& 1 & $\lnl ij$ \\
& & $\mathcal{W}_{ij}$ \\
& 2 & $\lnlxnl ijk$ \\
& & $\mathcal{W}_{ijk}$ \\
& 3 & $\lnldnlxnl ijk\ell$ \\
& & $\mathcal{W}_{ijk\ell}$ \\
\end{tabular}
\caption{For meshes of convex $n$-dimensional polytopes in $\R^n$, $n=2$ or $3$, computational basis functions for each differential form order $0\leq k\leq n$ are listed.
The notation is defined in Section~\ref{sec:bkgd}.
}
\label{tab:comp-bases}
\end{table}
Our expansion in this paper to vector-valued methods is inspired by Whitney differential forms, first defined in~\cite{W1957}.
Bossavit recognized that Whitney forms could be used to construct basis functions for computational electromagnetics~\cite{B1988a}.
The theory of finite element exterior calculus unified subsequent research in this area~\cite{AFW2006}.
In particular, Arnold, Falk and Winther showed how functions like those appearing in Table~\ref{tab:comp-bases} can be used to build spanning sets and bases for any the ${\mathcal P}_r\Lambda^k$ and ${\mathcal P}_r^-\Lambda^k$ spaces on simplices~\cite{AFW2009}.
The FENiCS Project~\cite{Aetal2015} has implemented these functions on simplices as part of a broadly applicable open source finite element software package.
Some prior work has explored the possibility of Whitney functions over non-simplicial elements in specific cases of rectangular grids~\cite{G2002}, square-base pyramids~\cite{GH1999}, and prisms~\cite{B2008}.
Other authors have examined the ability of generalized Whitney functions to recover constant-valued forms in certain cases~\cite{ESW2006,KRS2011}, whereas here we show their ability to reproduce \textbf{\textit{all}} the elements of the spaces denoted ${\mathcal P}_1^-\Lambda^k$ in finite element exterior calculus.
Gillette and Bajaj considered the use of generalized Whitney forms on polytope meshes defined by duality from a simplicial mesh~\cite{GB2010,GB2011}, which illustrated potential benefits to discrete exterior calculus~\cite{H2003}, computational magnetostatics, and Darcy flow modeling.
Recent work~\cite{MRS2014} has also shown generalized barycentric coordinates to be effective when used in tandem with virtual element methods~\cite{dVBCMMR2013}, which are developed in a similar fashion to traditional mimetic methods~\cite{LMS2014}.
\begin{table}[h]
\centering
\begin{tabular}{c|c|l|l}
n & k & global continuity & polynomial reproduction\\
\hline
2 & 0 & $H^1(\mesh)$ & ${\mathcal P}_1 \Lambda^0(\mesh)$ \\
& 1 & $H(\textnormal{curl\,},\mesh)$, by Theorem~\ref{thm:hcurl-conf} & ${\mathcal P}_1 \Lambda^1(\mesh)$, by Theorem~\ref{thm:pr-lnl} \\
& & $H(\textnormal{curl\,},\mesh)$, by Theorem~\ref{thm:hcurl-conf} & ${\mathcal P}_1^- \Lambda^1(\mesh)$, by Theorem~\ref{thm:pr-whit-ij}\\
& & $H(\textnormal{div\,},\mesh)$, see Remark~\ref{rmk:rot-for-cty} & ${\mathcal P}_1 \Lambda^1(\mesh)$, by Corollary~\ref{cor:pr-lrnl} \\
& & $H(\textnormal{div\,},\mesh)$, see Remark~\ref{rmk:rot-for-cty} & ${\mathcal P}_1^- \Lambda^1(\mesh)$, by Corollary~\ref{cor:pr-rotWhit} \\
& 2 & none (piecewise linear) & ${\mathcal P}_1 \Lambda^2(\mesh)$, by Theorem~\ref{thm:pr-lnldrnl} \\
& & none (piecewise constant) & ${\mathcal P}_1^- \Lambda^2(\mesh)$, see Remark~\ref{rmk:topdim} \\
\hline
3 & 0 & $H^1(\mesh)$ & ${\mathcal P}_1 \Lambda^0(\mesh)$ \\
& 1 & $H(\textnormal{curl\,},\mesh)$, by Theorem~\ref{thm:hcurl-conf} & ${\mathcal P}_1 \Lambda^1(\mesh)$, by Theorem~\ref{thm:pr-lnl} \\
& & $H(\textnormal{curl\,},\mesh)$, by Theorem~\ref{thm:hcurl-conf} & ${\mathcal P}_1^- \Lambda^1(\mesh)$, by Theorem~\ref{thm:pr-whit-ij}\\
& 2 & $H(\textnormal{div\,},\mesh)$, by Theorem~\ref{thm:hdiv-conf} & ${\mathcal P}_1 \Lambda^2(\mesh)$, by Theorem~\ref{thm:pr-lnlxnl}\\
& & $H(\textnormal{div\,},\mesh)$, by Theorem~\ref{thm:hdiv-conf} & ${\mathcal P}_1^- \Lambda^2(\mesh)$, by Theorem~\ref{thm:pr-whit-ijk}\\
& 3 & none (piecewise linear) & ${\mathcal P}_1 \Lambda^3(\mesh)$, see Remark~\ref{rmk:topdim} \\
& & none (piecewise constant) & ${\mathcal P}_1^- \Lambda^3(\mesh)$, see Remark~\ref{rmk:topdim} \\
\hline
\end{tabular}
\caption{Summary of the global continuity and polynomial reproduction properties of the spaces considered.}
\label{tab:results-summary}
\end{table}
Using the bases defined in Table~\ref{tab:comp-bases}, our main results are summarized in Table~\ref{tab:results-summary}.
On a mesh of convex $n$-dimensional polytopes in $\R^n$ with $n=2$ or $3$, we construct computational basis functions associated to the polytope elements for each differential form order $k$ as indicated.
Each function is built from generalized barycentric coordinates, denoted $\lambda_i$, and their gradients; formulae for the Whitney-like functions, denoted $\mathcal{W}$, are given in Section~\ref{subset:whit-forms}.
In the vector-valued cases ($0<k<n$), we prove that the functions agree on tangential or normal components at inter-element boundaries, providing global continuity in $H(\textnormal{curl})$ or $H(\textnormal{div})$.
The two families of polynomial differential forms that are reproduced, ${\mathcal P}_r\Lambda^k$ and ${\mathcal P}_r^-\Lambda^k$, were shown to recover and generalize the classical simplicial finite element spaces mentioned previously, via the theory of finite element exterior calculus~\cite{AFW2006,AFW2010}.
The outline of the paper is as follows.
In Section~\ref{sec:bkgd}, we describe relevant theory and prior work in the areas of finite element exterior calculus, generalized barycentric coordinates, and Whitney forms.
In Section~\ref{sec:global-cnty}, we show how the functions listed in Table~\ref{tab:comp-bases} can be used to build piecewise-defined functions with global continuity in $H^1$, $H(\textnormal{curl})$ or $H(\textnormal{div})$, as indicated.
In Section~\ref{sec:poly-repro}, we show how these same functions can reproduce the requisite polynomial differential forms from ${\mathcal P}_1\Lambda^k$ or ${\mathcal P}_1^-\Lambda^k$, as indicated in Table~\ref{tab:comp-bases}, by exhibiting explicit linear combinations whose coefficients depend only on the location of the vertices of the mesh.
In Section~\ref{sec:polyg-fams}, we count the basis functions constructed by our approach on generic polygons and polyhedra and explain how the size of the basis could be reduced in certain cases.
\section{Background and prior work}
\label{sec:bkgd}
\subsection{Spaces from Finite Element Exterior Calculus}
\label{subsec:feec}
Finite element spaces can be broadly classified according to three parameters: $n$, the spatial dimension of the domain, $r$, the order of error decay, and $k$, the differential form order of the solution space.
The $k$ parameter can be understood in terms of the classical finite element sequence for a domain $\Omega\subset\R^n$ with $n=2$ or $3$, commonly written as
\[\xymatrix @R=.02in{
n=2: & {H^1} \ar[rr]^-{\text{grad}} && {H(\textnormal{curl})} \ar@{<->}[rr]^-{\text{rot}} && {H(\textnormal{div})} \ar[rr]^-{\text{div}} && {L^2}\\
n=3: & {H^1} \ar[rr]^-{\text{grad}} && {H(\textnormal{curl})} \ar[rr]^-{\text{curl}} && {H(\textnormal{div})} \ar[rr]^-{\text{div}} && {L^2}
}\]
Note that for $n=2$, given $\vec F(x,y) := \twovec{F_1(x,y)}{F_2(x,y)}$, we use the definitions:
\[\textnormal{curl\,}\,\vec F := \frac{\partial F_1}{\partial y}- \frac{\partial F_2}{\partial x},\quad\textnormal{rot}\,\vec F := \begin{bmatrix} 0 & {-1} \\ 1 & 0 \end{bmatrix}\vec F\quad\text{and}\quad \textnormal{div\,}\vec F := \frac{\partial F_1}{\partial x}+ \frac{\partial F_2}{\partial y}.\]
\text{}\\
Thus, in $\R^2$, we have both $\textnormal{curl\,}\nabla\phi =0$ and $\textnormal{div\,}\textnormal{rot}\,\nabla\phi=0$ for any $\phi\in H^2$.
Put differently, $\textnormal{rot}$ gives an isomorphism from $H(\textnormal{curl})$ to $H(\textnormal{div})$ in $\R^2$.
In some cases we will write $H(\textnormal{curl\,},\Omega)$ and $H(\textnormal{div\,},\Omega)$ if we wish to emphasize the domain in consideration.
In the terminology of differential topology, the applicable sequence is described more simply as the $L^2$ deRham complex of $\Omega$.
The spaces are re-cast as differential form spaces $H\Lambda^k$ and the operators as instances of the exterior derivative $d_k$, yielding
\[\xymatrix @R=.02in{
n=2: & {H\Lambda^0} \ar[rr]^-{d_0} && {H\Lambda^1} \ar[rr]^-{d_1} && {H\Lambda^2} \\
n=3: & {H\Lambda^0} \ar[rr]^-{d_0} && {H\Lambda^1} \ar[rr]^-{d_1} && {H\Lambda^2} \ar[rr]^-{d_2} && {H\Lambda^3}
}\]
Finite element methods seek approximate solutions to a PDE in finite dimensional subspaces $\Lambda^k_h$ of the $H\Lambda^k$ spaces, where $h$ denotes the maximum diameter of a domain element associated to the subspace.
The theory of finite element exterior calculus classifies two families of suitable choices of $\Lambda^k_h$ spaces on meshes of simplices, denoted ${\mathcal P}_r\Lambda^k$ and ${\mathcal P}_r^-\Lambda^k$~\cite{AFW2006,AFW2010}.
The space ${\mathcal P}_r\Lambda^k$ is defined as ``those differential forms which, when applied to a constant vector field, have the indicated polynomial dependence''~\cite[p. 328]{AFW2010}.
This can be interpreted informally as the set of differential $k$ forms with polynomial coefficients of total degree at most $r$.
The space ${\mathcal P}_r^-\Lambda^k$ is then defined as the direct sum
\begin{equation}
\label{eq:prminus-decomp}
{\mathcal P}_r^-\Lambda^k := {\mathcal P}_{r-1}\Lambda^k \oplus \kappa{\mathcal H}_{r-1}\Lambda^{k+1},
\end{equation}
where $\kappa$ is the Koszul operator and ${\mathcal H}_r$ denotes homogeneous polynomials of degree $r$~\cite[p. 331]{AFW2010}.
We will use the coordinate formulation of $\kappa$, given in~\cite[p. 329]{AFW2010} as follows.
Let $\omega\in\Lambda^k$ and suppose that it can be written in local coordinates as $\omega_x=a(x)dx_{\sigma_1}\wedge\cdots\wedge dx_{\sigma_k}$.
Then $\kappa\omega$ is written as
\begin{equation}
\label{eq:def-kappa}
(\kappa\omega)_x := \sum_{i=1}^k (-1)^{i+1}a(x)x_{\sigma(i)}dx_{\sigma_1}\wedge\cdots\wedge\widehat{dx_{\sigma_i}}\wedge\cdots\wedge dx_{\sigma_k},
\end{equation}
where $\wedge$ denotes the wedge product and $\widehat{dx_{\sigma_i}}$ means that the term is omitted.
For example, let $n=3$ and write $x,y,z$ for $x_1,x_2,x_3$.
Then $dydz\in{\mathcal H}_0\Lambda^2$ and $\kappa dydz = ydz-zdy\in{\mathcal H}_1\Lambda^1$.
We summarize the relationship between the spaces ${\mathcal P}_1\Lambda^k$, ${\mathcal P}_1^-\Lambda^k$ and certain well-known finite element families in dimension $n=2$ or $3$ in Table~\ref{tab:fe-spaces}.
\begin{table}[ht]
\begin{center}
\begin{tabular}{c|c|c|c|l|c}
n & k & dim & space & classical description & reference\\ \hline
2 & 0 & 3 & ${\mathcal P}_1 \Lambda^0(\mathcal{T})$ & Lagrange, degree $\leq 1$ &\\
& & 3 & ${\mathcal P}_1^- \Lambda^0(\mathcal{T})$ & Lagrange, degree $\leq 1$ &\\
& 1 & 6 & ${\mathcal P}_1 \Lambda^1(\mathcal{T})$ & Brezzi-Douglas-Marini, degree $\leq 1$ & \cite{BDM85}\\
& & 3 & ${\mathcal P}_1^- \Lambda^1(\mathcal{T})$ & Raviart-Thomas, order $0$ & \cite{RT1977}\\
& 2 & 3 & ${\mathcal P}_1 \Lambda^2(\mathcal{T})$ & discontinuous linear &\\
& & 1 & ${\mathcal P}_1^- \Lambda^2(\mathcal{T})$ & discontinuous piecewise constant &\\ \hline
3 & 0 & 4 & ${\mathcal P}_1 \Lambda^0(\mathcal{T})$ & Lagrange, degree $\leq 1$ &\\
& & 4 & ${\mathcal P}_1^- \Lambda^0(\mathcal{T})$ & Lagrange, degree $\leq 1$ &\\
& 1 & 12 & ${\mathcal P}_1 \Lambda^1(\mathcal{T})$ & N{\'e}d{\'e}lec~second kind $H(\textnormal{curl})$, degree $\leq 1$ & \cite{N1986,BDDM87}\\
& & 6 & ${\mathcal P}_1^- \Lambda^1(\mathcal{T})$ & N{\'e}d{\'e}lec~first kind $H(\textnormal{curl})$, order $0$ & \cite{N1980}\\
& 2 & 12 & ${\mathcal P}_1 \Lambda^2(\mathcal{T})$ & N{\'e}d{\'e}lec~second kind $H(\textnormal{div})$, degree $\leq 1$ &\cite{N1986,BDDM87}\\
& & 4 & ${\mathcal P}_1^- \Lambda^2(\mathcal{T})$ & N{\'e}d{\'e}lec~first kind $H(\textnormal{div})$, order $0$ & \cite{N1980}\\
& 3 & 4 & ${\mathcal P}_1 \Lambda^3(\mathcal{T})$ & discontinuous linear & \\
& & 1 & ${\mathcal P}_1^- \Lambda^3(\mathcal{T})$ & discontinuous piecewise constant & \\ \hline
\end{tabular}
\end{center}
\caption{Correspondence between ${\mathcal P}_1\Lambda^k(\mathcal{T})$, ${\mathcal P}_1^-\Lambda^k(\mathcal{T})$ and common finite element spaces associated to a simplex $\mathcal{T}$ of dimension $n$.
Further explanation of these relationships can be found in~\cite{AFW2006,AFW2010}.
Our constructions, when reduced to simplices, recover known local bases for each of these spaces.}
\label{tab:fe-spaces}
\end{table}
A crucial property of ${\mathcal P}_r\Lambda^k$ and ${\mathcal P}_r^-\Lambda^k$ is that each includes in its span a sufficient number of polynomial differential $k$-forms to ensure an \textit{a priori} error estimate of order $r$ in $H\Lambda^k$ norm.
In the classical description of finite element spaces, this approximation power is immediate; any computational or `local' basis used for implementation of these spaces must, by definition, span the requisite polynomial differential forms.
The main results of this paper are proofs that generalized barycentric coordinates can be used as local bases on polygonal and polyhedral element geometries to create analogues to the lowest order ${\mathcal P}_r\Lambda^k$ and ${\mathcal P}_r^-\Lambda^k$ spaces with the same polynomial approximation power and global continuity properties.
In the remainder of the paper, we will frequently use standard vector proxies~\cite{AMR1988} in place of differential form notation, as indicated here:
\begin{align*}
\twovecT {u_1}{u_2} & \;\;\longleftrightarrow \;\; u_1dx_1+u_2dx_2\in\Lambda^1(\R^2), \\
\threevecT {v_1}{v_2}{v_3} & \;\;\longleftrightarrow\;\; v_1dx_1+v_2dx_2+v_3dx_3\in\Lambda^1(\R^3),\\
\threevecT {w_1}{w_2}{w_3} & \;\;\longleftrightarrow\;\; w_1dx_2dx_3+w_2dx_3dx_1+w_3dx_1dx_2\in\Lambda^2(\R^3).
\end{align*}
\subsection{Generalized Barycentric Coordinates}
\label{subsec:gbcs}
Let $\mathfrak m$ be a convex $n$-dimensional polytope in $\R^n$ with vertex set $\{\textbf{v}_i\}$, written as column vectors.
A set of non-negative functions $\{\lambda_i\}:\mathfrak m\rightarrow\R$ are called \textbf{generalized barycentric coordinates} on $\mathfrak m$ if for any linear function $L:\mathfrak m\rightarrow\R$, we can write
\begin{equation}
\label{eq:lin-comp}
L = \sum_i L(\textbf{v}_i)\lambda_i,
\end{equation}
We will use the notation $\mathbb I$ to denote the $n\times n$ identity matrix and $\textbf{x}$ to denote the vector ${\begin{bmatrix} x_1 & x_2 & \cdots & x_n\end{bmatrix}}^T$ where $x_i$ is the $i$th coordinate in $\R^n$.
We have the following useful identities:
\begin{align}
\sum_i\lambda_i(\textbf{x}) &= 1 \label{eq:pof1} \\
\sum_i\textbf{v}_i\lambda_i(\textbf{x}) &= \textbf{x} \label{eq:pofx}\\
\sum_i\nabla\lambda_i (\textbf{x})& = 0 \label{eq:gradsum0} \\
\sum_i\textbf{v}_i\nabla\lambda_i^T(\textbf{x}) & = \mathbb I \label{eq:vgradsumI}
\end{align}
Equations (\ref{eq:pof1}) and (\ref{eq:pofx}) follow immediately from (\ref{eq:lin-comp}) while (\ref{eq:gradsum0}) and (\ref{eq:vgradsumI}) follow by taking the gradient of equations (\ref{eq:pof1}) and (\ref{eq:pofx}), respectively.
If $\textbf{x}$ is constrained to an $n-1$ dimensional facet of $\mathfrak m$ and the index set of the summations are limited to those vertices that define $\mathfrak m$, then (\ref{eq:pof1})-(\ref{eq:vgradsumI}) still hold; in particular, this implies that generalized barycentric coordinates on a polyhedron restrict to generalized barycentric coordinates on each of its polygonal faces.
As mentioned in the introduction, there are many approaches to defining generalized barycentric coordinates.
In regards to applications in finite element methods, the Wachspress coordinates~\cite{W1975,W2011} are commonly used as they are rational functions in both 2D and 3D with explicit formulae; code for their implementation in MATLAB is given in the appendix of~\cite{FGS2013}.
Other practical choices of generalized barycentric coordinates for finite elements include mean value~\cite{F2003}, maximum entropy~\cite{HS2008,Su04}, and moving least squares~\cite{MS10}.
The results of this work do not rely on any properties of the coordinates other than their non-negativity and linear reproduction property (\ref{eq:lin-comp}).
\subsection{Whitney forms}
\label{subset:whit-forms}
Let $\mathfrak m$ be a convex $n$-dimensional polytope in $\R^n$ with vertex set \textnormal{$\{\textbf{v}_i\}$} and an associated set of generalized barycentric coordinates $\{\lambda_i\}$.
Define associated sets of index pairs and triples by
\begin{align}
E & := \{(i,j) ~:~ \textbf{v}_i,\textbf{v}_j\in\mathfrak m\} \label{eq:Em-def},\\
T & := \{(i,j,k) ~:~ \textbf{v}_i,\textbf{v}_j,\textbf{v}_k\in\mathfrak m\} \label{eq:Tm-def}.
\end{align}
If $\mathfrak m$ is a \textit{simplex}, the elements of the set
\[\left\{\lnl ij - \lnl ji~:~(i,j)\in E\right\}\]
are called Whitney 1-forms and are part of a more general construction~\cite{W1957}, which we now present.
Again, if $\mathfrak m$ is a $\textit{simplex}$, the Whitney $k$-forms are elements of the set
\begin{equation}
\label{eq:whit-def}
\left\{k!\sum_{i=0}^k(-1)^i\; \lambda_{j_i} \;d\lambda_{j_0}\wedge\ldots\wedge\widehat{d\lambda_{j_i}}\wedge\ldots\wedge d\lambda_{j_k}\right\},
\end{equation}
where $j_0,\ldots,j_k$ are indices of vertices of $\mathfrak m$.
As before, $\wedge$ denotes the wedge product and $\widehat{dx_{\sigma_i}}$ means that the term is omitted.
Up to sign, this yields a set of $n+1\choose k+1$ distinct functions and provides a local basis for ${\mathcal P}_1^-\Lambda^k$~\cite{AFW2009}.
We now generalize these definitions to the case where $\mathfrak m$ is non necessarily a simplex.
For any $(i,j)\in E$, define a generalized Whitney 1-form on $\mathfrak m$ by
\begin{align}
\label{eq:whit-edge-def}
\mathcal{W}_{ij} & := \lambda_i\nabla\lambda_j-\lambda_j\nabla\lambda_i.
\end{align}
If $n=3$, then for any $(i,j,k)\in T$, define a generalized Whitney 2-form on $\mathfrak m$ by
\begin{align}
\label{eq:whit-tri-def}
\mathcal{W}_{ijk} & := (\whitC ijk) + (\whitC jki) + (\whitC kij).
\end{align}
Note that $\mathcal{W}_{ii}=0$ and if $i$, $j$, and $k$ are not distinct then $\mathcal{W}_{ijk}=0$.
Whitney forms have natural interpretations as vector fields when $k=1$ or $n-1$.
Interpolation of vector fields requires less data regularity than the canonical scalar interpolation theory using nodal values.
Averaged interpolation developed for scalar spaces~\cite{Cl75,SZ90} has been extended to families of spaces from finite element exterior calculus~\cite{CW08}.
Recent results on polygons and polyhedra can be extended to less regular data with average interpolation following the framework in~\cite{Ra12}, based on affine invariance of the coordinates.
\section{Global Continuity Results}
\label{sec:global-cnty}
We first present results about the global continuity properties of vector-valued functions defined in terms of generalized barycentric coordinates and their gradients over a mesh of $n$-dimensional polytopes in $\R^n$ with $n=2$ or $3$.
By `mesh' we mean a cellular complex in which each cell is a polygon (for $n=2$) or polyhedron (for $n=3$); for more on cellular complexes see e.g.~\cite{C2008}.
Voronoi meshes are examples of cellular complexes since they are composed of $n$-dimensional polytopes that meet along their $n-1$ dimensional facets.
We say that a function is defined `piecewise with respect to a mesh' when the definition of the function on the interior of a mesh element depends only on geometrical properties of the element (as opposed to depending on adjacent elements, for instance).
We begin with a general result about global continuity in such a setting.\\
\begin{prop}
\label{prop:trace2conf}
Fix a mesh $\mesh$ of $n$-dimensional polytopes in $\R^n$ with $n=2$ or $3$.
Let $\textnormal{$\textbf{u}$}$ be a vector field defined piecewise with respect to $\mesh$.
Let $\mathfrak f$ be a face of codimension 1 with $\textnormal{$\textbf{u}$}_1$, $\textnormal{$\textbf{u}$}_2$ denoting the values of $\textnormal{$\textbf{u}$}$ on $\mathfrak f$ as defined by the two $n$-dimensional mesh elements sharing $\mathfrak f$.
Write $\textnormal{$\textbf{u}$}_i = T_\mathfrak f(\textnormal{$\textbf{u}$}_i) + N_\mathfrak f(\textnormal{$\textbf{u}$}_i)$ where $T_\mathfrak f(\textnormal{$\textbf{u}$}_i)$ and $N_\mathfrak f(\textnormal{$\textbf{u}$}_i)$ are the vector projections of $\textnormal{$\textbf{u}$}_i$ onto $\mathfrak f$ and its outward normal, respectively.
\renewcommand{\labelenumi}{(\roman{enumi}.)}
\begin{enumerate}
\item If $T_\mathfrak f(\textnormal{$\textbf{u}$}_1)=T_\mathfrak f(\textnormal{$\textbf{u}$}_2)$ for all $\mathfrak f\in \mesh$ then $\textnormal{$\textbf{u}$}\in H(\textnormal{curl\,},\mesh)$.
\item If $N_\mathfrak f(\textnormal{$\textbf{u}$}_1)=N_\mathfrak f(\textnormal{$\textbf{u}$}_2)$ for all $\mathfrak f\in \mesh$ then $\textnormal{$\textbf{u}$}\in H(\textnormal{div\,},\mesh)$.
\end{enumerate}
\end{prop}
\text{}\\
The results of Proposition~\ref{prop:trace2conf} are well-known in the finite element community; see e.g.~Ern and Guermond~\cite[Section 1.4]{EG04}.\\
\begin{prop}
\label{prop:bctrcfaceprop}
Let $\mathfrak m$ be a convex $n$-dimensional polytope in $\R^n$ with vertex set \textnormal{$\{\textbf{v}_i\}_{i\in I}$} and an associated set of generalized barycentric coordinates $\{\lambda_i\}_{i\in I}$.
Let $\mathfrak f$ be a face of $\mathfrak m$ of codimension 1 whose vertices are indexed by $J\subsetneq I$.
If $k\not\in J$ then $\lambda_k\equiv 0$ on $\mathfrak f$ and $\nabla\lambda_k$ is normal to $\mathfrak f$ on $\mathfrak f$, pointing inward.
\end{prop}
\begin{proof}
Fix a point $\textbf{x}_0\in\mathfrak m$.
Observe that $\sum_{i\in I}\textbf{v}_i\lambda_i(\textbf{x}_0)$ is a point in $\mathfrak m$ lying in the interior of the convex hull of those $\textbf{v}_i$ for which $\lambda_i(\textbf{x}_0)>0$, since the $\lambda_i$ are non-negative by definition.
By (\ref{eq:pofx}), this summation is equal to $\textbf{x}_0$.
Hence, if $\textbf{x}_0\in \mathfrak f$, then $\lambda_k\equiv 0$ on $\mathfrak f$ unless $k\in J$, proving the first claim.
The same argument implies that for any $k\not\in $J, $\mathfrak f$ is part of the zero level set of $\lambda_k$.
Hence, for $k\not\in J$, $\nabla\lambda_k$ is orthogonal to $\mathfrak f$ on $\mathfrak f$.
In that case, $\nabla\lambda_k$ points inward since $\lambda_k$ has support inside $\mathfrak m$ but not on the other side of $\mathfrak f$.
\end{proof}
\begin{figure}
\begin{center}
\includegraphics[height=.3\textwidth]{fig/l3g2-right} ~\includegraphics[height=.3\textwidth]{fig/l4g1-left}
\end{center}
\caption{The $H(\textnormal{curl})$ conformity condition of Proposition~\ref{prop:trace2conf} is satisfied automatically by the $\lambda_i\nabla\lambda_j$ functions, as shown in the example above.
When the elements are brought together, the vector fields will agree on the projection to the shared edge at any point along the shared edge.
Here, $i$ and $j$ are the indices for the vertices at the top and bottom, respectively, of the shared edge. For this example, we used the Wachspress functions to compute the vector functions on each element and MATLAB to visualize the result.}
\label{fig:conf}
\end{figure}
We now show that generalized barycentric coordinates and their gradients defined over individual elements in a mesh of polytopes naturally stitch together to build conforming finite elements with global continuity of the expected kind.
Figure~\ref{fig:conf} presents an example of two vector functions agreeing on their tangential projections along a shared edge.
To be clear about the context, we introduce notation for generalized barycentric hat functions, defined piecewise over a mesh of polytopes $\{\mathfrak m\}$ by
\[\hat\lambda_i(\textbf{x})= \begin{cases} \lambda_i(\textbf{x})~\text{as defined on $\mathfrak m$} & \text{if $\textbf{x}\in\mathfrak m$ and $\textbf{v}_i\in \mathfrak m$;} \\ 0 & \text{if $\textbf{x}\in\mathfrak m$ but $\textbf{v}_i \not\in \mathfrak m$.} \end{cases}\]
Note that generalized barycentric coordinates $\lambda_i$ are usually indexed locally on a particular polytope while the $\hat\lambda_i$ require a global indexing of the vertices to consistently identify matching functions across element boundaries.
Further, $\hat\lambda_i$ is well-defined at vertices and edges of the mesh as any choice of generalized barycentric coordinates on a particular element will give the same value at such points.
If $\textbf{x}$ belongs to the interior of shared faces between polyhedra in $\R^3$ (or higher order analogues), $\hat\lambda_i(\textbf{x})$ is well-defined so long as the same \textit{kind} of generalized barycentric coordinates are chosen on each of the incident polyhedra (e.g.\ Wachspress or mean value).
Our first result about global continuity concerns functions of the form $\hlnhl ij$, where $i$ and $j$ are indices of vertices belonging to at least one fixed mesh element $\mathfrak m$.
Note that the vertices $\textbf{v}_i$ and $\textbf{v}_j$ need not define an edge of $\mathfrak m$.\\
\begin{theorem}
\label{thm:hcurl-conf}
Fix a mesh $\mesh$ of $n$-dimensional polytopes $\{\mathfrak m\}$ in $\R^n$ with $n=2$ or $3$ and assign some ordering $\textnormal{$\textbf{v}_1,\ldots,\textbf{v}_p$}$ to all the vertices in the mesh.
Fix an associated set of generalized barycentric coordinate hat functions $\textnormal{$\hat\lambda_1,\ldots,\hat\lambda_p$}$.
Let
\[\textnormal{$\textbf{u}$}\in\textnormal{span}~\left\{\hlnhl ij~:~\textnormal{$\exists~\mathfrak m\in\mesh$ such that $\textbf{v}_i,\textbf{v}_j\in\mathfrak m$}\right\}.\]
Then $\textnormal{$\textbf{u}$}\in H(\textnormal{curl\,},\mesh)$.
\end{theorem}
\begin{proof}
Following the notation of Proposition~\ref{prop:trace2conf}, it suffices to show that $T_\mathfrak f(\textnormal{$\textbf{u}$}_1)=T_\mathfrak f(\textnormal{$\textbf{u}$}_2)$ for an arbitrary face $\mathfrak f\in \mesh$ of codimension 1.
Consider an arbitrary term $c_{ij}\hlnhl ij$ in the linear combination defining $\textnormal{$\textbf{u}$}$.
Observe that if $\textbf{v}_i\not\in \mathfrak f$, then by Proposition~\ref{prop:bctrcfaceprop}, $\hat \lambda_i\equiv 0$ on $\mathfrak f$ and hence $\textnormal{$\textbf{u}$}\equiv 0$ on $\mathfrak f$.
Further, if $\textbf{v}_j\not\in \mathfrak f$, then $\nabla\hat \lambda_j$ is orthogonal to $\mathfrak f$.
Therefore, without loss of generality, we can reduce to the case where $\textbf{v}_i,\textbf{v}_j\in \mathfrak f$.
Since $\hat \lambda_i$ and $\hat \lambda_j$ are both $C^0$ on $\mesh$, their well-defined values on $\mathfrak f$ suffice to determine the projection of $\hlnhl ij$ to $\mathfrak f$.
Since the choice of pair $ij$ was arbitrary, we have $T_\mathfrak f(\textnormal{$\textbf{u}$}_1)=T_\mathfrak f(\textnormal{$\textbf{u}$}_2)$, completing the proof.
\end{proof}
\text{}
\begin{remark}
\label{rmk:rot-for-cty}
{\em
When $n=2$, we may replace $\hlnhl ij$ in the statement Theorem~\ref{thm:hcurl-conf} by $\textnormal{rot}~\hlnhl ij$ and conclude that $\textnormal{$\textbf{u}$}\in H(\textnormal{div\,},\mesh)$.
This is immediate since $\textnormal{rot}$ gives an isomorphism between $H(\textnormal{curl})$ and $H(\textnormal{div})$ in $\R^2$, as discussed in Section~\ref{subsec:feec}.
When $n=3$, we construct functions in $H(\textnormal{div\,},\mesh)$ using triples of indices associated to vertices of mesh elements, according to the next result.
}
\end{remark}
\text{}
\begin{theorem}
\label{thm:hdiv-conf}
Fix a mesh $\mesh$ of polyhedra $\{\mathfrak m\}$ in $\R^3$ and assign some ordering $\textnormal{$\textbf{v}_1,\ldots,\textbf{v}_p$}$ to all the vertices in the mesh.
Fix an associated set of generalized barycentric coordinate hat functions $\textnormal{$\hat\lambda_1,\ldots,\hat\lambda_p$}$.
Let
\[\textnormal{$\textbf{u}$}\in\textnormal{span}~\left\{\hatlnlxnl ijk~:~\textnormal{$\exists~\mathfrak m\in\mesh$ such that $\textbf{v}_i,\textbf{v}_j,\textbf{v}_k\in\mathfrak m$}\right\}.\]
Then $\textnormal{$\textbf{u}$}\in H(\textnormal{div\,},\mesh)$.
\end{theorem}
\begin{proof}
Again following the notation of Proposition~\ref{prop:trace2conf}, it suffices to show that $N_\mathfrak f(\textnormal{$\textbf{u}$}_1)=N_\mathfrak f(\textnormal{$\textbf{u}$}_2)$ for an arbitrary face $\mathfrak f\in \mesh$ of codimension one whose vertices are indexed by $J$.
We will use the shorthand notation
\[\xi_{ijk} := \hatlnlxnl ijk. \]
Consider an arbitrary term $c_{ijk}\xi_{ijk}$ in the linear combination defining $\textnormal{$\textbf{u}$}$.
We will first show that $\xi_{ijk}$ has a non-zero normal component on $\mathfrak f$ only if $i,j,k\in J$.
If $i\not\in J$ then $\hat\lambda_i\equiv0$ on $\mathfrak f$ by Proposition~\ref{prop:bctrcfaceprop}, making $\xi_{ijk}\equiv 0$ on $\mathfrak f$, as well.
If $i\in J$ but $j,k\not\in J$, then $\nabla\hat\lambda_j$ and $\nabla\hat\lambda_k$ are both normal to $\mathfrak f$ on $\mathfrak f$ by Proposition~\ref{prop:bctrcfaceprop}.
Hence, their cross product is zero and again $\xi_{ijk}\equiv 0$ on $F$.
If $i,j\in J$ but $k\not\in J$ then again $\nabla\hat\lambda_k\perp \mathfrak f$ on $\mathfrak f$.
Since $\nabla\hat\lambda_j\times\nabla\hat\lambda_k\perp \nabla\hat\lambda_k$, we conclude that $\xi_{ijk}$ has no normal component on $\mathfrak f$.
The same argument holds for the case $i,k\in J$, $j\not\in J$.
The only remaining case is $i,j,k\in J$, proving the claim.
Thus, without loss of generality, we assume that $i,j,k\in J$.
Since $\hat\lambda_j$ and $\hat\lambda_k$ are both $C^0$ on $\mesh$, their well-defined values on $\mathfrak f$ suffice to determine the projection of $\nabla\hat\lambda_j$ and $\nabla\hat\lambda_k$ to $\mathfrak f$, which then uniquely defines the normal component of $\nabla\hat\lambda_j\times\nabla\hat\lambda_k$ on $\mathfrak f$.
Since $\hat\lambda_i$ is also $C^0$ on $\mesh$, and the choice of $i,j,k$ was arbitrary, we have $N_\mathfrak f(\textnormal{$\textbf{u}$}_1)=N_\mathfrak f(\textnormal{$\textbf{u}$}_2)$, completing the proof.
\end{proof}
\text{}\\
\section{Polynomial Reproduction Results}
\label{sec:poly-repro}
We now show how generalized barycentric coordinate functions $\lambda_i$ and their gradients can reproduce all the polynomial differential forms in ${\mathcal P}_1\Lambda^k$ and ${\mathcal P}_1^-\Lambda^k$ for $0\leq k\leq n$ with $n=2$ or $3$.
The results for the functions $\lnl ij$ and $\mathcal{W}_{ij}$ extend immediately to any value of $n\geq 2$ since those functions do not use any dimension-specific operators like $\times$ or $\textnormal{rot}$. \\
\begin{theorem}
\label{thm:pr-lnl}
Fix $n\geq 2$.
Let $\mathfrak m$ be a convex $n$-dimensional polytope in $\R^n$ with vertex set \textnormal{$\{\textbf{v}_i\}$}.
Given any set of generalized barycentric coordinates $\{\lambda_i\}$ associated to $\mathfrak m$,
\begin{equation}
\label{eq:lnl-id}
\textnormal{$
\sum_{i,j}\lnl ij(\textbf{v}_j-\textbf{v}_i)^T= \mathbb I,
$}
\end{equation}
where $\mathbb I$ is the $n\times n$ identity matrix.
Further, for any $n\times n$ matrix $\A$,
\begin{equation}
\label{eq:lnl-Ax}
\textnormal{$
\sum_{i,j}(\A\textbf{v}_i \cdot\textbf{v}_j)(\lnl ij) =\A\textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\lnl ij\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1\Lambda^1(\mathfrak m).$
\end{theorem}
\begin{proof}
From (\ref{eq:pof1}) - (\ref{eq:vgradsumI}), we see that
\begin{align*}
\sum_{i,j}\lnl ij(\textbf{v}_j-\textbf{v}_i)^T
& = \left(\sum_{i}\lambda_i\right)\left(\sum_j\nabla\lambda_j\textbf{v}_j^T\right) - \left(\sum_{j}\nabla\lambda_j\right)\left(\sum_i\lambda_i\textbf{v}_i^T\right) \\
& = 1(\mathbb I^T) - 0(\textbf{x}^T) = \mathbb I,
\end{align*}
establishing (\ref{eq:lnl-id}).
Similarly for (\ref{eq:lnl-Ax}), a bit of algebra yields
\begin{align*}
\sum_{i,j}(\A\textbf{v}_i \cdot\textbf{v}_j)(\lnl ij)
& = \sum_{i,j}(\lnl ij) \textbf{v}_j^T \A\textbf{v}_i
= \sum_{i,j} \nabla\lambda_j \textbf{v}_j^T \A\textbf{v}_i \lambda_i\\
& = \left(\sum_{j} \nabla\lambda_j \textbf{v}_j^T\right)\A\left(\sum_i \textbf{v}_i \lambda_i\right)
= \mathbb I^T\A\textbf{x}
= \A\textbf{x}
\end{align*}
We have shown that any vector of linear polynomials can be written as a linear combination of $\lnl ij$ functions, hence the span of these functions contains the vector proxies for all elements of ${\mathcal P}_1\Lambda^1(\mathfrak m)$.
\end{proof}
\text{}\\
\begin{cor}
\label{cor:pr-lrnl}
Let $\mathfrak m$ be a convex polygon in $\R^2$ with vertex set \textnormal{$\{\textbf{v}_i\}$}.
Given any set of generalized barycentric coordinates $\{\lambda_i\}$ associated to $\mathfrak m$,
\begin{equation}
\label{eq:lrnl-id}
\textnormal{$
\sum_{i,j}\textnormal{rot}\,\lnl ij(\textnormal{rot}(\textbf{v}_j-\textbf{v}_i))^T= \mathbb I,
$}
\end{equation}
where $\mathbb I$ is the $2\times 2$ identity matrix.
Further, for any $2\times 2$ matrix $\A$,
\begin{equation}
\label{eq:lrnl-Ax}
\textnormal{$
\sum_{i,j}(-\textnormal{rot}\,\A\,\textbf{v}_i \cdot\textbf{v}_j)(\textnormal{rot}\, \lnl ij) =\A\textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\textnormal{rot}\lnl ij\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1\Lambda^1(\mathfrak m).$
\end{cor}
\begin{proof}
For (\ref{eq:lrnl-id}), observe that for any $\textbf{w},\textbf{y}\in\R^2$, $\displaystyle\textbf{w}\textbf{y}^T = \begin{bmatrix} a & b \\ c & d\end{bmatrix}$ implies $\displaystyle(\textnormal{rot}~\textbf{w})(\textnormal{rot}~\textbf{y})^T = \begin{bmatrix} d & -c \\ -b & a \end{bmatrix}$.
Hence, the result follows immediately from (\ref{eq:lnl-id}).
For (\ref{eq:lrnl-Ax}), note $\textnormal{rot}^{-1}=-\textnormal{rot}$ and define $\B:=-\textnormal{rot}\,\A$.
Using $\B$ as the matrix in (\ref{eq:lnl-Ax}), we have
\[\sum_{i,j}(\B\textbf{v}_i \cdot\textbf{v}_j)(\lnl ij) =\B\textbf{x}\]
Applying $\textnormal{rot}$ to both sides of the above yields the result.
\end{proof}
\text{}\\
\begin{theorem}
\label{thm:pr-lnlxnl}
Let $\mathfrak m$ be a convex polyhedron in $\R^3$ with vertex set \textnormal{$\{\textbf{v}_i\}$}.
Given any set of generalized barycentric coordinates $\{\lambda_i\}$ associated to $\mathfrak m$,
\begin{equation}
\label{eq:lnlxnl-id}
\textnormal{$
\frac 12\sum_{i,j,k}\lnlxnl ijk\left((\textbf{v}_j-\textbf{v}_i)\times(\textbf{v}_k-\textbf{v}_i)\right)^T= \mathbb I ,
$}
\end{equation}
where $\mathbb I$ is the $n\times n$ identity matrix.
Further, for any $n\times n$ matrix $\A$,
\begin{equation}
\label{eq:lnlxnl-Ax}
\textnormal{$
\frac 12\sum_{i,j,k}(\A\textbf{v}_i \cdot(\textbf{v}_j\times\textbf{v}_k))(\lnlxnl ijk) =\A\textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\lnlxnl ijk\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j,\textbf{v}_k\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1\Lambda^2(\mathfrak m).$
\end{theorem}
\begin{proof}
We start with (\ref{eq:lnlxnl-id}).
First, observe that
\[(\textbf{v}_j-\textbf{v}_i)\times(\textbf{v}_k-\textbf{v}_i) = \textbf{v}_i\times\textbf{v}_j +\textbf{v}_j\times\textbf{v}_k + \textbf{v}_k\times\textbf{v}_i.\]
By (\ref{eq:gradsum0}), we have that
\begin{align*}
\sum_{i,j,k}\lnlxnl ijk\left(\textbf{v}_i\times\textbf{v}_j\right)^T & = \sum_{i,j} \lambda_i \left(\nabla \lambda_j \times \left( \sum_k \nabla \lambda_k \right)\right) \left(\textbf{v}_i\times\textbf{v}_j\right)^T
=0.
\end{align*}
A similar argument shows that replacing $\textbf{v}_i\times\textbf{v}_j$ with $\textbf{v}_k\times\textbf{v}_i$ also yields the zero matrix.
Hence,
\begin{align*}
\sum_{i,j,k}\lnlxnl ijk\left((\textbf{v}_j-\textbf{v}_i)\times(\textbf{v}_k-\textbf{v}_i)\right)^T &= \sum_{i,j,k}\lnlxnl ijk\left(\textbf{v}_j\times\textbf{v}_k\right)^T
\end{align*}
\begin{align*}
= \sum_i \lambda_i \sum_{j,k}\left(\nabla \lambda_j \times \nabla \lambda_k\right) \left(\textbf{v}_j\times\textbf{v}_k\right)^T
& =\sum_{j,k}\left(\nabla \lambda_j \times \nabla \lambda_k\right) \left(\textbf{v}_j\times\textbf{v}_k\right)^T.
\end{align*}
To simplify this further, we use the Kronecker delta symbol $\delta_{i_1i_2}$ and the 3D Levi-Civita symbol $\varepsilon_{i_1 i_2 i_3}$.
It suffices to show that the entry in row $r$, column $c$ of the matrix $\sum_{j,k}(\nlxnl jk)\left(\textbf{v}_j\times \textbf{v}_k\right)^T$ is $2\delta_{rc}$.
We see that
\begin{align*}
\left[\sum_{j,k}(\nlxnl jk)\left(\textbf{v}_j\times \textbf{v}_k\right)^T\right]_{rc}
& = \sum_{j,k}\varepsilon_{r\ell m}(\nl j)_\ell(\nl k)_m \varepsilon_{cpq} (\textbf{v}_j)_p(\textbf{v}_k)_q \\
& = \varepsilon_{r\ell m}\varepsilon_{cpq} \sum_{j}(\textbf{v}_j)_p(\nl j)_\ell \sum_{k} (\textbf{v}_k)_q(\nl k)_m \\
& = \varepsilon_{r\ell m}\varepsilon_{cpq} \delta_{\ell p}\delta_{mq}.
\end{align*}
The last step in the above chain of equalities follows from (\ref{eq:vgradsumI}).
Observe that $\varepsilon_{r\ell m}\varepsilon_{cpq} \delta_{\ell p}\delta_{mq}= \varepsilon_{r\ell m}\varepsilon_{c\ell m} = 2\delta_{r c}$, as desired.
For (\ref{eq:lnlxnl-Ax}), observe that
\begin{align*}
\sum_{i,j,k}(\A\textbf{v}_i \cdot(\textbf{v}_j\times\textbf{v}_k))(\lnlxnl ijk)
& = \left(\sum_{i}\A\textbf{v}_i\lambda_i\right)\cdot \sum_{j,k} (\textbf{v}_j\times\textbf{v}_k)(\nlxnl jk) \\
& = \sum_{j,k} (\nlxnl jk)(\textbf{v}_j\times\textbf{v}_k)^T \left(\A \sum_{i}\textbf{v}_i\lambda_i\right) \\
& = 2\;\mathbb I\;\A\textbf{x} = 2\A\textbf{x}.
\end{align*}
Note that we used the proof of (\ref{eq:lnlxnl-id}) to rewrite the sum over $j,k$ as $2\mathbb I$.
We have shown that any vector of linear polynomials can be written as a linear combination of $\lnlxnl ijk$ functions, hence the span of these functions contains the vector proxies for all elements of ${\mathcal P}_1\Lambda^2(\mathfrak m)$.
\end{proof}
\text{}\\
\begin{theorem}
\label{thm:pr-lnldrnl}
Let $\mathfrak m$ be a convex polygon in $\R^2$ with vertex set \textnormal{$\{\textbf{v}_i\}$}.
Given any set of generalized barycentric coordinates $\{\lambda_i\}$ associated to $\mathfrak m$,
\begin{equation}
\label{eq:lnldrnl-one}
\textnormal{$
\frac 12\sum_{i,j,k}\lnldrnl ijk\left((\textbf{v}_j-\textbf{v}_i)\cdot \textnormal{rot}(\textbf{v}_k-\textbf{v}_i)\right)= 1.
$}
\end{equation}
Further, for any vector $\textbf{a} \in \R^2$,
\begin{equation}
\label{eq:lnldrnl-Ax}
\textnormal{$
\frac 12\sum_{i,j,k}(\textbf{a}^T\textbf{v}_i (\textbf{v}_j\cdot \textnormal{rot}\textbf{v}_k))(\lnldrnl ijk) =\textbf{a}^T\textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\lnldrnl ijk\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j,\textbf{v}_k\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1\Lambda^2(\mathfrak m).$
\end{theorem}
\begin{proof}
The proof is essentially identical to that of Theorem~\ref{thm:pr-lnlxnl}. First,
\[(\textbf{v}_j-\textbf{v}_i) \cdot \textnormal{rot}(\textbf{v}_k-\textbf{v}_i) = \textbf{v}_i\cdot \textnormal{rot}\textbf{v}_j +\textbf{v}_j\cdot \textnormal{rot}\textbf{v}_k + \textbf{v}_k\cdot \textnormal{rot}\textbf{v}_i,\]
and by (\ref{eq:gradsum0}),
\begin{align*}
\sum_{i,j,k}\lnldrnl ijk & \left(\textbf{v}_i\cdot \textnormal{rot}\textbf{v}_j\right) = \sum_{i,j} \lambda_i \left(\nabla \lambda_j \cdot \textnormal{rot} \left( \sum_k \nabla \lambda_k \right)\right) \left(\textbf{v}_i\cdot \textnormal{rot} \textbf{v}_j\right)
=0.
\end{align*}
A similar argument shows that replacing $\textbf{v}_i\cdot \textnormal{rot} \textbf{v}_j$ with $\textbf{v}_k\cdot \textnormal{rot}\textbf{v}_i$ also yields zero.
Hence as before,
\begin{align*}
\sum_{i,j,k}\lnldrnl ijk & \left((\textbf{v}_j-\textbf{v}_i)\cdot \textnormal{rot}(\textbf{v}_k-\textbf{v}_i)\right)^T =\sum_{j,k}\left(\nabla \lambda_j \cdot \textnormal{rot} \nabla \lambda_k\right) \left(\textbf{v}_j\cdot \textnormal{rot}\textbf{v}_k\right)^T.
\end{align*}
Finally, the same argument holds using the 2D Levi-Civita symbol:
\begin{align*}
\sum_{j,k}(\nldrnl jk)\left(\textbf{v}_j\cdot\textnormal{rot} \textbf{v}_k\right)
& = \sum_{j,k}\varepsilon_{\ell m}(\nl j)_\ell(\nl k)_m \varepsilon_{p q} (\textbf{v}_j)_p(\textbf{v}_k)_q \\
& = \varepsilon_{\ell m}\varepsilon_{pq} \sum_{j}(\textbf{v}_j)_p(\nl j)_\ell \sum_{k} (\textbf{v}_k)_q(\nl k)_m \\
& = \varepsilon_{\ell m}\varepsilon_{pq} \delta_{\ell p}\delta_{mq}
= \varepsilon_{\ell m}\varepsilon_{\ell m} = 2,
\end{align*}
establishing (\ref{eq:lnldrnl-one}). For (\ref{eq:lnldrnl-Ax}), observe that
\begin{align*}
\sum_{i,j,k}(\textbf{a}^T\textbf{v}_i & (\textbf{v}_j\cdot \textnormal{rot} \textbf{v}_k)) (\lnldrnl ijk) \\
& = \left(\sum_{i}\textbf{a}^T\textbf{v}_i\lambda_i\right) \sum_{j,k} (\textbf{v}_j\cdot \textnormal{rot} \textbf{v}_k)(\nldrnl jk) \\
& = \sum_{j,k} (\nldrnl jk)(\textbf{v}_j\cdot \textnormal{rot} \textbf{v}_k)^T \left(\textbf{a}^T \sum_{i}\textbf{v}_i\lambda_i\right)
= 2\textbf{a}^T\textbf{x}.
\end{align*}
\end{proof}
\begin{remark}
{\em
The proof of Theorem~\ref{thm:pr-lnldrnl} can also be obtained by augmenting the 2D vectors and matrices with zeros to make 3D vectors and matrices and recognizing (\ref{eq:lnldrnl-one}) as the element equality in the third row and third column of (\ref{eq:lnlxnl-id}).
}
\end{remark}
\text{}\\
We also have polynomial reproduction results using the Whitney-like basis functions (\ref{eq:whit-edge-def}) and (\ref{eq:whit-tri-def}).
Recall that ${\mathcal H}_r$ denotes homogeneous polynomials of degree $r$ and let $\M_{n\times n}$ denote $n\times n$ matrices.
We have the following theorems.\\
\begin{theorem}
\label{thm:pr-whit-ij}
Fix $n\geq 2$.
Let $\mathfrak m$ be a convex $n$-dimensional polytope in $\R^n$ with vertex set \textnormal{$\{\textbf{v}_i\}$} and an associated set of generalized barycentric coordinates $\{\lambda_i\}$.
Then
\begin{equation}
\label{eq:whit1-id}
\textnormal{$
\sum_{i<j}\mathcal{W}_{ij}(\textbf{v}_j-\textbf{v}_i)^T=\mathbb I.
$}
\end{equation}
Further, define a map $\Phi:{\mathcal H}_1\Lambda^1(\R^n)\rightarrow \M_{n\times n}$ by
\[\textnormal{$
\sum_{i=1}^n\left(\sum_{j=1}^n a_{ij}x_j\right) dx_i \longmapsto \left [ \text{sign }(a_{ij}) \right ] .
$}\]
Then for all $\omega\in {\mathcal H}_0\Lambda^2(\R^n)$,
\begin{equation}
\label{eq:whit1-x}
\textnormal{$
\sum_{i<j}\left(\Phi(\kappa\omega)\textbf{v}_i)\cdot \textbf{v}_j\right) \mathcal{W}_{ij} = (\Phi(\kappa\omega))\textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\mathcal{W}_{ij}\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1^-\Lambda^1(\mathfrak m).$
\end{theorem}
\begin{proof}
For (\ref{eq:whit1-id}), we reorganize the summation and apply (\ref{eq:lnl-id}) to see that
\begin{align*}
\sum_{i<j}\mathcal{W}_{ij}(\textbf{v}_j-\textbf{v}_i)^T
& = \sum_{i<j}\lnl ij(\textbf{v}_j-\textbf{v}_i)^T - \sum_{i<j}\lnl ji(\textbf{v}_j-\textbf{v}_i)^T \\
& = \sum_{i<j}\lnl ij(\textbf{v}_j-\textbf{v}_i)^T + \sum_{j<i}\lnl ij(\textbf{v}_j-\textbf{v}_i)^T \\
& =\sum_{i,j}\lnl ij(\textbf{v}_j-\textbf{v}_i)^T = \mathbb I.
\end{align*}
For (\ref{eq:whit1-x}), fix $\omega\in {\mathcal H}_0\Lambda^2(\R^n)$ and express it as
\[\omega = \sum_{i<j}a_{ij}dx_idx_j,\]
for some coefficients $a_{ij}\in\R$.
Then
\[\kappa\omega = \sum_{i<j}a_{ij}(x_idx_j-x_jdx_i).\]
The entries of the matrix $\Phi(\kappa\omega)$ are thus given by
\begin{equation}
\label{eq:phi-entries}
\left[\Phi(\kappa\omega)\right]_{ij} =
\begin{cases}
\text{sign }(a_{ij}) & \mbox{if } i<j, \\
-\text{sign }(a_{ij}) & \mbox{if } i>j, \\
0 & \mbox{if } i=j. \\
\end{cases}
\end{equation}
From (\ref{eq:lnl-Ax}), we have that
\[\sum_{i,j}\left(\Phi(\kappa\omega)\textbf{v}_i)\cdot \textbf{v}_j\right) \lnl ij = (\Phi(\kappa\omega))\textbf{x},\qquad \forall \omega\in {\mathcal H}_0\Lambda^2(\R^n)\]
Since $\Phi(\kappa\omega)$ is anti-symmetric by (\ref{eq:phi-entries}), we have that
\begin{align*}
\sum_{i,j}\left(\Phi(\kappa\omega)\textbf{v}_i)\cdot \textbf{v}_j\right) & \lnl ij \\
& = \sum_{i<j}\left(\Phi(\kappa\omega)\textbf{v}_i)\cdot \textbf{v}_j\right) \lnl ij + \sum_{j<i}\left(\Phi(\kappa\omega)\textbf{v}_i)\cdot \textbf{v}_j\right) \lnl ij \\
& = \sum_{i<j}\left(\Phi(\kappa\omega)\textbf{v}_i)\cdot \textbf{v}_j\right) \mathcal{W}_{ij}.
\end{align*}
We have shown that any vector proxy of an element of ${\mathcal P}_0\Lambda^1(\mathfrak m)$ or $\kappa{\mathcal H}_{0}\Lambda^{2}(\mathfrak m)$ can be written as a linear combination of $\mathcal{W}_{ij}$ functions.
By (\ref{eq:prminus-decomp}), we conclude that the span of the $\mathcal{W}_{ij}$ functions contains the vector proxies for all elements of ${\mathcal P}_1^-\Lambda^1(\mathfrak m)$.
\end{proof}
\text{}\\
\begin{cor}
\label{cor:pr-rotWhit}
Let $\mathfrak m$ be a convex polygon in $\R^2$ with vertex set \textnormal{$\{\textbf{v}_i\}$}.
Given any set of generalized barycentric coordinates $\{\lambda_i\}$ associated to $\mathfrak m$,
\begin{equation}
\label{eq:rotWhit-id}
\textnormal{$
\sum_{i<j}\textnormal{rot}~\mathcal{W}_{ij}~\textnormal{rot}(\textbf{v}_j-\textbf{v}_i)^T=\mathbb I,
$}
\end{equation}
where $\mathbb I$ is the $2\times 2$ identity matrix.
Further,
\begin{equation}
\label{eq:rotWhit-x}
\textnormal{$
\sum_{i<j}\left((\textnormal{rot}~\textbf{v}_i)\cdot \textbf{v}_j\right) \textnormal{rot}~\mathcal{W}_{ij} = \textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\textnormal{rot}\;\mathcal{W}_{ij}\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1^-\Lambda^1(\mathfrak m).$
\end{cor}
\begin{proof}
By the same argument as the proof of (\ref{eq:lrnl-id}) in Corollary~\ref{cor:pr-lrnl}, the identity (\ref{eq:rotWhit-id}) follows immediately from (\ref{eq:whit1-id}).
For (\ref{eq:rotWhit-x}), observe that setting $\omega:=1\in {\mathcal H}_0\Lambda^2(\R^2)$, we have that $\Phi(\kappa\omega)=\textnormal{rot}$.
Therefore, (\ref{eq:whit1-x}) implies that
\[\sum_{i<j}\left(\textnormal{rot}~\textbf{v}_i)\cdot \textbf{v}_j\right) \mathcal{W}_{ij} = \textnormal{rot}~\textbf{x}.\]
Applying $\textnormal{rot}$ to both sides of the above equation completes the proof.
\end{proof}
\text{}\\
\begin{theorem}
\label{thm:pr-whit-ijk}
Let $\mathfrak m$ be a convex polyhedron in $\R^3$ with vertex set \textnormal{$\{\textbf{v}_i\}$} and an associated set of generalized barycentric coordinates $\{\lambda_i\}$.
Then
\begin{equation}
\label{eq:whit3-id}
\textnormal{$
\sum_{i<j<k}\mathcal{W}_{ijk} \left((\textbf{v}_j-\textbf{v}_i)\times(\textbf{v}_k-\textbf{v}_i)\right)^T = \mathbb I,
$}
\end{equation}
and
\begin{equation}
\label{eq:whit3-x}
\textnormal{$
\sum_{i<j<k}(\textbf{v}_i\cdot(\textbf{v}_j\times\textbf{v}_k))\mathcal{W}_{ijk}=\textbf{x}.
$}
\end{equation}
Thus, $\displaystyle\textnormal{span}\left\{\mathcal{W}_{ijk}\;:\;\textnormal{$\textbf{v}_i,\textbf{v}_j,\textbf{v}_k\in\mathfrak m$}\right\}\supseteq {\mathcal P}_1^-\Lambda^2(\mathfrak m).$
\end{theorem}
\begin{proof}
We adopt the shorthand notations
\[\xi_{ijk} := \lambda_i\nabla\lambda_j\times\nabla\lambda_k,
\quad \textbf{z}_{ijk}:= (\textbf{v}_j-\textbf{v}_i)\times(\textbf{v}_k-\textbf{v}_i),
\quad \textbf{v}_{ijk}:= \textbf{v}_i\cdot(\textbf{v}_j\times\textbf{v}_k).\]
For (\ref{eq:whit3-id}), we re-write (\ref{eq:lnlxnl-id}) as
\[
\sum_{i,j,k}\xi_{ijk}{\textbf{z}_{ijk}}^T= 2\mathbb I.
\]
Observe that $\xi_{ijk}{\textbf{z}_{ijk}}^T=(-\xi_{ikj})(-\textbf{z}_{ikj})^T=\xi_{ikj}{\textbf{z}_{ikj}}^T$ and $\textbf{z}_{ijk}=0$ if $i$, $j$, $k$ are not distinct.
Thus,
\begin{align*}
2\mathbb I
& = \sum_{\substack{i<j<k \\ k<i<j \\ j<k<i}}\xi_{ijk}{\textbf{z}_{ijk}}^T + \sum_{\substack{i<k<j \\ k<j<i \\ j<i<k}}\xi_{ikj}{\textbf{z}_{ikj}}^T.
\end{align*}
The two summations have different labels for the indices but are otherwise identical.
Therefore,
\begin{align*}
\mathbb I
& = \sum_{i<j<k}\xi_{ijk}{\textbf{z}_{ijk}}^T + \sum_{k<i<j}\xi_{ijk}{\textbf{z}_{ijk}}^T + \sum_{j<k<i}\xi_{ijk}{\textbf{z}_{ijk}}^T \\
& = \sum_{i<j<k}\xi_{ijk}{\textbf{z}_{ijk}}^T + \xi_{jki}{\textbf{z}_{jki}}^T + \xi_{kij}{\textbf{z}_{kij}}^T \\
& = \sum_{i<j<k}(\xi_{ijk}+\xi_{jki}+\xi_{kij}){\textbf{z}_{ijk}}^T \\
& = \sum_{i<j<k}\mathcal{W}_{ijk}\left((\textbf{v}_j-\textbf{v}_i)\times(\textbf{v}_k-\textbf{v}_i)\right)^T.
\end{align*}
For (\ref{eq:whit3-x}), we take $\A$ as the identity, and re-write (\ref{eq:lnlxnl-Ax}) as
\[
\sum_{i,j,k}\textbf{v}_{ijk}\xi_{ijk} =2\textbf{x}.
\]
Observe that $\textbf{v}_{ijk}\xi_{ijk}=(-\textbf{v}_{ikj})(-\xi_{ikj})=\textbf{v}_{ikj}\xi_{ikj}$ and $\textbf{v}_{ijk}=0$ if $i$, $j$, $k$ are not distinct.
Thus,
\begin{align*}
2\textbf{x}
& = \sum_{\substack{i<j<k \\ k<i<j \\ j<k<i}}\textbf{v}_{ijk}\xi_{ijk} + \sum_{\substack{i<k<j \\ k<j<i \\ j<i<k}}\textbf{v}_{ikj}\xi_{ikj} .
\end{align*}
The rest of the argument follows similarly, yielding
\begin{align*}
\textbf{x}
& = \sum_{i<j<k}\textbf{v}_{ijk}\xi_{ijk} + \sum_{k<i<j}\textbf{v}_{ijk}\xi_{ijk} + \sum_{j<k<i}\textbf{v}_{ijk}\xi_{ijk} \\
& = \sum_{i<j<k}(\textbf{v}_i\cdot(\textbf{v}_j\times\textbf{v}_k))\mathcal{W}_{ijk}.
\end{align*}
Note that ${\mathcal H}_{0}\Lambda^{3}(\mathfrak m)$ is generated by the volume form $\eta=dxdydz$ and that $\kappa\eta$ has vector proxy $\textbf{x}$.
Thus, by (\ref{eq:prminus-decomp}), we have shown that the span of the $\mathcal{W}_{ijk}$ functions contains the vector proxy of any element of ${\mathcal P}_1^-\Lambda^2(\mathfrak m)$.
\end{proof}
\text{}\\
\begin{remark}
\label{rmk:topdim}
\em
There are some additional constructions in this same vein that could be considered.
On a polygon in $\R^2$, we can define $\mathcal{W}_{ijk}$ in the same way as (\ref{eq:whit-tri-def}), interpreting $\times$ as the two dimensional cross product.
Likewise, on a polyhedron in $\R^3$, we can define $\mathcal{W}_{ijk\ell}$ according to formula (\ref{eq:whit-def}), yielding functions that are summations of terms like $\lambda_i(\nabla\lambda_j\cdot(\nlxnl k\ell)$.
These constructions will yield the expected polynomial reproduction results, yet they are not of practical interest in finite element contexts, as we will see in the next section.
\end{remark}
\section{Polygonal and Polyhedral Finite Element Families}
\label{sec:polyg-fams}
Let $\mesh$ be a mesh of convex $n$-dimensional polytopes $\{\mathfrak m\}$ in $\R^n$ with $n=2$ or $3$ and assign some ordering $\textnormal{$\textbf{v}_1,\ldots,\textbf{v}_p$}$ to all the vertices in the mesh.
Fix an associated set of generalized barycentric hat functions $\textnormal{$\hat\lambda_1,\ldots,\hat\lambda_p$}$ as in Section~\ref{sec:global-cnty}.
In Table~\ref{tab:comp-bases}, we list all the types of scalar-valued and vector-valued functions that we have defined this setting.
When used over all elements in a mesh of polygons or polyhedra, these functions have global continuity and polynomial reproduction properties as indicated in Table~\ref{tab:results-summary}.
These two properties -- global continuity and polynomial reproduction -- are essential and intertwined necessities in the construction of $H\Lambda^k$-conforming finite element methods on \textit{any} type of domain mesh.
Global continuity of type $H\Lambda^k$ ensures that the piecewise-defined approximate solution is an element of the function space $H\Lambda^k$ in which a solution is sought.
Polynomial reproduction of type ${\mathcal P}_1\Lambda^k$ or ${\mathcal P}_1^-\Lambda^k$ ensures that the error between the true solution and the approximate solution decays linearly with respect to the maximum diameter of a mesh element, as measured in $H\Lambda^k$ norm.
On meshes of simplicial elements, the basis functions listed in Table~\ref{tab:comp-bases} are known and often used as local bases for the corresponding classical finite element spaces listed in Table~\ref{tab:fe-spaces}, meaning our approach recapitulates known methods on simplicial meshes.
\begin{table}[t]
\centering
\begin{tabular}{c|c|c|c|c|c}
n & k & space & \# construction & \# boundary & \# polynomial \\ \hline
&&&&&\\[-2mm]
2 & 0 & ${\mathcal P}_1 \Lambda^0(\mathfrak m)$ & $v$ & $v$ & $3$\\
& & ${\mathcal P}_1^- \Lambda^0(\mathfrak m)$ & $v$ & $v$ & $3$\\
&&&&&\\[-2mm]
& 1 & ${\mathcal P}_1 \Lambda^1(\mathfrak m)$ & $\displaystyle 2{v \choose 2}$ & $2e$ & $6$\\
&&&&&\\[-2mm]
& & ${\mathcal P}_1^- \Lambda^1(\mathfrak m)$ & $\displaystyle {v \choose 2}$ & $e$ & $3$\\
&&&&&\\[-2mm]
& 2 & ${\mathcal P}_1 \Lambda^2(\mathfrak m)$ & $\displaystyle 3{v \choose 3}$ & $0$ & $3$\\
&&&&&\\[-2mm]
& & ${\mathcal P}_1^- \Lambda^2(\mathfrak m)$ & $\displaystyle {v \choose 3}$ & $0$ & $1$\\ [3mm]
\hline
\end{tabular}\\
\caption{Dimension counts relevant to serendipity-style reductions in basis size are shown.
Here, $v$ and $e$ denote the number of vertices and edges in the polygonal element $\mathfrak m$.
The column `\# construction' gives the number of basis functions we define
(cf.\ Table~\ref{tab:comp-bases}), `\# boundary' gives the number of basis functions related to inter-element continuity, and `\# polynomial' gives the dimension of the contained space of polynomial differential forms.}
\label{tab:sizes-2d}
\end{table}
\paragraph*{\underline{Relation to quadrilateral and serendipity elements}}
Consider the scalar bi-quadratic element on \textit{rectangles}, which has nine degrees of freedom: one associated to each vertex, one to each edge midpoint, and one to the center of the square.
It has long been known that the `serendipity' element, which has only the eight degrees of freedom associated to the vertices and edge midpoints of the rectangle, is also an $H^1$-conforming, quadratic order method.
In this case, polynomial reproduction requires the containment of ${\mathcal P}_2\Lambda^0(\mathfrak m)$ in the span of the basis functions, meaning at least six functions are required per element $\mathfrak m\in\mesh$.
To ensure global continuity of $H^1$, however, the method must agree `up to quadratics' on each edge, which necessitates the eight degrees of freedom associated to the boundary.
Therefore, the serendipity space associated to the scalar bi-quadratic element on a rectangle has dimension eight.
In a previous paper~\cite{RGB2011a}, we generalized this `serendipity' reduction to ${\mathcal P}_2\Lambda^0(\mesh)$ where $\mesh$ is a mesh of strictly convex polygons in $\R^2$.
For a simple polygon with $n$ vertices (and thus $n$ edges), polynomial reproduction still only requires $6$ basis functions, while global continuity of $H^1$ still requires reproduction of quadratics on edges, leading to a total of $2n$ basis functions required per element $\mathfrak m\in\mesh$.
Given a convex polygon, our approach takes the $n+{n \choose 2}$ pairwise products of all the $\lambda_i$ functions and forms explicit linear combinations to yield a set of $2n$ basis functions with the required global $H^1$ continuity and polynomial reproduction properties.
\paragraph*{\underline{Reduction of basis size}} A similar reduction procedure can be applied to the polygonal and polyhedral spaces described in Table~\ref{tab:comp-bases}.
A key observation is that the continuity results of Theorems \ref{thm:hcurl-conf} and \ref{thm:hdiv-conf} only rely on the agreement of basis functions whose indices are of vertices on a shared boundary edge (in 2D) or face (in 3D).
For example, if vertices $\textbf{v}_i$ and $\textbf{v}_j$ form the edge of a polygon in a 2D mesh, $H(\textnormal{curl\,},\mesh)$ continuity across the edge comes from identical tangential contributions in the $\lnl ij$ and $\lnl ji$ functions from either element containing this edge and zero tangential contributions from all other basis functions.
Thus, basis functions whose indices do not belong to a single polygon edge (in 2D) or polyhedral face (in 3D) do not contribute to inter-element continuity, allowing the basis size to be reduced.
\begin{table}[ht]
\centering
\begin{tabular}{c|c|c|c|c|c}
n & k & space & \# construction & \# boundary & \# polynomial \\ \hline
&&&&&\\[-2mm]
3 & 0 & ${\mathcal P}_1 \Lambda^0(\mathfrak m)$ & $v$ & $v$ & $4$\\
& & ${\mathcal P}_1^- \Lambda^0(\mathfrak m)$ & $v$ & $v$ & $4$\\
& 1 & ${\mathcal P}_1 \Lambda^1(\mathfrak m)$ & $\displaystyle 2 {v \choose 2}$ & $\displaystyle \left(\sum_{a=1}^f v_a(v_a-1)\right) - 2e$ & $12$\\
& & ${\mathcal P}_1^- \Lambda^1(\mathfrak m)$ & $\displaystyle {v \choose 2}$ & $\displaystyle \left(\sum_{a=1}^f {v_a\choose 2}\right) - e$ & $6$\\
&&&&&\\[-3mm]
& 2 & ${\mathcal P}_1 \Lambda^2(\mathfrak m)$ & $\displaystyle 3{v \choose 3}$ & $\displaystyle \sum_{a=1}^f \frac{v_a(v_a-1)(v_a-2)}{2}$ & $12$\\
&&&&&\\[-3mm]
& & ${\mathcal P}_1^- \Lambda^2(\mathfrak m)$ & $\displaystyle {v \choose 3}$ & $\displaystyle \sum_{a=1}^f {v \choose 3}$ & $4$\\
& 3 & ${\mathcal P}_1 \Lambda^3(\mathfrak m)$ & $\displaystyle 4{v \choose 4}$ & $0$ & $4$\\
&&&&&\\[-2mm]
& & ${\mathcal P}_1^- \Lambda^3(\mathfrak m)$ & $\displaystyle {v \choose 4}$ & $0$ & $1$\\[3mm]
\hline
\end{tabular}\\
\caption{The $n=3$ version of Table~\ref{tab:sizes-2d}.
Here, $f$ denotes the number of faces on a polyhedral element $\mathfrak m$ and $v_a$ denotes the number of vertices on a particular face $\mathfrak f_a$.
The entries of the `\# boundary' column are determined by counting functions associated to each face of the polyhedron and, in the $k=1$ cases, accounting for double-counting by subtraction.
}
\label{tab:sizes-3d}
\end{table}
To quantify the extent to which the bases we have defined could be reduced without affecting the global continuity properties, we count the number of functions associated with codimension 1 faces for each space considered.
For a polygon in 2D, the results are summarized in Table~\ref{tab:sizes-2d}.
The $k=0$ case is optimal in the sense that every basis function $\lambda_i$ contributes to the $H^1$-continuity in some way, meaning no basis reduction is available.
In the $k=1$ cases, the number of basis functions we construct is quadratic in the number of vertices, $v$, of the polygon, but the number associated with the boundary is only linear in the number of edges, $e$.
Since $e=v$ for a simple polygon, this suggests a basis reduction procedure would be both relevant and useful; the description of such a reduction will be the focus of a future work.
In the $k=2$ cases, our procedure constructs $O(v^3)$ basis functions but no inter-element continuity is required; in these cases, a discontinuous Galerkin or other type of finite element method would be more practical.
For a polyhedron $\mathfrak m$ in 3D, the results are summarized in Table~\ref{tab:sizes-3d}.
As in 2D, the basis for the $k=0$ case cannot be reduced while the bases for the $k=n$ cases would not be practical for implementation since no inter-element continuity is required.
In the $k=1$ cases, the number of basis functions we construct is again quadratic in $v$, while the number of basis functions required for continuity can be reduced for non-simplicial polyhedra.
For instance, if $\mathfrak m$ is a hexahedron, our construction for ${\mathcal P}_1\Lambda^1$ gives 56 functions but only 48 are relevant to continuity; in the ${\mathcal P}_1^-\Lambda^1$ case, we construct 28 functions but only 20 are relevant to continuity.
In the $k=2$ cases, a similar reduction is possible for non-simplicial polyhedra.
Again in the case of a hexahedron, we construct 168 functions for ${\mathcal P}_1\Lambda^1$ and 56 functions for ${\mathcal P}_1^-\Lambda^1$, but the elements require only 72 and 24 functions, respectively, for inter-element continuity.
\paragraph*{\underline{Current and future directions}}
It remains to discover additional properties of Whitney-like basis functions built from generalized barycentric coordinates and their use in finite element methods.
In the time since this manuscript first appeared online, Chen and Wang~\cite{CW2015} have presented an approach for constructing `minimal dimension' local basis sets based on the results of this paper.
Their theoretical and numerical results indicate that minimal spaces can, indeed, be constructed using the methods presented here with expected rates of convergence on certain classes of polygons and polyhedra.
We expect that the ideas introduced here will continue to influence the rapidly expanding use of polytopal finite element methods in scientific and engineering applications.\\
\noindent
\textbf{Acknowledgements.}
The authors would like to thank the anonymous referees for their helpful suggestions to improve the paper.
AG was supported in part by NSF Award 1522289 and a J.\ Tinsley Oden Fellowship.
CB was supported by was supported in part by a grant from NIH (R01-GM117594) and contract (BD-4485) from Sandia National Labs.
Sean Stephens helped produce the figure.
\bibliographystyle{abbrv}
|
\section{Introduction}
That black holes have no hair is a long-standing dictum of classical general
relativity \cite{nohair}, one whose content is highly contingent upon
assumed conditions. Although the original no-hair theorems were more
about limiting charges a black hole could carry, they have come to be
taken more widely as meaning {\it black holes cannot support nontrivial
fields on their event horizon}. This outlook is supported by the original
no hair theorems for gauge fields and scalars \cite{ChaseAdler}, which
placed what were regarded as eminently reasonable conditions on
matter fields. In the intervening years, however, it has become clear that
these conditions are not only too restrictive \cite{nohair2}, but in fact there
are many situations of physical interest in which black holes can
support nontrivial field configurations. Most of these are concerned
with asymptotically flat space times \cite{otherhair} whose hair
falls off sufficiently rapidly at large distances from the black hole, though
there are examples of nonsingular cosmological solutions with time
dependence \cite{coshair}, or indeed scalar condensates around
Kerr black holes \cite{Kerrscalar}.
Topological defects form an interesting class of alternative examples
of black hole hair outside of the asymptotically flat class.
Both domain walls and cosmic strings \cite{Vilenkin}, topologically
stable objects with a nontrivial quantum-field-theoretic vacuum
structure, can have significant gravitational influence, and were
originally expected to be antipathetic to black holes,
in part because of the problem of how
to have the associated fields end on the event horizon, but
also because of the strong global gravitational impact of the black
hole. Domain walls provide a `mirror' to spacetime (effectively compactifying
space \cite{IpS}) and cosmic strings yield a conical deficit that
generates a gravitational lens \cite{Vilenkin2}.
It is now known that both can ``pierce'' the black hole \cite{EGS,AGK}:
in the former case, the field theoretic wall provides a smooth transition
between mirror images of the northern hemisphere of the
C-metric\footnote{An accelerating black hole metric \cite{KinW}.},
whereas in the latter case a smooth version of the Aryal-Ford-Vilenkin
metric \cite{AFV} represents a black hole with a conical
deficit through its poles. The original solution \cite{AGK} has
been generalized in a number of ways to include vortices ending
on black holes \cite{RGMH}, charged black holes
\cite{CCESa,BEG}, dilatonic black holes \cite{ROG},
rotating black holes \cite{GB,GKW}, and asymptotically
dS \cite{GMdS} and AdS black holes \cite{VH,DGM1}.
Fields typically terminate on the event horizon or, in the case of
extremal black holes, be expelled from the horizon if the width
of the string is comparable to the size of the black hole.
Most recently, the rotating black hole has been subject to a thorough
study \cite{GKW}, which analysis corrected earlier work
that had a flawed ansatz \cite{GB}. There is now a detailed understanding
of how the core fields of a vortex accommodate the rotation of asymptotically
flat black holes and their associated `electric' field generation.
The vortex cuts out a local co-rotating deficit azimuthal angle, which leads
to some novel features, shifting the ergosphere of the black hole and altering
the innermost stable circular orbit (ISCO). As with charged black holes,
flux expulsion can indeed take place under certain circumstances.
However unlike the charged case the phase transition is of first order
and numerical evidence suggests that the flux-expelled solution is
not dynamically stable.
Here, we investigate the impact of a negative cosmological constant
on the problem of a vortex piercing a black hole. Specifically, we
obtain vortex solutions for an Abelian Higgs model minimally coupled
to Einstein gravity in four dimensions with a negative cosmological
constant. We obtain both approximate and numerical vortex solutions
to the field equations of the Abelian Higgs model in the background
of a Kerr-Newman-AdS black hole. We find that as the AdS length,
$\ell$, becomes comparable to the size of the vortex, the core of the
vortex increasingly narrows and the fields exhibit asymptotic power-law
falloff instead of exponential. We find that the Meissner effect, observed
previously for extremal Kerr and Reissner-Nordstrom black holes,
persists here as well, and is first
order if there is non-zero rotation but is otherwise 2nd order.
We find that the flux can pierce the horizon provided
the AdS length is sufficiently large, and numerically obtain the
critical radius for the transition from piercing to expulsion.
Our work may have interesting astrophysical implications.
It has long been known \cite{Wald,Expulsion1} that
spinning black holes tend to expel magnetic fields in a continuous
way as the black hole is spun up. Indeed, it has
been argued that all stationary, axisymmetric magnetic fields are
expelled from the Kerr horizon in the extremal limit
\cite{Expulsion2}. Since a Killing vector in the vacuum
spacetime can act as a vector potential for a Maxwell test field,
as the hole is `spun up' toward extremality, the component of
the magnetic field normal to the horizon approaches zero,
and so the flux lines are expelled (a phenomenon that also
occurs for black strings and $p$-branes \cite{Expulsion3}).
This Meissner-like effect could quench the power of astrophysical
jets, since the magnetic fields need to pierce the horizon to extract
rotational energy from the black hole, though it has been recently
argued \cite{Penna} that split-monopole magnetic fields may
continue to power black hole jets, with the fields becoming
entirely radial near the horizon, avoiding expulsion.
In contrast to this we find (as for the asymptotically flat case
\cite{GKW}) in the Abelian Higgs model that for large AdS
black holes the vortex pierces the event horizon, whereas
flux is expelled if the black hole is sufficiently small.
From a holographic perspective, a vortex in the bulk has an
interpretation as a defect in the the dual CFT \cite{VH,Dias:2013bwa},
corresponding in the dual superfluid to heavy pointlike excitations
around which the phase of the condensate winds.
We comment briefly at the end of our paper on a holographic
interpretation of our results.
\section{Abelian Higgs model for a cosmic string}
The abelian Higgs model is the canonical toy model for a cosmic
string, as it has the simplest action with the requisite vacuum structure
to allow a vortex to form. We write the action as\footnote{We use
units in which $\hbar=c=1$ and a mostly minus signature.}
\begin{equation} \label{abhact}
S = \int d^4x \sqrt{-g} \left [ D_{\mu}\Phi ^{\dagger}D^{\mu}\Phi -
{\textstyle{\frac{1}{4}}} {\tilde F}_{\mu \nu}{\tilde F}^{\mu \nu} - {\textstyle{\frac{1}{4}}}\lambda
(\Phi ^{\dagger} \Phi - \eta ^2)^2 \right ]\,,
\end{equation}
where $\Phi$ is the Higgs field, and $A_\mu$ the U(1) gauge
boson with field strength ${\tilde F}_{\mu\nu}$.
As per usual, we rewrite the field content as:
\begin{eqnarray}
\Phi (x^{\alpha}) &=& \eta X (x^{\alpha}) e^{i\chi(x^{\alpha}) }\,, \\
A_{\mu} (x^{\alpha}) &=& \frac{1}{e} \left [ P_{\mu} (x^{\alpha}) -
\nabla_{\mu} \chi (x^{\alpha}) \right ]\,.
\end{eqnarray}
These fields extract the physical degrees of freedom of the broken symmetric
phase, with $X$ representing the residual massive Higgs field, and $P_\mu$ the
massive vector boson. The gauge degree of freedom,
$\chi$, is explicitly subtracted, although any non-integrable phase
factors have a physical interpretation as a vortex.
In terms of these new variables, the equations of motion are
\begin{eqnarray} \label{vorteqn}
\nabla _{\mu}\nabla ^{\mu} X - P_{\mu}P^{\mu}X + \frac{\lambda\eta^2}{2}
X(X^2 -1) &=& 0\,, \\
\nabla _{\mu}F^{\mu \nu} + 2e^2\eta^2 X^2 P^{\nu}&=& 0\,.
\end{eqnarray}
Because we have not set $G\equiv 1$, we still have the freedom to
fix the units of energy, or $\eta$. We therefore choose to set
$\sqrt{\lambda}\eta=1$, effectively stating our Higgs field has order
unity mass.
For further use we also introduce the Bogomol'nyi parameter
\cite{Bog}:
\begin{equation}
\beta = \lambda/2e^2\,,
\end{equation}
indicating the gauge field has mass of order $1/\sqrt{\beta}$.
Alternately, we can rescale the dimensionful parameters $t$ and $r$
in the equations of motion: $t\to\sqrt{\lambda}\eta t$, etc.\ and
their corresponding gauge field components $P_t\to P_t/
\sqrt{\lambda}\eta$ -- note $P_\phi$ remains unrescaled however.
A straight static vortex solution will then have the Higgs profile,
$X_{NO}$, dependent on a single radial variable, $R$ say,
and the gauge field will have a single angular component,
$P_\phi = P_{NO}(R)$, where in flat spacetime $X_{NO}$
and $P_{NO}$ satisfy the Nielsen-Olesen equations \cite{NO}
\begin{equation}
\begin{aligned}
X_{NO}'' + \frac{X_{NO}'}{R} &= \frac{P_{NO}^2X_{NO}}{R^2}
+ \frac12 X_{NO}(X_{NO}^2-1)\,, \\
P_{NO}'' - \frac{P_{NO}'}{R} &= \frac{X_{NO}^2P_{NO}}{\beta}\,.
\end{aligned}
\end{equation}
The profiles of the $X_{NO}$ and $P_{NO}$ fields are highly localized
around $R=0$, and represent a Higgs core in which the U(1)
symmetry is restored with (in this case) a unit of magnetic flux
threading through. Higher winding strings can be obtained by
replacing $P_{NO} \to NP_{NO}$, although
these are unstable to splitting for $\beta>1$.
Since we are interested in vortices in an anti-de Sitter black
hole background, for future reference we now discuss the vortex
solution in the pure AdS geometry:
\begin{equation}
\begin{aligned}
ds^2 &= \Bigl(1+\frac{r^2}{\ell^2}\Bigr)dt^2
-\frac{dr^2}{\left(1+\frac{r^2}{\ell^2}\right )}
-r^2 d\theta^2-r^2 \sin^2\!\theta d\phi^2 \\
&=\frac{\ell^2 + R^2}{\ell^2(1-Z^2)} dt^2
- \frac{\ell^2+R^2}{(1-Z^2)^2}dZ^2
-\frac{\ell^2 dR^2}{\ell^2+R^2}
-R^2 d\phi^2 \,.
\end{aligned}
\end{equation}
By writing the AdS metric in this second, cylindrical, form we can
see that if we align the vortex in the $\{R,\phi\}$ plane, the equations
of motion will be independent of $Z$, and hence our vortex can once
again be represented by a set of ordinary differential equations:
\begin{equation}
\begin{aligned}
\left ( 1+\frac{R^2}{\ell^2}\right)P_{0}''
+\left(\frac{2R^2}{\ell^2}-1\right)\frac{P_{0}'}{R}
&= \frac{X_{0}^2P_{0}}{\beta}\,,\\
\left(1+\frac{R^2}{\ell^2}\right)X_{0}''
+\left(\frac{4R}{\ell^2}+\frac{1}{R}\right)X_{0}'
-\frac{P_{0}^2X_{0}}{R^2} -\frac{1}{2}X_{0}(X_{0}^2-1)&=0\,.
\end{aligned}
\label{AdSNO}
\end{equation}
As $R\to0$, the additional terms dependent on the AdS background
drop away, and we have a very similar field structure on axis to the
Nielsen-Olesen vortex. For $R\gtrsim\ell$ however, the functions are
modified, and the asymptotic fall-off of the fields becomes power
law rather than exponential.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\textwidth]{adsnov.pdf}
\caption{{\bf AdS-NO vortex:}
The values of $X$ and $P$ for the AdS NO vortex are
depicted as functions of $R$.}
\label{fig:adsNO}
\end{center}
\end{figure}
In figure \ref{fig:adsNO} we show the Higgs and gauge profiles for
the AdS vortex. At large $\ell$, the profile is essentially the same as
the pure NO-vortex. However as $\ell$ approaches the scale of the
vortex, the core is seen to narrow, and the power law fall-off becomes
more apparent. Although we can formally integrate these equations
for $\ell \lesssim 1$, it is unclear that such solutions with our boundary
conditions are physically
relevant, as the false vacuum $X=0$ becomes stable for Compton
wavelengths above the AdS scale \cite{BF}.
\section{Vortices in Kerr-AdS: Analytics}
Although the full exact solution of a vortex in a black hole background
must be found numerically, there are two ways in which
we can gain insight into the system analytically. The first is by
construction of an approximate solution, and the second is the
case of extremal black holes in which we can prove the existence
(or not) of a piercing solution on the event horizon.
We start by writing down the charged rotating black hole solution
\cite{CarterCMP}
\begin{equation}
ds^2=\frac{\Delta}{\Sigma}\left[dt-\frac{a\sin^2\!\theta}{\Xi}d\phi\right]^2
-\frac{\Sigma}{\Delta} dr^2-\frac{\Sigma}{S}d\theta^2
-\frac{S\sin^2\!\theta}{\Sigma}\left[a dt-\frac{r^2+a^2}{\Xi}d\phi\right]^2\,,
\label{KNADS}
\end{equation}
where
\begin{eqnarray}
\Sigma&=&r^2+a^2\cos^2\!\theta\,,\quad \Xi=1-\frac{a^2}{\ell^2}\,,
\quad S=1-\frac{a^2}{\ell^2}\cos^2\!\theta\,,\nonumber\\
\Delta&=&(r^2+a^2)\Bigl(1+\frac{r^2}{\ell^2}\Bigr)-2mr+q^2\,,
\end{eqnarray}
and the $U(1)$ potential is
\begin{equation}
A=-\frac{qr}{\Sigma}\left(dt-\frac{a\sin^2\!\theta}{\Xi}d\phi\right)\,.
\end{equation}
The mass $M$, the charge $\mathcal{Q}$, and the angular momentum $J$
are related to the parameters $m$, $q$, and $a$ as follows:
\begin{equation}\label{physical}
GM=\frac{m}{\Xi^2}\,,\quad G\mathcal{Q}
=\frac{q}{\Xi}\,,\quad GJ=\frac{am}{\Xi^2}\,.
\end{equation}
The ergosphere is located at $\Delta = a^2 S \sin^2\theta$,
and the horizon at $\Delta=0$. For large $\ell$, the horizon
is just slightly perturbed from its Kerr-Newman value. As
$\ell$ decreases, the horizon radius drops, and for small $\ell$
asymptotes to $m^{1/3} \ell^{2/3}$ (or $\sqrt{q\ell}$ for nonzero charge).
We see therefore that for smaller values of $\ell$, the fact that $m\gg1$
is no guarantee that the horizon radius must also be similarly
large in general. However, as we have already remarked, we do
not expect $\ell\lesssim1$ to be physically relevant. Therefore
in any analytic approximation, we will assume $\ell>1$.
Before moving to the vortex equations and analytic results,
it is worth remarking on the behaviour of the horizon radius
in a little more detail, and how this depends on $\ell$. This
is most succinctly captured by the extremal horizon radius,
when $\Delta = \Delta'=0$, which implies
\begin{equation}
r_+ = \frac{\ell}{\sqrt{6}} \left [
\left ( \left ( 1 + \frac{a^2}{\ell^2} \right )^2
+ 12 \left (\frac{a^2+q^2}{\ell^2} \right)
\right)^{1/2} - \left ( 1 + \frac{a^2}{\ell^2} \right ) \right] ^{1/2}\,.
\label{rplusext}
\end{equation}
We see therefore that $r_+(a,q,\ell) < \sqrt{a^2+q^2}$, the
Kerr-Newman value. Moreover, as $\ell$ drops, it is easy to
see that $r_+$ also drops, and for $\ell \lesssim 10$
drops quite sharply. Therefore, for the purposes of finding
an approximate solution for the vortex functions, which
typically assumes the black hole is large, we must consider
$\ell \gtrsim 10$, and for considerations of flux expulsion,
which typically happens for small black holes, we would expect
any argument to be sensitive to the value of $\ell$.
To find the vortex equations, we must consider not only the $X$
and $P_\phi$ functions, but also a nonzero $P_t$:
\begin{eqnarray}
0&=&\Delta X_{,rr}+\Delta' X_{,r}+S X_{,\theta\theta}
+\cot\theta\Bigl(S +\frac{2a^2}{\ell^2}\sin^2\theta\Bigr)
X_{,\theta}\nonumber\\
&&+\Sigma P_\mu^2X-\frac{\Sigma}{2}X(X^2-1)\,, \label{PXeq}\\
&&\nonumber\\
\frac{X^2}{\beta}P_t&=&
\frac{\triangle}{\Sigma}P_{t,rr}
+\frac{S}{\Sigma}P_{t,\theta\theta}
+\frac{2a\Xi\cot\theta}{\Sigma^3}\Bigl(\rho^2 S -\Delta
+\frac{a^2}{\ell^2}\Sigma\sin^2\!\theta\Bigr)P_{\phi,\theta}\nonumber\\
&&-\frac{a\Xi}{\Sigma^3}\Bigl(2r(Sa^2\sin^2\!\theta-\Delta)
+\Sigma\Delta'\Bigr)P_{\phi,r}\nonumber\\
&&+\frac{\cot\theta}{\Sigma^3}\left(S\bigl(\rho^4+a^4\sin^4\!\theta\bigr)
-2a^2\sin^2\!\theta\Bigl(\Delta-\frac{\rho^2\Sigma}{\ell^2}\Bigr)\right)
P_{t,\theta}\nonumber\\
&&-\frac{\sin^2\!\theta}{\Sigma^3}\Bigl(a^2\bigl(2r\rho^2S+\Sigma\Delta'\bigr)
-\frac{2r\rho^2\Delta}{\sin^2\!\theta}\Bigr)P_{t,r}\,, \label{Pteq}\\
&&\nonumber\\
\frac{X^2}{\beta} P_\phi&=&
\frac{\Delta}{\Sigma}P_{\phi,rr}
+\frac{S}{\Sigma}P_{\phi,\theta\theta}
+\frac{\rho^2}{\Sigma^3}\bigl(2r S a^2\sin^2\!\theta +\Sigma \Delta'
-2r\Delta\bigr)P_{\phi,r}\nonumber\\
&&+\frac{\cot\theta}{\Sigma^3}\left( 2a^2\sin^2\!\theta
\Bigl(\Delta-\frac{a^2}{\ell^2}\Sigma \sin^2\!\theta\Bigr)
-S\Bigl(a^2\sin^2\!\theta(\rho^2-\Sigma)+\rho^4\Bigr)\right)
P_{\phi,\theta}\nonumber\\
&&+\frac{2\cot\theta a^3\sin^4\!\theta}{\Xi\Sigma^3}
\left(\Delta-\rho^2\Bigl(1+\frac{r^2}{\ell^2}\Bigr)\right) P_{t,\theta}\nonumber\\
&&+\frac{a\sin^2\!\theta}{\Xi\Sigma^3} \Bigl( 2r
\bigl(\rho^4 S-\Delta(\Sigma+\rho^2)\bigr)
+\rho^2\Sigma\Delta'\Bigr)P_{t,r}\,, \label{Ppeq}
\end{eqnarray}
where $\rho^2=r^2+a^2$ has been introduced for visual clarity,
$\Delta'= d\Delta/dr$, and
\begin{equation}\label{Psquared}
P_\mu^2=\frac{(\rho^2P_t+a\Xi P_\phi)^2}{\Sigma \Delta}
-\frac{(\Xi P_\phi+a\sin^2\!\theta P_t)^2}{\Sigma S\sin^2\!\theta}\,.
\end{equation}
\subsection{Approximate solution}
As with the original Schwarzschild, Reissner-Nordstrom and Kerr
black holes, it is useful to develop an analytic approximate solution.
Clearly we expect this to make use of the (possibly AdS) Nielsen
Olesen solutions, and to depend on a single function of $r$ and $\theta$.
Consider the function
\begin{equation}
R\equiv \frac{\rho}{\sqrt{\Xi}}\sin\theta\,,
\end{equation}
which tends to the Kerr expression $\rho\sin\theta$ as $\ell\to\infty$.
Then, assuming that the vortex is much thinner than the black hole
horizon radius means that $\rho$ is always much greater than one,
and focusing on the core region of the vortex [$R<{\cal O}(10)$]
means that $\sin\theta\ll1$. We can therefore expand the
metric functions
\begin{equation}
\Sigma = \rho^2 \left ( 1 - \frac{a^2 R^2 \Xi}{\rho^4}\right)\simeq \rho^2\,,\quad
S= \Xi \left ( 1 + \frac{a^2 R^2}{\ell^2 \rho^2}\right)\simeq \Xi\,,
\end{equation}
and derivatives as
\begin{eqnarray}
\frac{\partial~}{\partial r} &=& \frac{Rr}{\rho^2} \frac{d~}{dR}\,,\qquad
\frac{\partial~}{\partial\theta} =\frac{\rho}{\sqrt{\Xi}}
\left ( 1 - \frac{\Xi R^2}{\rho^2} \right)^{1/2} \frac{d~}{dR}
\simeq \frac{\rho}{\sqrt{\Xi}} \frac{d~}{dR}\,, \nonumber \\
\Delta \frac{\partial^2~}{\partial r^2}
+ S \frac{\partial^2~}{\partial\theta^2}
&=& \left [ S\Bigl( \frac{\rho^2}{\Xi} - R^2 \Bigr)
+ \frac{\Delta R^2 r^2}{\rho^4} \right] \frac{d^2~}{dR^2}
+ \Bigl( \frac{\Delta a^2}{\rho^4}-S\Bigr) R \frac{d~}{dR} \\
&\simeq& \rho^2 \Bigl( 1 + \frac{\Delta R^2}{\rho^4}\Bigr)
\frac{d^2~}{dR^2}\,, \nonumber
\end{eqnarray}
to leading order in $R/\rho$. This already leads to significant
simplification of several of the terms in \eqref{PXeq}-\eqref{Ppeq}.
Then a little experimentation suggests the following approximate
functions
\begin{equation}\label{ApproxSol}
X \simeq X_{0}(R)\,,\quad
P_\phi \simeq P_{0}(R)\,,\quad
P_t \simeq \frac{a}{\rho^2}\Bigl(\frac{\Delta}{\rho^2}
-\Xi\Bigr) P_{0}(R)\,,
\end{equation}
which to leading order give the approximate equations:
\begin{equation}
\begin{aligned}
0&= \left ( 1 + \frac{\Delta R^2}{\rho^4}\right ) X_0''
+ \left ( 1+ \frac{4R^2}{\ell^2} \right )\frac{X_0'}{R}
- \frac{P_0^2X_0}{R^2}-\frac{X_0}{2}(X_0^2-1)\,, \\
\frac{X_0^2}{\beta} P_0 &=
\left ( 1 + \frac{\Delta R^2}{\rho^4}\right )P_0'' -
\left ( 1 + \frac{\left( 2\Delta - r\Delta'\right) R^2}{\rho^4}
\right) \frac{P_0'}{R}\,.
\end{aligned}
\end{equation}
Away from the horizon, $\Delta \sim \rho^4/\ell^2$ to leading
order, and we recover the AdS Nielsen-Olesen equations
\eqref{AdSNO}. However retaining the $R^2/\ell^2$ terms
is perhaps misleading, as we require $\ell>{\cal O}(10)$
in order for the horizon radius of an extremal black hole not to be
too small. We also see that on (or near) the horizon, the
${\cal O}(R^2/\ell^2)$ corrections to the Nielsen-Olesen
equations fail to have the precise AdS form. This implies that
while we can use the analytic approximation to good effect away
from the black hole, near the horizon we would expect
corrections to our solution at order ${\cal O}(\ell^{-2})$.
Note that because of the behaviour of $\Delta$ at large $r$,
the approximation for $P_t$ in \eqref{ApproxSol} actually
becomes proportional to $P_\phi$ at large $r$:
$P_t\sim a P_\phi/\ell^2$. Our gauge field is thus
\begin{equation}
\mathbi{P} = P_\phi d\phi+ P_t dt \sim
P_0(R) \left (d\phi+ \frac{a}{\ell^2} dt \right)\,,
\end{equation}
therefore it would appear that we have an electric field
inside our vortex far from the black hole. In fact, this
is simply an artifact of the Boyer-Lindquist style
coordinates we have used in \eqref{KNADS}, which
asymptote AdS$_4$ in a rotating frame with angular
momentum $\Omega_\infty=a/\ell^2$ \cite{GPP}.
One may remove this rotation by introducing new variables
\begin{equation}
\varphi=\phi+\frac{a}{l^2}t\,,\quad T=t\,.
\end{equation}
It is then easy to check that $\mathbi{P}$ in \eqref{ApproxSol} now reads
\begin{equation}
\mathbi{P}=P_0(R)\Bigl(d\varphi-\frac{a(2mr-q^2)}{\rho^4}dT\Bigr)\,.
\end{equation}
The $P_T$ component is now negative definite and falls
off appropriately at large $r$. The form of this solution is
now identical to that used in \cite{GKW}.
Figure \ref{fig:compare} shows a comparison of this
pseudo-analytic approximation with a numerically obtained
solution for an extremal low mass lowish $\ell$ black hole.
We take the values $m=3, \ell=20, q=0$, and with $a\simeq2.939$
at its extremal value in order to draw a parallel with the plot
in \cite{GKW}.
What is clearly shown is that the approximation is extremely good
almost everywhere, the only slight discrepancy appearing
near the event horizon -- as expected given the structure of
the corrections to the approximation there.
\begin{figure}[hb]
\begin{center}
\includegraphics[width=4cm]{compareX.pdf}\nobreak
\includegraphics[width=4cm]{compareP.pdf}\nobreak
\includegraphics[width=4cm]{compareQ.pdf}\\
\caption{{\bf Approximate vs. numerical solution:}
In each case the numerical solution
is shown in solid colour, and the approximation in dashed black.
Contours of $0.1-0.9$ (in steps of $0.2$) of the range of each
field are shown. From left to right: The Higgs field in blue,
the $P_\phi$ field in red, and $P_T$ (the component
with respect to the nonrotating frame at infinity) in brown.
For $P_T$,
we show contours of $0.1-0.9$ of the maximal negative
value, which is attained on the poles of the horizon.
The outer grey curve represents the boundary of the ergosphere.
}
\end{center}
\label{fig:compare}
\end{figure}
\subsection{Extremal black holes}
The extremal horizon exhibits
a Meissner effect for the cosmic string, in which if the black
hole becomes too `small' the cosmic string magnetic flux
is expelled from the black hole, and the horizon remains in
the false vacuum. For both Reissner-Nordstrom \cite{BEG}
and Kerr \cite{GKW} black holes, the existence of this phase
transition has been proven analytically, as well as
demonstrated numerically. The Reissner-Nordstrom
transition is second order, corresponding to a continuous
change in the order parameter (the magnitude of the Higgs
field) between piercing and expelling solutions. For the
Kerr black hole however, the phase transition was first order,
corresponding to a discontinuous change in the value
of the gradient of the zeroth component of the gauge field
between piercing and expelling solutions.
We will now argue for the existence of a Meissner effect in
the AdS-Kerr-Newman black holes; the Kerr-Newman situation
follows from taking the large-$\ell$ limit. Begin by defining new
variables $P$ and $Q$:
\begin{equation}
S P = \Xi P_\phi + a \sin^2\!\theta P_t\,,
\quad (r-r_+)Q = \rho^2 P_t + a \Xi P_\phi\,,
\end{equation}
where the factors have been chosen so that the
horizon equations are clearly identifiable, and the
range of $P$ is $P\in[0,1]$.
The field equations \eqref{PXeq}-\eqref{Pteq} become
\begin{eqnarray}
0 &=& \frac{\Delta}{\Sigma}X,_{rr}+\frac{\Delta'}{\Sigma}X,_r
+\frac{1}{\Sigma\sin\theta}
\Bigl(S\sin\theta X,_{\theta}\Bigr),_{\theta}
\nonumber \\
&& +\left(\frac{(r-r_+)^2 Q^2}{\Sigma\,\Delta}
-\frac{P^2}{\Sigma\,S\,\sin^2\theta}\right)X -\frac{X}{2}(X^2-1) ,
\label{bulkextX} \\
\frac{X^2P}{\beta} &=& \frac{\Delta}{\Sigma}P,_{rr}
+\frac{S}{\Sigma}P,_{\theta\theta}+\frac{\Sigma\Delta'
-2r\Delta}{\Sigma^2}P_{,r}
+\frac{\cot \theta}{\Sigma}\Bigl(4\frac{a^2}{l^2}\sin^2\!\theta
-\frac{S}{\Sigma}\bigl(\Sigma-2a^2\sin^2\!\theta\bigr)\Bigr)P,_{\theta}
\nonumber \\
&& +\frac{2a \sin^2\theta}{\Sigma^2}
\left((r-r_+)\left (r Q_{,r}-\cot\theta\,Q_{,\theta}- Q\right)
+aP\Bigl(1-\frac{r^2}{l^2}\Bigr) + rQ\right)\,,
\label{bulkextP} \\
\frac{X^2 Q}{\beta} &=& \frac{\Delta}{\Sigma}
\frac{\left[(r-r_+)Q\right]_{,rr} }{(r-r_+)}
+\frac{S}{\Sigma}Q_{,\theta\theta}
+ \frac{\cot\theta}{\Sigma^2}(2a^2\sin^2\theta
(1 + \frac{r^2}{\ell^2}) +S\,\Sigma ) Q_{,\theta}\nonumber \\
&& + \frac{2\Delta}{\Sigma^2} \left (\frac{a}{(r-r_+)}
(r S P_{,r}-S\cot\theta P_{,\theta}-(2-S)P) -rQ_{,r} -\frac{r_+Q}{(r-r_+)}
\right)\,, \label{bulkextQ}
\end{eqnarray}
which in the extremal limit and on the horizon reduce to
\begin{eqnarray}
\left (S\sin\theta X'\right)' &=& X \sin\theta \left [
\frac{S P^2}{\sin^2\theta}-\frac{2Q^2}{\Delta''_+}
-\frac{\Sigma_+}{2}(1-X^2) \right] ,
\label{extX} \\
\left ( \frac{S^2 P'}{\Sigma_+\sin\theta}\right)' &=&
P S \sin\theta \left [
\frac{X^2}{\beta \sin^2\theta}
-\frac{2a^2}{\Sigma_+^2}
\left ( 1 - \frac{r_+^2}{\ell^2}\right) \right]
-\frac{2a r_+S Q \sin\theta}{\Sigma_+^2}
\,, \label{extP} \\
\left ( \frac{S \sin\theta Q'}{\Sigma_+} \right)'
&=& \frac{X^2 Q}{\beta} \sin\theta
\,,\label{extQ}
\end{eqnarray}
where a prime now denotes $d/d\theta$, and the ``$+$''
subscript indicates the function is evaluated at $r=r_+$,
given by \eqref{rplusext}. Note that unlike the vacuum
Kerr case, in which $r_+=a$, there is no simple
factorization of $\Sigma_+$ leading to a clean
$\theta$-dependence in these equations.
Note that if $a=0$, $Q\equiv0$, and $S\equiv1$ and
our system of horizon equations reduces {\it precisely}
to the Reissner-Nordstrom horizon equations studied
in \cite{BEG}. Therefore we expect essentially the same
analytic arguments to hold here (which is the case as we shall
see below). Further, since $Q$ vanishes, we expect
a second order phase transition governed by the continuous
order parameter $X$. On the other hand, if $a\neq0$, $Q$ is
nonzero in the bulk of the spacetime and so we must
examine the full system of horizon equations.
Let us look first at the behaviour of the horizon function $Q$,
as this will give us the order of the phase transition.
For a piercing solution, $X$ is nontrivial on the horizon.
Hence
\begin{equation}
S \beta \sin\theta Q'(\theta) = \Sigma_+ \int_0^\theta
{X^2 Q\sin\theta} d\theta\,, \label{Qcontra}
\end{equation}
upon integrating \eqref{extQ}.
We can easily see this cannot be true unless
$Q\equiv0$. Evaluating \eqref{Qcontra} at the first point
at which $Q'=0$ tells us that $\int_0^\theta
{X^2 Q\sin\theta}=0$, but $Q$ is either positive and
increasing on this range, or negative and decreasing:
in either case, the integrand is positive or negative
definite, thus cannot be zero.
Therefore $Q\equiv0$ for a piercing solution. On the other
hand, an expelling solution has $X\equiv0$, with $P_\phi=1$,
hence
\begin{equation}
\label{expelsol}
P=\frac{\Xi \Sigma_+}{\rho_+^2 S}\,, \qquad
Q \equiv - \frac{2ar_+\Xi}{\rho_+^2}\,.
\end{equation}
Given that $Q$ changes in a discontinuous fashion,
we see that the phase transition is first order for nonzero $a$.
It is clear that a flux expelling solution to the horizon
system of equations \eqref{extX}-\eqref{extQ} can exist. However to
prove flux expulsion happens, this solution must be extendable to a
bulk solution. To demonstrate this, we follow the argument
of \cite{BEG}. If flux is expelled, $X\equiv0$ on the horizon,
and must become nonzero and positive a small distance from
the horizon, implying $(\Delta X_{,r})_{,r} >0$ just outside
the horizon. Referring to \eqref{bulkextX}, we see therefore
that
\begin{equation}
(S \sin\theta X_{,\theta})_{,\theta} +
\frac{(r_+^2+a^2\cos^2\theta) X}{2}\sin\theta
< \frac{SP^2}{\sin\theta}X <\frac{SX}{\sin\theta}
\end{equation}
is required if a flux expelling solution is to exist.
Integrating this inequality on $[\theta_0,\pi/2]$ gives
\begin{equation}
S \sin\theta_0 X_{,\theta_0} > \int_{\theta_0}^{\pi/2} \left (
\frac{(r_+^2+a^2\cos^2\theta)\sin\theta}{2}
- \frac{S}{\sin\theta}\right)Xd\theta \,.
\end{equation}
Defining $\alpha$ so that $\Sigma_+
\sin^2\alpha/S =2$,
by taking $\theta_0>\alpha$ we can bound this integral from
below using $X(\theta)>X(\theta_0)$. We can also
bound the derivative of $X$ by
$X_{,\theta_0}<\frac{X(\theta_0)-X(\alpha)}{\theta_0-\alpha}
< \frac{X(\theta_0)}{\theta_0-\alpha}$, leading to
\begin{equation}
S \sin\theta_0 \frac{X(\theta_0)}{\theta_0-\alpha} >
S \sin\theta_0 X_{,\theta_0}
> X(\theta_0) \int_{\theta_0}^{\pi/2} \left (
\frac{(r_+^2+a^2\cos^2\theta)\sin\theta}{2}
- \frac{S}{\sin\theta}\right) d\theta\,,
\end{equation}
which implies
\begin{equation}
\frac{({\theta_0-\alpha})}{S(\theta_0)\sin\theta_0} \left (
\frac{r_+^2\cos\theta_0}{2}
+\frac{a^2\cos^3\theta_0}{6}
+\Xi\log\tan\Bigl(\frac{\theta_0}{2}\Bigr)
-\frac{a^2}{l^2}\cos\theta_0
\right)<1
\label{expelinequality}
\end{equation}
on the interval $[\alpha,\pi/2]$. If this inequality is violated,
then we cannot have flux expulsion, and the vortex {\it must}
pierce the black hole. Note, if $a=0$, then \eqref{expelinequality}
is independent of $\ell$, and reduces to the previously
explored Reissner-Nordstrom
relation \cite{BEG}, giving the same upper bound on the
horizon radius for flux expulsion of $\sqrt{8.5}$. For $a\neq0$,
we must explore the $\{a,\ell\}$ phase plane (having ensured
that a solution $\alpha$ exists) to determine the upper bound
on the horizon radius. Clearly if $\ell$ drops too low, we require
a large charge to allow for a solution to $\alpha$. Hence for a given
$q$, we expect a minimal value of $\ell$ for this upper bound to exist.
This is shown most clearly for $q=0$, in figure \ref{fig:rcritical}.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\textwidth]{Qzerophasetrans2.pdf}
\caption{{\bf Meissner effect:}
An illustration of the analytic bounds on the critical horizon radius
for the Meissner effect for $q=0$. In the shaded regions, the vortex
should either pierce the horizon, or be expelled as indicated. The
critical radius therefore lies between these two bounds. For sufficiently
low $\ell$, flux is always expelled. Numerically obtained transition
radii are indicated. The solid $r_+=\ell/\sqrt{3}$ line on the left indicates
the $a=\ell$ singular limit.}
\label{fig:rcritical}
\end{center}
\end{figure}
To argue that a Meissner effect should exist for sufficiently low
horizon scales, we assume a piercing solution to
\eqref{extX}-\eqref{extQ} exists, in which
$X$ and $P$ will have nontrivial profiles symmetric around
$\theta=\pi/2$, with $X$ maximised and $P$ minimised (at
least for large $\ell$ or small $a<q$) at $\pi/2$.
If $a=0$, the argument of \cite{BEG} can be used to
deduce that for $r_+ \lesssim 0.7$ the flux
must be expelled, and this argument can be
extended to include small $a$ (see appendix).
For $q=0$, or dominant $a$, an alternate argument must
be used. At large $\ell$, $P$ is minimised at $\pi/2$, which
implies a constraint on $r_+$ given by (writing $X_m=X(\pi/2)$):
\begin{equation}
\begin{aligned}
P''\left(\frac{\pi}{2}\right) &= P \left ( \frac{X_m^2r_+^2}{\beta}
- \frac{2a^2}{r_+^2} \Bigl( 1 - \frac{r_+^2}{\ell^2}
\Bigr) \right ) \geq0 \;\;\;
\Rightarrow \;\;\; r_+^4 + 2r_+^2
\frac{a^2\beta}{\ell^2} > 2 a^2\beta\,.
\end{aligned}
\end{equation}
However, for low values of $\ell$, we cannot show that
$P$ is minimised at $\pi/2$, and indeed scrutiny of piercing
solutions near the phase transition indicates a tiny
modulation in $P$. What we can say however, is that $P$
has at most one additional turning point on $[0,\pi/2]$,
as the source term on the RHS of \eqref{extP} is
monotonically decreasing on $[0,\pi/2]$, hence
$S^2P'/\Sigma_+\sin\theta$ has at most one turning
point where $X^2 \Sigma_+^2 = 2 a^2\beta \sin^2\theta
(1-r_+^2/\ell^2)$.
Suppose therefore that we are at low $\ell$ and $P$
has such a turning point on $[0,\pi/2]$.
Now consider $S^2/\Sigma_+\sin\theta$; the derivative
\begin{equation}
\left ( \frac{S^2}{\Sigma_+\sin\theta}\right)'
= -\frac{S\cot\theta}{\Sigma_+^2 \sin\theta} \left [
(r_+^2+a^2)\Xi - 3 a^2 \Bigl( 1 + \frac{r_+^2}{\ell^2} \Bigr)\sin^2\theta
+ \frac{a^4}{\ell^2} \sin^4\theta\right]
\end{equation}
has a zero at $\theta_0$, where
\begin{equation}
\frac{a^2}{\ell^2} \sin^2\theta_0 = \frac32
\left ( 1 + \frac{r_+^2}{\ell^2} \right) - \frac12
\sqrt{9\left ( 1 + \frac{r_+^2}{\ell^2} \right)^2 - 4 \Xi
\frac{r_+^2 + a^2}{\ell^2}}\,.
\end{equation}
For $q=0$, $\sin\theta_0\in[0,\sqrt{2/3}]$, as
$\ell$ ranges from $a$ to $\infty$, whereas the node
in $P$ only switches on for lower $\ell$, and initially
appears at $\pi/2$. Therefore at $\theta_0$
we expect $S^2P'/\Sigma_+\sin\theta>0$,
and hence
\begin{equation}
(r_+^2 + a^2 \cos^2\theta_0)^2 > X^2(\theta_0)
\Sigma_+^2(\theta_0) > 2 a^2\beta \sin^2\theta_0
\Bigl(1-\frac{r_+^2}{\ell^2}\Bigr)\,.
\end{equation}
Thus, if this equality is not satisfied at $\theta_0$, we
deduce that a piercing solution is not possible, and
expulsion must occur. Figure \ref{fig:rcritical} shows
this lower bound for $q=0$.
The full details of the phase transition must be determined
numerically, and figure \ref{fig:rcritical} shows the
numerically obtained critical horizon radius as a function of
$\ell$ for $q=0$ together with the analytic lower and upper
bounds on $r_{+,crit}$.
We discuss the phase transition further
in section \ref{discuss}.
\section{Numerical solution}
In order to obtain numerical solutions of the vortex equations
\eqref{PXeq}-\eqref{Ppeq}, which form an elliptic system,
we follow references \cite{AGK} and \cite{GKW}, employing
a gradient flow technique on a two-dimensional polar grid.
Briefly, this method introduces a fictitious time variable,
with the `rate of change' of our functions being proportional to
the actual elliptic equations we wish to solve:
\begin{equation}
\dot{Y}^i = \Delta Y^i + F^i (\bf{Y}, \nabla \bf{Y})\,,
\end{equation}
where $\Delta^i$ represents a second order (linear)
elliptic operator and $F$ is a (possibly nonlinear) function
of the variables $Y^i$ and their gradients, such that the RHS
is our system of elliptic equations. We now have a
diffusion problem, and solutions to
this new equation eventually ``relax'' to a steady state,
in which the variables are no longer changing
with each time step, and the solutions $Y^i$ satisfy our
elliptic equations. The only subtlety with the given set-up is
that our elliptic system has one boundary (the event horizon)
on which our equations become parabolic. This was
discussed in detail in \cite{AGK}, with the result that on each
grid update, we update the event horizon, using the horizon
equations, and fixing
\begin{equation}
P_t=-\frac{a\Xi P_\phi}{r_+^2+a^2}\,
\end{equation}
on the horizon, which is mandated by finiteness of the
energy-momentum tensor.
\begin{figure}
\begin{center}
\includegraphics[width=4.9cm]{contoursL100X.pdf}\nobreak
\includegraphics[width=4.9cm]{contoursL100P.pdf}\\
\includegraphics[width=4.9cm]{contoursL10X.pdf}\nobreak
\includegraphics[width=4.9cm]{contoursL10P.pdf}\\
\caption{{\bf AdS-Kerr vortex:} A depiction of the numerical
solution for the
AdS-Kerr vortex for an extremal uncharged rotating black hole.
The upper plots have $\ell = 100$, the lower plots
$\ell=10$. In each case, the contours of the
Higgs field are shown on the left in blue ($X=0.1-0.9$ in steps
of $0.2$), and on the right,
the angular component of the gauge field, $P_\varphi$
in red (with the same contour steps as for $X$), and
$P_T$ in dashed black with contours of $0.1-0.9$ of
$P_{T,min}= -0.0519, -0.116$ for the $\ell=100$ and
$\ell=10$ cases respectively.}
\end{center}
\label{fig:q0}
\end{figure}
As an initial condition for
the integration, we use the approximate solutions for the
functions $X$, $P_{\phi}$ and $P_{t}$ given in equations
\eqref{ApproxSol}, where we obtain the forms for $P_0(R)$
and $X_0(R)$ by numerically integrating \eqref{AdSNO}
on a one-dimensional grid. The approximate solution is
accurate to order $r^{-2}$, thus we choose our outer
boundary to be sufficiently far from the horizon that our
analytic approximation is extremely accurate near this
outer radial boundary, which is not updated in our code.
On axis we impose the standard vortex boundary conditions,
($X=0$, $P_\phi=1$) while leaving $P_t$ to relax
by continuity. As pointed out in \cite{GKW}, the fact that $P_t$ is
not restricted can be understood by noting that there is a
dyonic degree of freedom that is introduced into the solution
due to the presence of the black hole.
\begin{figure}
\begin{center}
\includegraphics[width=4.9cm]{contoursL500X.pdf}\nobreak
\includegraphics[width=4.9cm]{contoursL500P.pdf}\\
\includegraphics[width=4.9cm]{contoursL505X.pdf}\nobreak
\includegraphics[width=4.9cm]{contoursL505P.pdf}
\caption{{\bf AdS-Kerr-Newman vortex:} Numerical solutions
for the AdS-Kerr-Newman vortex with
$\ell = 50$ and $q = 0$, (upper) and $q=5$ (lower) with
the same contour conventions as for figure \ref{fig:q0},
with $P_{T,min} = -0.0569$ for $q=0$, and $P_{T,min}
=-0.0563$ for $q=5$.
}\end{center}
\label{fig:50}
\end{figure}
Figures \ref{fig:q0} and \ref{fig:50} show a selection
of the solutions obtained
from the integration method above which highlight the
effects of the parameters $\ell$ and $q$ on the rotating
black hole vortex.
In all plots, we have chosen to illustrate the solution
by plotting contour lines for each field of $0.1-0.9$
of the full range of the field in steps of $0.2$. Thus,
for the $X$ and $P_\phi$ fields, we have shown the
$0.1, 0.3, 0.5, 0.7$, and $0.9$ contours, but for
the $P_T$ field (note -- this is the gauge field component
with respect to a {\it non-rotating} frame at infinity)
the maximally negative value of $P_T$ is attained on
the horizon at the poles. The numerical values of these
contours therefore vary from plot to plot. The
actual value of $P_{T,min}$ is given in the captions.
Figure \ref{fig:q0} shows the vortex solution for the
case of $\ell =100$ and $\ell=10$ respectively, at the extremal
limit with the charge parameter $q$ set to zero.
The solution away from the extremal limit is similar (see
\cite{GKW}), the main difference being that the actual
numerical values of the $P_T$ contours are lower.
For $\ell=100$, the plots are almost indistinguishable
from the vacuum Kerr vortex solution analysed in \cite{GKW},
however, for $\ell=10$, the effect of the cosmological
constant can be easily seen. Comparing the figures, one notes
that dropping the value of $\ell$ strongly impacts the size of
both the black hole horizon as well as the vortex, causing the
vortex width to tighten, the $P_T$ fields to shrink closer
to the horizon, which itself shrinks significantly.
Figure \ref{fig:50} then demonstrates the effect of adding a
non-zero charge to the AdS-Kerr vortex. As can be seen,
this does not significantly impact the vortex, and appears to
merely shift the horizon and ergosphere inwards,
while slightly causing the
$P_T$ contour lines to creep closer to the horizon, as is
expected since the rotation parameter $a=a_{ex}$ will
be lower with the charged black hole at the same mass.
\section{Discussion}\label{discuss}
We have examined the behaviour and interactions of vortices
with asymptotically AdS charged and rotating black holes.
We first obtained an approximate solution to the abelian Higgs Model
in the background of a Kerr-Newman AdS black hole,
and showed that the Nielsen-Olesen equations retain their
AdS form up to corrections of order $R^2/\ell^2$. Consequently
we found that our approximation was extremely good everywhere
except near the event horizon as expected. The comparison
illustrated in figure \ref{fig:compare} shows that the actual solution
has a stronger expulsion of flux than the approximation.
Upon transforming to a frame that is non-rotating at the boundary,
the form of our solution is very close to its asymptotically
flat counterpart.
For extremal black holes we explored the existence of a
Meissner effect with the cosmic string flux being expelled
from the black hole at small horizon radii (although one
should be cautious about the stability of such small
black holes \cite{CarDias}). We presented analytic arguments to
show that such a phase transition exists, showing that
in the presence of rotation it is a first order transition.
We numerically explored the phase space to confirm
this expectation, and figure \ref{fig:phasetrans} shows the
numerical results for the phase transition at several
values of $\ell$ and $\beta$. The existence of the first
order transition is confirmed, and the effect of $\ell$ is to
lower the critical value of $r_+$ at which the transition
occurs. This is also reflected in a drop of both analytic
bounds for expulsion and piercing of the vortex. We also
notice that the value of the order parameter ($X(\pi/2)$)
rises with decreasing $\ell$, seen in the right
plot of figure \ref{fig:phasetrans}. The left
plot of figure \ref{fig:phasetrans}
shows the effect of changing $\beta$, and is similar to
the corresponding plot for the vacuum Kerr solution
in \cite{GKW}. However the effect of changing $\beta$
is far more pronounced at the relatively low value of
$\ell=10$ illustrated. Note that, unlike pure Kerr, the
plots do not extend to $r_+^{-1}\to0$: there is an
upper limit on the angular momentum, and hence
horizon radius.
\begin{figure}
\begin{center}
\includegraphics[width=0.49\textwidth]{phtransbetas.pdf}\nobreak
\includegraphics[width=0.49\textwidth]{phtransells.pdf}
\caption{{\bf Flux expulsion behavior:}
Plots illustrating features of the flux expulsion phase
transition on the event horizon of the black hole. The
maximal value of the Higgs field $X_m=X(\pi/2)$ is shown
as a function of $r_+^{-1}$ for varying $\beta$ (left) and
$\ell$ (right).
}
\label{fig:phasetrans}
\end{center}
\end{figure}
The numerical integrations are considerably more sensitive
with the addition of the cosmological constant, mainly because
an additional scale has been added which causes the vortex to
contract, as well as the black hole. Unfortunately this has prevented
us from investigating the small-$\ell$ case in significant detail.
This is the region of interest for a holographic interpretation of
our results, though our solution would only be relevant in the IR
as it does not have the requisite boundary conditions.
Vortices in the bulk can be interpreted as defects in the dual CFT
\cite{VH,Dias:2013bwa}, where in the IR they are heavy pointlike
excitations in a superfluid around which the phase of the
condensate winds. A vortex must have a core radius since the
vanishing of the condensate at its location is energetically costly and
so must happen over some finite region. A recent study
\cite{Dias:2013bwa} of vortices with planar black holes has
indicated that their IR physics can be understood from the
viewpoint of a defect or boundary CFT \cite{Cardy:2004hm}.
A study of holographic superconductivity in the context of
(topologically spherical) rotating black holes \cite{Sonner}
found that the superconducting state in the dual theory
(for certain choices of parameters) can be destroyed for
sufficiently large rotation. The localization of the condensate
depends on the sign of the mass-squared term of the scalar,
with a droplet/ring-like structure appearing for positive/negative
values of this term. The instability towards forming vortex
anti-vortex pairs depends on this sign \cite{Sonner}.
It would be interesting to study these effects further in light of
our results. Even without rotation we obtain a Meissner effect,
and so very small black holes with expelled flux can exist.
Their interpretation in the context of the boundary theory (as well
as distinguishing them from the flux-pierced case) remains to be
understood, perhaps in
terms of the absence of a mass gap for the flux-expelled case.
\acknowledgments
RG is supported in part by STFC (Consolidated Grant ST/J000426/1),
in part by the Wolfson Foundation and Royal Society, and in part
by Perimeter Institute for Theoretical Physics. DK is supported by
Perimeter Institute. DW is supported by an STFC studentship. DW
would also like to thank Perimeter Institute for hospitality.
Research at Perimeter Institute is supported by the Government of
Canada through Industry Canada and by the Province of Ontario through the
Ministry of Research and Innovation. This work was supported in part
by the Natural Sciences and Engineering Research Council of Canada.
|
\section{Introduction}\label{introd}
The modeling of fracture in materials leads naturally to functional spaces with
discontinuities, in particular functions of bounded variation ($BV$) and of
bounded deformation ($BD$). In variational models, the key ingredients are a
volume term, corresponding to the stored energy and depending on the
diffuse part of the deformation gradient, and a surface term, modeling the
fracture energy and depending on the jump part of the deformation
gradient \cite{fra-mar,BFM,barenblatt,dug}.
For antiplane shear models one can consider a scalar displacement
$u\in BV(\Omega)$, typical models take the form
\begin{equation}\label{eqfrattura}
\int_\Omega h(|\nabla u|) dx + \kappa\,|D^cu|(\Omega)
+ \int_{\Omega\cap J_u} g(|[u]|) d{\mathcal H}^{n-1}\,.
\end{equation}
Here $h$ represents the strain energy density, quadratic near the origin; $g$ is
the surface energy density depending on the opening $[u]$ of the crack, and $\kappa\in[0,+\infty]$
is a constant related to the slope of $g$ at 0 and the slope of $h$ at $\infty$.
In models of brittle fracture one usually considers $g$ to be a constant,
given by twice the energy required to generate a free surface
\cite{fra-mar,BFM}.
Correspondingly $\kappa=\infty$ and the Cantor part $D^cu$ disappears, so that one
can assume $u\in SBV$.
Physically this represents a situation in which already for the smallest
opening there is no interaction between the two sides of the fracture,
surface reconstruction is purely local. Analytically, the resulting functional
coincides with the Mumford-Shah functional from image segmentation.
In ductile materials fracture proceeds through the opening of a series of
voids, separated by thin filaments which produce a weak bound between the
surfaces at moderate openings \cite{barenblatt,dug,FokouaContiOrtiz2014}. The
function $g$
grows then continuously from
$g(0)=0$ to some finite value $g(\infty)$, representing the energetic cost of
total fracture. The constant $\kappa$ is its slope at 0, and $D^cu$ represents the
distribution of microcracks. For the same reason the volume energy density $h$
becomes linear at $\infty$.
A large literature was devoted to the derivation of models like
(\ref{eqfrattura}) from more regular models, like damage or phase field
models, mainly within the framework of $\Gamma$-convergence.
These regularizations can be interpreted as microscopic physical models,
so that the $\Gamma$-limit justifies the macroscopic model
(\ref{eqfrattura}), or as regularizations used for example to approximate
(\ref{eqfrattura}) numerically.
Ambrosio and Tortorelli \cite{amb-tort1,amb-tort2} have shown that
\begin{equation}\label{eqAT}
\int_\Omega \left( (v^2+o(\varepsilon)) |\nabla u|^2 + \frac{(1-v)^2}{4 \varepsilon} +
\varepsilon|\nabla v|^2 \right) dx
\end{equation}
converges to the Mumford-Shah functional, which coincides with
(\ref{eqfrattura}) with $h(t)=t^2$, $\kappa=\infty$, $g(t)=1$.
This result was extended in many directions, for example to
vector-valued functions \cite{focardi,focardi_tesi}, to
linearized elasticity \cite{chambolle,addendum,iur12}, to second-order
problems \cite{AFM}, vectorial problems \cite{Sh},
and models with nonlinear injectivity constraints
\cite{HenaoMoracorralXu}; for numerical simulations
we refer to \cite{bel-cos,bou-fra-mar2,bou,bur-ort-sul,bur-ort-sul2}.
There is also a large numerical literature on the application to computer
vision, see for example \cite{Fusco2003,BarSochenKiryati2006} and references therein.
Models like (\ref{eqfrattura}) with a linear $h$ were obtained in \cite{AlicBrShah, AlFoc}.
The case with a quadratic $h$ and an affine $g$, i.e. $h(t)=t^2$, $g(t)=t+c$, and $\kappa=+\infty$, described
in \cite{amb-lem-roy} a strain localization plastic process. From the mathematical point of view
the functional was limit of models like (\ref{eqAT}) with an additional term linear in $|\nabla u|$.
In \cite{dm-iur,iur} the asymptotic behavior of a generalization of \eqref{eqAT} with different
scalings of the three terms was analyzed. In one of the several regimes identified, the limiting model
again exhibited an affine $g$. The result was then extended to the vectorial case in \cite{foc-iur}.
In \cite{iur} a different scaling of the parameters led to the Hencky's diffuse plasticity, i.e. to a model
like (\ref{eqfrattura}) with a linear $g$. This functional can be used to describe ductile fracture only at small openings.
Discrete models for fracture were studied for example in \cite{bra-gar-dm,BraidesGelli2006}.
Up to now, we are unaware of any result in which a ductile fracture model with $g$
continuous and bounded, as described above, has been derived.
In this work we study a damage model as proposed by Pham and Marigo
\cite{PM1,PM2} (cp. Remark~\ref{r:gen}), namely,
\begin{equation}\label{eqPM}
F_\varepsilon(u,v):=\int_\Omega \left( f_\varepsilon^2(v) |\nabla u|^2 + \frac{(1-v)^2}{4 \varepsilon} +
\varepsilon|\nabla v|^2 \right) dx,
\end{equation}
with $u,v\in H^1(\Omega)$, $0\leq v\leq 1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$, and $F_\varepsilon(u,v):=\infty$ otherwise,
and show that it converges to a cohesive fracture model like
(\ref{eqfrattura}), where $g$ is a continuous bounded function with $g(0)=0$,
which is linear close to the origin. The potential $f_\varepsilon:[0,1)\to[0,+\infty]$ in \eqref{eqPM} is defined by
\begin{equation}\label{e:feintro}
f_\varepsilon(s):=1\wedge \varepsilon^{1/2}f(s),
\end{equation}
where $f\in C^0([0,1),[0,+\infty))$ is nondecreasing, $f^{-1}(0)=\{0\}$, and it
satisfies
\begin{equation}\label{eqdivergencefintro}
\lim_{s\to1}(1-s)f(s)=\ell, \quad \ell\in(0,+\infty).
\end{equation}
Our main result describes the asymptotic of $(F_\varepsilon)$ as follows.
\begin{theorem}\label{t:mainintro}
Let $\Omega\subset{\mathbb R}^n$ be a bounded Lipschitz set.
Then, the functionals $F_\varepsilon$ $\Gamma$-converge in $L^1(\Omega){\times}L^1(\Omega)$
to the functional $F$ defined by
\begin{equation*}
F(u,v):=\begin{cases}
\displaystyle\int_\Omega h(|\nabla u|)dx+\int_{J_{u}}g(|[u]|)d{\mathcal H}^{n-1}+\ell|D^cu|(\Omega) & \textrm{if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$, $u\in GBV(\Omega)$}\cr
+ \infty & \text{otherwise}.
\end{cases}
\end{equation*}
Here the volume energy density $h$ is set as $h(s):=s^2$ if $s\leq\ell/2$ and as $h(s):=\ell s-\ell^2/4$ otherwise,
while the surface energy density $g$ is given by
\begin{alignat}1\nonumber
g(s):=\inf &
\left\{
\int_0^1|1-\beta|\sqrt{f^2(\beta)|\alpha'|^2+|\beta'|^2}\,dt:\,(\alpha,\beta)
\in H^1\big((0,1)\big),\right.
\\&\left. \phantom{\int}\hskip25mm \alpha(0)=0,\ \alpha(1)=s,\ \beta(0)=\beta(1)=1
\right\}.\label{eqintrodefg}
\end{alignat}
\end{theorem}
The key difference with the previously discussed work is that in our case the
optimal profiles for the damage variable $v$ and the elastic displacement $u$
cannot be determined separately. They instead arise from a joint vectorial
minimization problem which defines the cohesive energy $g$, specified in \eqref{eqintrodefg}.
Figures \ref{fig1} and \ref{fig2} show the behavior of $f_\varepsilon$ and $g$ in the case $f(s)=s/(1-s)$,
Figure \ref{fig3} shows the profiles $\alpha$, $\beta$ entering (\ref{eqintrodefg}).
Theorem~\ref{t:mainintro}, in the equivalent formulation given in Theorem \ref{t:gamma-lim} below,
is proved first in the one-dimensional case in Section~\ref{s:onedim},
relying on elementary arguments in which we estimate separately the diffuse and jump contributions,
and then extended to the general $n$-dimensional setting in Section~\ref{s:ndim}. This
extension is obtained by means of several tools. A slicing technique and the above mentioned
one-dimensional result are the key for the lower bound inequality.
Instead, the upper bound inequality is proved through the direct methods of $\Gamma$-convergence
on $SBV$, i.e. abstract compactness results and integral representation of the corresponding
$\Gamma$-limits. The latter methods are complemented with an ad-hoc one-dimensional construction
to match the lower bound on $SBV$ and a relaxation procedure to prove the result on $BV$.
Finally, the extension to $GBV$ is obtained via a simple truncation argument.
The issues of equi-coercivity of $F_\varepsilon$ and the convergence of the related minima are dealt with
in Theorem~\ref{t:comp} and Corollary~\ref{c:min} below, respectively.
Qualitative properties of the surface energy density $g$ defined in \eqref{eqintrodefg} are analyzed
in Section~\ref{s:gprop}. Its monotonicity, sublinearity, boundedness and linear behavior in the
origin are established in Proposition~\ref{11}. Proposition~\ref{p2} characterizes $g$ by means of
an asymptotic cell formula particularly convenient in the proof of the $\Gamma$-limsup inequality.
Furthermore, the dependence of $g$ on $f$ is analyzed in detail in Proposition~\ref{p:gl}.
The latter results on one hand show the variety of such a class of functions, and on the other hand
are instrumental to handle the approximation of other models.
In Section~\ref{s:further} we discuss how the phase field approximation scheme
can be used to approximate different fracture models.
We first consider damage functions of the form
\begin{equation*}
f_k(s):=\min\{1, \varepsilon_k^{1/2} \max\{f(s), a_ks\}\}
\end{equation*}
and show that if $a_k\to\infty$ and $a_k \varepsilon_k^{1/2}\to0$ then the a similar result holds with the limiting
surface energy $g(s)=1\wedge (\ell s)$, so that (\ref{eqfrattura})
reduces to Dugdale's fracture model
(Theorem~\ref{t:Dugdale} in Section \ref{subsecdugdale}).
Secondly we consider a situation in which $f$ diverges with exponent $p>1$ close to $s=1$, so that
(\ref{eqdivergencefintro}) is replaced by
\begin{equation*}
\lim_{s\to1}(1-s)^p f(s)=\ell\,.
\end{equation*}
Also in this case the functionals $\Gamma$-converge to a problem of the form
of (\ref{eqfrattura}), in this case however the fracture energy $g$ turns out
to be proportional to the opening $s$ to the power $2/(p+1)$ at small $s$. Correspondingly the coefficient $\kappa$ of the diffuse part is infinite, so that the limiting problem is framed in the space GSBV, see Theorem \ref{t:sublin} in
Section \ref{subsectpowerlaw}.
Finally we show that if $f_k(s)$ diverges as $\ell_k/(1-s)$, with $\ell_k\to\infty$, then Griffith's fracture model is recovered in the limit, see Theorem \ref{t:MS} in Section \ref{subsecgriffith} below.
We finally resume the structure of the paper. In Section~\ref{s:nota} we introduce some notations,
some preliminaries, and the functional setting of the problem.
The main result of the paper is stated in Section~\ref{s:stat},
where we also discuss the convergence of related minimum problems and
minimizers.
Our $\Gamma$-convergence result relies on several properties of the
surface energy density $g$ that are established in Section~\ref{s:gprop}.
The proof is then given first in the one-dimensional case in
Section~\ref{s:onedim} and then in $n$ dimensions in Section \ref{s:ndim}.
The three generalizations are discussed and proven in Section~\ref{s:further}.
\begin{figure}
\includegraphics[width=5cm]{fig1-crop}
\caption{Sketch of the function $f_\varepsilon(s)$ for the prototypical case $f(s)=s/(1-s)$.}
\label{fig1}
\end{figure}
\begin{figure}
\includegraphics[width=5cm]{fig2-crop}
\caption{Sketch of the function $g(s)$ defined in (\ref{eqintrodefg}), obtained by numerical
minimization using $f(s)=s/(1-s)$ (cp. Proposition~\ref{11} and Remark~\ref{r:g}).}
\label{fig2}
\end{figure}
\begin{figure}
\includegraphics[width=5cm]{fig3-crop}
\caption{Optimal profiles $(\alpha/s,\beta)$ obtained numerically from the minimization in the definition of
$g(s)$, see (\ref{eqintrodefg}), for $f(s)=s/(1-s)$ and for $s=0.1, 0.3, 0.5, 1,$ and $1.5$ (from top to bottom).
All curves remain inside the rectangle $(0,1)\times (0,1)$ except for the two endpoints.}
\label{fig3}
\end{figure}
\section{Notation and preliminaries}\label{s:nota}
Let $n\geq1$ be a fixed integer. We denote the Lebesgue measure and the $k$-dimensional Hausdorff measure in
$\mathbb{R}^n$ by $\mathcal{L}^n$ and $\mathcal{H}^k$, respectively. Given $\Omega\subset{\R}^n$ an open bounded
set with Lipschitz boundary, we define $\mathcal{A}(\Omega)$ as the set of all open subsets of $\Omega$.
Throughout the paper $c$ denotes a generic positive constant that can vary from line to line.
\subsection{\texorpdfstring{\boldmath{$\Gamma$}}{Gamma}- and {\texorpdfstring{\boldmath{$\overline{\Gamma}$}}{Gamma bar}-convergence}}
Given an open set $\Omega\subset{\R}^n$ and a sequence of functionals
$\mathscr{F}_k:X\times\mathcal{A}(\Omega)\to [0,+\infty]$, $(X,d)$ a separable metric space,
such that the set function $\mathscr{F}_k(u;\cdot)$ is nondecreasing on the family $\mathcal{A}(\Omega)$
of open subsets of $\Omega$, set
\[
\mathscr{F}'(\cdot; A):=\Gamma\hbox{-}\liminf_{k\to+\infty} \mathscr{F}_k(\cdot; A),\quad
\mathscr{F}''(\cdot; A):=\Gamma\hbox{-}\limsup_{k\to+\infty} \mathscr{F}_k(\cdot; A)
\]
for every $A\in\mathcal{A}(\Omega)$.
If $A=\Omega$ we drop the set dependence in the above notation. Moreover we recall that if
$\mathscr{F}=\mathscr{F}'=\mathscr{F}''$ we say that $\mathscr{F}_k$ $\Gamma$-converges to $\mathscr{F}$
(with respect to the metric $d$).
Next we recall the notion of \emph{$\overline\Gamma$-convergence},
useful in particular to deal with the integral representation of
$\Gamma$-limits of families of integral functionals.
We say that $(\mathscr{F}_k)$ $\overline\Gamma$
-\emph{converges} to $\mathscr{F}:X\times \mathcal{A}(\Omega)\to [0,+\infty]$ if $\mathscr{F}$ is the
inner regular envelope of both functionals $\mathscr{F}'$ and $\mathscr{F}''$, i.e.,
\[
\mathscr{F}(u;A)=\sup\{\mathscr{F}'(u;A^\prime):\, A^\prime\in \mathcal{A}(\Omega),\, A^\prime \subset\subset A\}
=\sup \{ \mathscr{F}''(u;A^\prime): A^\prime\in \mathcal{A}(\Omega),\,A^\prime\subset\subset A\},
\]
for every $(u,A)\in X\times \mathcal{A}(\Omega)$.
\subsection{Functional setting of the problem}
All the results we shall prove in what follows will be set in the spaces $BV$ and $SBV$
and in suitable generalizations. For the definitions, the notations and the main properties
of such spaces we refer to the book \cite{ambrosio}. Below we just recall the definition of
$SBV^2(\Omega)$ that we shall often use in the sequel:
\[
SBV^2(\Omega):=
\big\{u\in SBV(\Omega):\nabla u\in L^2(\Omega)\text{ and }\mathcal{H}^{n-1}(J_u)<+\infty\big\}.
\]
Moreover, a function $u:\Omega\to{\mathbb R}$ belongs to $GBV(\Omega)$ (respectively to $GSBV(\Omega)$)
if the truncations $u^M:=-M\vee(u\wedge M)$ belong to $BV_{\textsl{loc}}(\Omega)$ (respectively
to $SBV_{\textsl{loc}}(\Omega)$), for every $M>0$.
For fine properties of $GBV$ and $GSBV$ again we refer to \cite{ambrosio}.
The prototype of the asymptotic result we shall prove in Sections~\ref{s:onedim}, \ref{s:ndim},
and \ref{s:further} concerns the Mumford-Shah functional of image segmentation
\be{\label{e:MS}
{M\!S}(u):=\begin{cases}
\displaystyle{\int_{\Omega}|\nabla u|^2dx+{\mathcal H}^{n-1}(J_u)} & \textrm{if $u\in GSBV(\Omega)$,}\cr
+\infty & \textrm{otherwise in $L^1(\Omega)$}.
\end{cases}
}
Let $\psi:[0,1]\to[0,1]$ be any nondecreasing lower-semicontinuous function such that
$\psi^{-1}(0)=0$ and $\psi(1)=1$. Then the classical approximation by Ambrosio and Tortorelli
(cp. \cite{amb-tort1, amb-tort2}, and also \cite{focardi}) establishes that the two fields
functionals $AT_k^\psi:L^1(\Omega)\times L^1(\Omega)\to[0,+\infty]$
\begin{equation}\label{e:ATk}
AT_k^\psi(u,v):=\begin{cases}
\displaystyle{\int_\Omega{\Big(\psi^2(v)|\nabla u|^2+\frac{(1-v)^2}{4\varepsilon_k}
+\varepsilon_k|\nabla v|^2\Big) dx}} & \textrm{if $(u,v)\in H^1(\Omega){\times}H^1(\Omega)$}\cr
& \text{and $0\leq v\leq1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$,}\\
+\infty &\text{otherwise}
\end{cases}
\end{equation}
$\Gamma$-converge in $L^1(\Omega){\times}L^1(\Omega)$ to
\[
\widetilde{{M\!S}}(u,v):=\begin{cases}
{M\!S}(u) & \textrm{if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$,}\cr
+\infty & \text{otherwise},
\end{cases}
\]
that is equivalent to the Mumford-Shah functional ${M\!S}$ for minimization purposes.
We finally introduce the notation related to slicing.
Fixed $\xi\in \mathbb{S}^{n-1}:=\{\xi\in\mathbb{R}^n:|\xi|=1\}$, let $\Pi^\xi:=\big\{y\in\mathbb{R}^n:y\cdot\xi=0\big\}$,
and for every subset $A\subset {\R}^n$ set
\bes{
& A_y^\xi:=\big\{t\in\mathbb{R}:y+t\xi\in A\big\}\quad
\text{ for $y\in \Pi^\xi$},\\
& A^\xi:=\{y\in\Pi^\xi:A^\xi_y\neq\varnothing\}.
}
For $u:\Omega\to{\mathbb R}$ we define the slices
$u^\xi_y:\Omega_y^\xi\to\mathbb{R}$ by
$
u^\xi_y(t):=u(y+t\xi).
$
Observe that if $u_k,u\in L^1(\Omega)$ and $u_k\to u$ in $L^1(\Omega)$,
then for every $\xi\in \mathbb{S}^{n-1}$ there exists a subsequence $(u_{k_j})$ such that
\[
(u_{k_j})^\xi_y\to u^\xi_y\text{ in }L^1(\Omega^\xi_y)\quad
\textrm{for $\mathcal{H}^{n-1}$-a.e.\ $y\in\Omega^\xi$}.
\]
\medskip
\section{The main results: approximation, compactness and convergence of minimizers}
\label{s:stat}
Given a bounded open set $\Omega\subset{\R}^n$ with Lipschitz boundary and an infinitesimal sequence
$\varepsilon_k>0$, we consider the sequence of functionals $F_k\colon L^1(\Omega){\times}L^1(\Omega)\to [0,+\infty]$
\begin{equation}\label{Fk}
F_k(u,v):=\left\{ \begin{array}{ll}
\ {\displaystyle\int_\Omega{\Big(f_k^2(v)|\nabla u|^2+\frac{(1-v)^2}{4\varepsilon_k}
+\varepsilon_k|\nabla v|^2\Big) dx}}
& \textrm{if $(u,v)\in H^1(\Omega){\times}H^1(\Omega)$}\\
& \text{and $0\leq v\leq1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$},\\
\\
+ \infty & \text{otherwise},
\end{array} \right.
\end{equation}
where
\begin{equation}\label{fk}
f_k(s):=1\wedge\varepsilon_k^{1/2} f(s) \,,\hskip1cm f_k(1)=1\,,
\end{equation}
and
\begin{equation}\label{f0}
\text{$f\in C^0([0,1),[0,+\infty))$ is a nondecreasing function satisfying $f^{-1}(0)=\{0\}$}
\end{equation}
with
\be{\label{f1}
\displaystyle\lim_{s\to1}(1-s)f(s)=\ell, \quad \ell\in(0,+\infty).
}
In particular, the function $[0,1)\mapsto(1-s)f(s)$ can be continuously extended to $s=1$ with
value $\ell$. One can consider $f(s):=\frac{s}{1-s}$ as prototype.
It is also useful to introduce a localized version $F_k(\cdot;A)$ of $F_k$
simply obtained by substituting the domain of integration $\Omega$ with any measurable subset $A$
of $\Omega$ itself. In particular, to be consistent with \eqref{Fk}, for $A=\Omega$ we shall
not indicate the dependence on the domain of integration.
Let now $\Phi\colon L^1(\Omega)\to [0,+\infty]$ be defined by
\begin{equation}\label{Phi}
\Phi(u):=\left\{ \begin{array}{ll}
\ {\displaystyle\int_\Omega h(|\nabla u|)dx+\int_{J_{u}}g(|[u]|)d{\mathcal H}^{n-1}+\ell|D^cu|(\Omega)}
& \textrm{if $u\in GBV(\Omega)$,}\\
+ \infty & \text{otherwise},
\end{array} \right.
\end{equation}
where we recall that $h,g\colon[0,+\infty)\to[0,+\infty)$ are given by
\be{\label{h} h(s):=\left\{ \begin{array}{ll}
s^2 & \textrm{if $s\leq \ell/2$,}\\
\ell s-{\ell}^2/4 & \textrm{if $s\geq \ell/2$},
\end{array} \right.}
and
\be{\label{g}
g(s):=\inf_{(\alpha,\beta)\in\mathcal{U}_s}
\int_0^1|1-\beta|\sqrt{f^2(\beta)|\alpha'|^2+|\beta'|^2}\,dt,
}
where $\mathcal{U}_s:=\mathcal{U}_s(0,1)$ and for all $T>0$
\begin{equation}\label{e:admfnctns}
\mathcal{U}_s(0,T):=\{\alpha,\beta\in H^1\big((0,T)\big):\, 0\le \beta\le 1, \, \alpha(0)=0,\, \alpha(T)=s,\, \beta(0)=\beta(T)=1\}.
\end{equation}
At the points $t$ with $\beta(t)=1$ the integrand in (\ref{g}) reduces to $\ell |\alpha'|(t)$, in agreement with (\ref{f1}).
Our main result is the following.
\begin{theorem}\label{t:gamma-lim}
Under the assumptions above, the functionals $F_k$ $\Gamma$-converge in $L^1(\Omega){\times}L^1(\Omega)$ to the functional $F$ defined by
\begin{equation}\label{F}
F(u,v):=\begin{cases}
\Phi(u) & \textrm{if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$,}\cr
+ \infty & \text{otherwise}.
\end{cases}
\end{equation}
\end{theorem}
\begin{remark}\label{r:gen}
The assumption that $f^{-1}(0)=0$ is not restrictive and changes only the detailed properties of
$g$. Indeed, standing all the other hypotheses, defining $\lambda:=\sup\{s\in[0,1):\,f(s)=0\}\in[0,1)$,
we would get that $g(s)\leq (1-\lambda)^2\wedge\ell s$ (cp.~Proposition~\ref{11} below).
In addition, the function $(1-v)^2$ in \eqref{Fk} can be replaced by any continuous, decreasing
function $d(v)$ with $d(1)=0$. In this case $d^{1/2}(s)$ and $d^{1/2}(\beta)$ appear in formulas
\eqref{f1} and \eqref{g} in place of $1-s$ and $1-\beta$ respectively, and we obtain
$g(s)\leq 2\int_0^1d^{1/2}(t)dt\wedge\ell s$ (see Proposition~\ref{11}).
Finally the definition of $f_k$ in \eqref{fk} can be given in the following more general form
$f_k:=\psi_k\wedge\varepsilon^{1/2}f$. Here the truncation of $f$ is performed with any continuous
nondecreasing function $\psi_k:[0,1]\to[0,1]$ satisfying $\psi_k\geq c>0$, $\lim_k\psi_k(1)=1$,
and converging uniformly in a neighborhood of $1$.
\end{remark}
We next address the issue of equi-coercivity for the $F_k$'s.
\begin{theorem}\label{t:comp}
Under the assumptions above, if $(u_k,v_k)\in H^1(\Omega){\times}H^1(\Omega)$ is such that
$$\sup_k \left(F_k(u_k,v_k)+||u_k||_{L^1(\Omega)}\right)<+\infty,$$
then there exists a subsequence $(u_j,v_j)$ of $(u_k,v_k)$ and a function $u\in GBV\cap L^1(\Omega)$
such that $u_j\to u$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$ and $v_j\to 1$ in $L^1(\Omega)$.
\end{theorem}
We shall prove Theorem~\ref{t:gamma-lim} in Sections~\ref{s:onedim} and \ref{s:ndim},
Theorem~\ref{t:comp} shall be established in Section~\ref{s:ndim}.
In the rest of this section instead we address the issue of convergence of minimum problems.
Minimum problems related to the functional $F_k$ could have no solution due to a lack of coercivity.
Therefore we slightly perturb the $f_k$'s to guarantee the existence of a minimum point for each $F_k$.
This together with Theorems~\ref{t:gamma-lim} and \ref{t:comp} shall in turn imply the convergence of
minima and minimizers as $k\uparrow\infty$.
Let $\eta_k,\varepsilon_k$ be positive infinitesimal sequences such that $\eta_k=o(\varepsilon_k)$ and let
$\zeta\in L^q(\Omega)$,
with $q>1$. Let us consider the sequence of functionals $G_k\colon L^1(\Omega){\times}L^1(\Omega)\to [0,+\infty]$ defined by
$$
G_k(u,v):=\left\{ \begin{array}{ll}
\ {\displaystyle\int_\Omega{\Big(\big(f_k^2(v)+\eta_k\big)|\nabla u|^2+\frac{(1-v)^2}{4\varepsilon_k}
+\varepsilon_k|\nabla v|^2+|u-\zeta|^q\Big)dx}} & \textrm{if $(u,v)\in H^1(\Omega){\times}H^1(\Omega)$}\\
& \text{and $0\leq v\leq1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$},\\
\\
+ \infty & \text{otherwise},
\end{array} \right.
$$
where $f_k$ is as in \eqref{fk}.
Let now $\mathscr{G}\colon L^1(\Omega)\to [0,+\infty]$ be defined by
$$
\mathscr{G}(u):=\left\{ \begin{array}{ll}
\ {\displaystyle\int_\Omega h(|\nabla u|)dx+\int_{J_{u}}g(|[u]|)d{\mathcal H}^{n-1}+\ell|D^cu|(\Omega)+\int_{\Omega}|u-\zeta|^qdx}
& \textrm{if $u\in GBV(\Omega)$,}\\
+ \infty & \text{otherwise},
\end{array} \right.
$$
where $h,g,$ and $\ell$ are as in \eqref{h}, \eqref{g} and \eqref{f1} respectively.
Then the following corollary holds true.
\begin{corollary}\label{c:min}
For every $k$,
let $(u_k,v_k)\in H^1(\Omega){\times} H^1(\Omega)$ be a minimizer of the problem
\begin{equation}\label{intro43}
\min_{(u,v)\in H^1(\Omega){\times} H^1(\Omega)}G_k(u_k,v_k).
\end{equation}
Then $v_k\to 1$ in $L^1(\Omega)$ and a subsequence of $u_k$ converges in $L^q(\Omega)$
to a minimizer $u$ of the problem
$$
\min_{u\in GBV(\Omega)}\mathscr{G}(u).
$$
Moreover the minimum values of (\ref{intro43}) tend to the minimum value of the limit problem.
\end{corollary}
\begin{proof}
We shall only sketch the main steps to establish the conclusion, being the arguments quite standard.
One first proves that in fact the functionals $F_k$ $\Gamma$-converge to $F$ in
$L^q(\Omega){\times}L^1(\Omega)$, where $F_k$ and $F$ are the
functionals in Theorem~\ref{t:gamma-lim}.
Indeed, when $q>1$ the $\Gamma$-limsup inequality works exactly as in the case $q=1$,
whereas the $\Gamma$-liminf inequality is an immediate consequence of the comparison
with the case $q=1$ and of \cite[Proposition 6.3]{dalmaso}. Let us observe that the presence of $\eta_k$
in the functional $F_k$ does not modify the $\Gamma$-convergence result and that the proofs still hold analogously.
As a consequence $G_k$ $\Gamma$-converges to $G$ in $L^1(\Omega){\times}L^1(\Omega)$ for every $q\geq 1$,
where $G(u,v):=\mathscr{G}(u)$ if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$ and $\infty$ otherwise.
Indeed, \cite[Proposition 6.3]{dalmaso} yields that the $\Gamma$-limsup
of $G_k$ in $L^1(\Omega){\times}L^1(\Omega)$ is less than or equal to the one in $L^q(\Omega){\times}L^1(\Omega)$. In addition,
$\int_{\Omega}|\cdot-\zeta|^qdx$ is continuous in $L^q(\Omega){\times}L^1(\Omega)$, so that the conclusion
follows from \cite[Propositions 6.17 and 6.21]{dalmaso}.
The previous result combined with a general result of $\Gamma$-convergence technique
\cite[Corollary 7.20]{dalmaso} concludes the proof of Corollary \ref{c:min} through
the compactness result Theorem \ref{t:comp}.
\end{proof}
\section{Properties of the surface energy density}\label{s:gprop}
In this section we shall establish several properties enjoyed by the surface energy density $g$
defined in \eqref{g}.
To this aim we shall often exploit that, in computing
$g(s)$, $s\geq0$, we may assume that the admissible functions $\alpha$ satisfy $0\leq\alpha\leq s$ by a
truncation argument (whereas $0\leq\beta\leq1$ by definition). Further, given a curve $(\alpha,\beta)\in \mathcal{U}_s(0,T)$,
note that the integral appearing in the definition of $g$ is invariant under reparametrizations of $(\alpha,\beta)$.
\begin{proposition}\label{11}
The function $g$ defined in (\ref{g}) enjoys the following properties:
\begin{enumerate}[label=(\roman*)]
\item\label{e21} $g(0)=0$, and $g$ is subadditive, \textsl{i.e.},
$g(s_1+s_2)\leq g(s_1)+g(s_2)$, for every $s_1,s_2\in{\mathbb R}^+$;
\item\label{e19} $g$ is nondecreasing, $0\leq g(s)\leq 1\wedge\ell s$ for all $s\in{\mathbb R}^+$,
and $g$ is Lipschitz continuous with Lipschitz constant $\ell$;
\item\label{e20}
\be{\label{e23}\lim_{s\uparrow\infty}g(s)=1;}
\item\label{e22}
\be{\label{e24}\lim_{s\downarrow 0}\frac{g(s)}{s}=\ell.}
\end{enumerate}
\end{proposition}
\begin{proof}
Proof of \ref{e21}.
The couple $(\alpha,\beta)=(0,1)$ is admissible for the minimum problem defining $g(0)$, so that $g(0)=0$.
In order to prove that $g$ is subadditive we fix $s_1,s_2\in{\mathbb R}^+$ and we consider the minimum problems for
$g(s_1)$ and $g(s_2)$, respectively. Let $\eta>0$ and let $(\alpha_1,\beta_1),(\alpha_2,\beta_2)$ be admissible
couples respectively for $g(s_1)$ and $g(s_2)$ such that for $i=1,2$
\be{\label{e27}
\int_0^1|1-\beta_i|\sqrt{f^2(\beta_i)|\alpha_i'|^2+|\beta_i'|^2}dt<g(s_i)+\eta.
}
Next define $\alpha:=\alpha_1$ in $[0,1]$, $\alpha:=\alpha_2(\cdot-1)+s_1$ in $[1,2]$, $\beta:=\beta_1$ in
$[0,1]$, and $\beta:=\beta_2(\cdot-1)$ in $[1,2])$.
An immediate computation and the reparametrization property mentioned above
entail the subadditivity of $g$ since $\eta$ is arbitrary.
Proof of \ref{e19}.
In order to prove that $g$ is nondecreasing we fix $s_1,s_2$ with $s_1< s_2$ and $\eta>0$, and we consider
$(\alpha,\beta)$ satisfying a condition analogous to
\eqref{e27} for $g(s_2)$.
Then $(\frac{s_1}{s_2}\alpha,\beta)$ is admissible for $g(s_1)$, thus we infer
$$g(s_1)\leq \int_0^1|1-\beta|\sqrt{\Big(\frac{s_1}{s_2}\Big)^2f^2(\beta)|\alpha'|^2+|\beta'|^2}dt< g(s_2)+\eta,$$
since $s_1/{s_2}< 1$. As $\eta\to 0$ we find $g(s_1)\leq g(s_2)$.
Next we prove that $g(s)\leq 1\wedge \ell s$.
Indeed, inequality $g\leq 1$ straightforwardly comes from the fact that for every $s\geq 0$ the following couple
$(\alpha,\beta)$ is admissible for $g(s)$:
$\alpha:=0$ in $(0,1/3)$, $\alpha:=s$ in $(2/3,1)$, and linearly linked in $(1/3,2/3)$, and $\beta:=0$ in $(1/3,2/3)$ and linearly
linked to $1$ in $(0,1/3)$ and in $(2/3,1)$.
Moreover, $g(s)\leq \ell s$ for every $s\geq0$ since the couple $(st,1)$ is admissible for $g(s)$.
The Lipschitz continuity of $g$ is an obvious consequence of the facts
that $g$ is nondecreasing, subadditive and $g(s)\leq \ell s$ for $s\geq0$.
Proof of \ref{e20}.
Let $s_k$, $k\in{\mathbb N}$, be a diverging sequence and let
$(\alpha_k,\beta_k)$ be an admissible couple for $g(s_k)$ such that
\be{\label{e25}\int_0^1|1-\beta_k|\sqrt{f^2(\beta_k)|\alpha'_k|^2+|\beta'_k|^2}dt<g(s_k)+\frac{1}{k}.}
If $\inf_{(0,1)}\beta_k\geq\delta$ for some $\delta>0$ and for every $k$, then there exists a constant $c(\delta)>0$ such that
$f(\beta_k)(1-\beta_k)>c(\delta)$, since $f(s)(1-s)\to 0$ if and only if $s\to0$.
Therefore by \eqref{e25} one finds
$$c(\delta)s_k\leq g(s_k)+\frac{1}{k},$$
so that $g(s_k)\to+\infty$ as $k\to+\infty$ and this contradicts the fact that $g\leq 1$. Therefore there exists a sequence $x_k\in(0,1)$
such that $\beta_k(x_k)\to0$ up to subsequences. Since we have already shown that $g\leq 1$, we conclude the proof of \eqref{e23} noticing
that \eqref{e25} yields \be{\label{e26}(1-\beta_k(x_k))^2 \leq\int_0^{x_k}|1-\beta_k||\beta'_k|dt+\int_{x_k}^1|1-\beta_k||\beta'_k|dt\leq g(s_k)+\frac{1}{k}.}
Proof of \ref{e22}.
Let $s_k$, $k\in{\mathbb N}$, be an infinitesimal sequence and let $(\alpha_k,\beta_k)$ be an admissible
couple for $g(s_k)$ satisfying \eqref{e25} with $s_k/k$ in place of $1/k$. If there exists $\delta>0$,
a not relabeled subsequence of $k$, and a sequence $x_k\in[0,1]$ such that $\beta_k(x_k)<1-\delta$,
then the same computation as in \eqref{e26} leads to
\[\delta^2\leq g(s_k)+\frac{s_k}{k}.
\]
As $k\to+\infty$ this contradicts the fact that $g(s)\le \ell s$. Therefore, $\beta_k$ converges uniformly to $1$ and fixing $\delta>0$
$$(\ell-\delta)s_k\leq\int_0^1(1-\beta_k)f(\beta_k)|\alpha'_k|dt\leq g(s_k)+\frac{s_k}{k}$$
holds for $k$ large by \eqref{f1}. Formula \eqref{e24} immediately follows dividing both sides of the last inequality by $s_k$, taking
first $k\to+\infty$ and then $\delta\to0$, and using the fact that $g(s)\leq \ell\,s$, for $s\geq0$.
\end{proof}
\begin{remark}\label{r:g}
We can actually show that $g$ does not coincide with the function $1\wedge\ell\,s$ at least in the model case
$f(s)=\frac{\ell s}{1-s}$ by slightly refining the construction used in \ref{e19} above.
With fixed $s>0$, let $\lambda\in[0,1]$ and set $\alpha:=0$ on $[0,1/3]$, $\alpha:=s$ on $[1/3,2/3]$, and
the linear interpolation of such values on $[1/3,2/3]$; moreover, set $\beta_\lambda:=\lambda$ on $[1/3,2/3]$
and the linear interpolation of the values $1$ and $\lambda$ on each interval $[0,1/3]$ and $[2/3,1]$ in order
to match the boundary conditions. Straightforward calculations lead to
\[
g(s)\leq(1-\lambda)^2+(1-\lambda)f(\lambda)\,s.
\]
Thus, minimizing over $\lambda\in[0,1]$ yields in turn
\[
g(s)\leq \ell s-\frac{(\ell s)^2}{4}<1\wedge\ell s\quad\text{ for all $s\in(0,2/\ell)$.}
\]
\end{remark}
In what follows it will be convenient to provide an alternative representation of $g$ by means of
a cell formula more closely related to the one-dimensional version of the energies $F_k$'s.
To this aim we introduce the function $\hat{g}\colon [0,+\infty)\to[0,+\infty)$ defined by
\begin{equation}\label{hatg}
\hat{g}(s):=\lim_{T\uparrow\infty}\inf_{(\alpha,\beta)\in\mathcal{U}_s(0,T)}
\int_0^T \left(f^2(\beta)|\alpha'|^2+\frac{|1-\beta|^2}{4}+|\beta'|^2\right)dt,
\end{equation}
the class $\mathcal{U}_s(0,T)$ has been introduced in \eqref{e:admfnctns}.
We note that $\hat{g}$ is well-defined as the minimum problems appearing in its definition are decreasing
with respect to $T$.
\begin{proposition}\label{p2}
For all $s\in [0,+\infty)$ it holds $g(s)=\hat{g}(s)$.
\end{proposition}
\begin{proof}
Let $\alpha,\beta \in H^1\big((0,T)\big)$, $T>0$, be admissible functions in the definition of $\hat g(s)$.
By Cauchy inequality we obtain
\begin{equation*}
\sqrt{f^2(\beta)|\alpha'|^2 + |\beta'|^2}\, |1-\beta|\le f^2(\beta)|\alpha'|^2 + |\beta'|^2+\frac{(1-\beta)^2}{4}
\end{equation*}
and integrating
\begin{equation*}
\int_0^T |1-\beta| \, \sqrt{f^2(\beta)|\alpha'|^2 + |\beta'|^2}\, dt\le \int_0^T
\left(f^2(\beta)|\alpha'|^2+\frac{|1-\beta|^2}{4}+|\beta'|^2\right)dt\,.
\end{equation*}
The first integral is one-homogeneous in the derivatives, therefore we can reparametrize from $(0,T)$
to $(0,1)$. Taking the infimum over all such $\alpha$, $\beta$, and $T$ we obtain $g(s)\le \hat g(s)$.
To prove the converse inequality, we first show that
$\alpha$ and $\beta$ in the infimum problem defining $g$ can be taken in $W^{1,\infty}\big((0,1)\big)$. Let $\eta>0$ small and let $\alpha,\beta\in H^1\big((0,1)\big)$
be competitors for $g(s)$ such that
\be{\label{e:quasimin}\int_{0}^{1} |1-\beta| \, \sqrt{f^2(\beta)|\alpha'|^2 + |\beta'|^2}\, dt< g(s)+\eta.}
By density we find two sequences $\alpha_j,\beta_j\in W^{1,\infty}\big((0,1)\big)$ (actually in $C^\infty([0,1])$) such that
$\alpha_j(0)=0$, $\alpha_j(1)=s$, $\beta_j(0)=\beta_j(1)=1$, $0\leq\beta_j\leq 1$, and converging respectively
to $\alpha$ and $\beta$ in $H^1\big((0,1)\big)$. Since the function $(1-s)f(s)$ is uniformly continuous and $\beta_j\to\beta$ also uniformly, we deduce that
$$\int_{0}^{1} |1-\beta_j| \, \sqrt{f^2(\beta_j)|\alpha_j'|^2 + |\beta_j'|^2}\, dt< g(s)+\eta$$
for $j$ large, and this concludes the proof of the claim.
Let us prove now that $\hat{g}\leq g$.
We fix a small parameter $\eta>0$ and
consider competitors $\alpha,\beta\in W^{1,\infty}\big((0,1)\big)$
for $g(s)$ satisfying \eqref{e:quasimin}.
We define, for $t\in [0,1]$,
\begin{equation*}
\beta^\eta(t):=\beta(t)\wedge (1-\eta)\quad \text{and}\quad \psi_\eta(t):=\int_{0}^t \frac{2}{1-\beta^\eta}
\sqrt{\eta+f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2} dt' \,.
\end{equation*}
The function $\psi_\eta:[0,1]\to[0,M_\eta:=\psi_\eta(1)]$ is bilipschitz and in particular invertible.
We define $\bar\alpha^\eta, \bar\beta^\eta\in W^{1,\infty}\big((0,M_\eta)\big)$ by
\begin{equation*}
\bar\alpha^\eta := \alpha \circ \psi_\eta^{-1} \text{ and }
\bar\beta^\eta := \beta^\eta \circ \psi_\eta^{-1}\,.
\end{equation*}
We compute, using the definition and the change of variables $x=\psi_\eta(t)$,
\begin{alignat*}1
\int_0^{M_\eta} \frac{(1-\bar\beta^\eta)^2}4 dx &=
\int_0^{M_\eta} \frac{(1-\beta^\eta(\psi_\eta^{-1}(x)))^2}4 dx =
\int_{0}^1 \frac{(1-\beta^\eta(t))^2}4 \psi_\eta'(t) dt \\
&=\int_{0}^1 \frac{1-\beta^\eta}2 \sqrt{\eta+
f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2} dt \\
&\le \sqrt\eta+ \int_{0}^1 \frac{1-\beta^\eta}2 \sqrt{f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2} dt\,,
\end{alignat*}
where we inserted $\psi_\eta'$ from the definition of $\psi_\eta$ and used
$\sqrt{\eta+A}\le \sqrt \eta + \sqrt A$. Analogously,
\begin{alignat*}1
\int_0^{M_\eta} \left( f^2(\bar\beta^\eta)|(\bar\alpha^\eta)'|^2+|(\bar\beta^\eta)'|^2\right) dx &=
\int_{0}^1 \left( f^2(\beta^\eta)|\alpha'|^2+|(\beta^\eta)'|^2\right) \frac{1}{\psi'_\eta} dt \\
&= \int_{0}^1 \left( f^2(\beta^\eta)|\alpha'|^2+|(\beta^\eta)'|^2\right)
\frac{1-\beta^\eta}{2 \sqrt{\eta+
f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2}}dt \\
&\le \int_{0}^1
\frac{1-\beta^\eta}{2} \sqrt{f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2}dt \,.
\end{alignat*}
We extend $\bar\alpha^\eta$ and $\bar\beta^\eta$ to $(-1, M_\eta+1)$ setting
$\bar\alpha^\eta:=0$ in $(-1,0)$, $\bar\alpha^\eta:=s$ in $(M_\eta,M_\eta+1)$, and $\bar\beta^\eta$ the linear interpolation between $1-\eta$ and 1 in each of the
two intervals, so that they obey the required boundary conditions for $\hat g$ in the larger interval.
Collecting terms, we obtain
\begin{align}
\hat g(s)&\le \int_{-1}^{M_\eta+1} \left(\frac{(1-\bar\beta^\eta)^2}4
+ f^2(\bar\beta^\eta)|\bar\alpha'_\eta|^2+|\bar(\beta^\eta)'|^2\right) dx\nonumber\\
&\le \sqrt \eta + 3\eta^2 + \int_{0}^1
(1-\beta^\eta) \sqrt{f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2}dt \,,\label{al:p1}
\end{align}
where the $3\eta^2$ term comes from an explicit computation on the two boundary intervals.
It remains to replace $\beta^\eta$ by $\beta$ in the last integral.
We observe that $(\beta^\eta)'=0$ almost everywhere on the set where $\beta\ne \beta^\eta$ (which coincides with the set $\{\beta>1-\eta\}$).
Therefore
\begin{alignat*}1
\int_{\{\beta\ne \beta^\eta\}}
(1-\beta^\eta) \sqrt{f^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2}dt &=
\int_{\{\beta\ne \beta^\eta\}}
(1-\beta^\eta) f(\beta^\eta)\, |\alpha'| dt \\
&\le \int_{\{\beta\ne \beta^\eta\}}
(1-\beta) f(\beta)\, |\alpha'| dt
+ \omega(\eta) \int_0^1 |\alpha'| dt
\end{alignat*}
where $\omega(\eta)$ is the continuity modulus of $(1-s)f(s)$ near $s=1$, and therefore
\begin{alignat*}1
\hat g(s)
&\le \sqrt \eta + 3\eta^2
+\omega(\eta)\int_0^1|\alpha'| dt+
\int_{0}^1
(1-\beta) \sqrt{f^2(\beta)\, |\alpha'|^2 + |\beta'|^2}dt \,.
\end{alignat*}
Since the last integral is less than $g(s)+\eta$ and $\eta$ can be made arbitrarily small, this concludes the proof.
\end{proof}
For the proof of the lower bound
we also need to introduce the auxiliary functions $g^{(\eta)}\colon [0,+\infty)\to[0,+\infty)$,
for $\eta>0$, defined by
\be{\label{geta}
g^{(\eta)}(s):=\inf_{(\alpha,\beta)\in\mathcal{U}^{(\eta)}_s}
\int_0^1|1-\beta|\sqrt{f^2(\beta)|\alpha'|^2+|\beta'|^2}\,dt,
}
where
$$
\mathcal{U}^{(\eta)}_s:=\{\alpha,\beta\in H^1\big((0,1)\big):\, \alpha(0)=0,\, \alpha(1)=s,\, \beta(0)=\beta(1)=1-\eta\}.
$$
\begin{proposition}\label{p1}
For all $s\in [0,+\infty)$ it holds
\[
|g(s)-g^{(\eta)}(s)|\leq\eta^2.
\]
\end{proposition}
\begin{proof}
We consider the minimum problems for $g$ and $g^{(\eta)}$ respectively in the intervals $(-1,2)$ and $(0,1)$.
Let $(\alpha_{\eta},\beta_{\eta})$ be an admissible couple for $g^{(\eta)}(s)$ and let $\alpha:=0$ in $(-1,0)$,
$\alpha:=\alpha_{\eta}$ in $(0,1)$, and $\alpha:=s$ in $(1,2)$;
we also set $\beta:=\beta_{\eta}$ in $(0,1)$ and linearly linked to $1$ in $(-1,0)$ and in $(1,2)$.
Then an easy computation shows that
$$g(s)\leq \int_0^1|1-\beta_{\eta}|\sqrt{f^2(\beta_{\eta})|\alpha'_{\eta}|^2+|\beta'_{\eta}|^2}dt+\eta^2.$$
By taking the infimum on $(\alpha_{\eta},\beta_{\eta})$ we infer that
$$g(s)\leq g^{(\eta)}(s)+\eta^2.$$
Reversing the roles of $g$ and $g^{(\eta)}$ we conclude.
\end{proof}
Finally, we study the dependence of $g$ on the function $f$ in detail. The results in the next proposition
provide a first insight on the class of functions $g$ that arise as surface energy densities in our
analysis. Moreover, they will be instrumental to get in the limit different energies by slightly changing
the functionals $F_k$'s in \eqref{Fk} (cp. Theorems~\ref{t:Dugdale}, \ref{t:sublin}, and \ref{t:MS} below).
\begin{proposition}\label{p:gl}
Let $(\f{j})$ be a sequence of functions satisfying \eqref{f0} and \eqref{f1}. Denote by $\ell_j$, $g_j$
the value of the limit in \eqref{f1} and the function in \eqref{g} corresponding to $\f{j}$, respectively.
Then,
\begin{itemize}
\item[(i)] if $\ell_j=\ell$ for all $j$, $\f{j}\geq \f{j+1}$, and $\f{j}(s)\downarrow 0$ for all
$s\in[0,1)$, then $g_j\geq g_{j+1}$ and $g_j(s)\downarrow 0$ for all $s\in[0,+\infty)$;
\item[(ii)] if $\ell_j=\ell$ for all $j$, $\f{j}\leq \f{j+1}$, and $\f{j}(s)\uparrow \infty$ for all
$s\in(0,1)$, then $g_j\leq g_{j+1}$ and $g_j(s)\uparrow 1\wedge\ell s$ for all $s\in[0,+\infty)$;
\item[(iii)] if $\ell_j\uparrow\infty$, $\f{j}\leq \f{j+1}$, and $\f{j}(s)\uparrow \infty$ for all
$s\in(0,1)$, then $g_j\leq g_{j+1}$ and $g_j(s)\to \chi_{(0,+\infty)}(s)$ for all $s\in[0,+\infty)$.
\end{itemize}
\end{proposition}
\begin{proof}
To prove item (i) we note that the monotonicity of the sequence $(\f{j})$ and the
pointwise convergence to a continuous function on $[0,1)$ yield that the sequence
$(\f{j})$ actually converges uniformly on compact subsets of $[0,1)$ to $0$.
Therefore, for all $\delta\in(0,1)$ we have for some $j_\delta$
\[
\max_{[0,1-\delta]}\f{j}\leq\delta\qquad\text{for all $j\geq j_\delta$.}
\]
Then, consider $\alpha_j,\beta_j$ defined as follows: $\alpha_j(t):=3s(t-1/3)$ on
$[1/3,2/3]$, $\alpha_j:=0$ on $[0,1/3]$, and $\alpha_j:=s$ on $[2/3,1]$; $\beta_j:=1-\delta$
on $[1/3,2/3]$ and a linear interpolation between the values $1$ and $1-\delta$ on each interval
$[0,1/3]$ and $[2/3,1]$. Straightforward calculations give
\[
g_j(s)\leq\delta^2\,s+\delta^2 \qquad\text{for all $j\geq j_\delta$,}
\]
from which the conclusion follows by passing to
the limit first in $j\uparrow\infty$ and finally letting $\delta\downarrow 0$.
We now turn to item (ii). We first note that $(g_j)$ is nondecreasing and that
\begin{equation}\label{e:geasy}
\lim_jg_j(s)\leq \ell s\wedge 1
\end{equation}
in view of item \ref{e19} in Proposition~\ref{11}.
Next we show the following: for all $\delta>0$
\begin{equation}\label{e:dmlim}
\lim_j\min_{t\in[\delta,1]}(1-t)\f{j}(t)=\ell.
\end{equation}
Let $s_j\in\textrm{argmin}_{[\delta,1]}(1-t)\f{j}(t)$, and denote by $j_k$
a subsequence such that
\[
\lim_k\min_{t\in[\delta,1]}(1-t)\f{j_k}(t)=\liminf_j\min_{t\in[\delta,1]}(1-t)\f{j}(t).
\]
Either $\limsup_ks_{j_k}<1$ or $\limsup_ks_{j_k}=1$. We exclude the former possibility:
suppose that, up to further subsequences not relabeled, $\lim_ks_{j_k}=s_\infty\in[\delta,1)$,
then for all $i\in{\mathbb N}$
\[
\liminf_k(1-s_{j_k})\f{j_k}(s_{j_k})=(1-s_\infty)\liminf_k\f{j_k}(s_{j_k})
\geq(1-s_\infty)f^{(i)}(s_\infty),
\]
that gives a contradiction by letting $i\uparrow\infty$ since by minimality of $s_j$
\be{\label{e:boundell}
(1-s_j)\f{j}(s_j)\leq\ell\qquad\text{for all $j$}.
}
Therefore, $\limsup_ks_{j_k}=1$, and thus we get
\[
\liminf_j(1-s_j)\f{j}(s_j)\geq \liminf_k(1-s_{j_k})\f{1}(s_{j_k})=\ell.
\]
Formula \eqref{e:dmlim} follows straightforwardly by this and \eqref{e:boundell}.
If $s=0$, clearly we conclude as $g_j(0)=0$ for all $j$.
Let then $s\in(0,+\infty)$ and $\alpha_j$, $\beta_j\in H^1\big((0,1)\big)$ be such that
$\alpha_j(0)=0$, $\alpha_j(1)=s$, $\beta_j(0)=\beta_j(1)=1$ and
\[
g_j(s)+\frac1j\geq
\int_0^1|1-\beta_j|\sqrt{(\f{j})^2(\beta_j)|\alpha'_j|^2+|\beta'_j|^2}dt.
\]
There are now two possibilities: either there exists $\delta>0$ and a subsequence $j_k$ such
that $\inf_{[0,1]}\beta_{j_k}\geq\delta$, or $\inf_{[0,1]}\beta_j\to0$. In the former case the
subsequence satisfies
\begin{equation*}
g_{j_k}(s)+\frac1{j_k}\geq\big(\min_{t\in[\delta,1]}(1-t)\f{j_k}(t)\big)\,s.
\end{equation*}
Taking the $\limsup_{k}$ and using (\ref{e:dmlim}) we
obtain
\begin{equation}\label{eqcaso1}
\limsup_j g_j(s) \ge \ell s\,.
\end{equation}
In the other case for every $\delta>0$ definitively it holds
\begin{equation}\label{e:altb}
g_{j}(s)+\frac1{j}\geq\int_0^1(1-\beta_{j})|\beta_{j}^\prime|dt\geq (1-\delta)^2.
\end{equation}
Taking again the $\limsup$ we obtain
\begin{equation}\label{eqcaso2}
\limsup_j g_j(s) \ge (1-\delta)^2\,.
\end{equation}
Since $\delta$ was arbitrary,
from (\ref{eqcaso1}) and (\ref{eqcaso2}) we obtain $\limsup_j g_j(s)\ge 1\wedge \ell s$ and,
recalling (\ref{e:geasy}), conclude the proof of (ii).
Let us now prove item (iii).
First we observe that $g_j(s)\leq 1$ for all $j$.
To prove the lower bound, we notice that arguing similarly as in the proof of \eqref{e:dmlim}
one obtains
\begin{equation}\label{e:dmlim2}
\lim_j\min_{t\in[\delta,1]}(1-t)\f{j}(t)=\infty \text{ for all $\delta>0$}.
\end{equation}
For any $s\in(0,+\infty)$ we choose $(\alpha_j, \beta_j)\in \mathcal{U}_s$ such that
\[
g_j(s)+\frac1j\geq\int_0^1|1-\beta_j|\sqrt{(\f{j})^2(\beta_j)|\alpha'_j|^2+|\beta'_j|^2}dt.
\]
If there is $\delta>0$ such that $\inf \beta_j\ge \delta$ for infinitely many $j$ then for the same
indices
\begin{equation*}
g_j(s)+\frac1j\ge \min_{t\in[\delta,1]} (1-t) f^{(j)}(t) s \,,
\end{equation*}
which in view of (\ref{e:dmlim2}) and the bound $g_j(s)\le 1$ is impossible.
Therefore $\inf_{[0,1]}\beta_j\to0$, which in view of (\ref{e:altb}) proves the assertion.
\end{proof}
\begin{remark}
The monotonicity assumption $\f{j}\leq \f{j+1}$ in items (ii) and (iii) above leads to simple proofs but
it is actually not needed. The same convergence results for $(g_j)$ would follow by using the uniform
convergence on compact subsets of $[0,1)$ of $(\f{j})$. The latter property is a consequence of the fact
that each $\f{j}$ is nondecreasing and that $f\in C^0([0,1))$.
\end{remark}
\section{Proof in the one-dimensional case}\label{s:onedim}
Let us study first the one-dimensional case $n=1$. As usual, we will prove a $\Gamma$-liminf inequality and a $\Gamma$-limsup inequality.
The following proposition gives the lower estimate.
\begin{proposition}[Lower bound]\label{p:liminfunidim}
For every $(u,v)\in L^1(\Omega){\times}L^1(\Omega)$ it holds
$$F(u,v)\leq F'(u,v).$$
\end{proposition}
\begin{proof}
The conclusion is equivalent to the following fact:
let $(u_k,v_k)$ be a sequence such that
\begin{equation}\label{ukvk}
(u_k,v_k)\to (u,v) \text{ in } L^1(\Omega){\times} L^1(\Omega),
\end{equation}
\begin{equation}\label{bounded}
\sup_k F_{k}(u_k,v_k)<+\infty,
\end{equation}
then $u\in BV(\Omega)$, $v=1$ $\mathcal{L}^1$-a.e.\ in $\Omega$,
and
\be{\label{1}\Phi(u)\leq
\liminf_{k\to \infty}F_k(u_k,v_k).}
Since the left-hand side of \eqref{1} is $\sigma$-additive and the right-hand side is $\sigma$-superadditive
with respect to $\Omega$, it is enough to prove the result when $\Omega$ is an interval.
For the sake of convenience in what follows we assume $\Omega=(0,1)$.
By \eqref{bounded} one deduces that $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$. Up to subsequences one can assume that the lower limit in \eqref{1} is in fact a limit
and that the convergences in \eqref{ukvk} are also $\mathcal{L}^1$-a.e.\ in $\Omega$.
For the first part of the proof we will use a discretization argument, following the lines of \cite{AlicBrShah}.
We fix $\delta\in(0,1)$ and for any
$N\in{\mathbb N}$ divide $\Omega$ into $N$ intervals
$$I^j_N:=\Big(\frac{j-1}{N},\frac{j}{N}\Big),\qquad j=1,\dots,N.$$
Up to subsequences we can assume that $\displaystyle\lim_{k\to+\infty} \inf_{I^j_N} v_k$ exists for every $j=1,\dots,N$.
We define
$$J_N:=\Big\{j\in\{1,\dots,N\}:\lim_{k\to+\infty} \inf_{I^j_N} v_k\le 1-\delta\Big\}.$$
Fixed $j\in J_N$, we denote by $x_k$ and $y$ two points in $I_N^j$ such that $v_k(x_k)<1-\delta/2$ and $v_k(y)\to 1$. Then by Cauchy's inequality we
deduce for $k$ large (assuming for instance $x_k\leq y$)
\be{\label{2}\int_{x_k}^{y}\Big(\frac{(1-v_k)^2}{4\varepsilon_k}+\varepsilon_k| v'_k|^2\Big) dx\geq \frac{1}{2}((1-v_k(x_k))^2-(1-v_k(y))^2)\geq \frac{\delta^2}{16}.}
The previous computation entails
$$\sup_N{\mathcal H}^0(J_N)<+\infty,$$
so that up to subsequences we can assume $J_N=\{j^N_1,\dots,j^N_L\}$, with $L$ independent on $N$,
and that all sequences $j^N_i/N$ converge. We denote by $S$ the set
of limits of these sequences,
$$
S=\{t_1,\dots,t_{L'}\}=\bigl\{\lim_{N\to+\infty}\frac{j^N_i}{N}\,,\hskip2mm i=1,\dots,L\bigr\} \subset\Omega\,.
$$
We claim now that there exists a modulus of continuity $\omega$, i.e., $\omega(\delta)\to0$ as $\delta\to0$,
depending only on $f$, such that for all $\eta$ sufficiently small and $k$ sufficiently large
(depending on $\eta$) one has
\be{\label{3}
(1-\omega(\delta))\int_{\Omega\setminus S_{\eta}}h(|u'_k|)dx\leq F_k(u_k,v_k,\Omega\setminus S_{\eta}),
}
where $S_{\eta}:=\bigcup_{i=1}^{L'}(t_i-\eta,t_i+\eta)$.
It suffices to prove (\ref{3}) in the case that $\eta$ is so small that
the intervals $(t_i-\eta,t_i+\eta)$ are pairwise disjoint.
In order to prove \eqref{3}, we observe that by definition of $f_k$ in \eqref{fk} and by Cauchy's inequality we obtain
\ba{\label{4}
F_k(u_k,v_k;\Omega\setminus S_{\eta})&\geq& \int_{\Omega\setminus S_{\eta}}\Big(f_k^2(v_k)|u'_k|^2+\frac{(1-v_k)^2}{4\varepsilon_k}\Big) dx\nonumber\\
&\geq& \int_{\Omega\setminus S_{\eta}}\Big(|u'_k|^2\wedge
\bigl(\varepsilon_kf^2(v_k)|u'_k|^2+\frac{(1-v_k)^2}{4\varepsilon_k}\bigr)\Big)dx\nonumber\\
&\geq& \int_{\Omega\setminus S_{\eta}}|u'_k|^2\wedge\bigl((1-v_k)f(v_k)|u'_k|\bigr)dx.
}
Let us note that $v_k> 1-\delta$ in $\Omega\setminus S_{\eta}$ for $k$ large. By \eqref{f1} there exists a modulus of continuity $\omega$ such that
\be{\label{5}|(1-s)f(s)-\ell|\leq \ell\omega(\delta),\qquad \textrm{for $s\geq1-\delta$}.}
Therefore by \eqref{4} and \eqref{5} we obtain
\be{\label{6}
F_k(u_k,v_k;\Omega\setminus S_{\eta})\geq(1-\omega(\delta))\int_{\Omega\setminus S_{\eta}}|u'_k|^2\wedge\ell|u'_k|dx
\geq(1-\omega(\delta))\int_{\Omega\setminus S_{\eta}}h(|u'_k|)dx.
}
The last inequality holds true as $h$ is the convex envelope of $t^2\wedge\ell t$. Formula (\ref{6}) proves the claim in \eqref{3}.
Notice that the boundedness assumption in \eqref{bounded} and formula \eqref{3} imply that
$$\sup_{k}\int_{\Omega\setminus S_{\eta}}|u'_k|dx<+\infty.$$
Therefore $u\in BV(\Omega\setminus S_{\eta})$, and actually the finiteness of $S$ ensures that $u\in BV(\Omega)$. In addition, the
$L^1$-lower semicontinuity of the functional $\Phi$ defined in \eqref{Phi} yields
\be{\label{e16}(1-\omega(\delta))\Phi(u;\Omega\setminus S_{\eta})\leq\liminf_k F_k(u_k,v_k;\Omega\setminus S_{\eta}).}
We now estimate the energy contribution on $S_{\eta}$. To this aim it is not restrictive to assume that $S\subseteq J_u$.
\medskip
Let us fix $i\in\{1,\dots,L'\}$ and consider $I^i_\eta:=(t_i-\eta,t_i+\eta)$. We claim that
\be{\label{8}(1-\omega(\delta))g(\esssup_{I^i_{\eta}} u-\essinf_{I^i_{\eta}} u)\leq\liminf_{k\to+\infty} F_k(u_k,v_k;I^i_{\eta})+O(\eta).}
Let us introduce a small parameter $\mu>0$ and $x_1,x_2\in I^i_{\eta}$ such that
\ba{ &\displaystyle v_k(x_1)\to 1, \qquad v_k(x_2)\to 1, \nonumber\\
\label{13} &\displaystyle u_k(x_1)\to u(x_1), \qquad u_k(x_2)\to u(x_2), \\
\label{7} &\displaystyle u(x_1)>\esssup_{I^i_{\eta}} u-\mu,\qquad u(x_2)<\essinf_{I^i_{\eta}} u+\mu.}
Assuming without loss of generality that $x_1< x_2$, we define $I:=(x_1,x_2)$.
There are just finitely many connected components of the set
$$\{x\in I: v_k(x)< 1-\eta\}$$
where $v_k$ achieves the value $1-\delta$, as a computation analogous to \eqref{2} easily shows (recall that $\eta\ll\delta$). Precisely one finds up to subsequences
that the number $N$ of these components is $$N\leq\frac{c}{\delta^2-\eta^2},$$
for some constant $c>0$ independent of $N$.
Let us now estimate the functional $F_k$ over each component $C_k^j$ of this type, $j=1,\dots,N$.
Since $v_k<1-\eta$ in $C_k^j$ one finds for $k$ large that $f_k(v_k)=\varepsilon^{1/2}_kf(v_k)$, so that for $j=1,\dots,N$ it follows
\begin{multline}\label{e15}
F_k(u_k,v_k;C_k^j)\geq \int_{C_k^j}{\Big(\varepsilon_kf^2(v_k)|u'_k|^2+\frac{(1-v_k)^2}{4\varepsilon_k}+\varepsilon_k|v'_k|^2\Big) dx}\\
\geq g^{(\eta)}\left(\left|\int_{C_k^j}u'_k dx\right|\right)\geq g\left(\left|\int_{C_k^j}u'_k dx\right|\right)-\eta^2,
\end{multline}
by Cauchy's inequality and Proposition \ref{p1}.
Outside the selected components $C_k^j$, $j=1,\dots,N$, one has $v_k\geq 1-\delta$, so that estimate \eqref{6} also holds with $I\setminus \bigcup_{j=1}^N C_k^j$
replacing $\Omega\setminus S_{\eta}$. Therefore
\ba{\label{9}
F_k\left(u_k,v_k;I\setminus \bigcup_{j=1}^N C_k^j\right)&\geq&(1-\omega(\delta))\int_{I\setminus \bigcup_{j=1}^N C_k^j}h(|u'_k|)dx\nonumber\\
&\geq& (1-\omega(\delta))\ell\int_{I\setminus \bigcup_{j=1}^N C_k^j}|u'_k|dx-(1-\omega(\delta))\frac{\ell^2}{4}{\mathcal L}^1(I\setminus \bigcup_{j=1}^N C_k^j)\nonumber\\
&\geq& (1-\omega(\delta))g\Big(\Big|\int_{I\setminus \bigcup_{j=1}^N C_k^j}u'_k dx\Big|\Big)-\frac{\ell^2}{2}\eta,}
where we have used the definition of $h$ and Proposition \ref{11} \ref{e19}.
By \eqref{e15}, \eqref{9}, and the subadditivity of $g$ one finds
$$F_k(u_k,v_k;I)+\frac{\ell^2}{2}\eta+\frac{c\,\eta^2}{\delta^2-\eta^2}\geq (1-\omega(\delta))g\left(\left|\int_I u'_k dx\right|\right)
=(1-\omega(\delta))g(|u_k(x_1)-u_k(x_2)|).$$
By property \eqref{13} and by the continuity of $g$, as $k\to+\infty$ one deduces
$$\liminf_{k\to+\infty} F_k(u_k,v_k;I^i_{\eta})+\frac{\ell^2}{2}\eta+\frac{c\eta^2}{\delta^2-\eta^2}\geq (1-\omega(\delta))g(|u(x_1)-u(x_2)|).$$
Finally property \eqref{7} concludes the proof of \eqref{8} as $\mu\to0$.
The thesis follows by summing \eqref{e16} and \eqref{8} for $i=1,\dots,L$ and taking first $\eta\to0$ and finally $\delta\to0$.
\end{proof}
\begin{proposition}[Upper bound]\label{p:limsupunidim}
For all $u\in BV(\Omega)$ there exists $(u_k,v_k)\to(u,1)$ in $L^1(\Omega){\times}L^1(\Omega)$ such that
$$\limsup_{k\to+\infty}F_k(u_k,v_k)\leq \Phi(u).$$
\end{proposition}
\begin{proof}
Let us consider first the case when $u\in SBV^2(\Omega)$.
By a localization argument it is not restrictive to assume that $J_{u}=\{x_0\}$ and to take $x_0=0$.
We also assume for a while that $u$ is constant in a neighborhood on both sides of $0$.
With fixed $\eta>0$, we consider $T_\eta>0$ and
$\alpha_\eta,\beta_\eta\in H^1\big((0,T_{\eta})\big)$ such that
$\alpha_\eta(0)=u^-(0),$ $\alpha_\eta(T_{\eta})=u^+(0)$, $0\leq\beta_\eta\leq 1$, $\beta_\eta(0)=\beta_\eta(T_\eta)=1$, and
\be{\label{e17}g\big(|[u](0)|\big)+\eta>\int_0^{T_{\eta}} \left(f^2(\beta_\eta)|\alpha_\eta'|^2+\frac{|1-\beta_\eta|^2}{4}+|\beta_\eta'|^2\right)dt.}
This choice is possible in view of Proposition \ref{p2}, up to a translation of the variable $\alpha_\eta$.
Let us define $A_k:=(-\frac{\varepsilon_kT_\eta}{2},\frac{\varepsilon_kT_\eta}{2})$ and
\ba
{u_k(x):=\left\{ \begin{array}{ll}
\ {\displaystyle \alpha_\eta\left(\frac{x}{\varepsilon_k}+\frac{T_\eta}{2}\right)}
& \textrm{if $x\in A_k$,}\\
u & \text{otherwise},
\end{array} \right.\nonumber\\
v_k(x):=\left\{ \begin{array}{ll}
\ {\displaystyle \beta_\eta\left(\frac{x}{\varepsilon_k}+\frac{T_\eta}{2}\right)}
& \textrm{if $x\in A_k$,}\\
1 & \text{otherwise}.
\end{array} \right.
}
An easy computation shows that $(u_k,v_k)\to(u,1)$ in $L^1(\Omega){\times}L^1(\Omega)$, that $u_k,v_k\in H^1(\Omega)$ for $k$ large, and that for the same $k$
$$F_k(u_k,v_k,\Omega\setminus A_k)\leq\int_{\Omega}|u'|^2dx,$$
being $f_k\leq 1$. Moreover using that $f_k\leq \varepsilon^{1/2}_kf$ and changing the variable $x$ with $y=\frac{x}{\varepsilon_k}+\frac{T_\eta}{2}$ one has
$$F_k(u_k,v_k,A_k)\leq g\big(|[u](0)|\big)+\eta,$$
where we have used \eqref{e17}. Therefore we find
$$F''(u,1)\leq \int_{\Omega}|u'|^2dx+\int_{J_u}(g(|[u]|) +\eta)d{\mathcal H}^0,$$
and then \be{\label{e18} F''(u,1)\leq \Phi(u),}
since $\eta$ is arbitrary.
Let us remove now the hypothesis that $u$ is constant near $0$. For a function $u\in SBV^2(\Omega)$ with $J_u=\{0\}$, one can consider
the sequence $u_j:=u$ in $\Omega\setminus (-1/j,1/j)$, with $u_j:=u(-1/j)$ in $(-1/j,0)$ and $u_j=u(1/j)$ in $(0,1/j)$. Then $u_j\to u$ in
$L^1(\Omega)$ and $|u'_j|\leq |u'|$ $\mathcal{L}^1$-a.e.\ in $\Omega$, so that by the lower semicontinuity of $F''$ and by the absolute continuity of $u$ on both sides of $0$ we conclude
as $j\to+\infty$ that $u$ still satisfies \eqref{e18}.
The extension of \eqref{e18} to each $u\in SBV^2(\Omega)$ with
${\mathcal H}^0(J_u)<+\infty$ is immediate and finally \cite[Propositions 3.3-3.5]{bou-bra-but} conclude the proof.
\end{proof}
\section[Proof in the \texorpdfstring{$n$}{n}-dimensional case]{Proof in the \texorpdfstring{$n$}{n}-dimensional case}\label{s:ndim}
In this section we establish the $\Gamma$-convergence result in the $n$-dimensional setting.
We recover the lower bound estimate by using a slicing technique thus reducing ourselves
to the one-dimensional setting of Proposition~\ref{p:liminfunidim}. Instead, the upper bound
inequality follows by an abstract approach based on integral representation results
(cp. Proposition~\ref{p:limsupndim} below).
\begin{proposition}\label{p:liminfndim}
For every $(u,v)\in L^1(\Omega){\times}L^1(\Omega)$ it holds
$$F(u,v)\leq F'(u,v).$$
\end{proposition}
\begin{proof}
Let us assume first that $u\in L^\infty(\Omega)$. We set $M:=||u||_{L^\infty(\Omega)}$.
Let $(u_k,v_k)$ be a sequence such that $(u_k,v_k)\to(u,v)$ in $L^1(\Omega){\times}L^1(\Omega)$ and
$\sup F_k(u_k,v_k)<+\infty$. Then it is straightforward that $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$.
We are going to show that $u\in BV(\Omega)$ and that
\be{\label{e34}\Phi(u)\leq \liminf_{k\to+\infty} F_k(u_k,v_k),}
that proves the thesis under the assumption of the boundedness of $u$.
Given $\xi\in\mathbb{S}^{n-1}$, we consider a subsequence $(u_r,v_r)$ of $(u_k,v_k)$ satisfying
$$((u_r)_y^\xi,(v_r)_y^\xi)\to (u_y^\xi,1) \text{ in }
L^1(\Omega_y^\xi){\times} L^1(\Omega_y^\xi) \text{ for } \mathcal{H}^{n-1}\text{-a.e.\ } y\in\Pi^\xi$$
and realizing the lower limit in \eqref{e34} as a limit.
By Fubini's theorem and Fatou's lemma one deduces that
\be{\label{e30}
\liminf_{r\to \infty}\int_{\Omega_y^\xi}\bigg(f_r((v_r)_y^\xi)\left|\nabla ((u_r)_y^\xi)\right|^2+
\frac{(1-(v_r)_y^\xi)^2}{4\varepsilon_r}+\varepsilon_r|\nabla ((v_r)^\xi_y)|^2\bigg)dt <+\infty}
holds for $\mathcal{H}^{n-1}\text{-a.e. } y\in \Omega^{\xi}$
The one-dimensional result Proposition \ref{p:liminfunidim} yields now that $u^\xi_y\in BV(\Omega^\xi_y)$ and that
\bm{\label{e29}\int_{\Omega^\xi_y} h(|\nabla (u^\xi_y)|)dt+\int_{J_{u^\xi_y}}g(|[u^\xi_y]|)d{\mathcal H}^0+\ell|D^cu^\xi_y|(\Omega^\xi_y)\leq\\
\leq\liminf_{r\to \infty}\int_{\Omega_y^\xi}\bigg(f_r((v_r)_y^\xi)\left|\nabla ((u_r)_y^\xi)\right|^2+
\frac{(1-(v_r)_y^\xi)^2}{4\varepsilon_r}+\varepsilon_r|\nabla ((v_r)^\xi_y)|^2\bigg)dt.}
Let us check that \eqref{e29} implies $u\in BV(\Omega)$ by estimating $\int_{\Omega^\xi}|D(u^\xi_y)|(\Omega^\xi_y)d{\mathcal H}^{n-1}$.
We first notice that
\be{\label{e32}\int_{\Omega^\xi_y} |\nabla (u^\xi_y)|dt\leq \frac{1}{\ell}\int_{\Omega^\xi_y} h(|\nabla (u^\xi_y)|)dt+\frac{\ell}{4}{\mathcal L}^1(\Omega^\xi_y),}
being $h(s)\geq \ell s-{\ell}^2/4$.
Since
$g(s)/s\to\ell$ as $s\to0$, with fixed $\eta>0$ one has
\be{\label{e33}g(s)>(\ell-\eta)s\qquad \textrm{for }s<\delta,}
for some $\delta$ sufficiently small.
Therefore \eqref{e32}, \eqref{e33}, and the boundedness of $u$ entail
\bm{|D(u^\xi_y)|(\Omega^\xi_y)\leq\frac{1}{\ell}\int_{\Omega^\xi_y} h(|\nabla (u^\xi_y)|)dt+\frac{\ell}{4}\diam{\Omega}+\frac{1}{\ell-\eta}\int_{\{t\in J_{u^\xi_y}:|[u^\xi_y]|<\delta\}}g(|[u^\xi_y]|)d{\mathcal H}^0\\
+\frac{2M}{g(\delta)}\int_{\{t\in J_{u^\xi_y}:|[u^\xi_y]|\geq\delta\}}g(|[u^\xi_y]|)d{\mathcal H}^0
+|D^cu^\xi_y|(\Omega^\xi_y)\nonumber\\
\leq c+c\left(\int_{\Omega^\xi_y} h(|\nabla (u^\xi_y)|)dt+\int_{J_{u^\xi_y}}g(|[u^\xi_y]|)d{\mathcal H}^0+\ell|D^cu^\xi_y|(\Omega^\xi_y)\right),
}
where $\diam\Omega$ denotes the diameter of $\Omega$ and $c:=\max\{\frac{1}{\ell},\frac{\ell}{4}\diam{\Omega},\frac{1}{\ell-\eta},\frac{2M}{g(\delta)}\}$.
Integrating the last inequality on $\Omega^\xi$ one deduces by \eqref{e29}
$$\int_{\Omega^\xi}|D(u^\xi_y)|(\Omega^\xi_y)d{\mathcal H}^{n-1}\leq c{\mathcal H}^{n-1}(\Omega^\xi)+c\sup_k F_k(u_k,v_k).$$
Taking $\xi=e_1,\dots,e_n$ one obtains $u\in BV(\Omega)$.
Let us prove now formula \eqref{e34} using localization. The integration on $\Omega^\xi$ of the one-dimensional estimate
in \eqref{e29} gives
\be{\label{e35}\int_\Omega h(|\nabla u\cdot\xi|)dx+\int_{J_{u}}|\nu_u\cdot\xi|g(|[u]|)d{\mathcal H}^{n-1}+\ell\int_\Omega |\gamma_u\cdot\xi|d|D^cu|\leq
\liminf_{k\to+\infty} F_k(u_k,v_k;\Omega),
}
where $\gamma_u:=\frac{d D^c u}{d|D^c u|}$ denotes the density of $D^cu$ with respect to $|D^c u|$.
Let $E\subset\Omega$ be a Borel set such that $D^a u(E)=0$ and $D^s u(\Omega\setminus E)=0$, and let
$$\lambda:={\mathcal L}^n\lfloor{\Omega\setminus E}+{\mathcal H}^{n-1}\lfloor{J_u}+|D^cu|\lfloor{E\setminus J_u}.$$
Let us consider a countable dense set $D\subset\mathbb{S}^{n-1}$ and the functions
$$\psi_\xi:=h(|\nabla u\cdot\xi|)\chi_{\Omega\setminus E}+|\nu_u\cdot\xi|g(|[u]|)\chi_{J_u}
+\ell|\gamma_u\cdot\xi|\chi_{E\setminus J_u},\qquad \xi\in D.$$
Then (\ref{e35}) gives $(\psi_\xi\lambda)(A)\le F'(u,1,A)$ for all open sets $A\subset\Omega$. Since
$F'(u,1,\cdot)$ is superadditive, this implies
$((\sup_\xi\psi_\xi)\lambda)(A)\le F'(u,1,A)$ (see \cite[Lemma 15.2]{braides})
and therefore the conclusion.
In the general case, if $u\in L^1\setminus L^\infty(\Omega)$ one considers $(u_k^M,v_k)$ and $(u^M,v)$,
where $u^M:=(-M\vee u)\wedge M$ denotes the truncation at level $M\in(0,+\infty)$. Since the functional
$F_k$ decreases by truncation and $u_k^M\to u^M$ in $L^1(\Omega)$,
we deduce that $u^M\in BV(\Omega)$ and
\be{\label{e37}\Phi(u^M)\leq\liminf_{k\to+\infty} F_k(u_k^M,v_k)\leq \liminf_{k\to+\infty} F_k(u_k,v_k).}
Therefore $u\in GBV(\Omega)$ and \eqref{e34} follows easily from \eqref{e37} as $M\to+\infty$.
\end{proof}
To prove the limsup inequality we follow an abstract approach. We first show that the
$\overline{\Gamma}$-limit is a Borel measure. The
only relevant property to be checked is the weak subadditivity of the $\Gamma$-limsup.
This is a consequence of De Giorgi's slicing and averaging argument as shown in the following lemma.
\begin{lemma}\label{l:wsub}
Let $(u,v)\inL^1(\Omega){\times}L^1(\Omega)$, let $A',A,B\in\mathcal{A}(\Omega)$ with $A'\subset\subset A$, then
\be{\label{e:subadd}
F''(u,1;A'\cup B)\leq F''(u,1;A)+F''(u,1;B).
}
\end{lemma}
\begin{proof}
We assume that the right-hand side of \eqref{e:subadd} is finite, so that $u\in GBV(A\cup B)$ and $v=1$ ${\mathcal L}^n$-a.e.\ in $A\cup B$.
We can reduce the problem to the case of functions $u\in BV\cap L^\infty(A\cup B)$. This is
a straightforward consequence of the fact that the energies $F_k$'s, and thus the $\Gamma$-limsup $F''$,
are decreasing by truncations. Actually, thanks to $L^1$ lower semicontinuity, they are continuous
under such an operation.
Under this assumption, let $(u_k^A,v_k^A)$, $(u_k^B,v_k^B)$ be recovery sequences for $(u,1)$ on $A$
and $B$ respectively, that is:
\begin{equation}\label{e:rec1}
(u_k^A,v_k^A), (u_k^B,v_k^B)\to (u,1) \text{ in } L^1(\Omega){\times} L^1(\Omega),
\end{equation}
and
\begin{equation}\label{e:rec2}
\limsup_{k\to+\infty} F_k(u_k^A,v_k^A;A)=F''(u,1;A),\quad \limsup_{k\to+\infty} F_k(u_k^B,v_k^B;B)=F''(u,1;B).
\end{equation}
Note that, again up to truncations, we may assume that
\begin{equation}\label{e:rec3}
(u_k^A,v_k^A),\, (u_k^B,v_k^B)\quad\text{are bounded in } L^\infty(\Omega).
\end{equation}
To simplify the calculations below we introduce the functionals
$G_k:L^1(\Omega)\times\mathcal{A}(\Omega)\to[0,+\infty]$ given by
\[
G_k(v;O):=\int_O\left(\frac{(1-v)^2}{4\varepsilon_k}+\varepsilon_k|\nabla v|^2\right)dx,\quad
\text{if $v\in H^1(\Omega)$,}
\]
$+\infty$ otherwise.
Notice that
\[
F_k(u,v;O)=\int_O f_k^2(v)|\nabla u|^2dx+G_k(v;O).
\]
Let $\delta:=\mathrm{dist}(A^\prime,\partial A)>0$, and with fixed $M\in{\mathbb N}$, we set for all
$i\in\{1,\ldots,M\}$
\[
A_i:=\left\{x\in \Omega:\,\mathrm{dist}(x,A^\prime)<\frac{\delta}{M}i\right\},
\]
and $A_0:=A^\prime$. Clearly, we have $A_{i-1}\subset\subset A_i\subset A$.
Denote by $\varphi_i\in C_c^1(\Omega)$ a cut-off function between $A_{i-1}$ and $A_i$,
i.e., $\varphi_i|_{A_{i-1}}=1$, $\varphi_i|_{A_{i}^c}=0$, and
$\|\nabla \varphi_i\|_{L^\infty(\Omega)}\leq \frac{2M}{\delta}$. Then, set
\begin{equation}\label{e:uki}
u_k^i:=\varphi_i\,u_k^A+(1-\varphi_i)u_k^B,
\end{equation}
and
\begin{equation}\label{e:vki}
v_k^i:=\begin{cases}
\varphi_{i-1}\,v_k^A+(1-\varphi_{i-1})(v_k^A\wedge v_k^B) & \text{ on } A_{i-1}\cr
v_k^A\wedge v_k^B & \text{ on } A_i\setminus {A}_{i-1} \cr
\varphi_{i+1}(v_k^A\wedge v_k^B)+(1-\varphi_{i+1})\,v_k^B & \text{ on } \Omega\setminus {A}_i.
\end{cases}
\end{equation}
With fixed $i\in\{2,\ldots,M-1\}$, by the very definitions in \eqref{e:uki} and \eqref{e:vki} above $(u_k^i,v_k^i)\in H^1(\Omega){\times}H^1(\Omega)$ and
the related energy $F_k$ on $A'\cup B$ can be estimated as follows
\begin{equation}\label{e:split}
F_k(u_k^i,v_k^i;A'\cup B)\leq F_k(u_k^A,v_k^A;A_{i-2})+F_k(u_k^B,v_k^B;B\setminus {A}_{i+1})+
F_k(u_k^i,v_k^i;B\cap(A_{i+1}\setminus {A}_{i-2})).
\end{equation}
Therefore, we need to bound only the last term. To this aim we further split the contributions in
each layer; in estimating each of such terms we shall repeatedly use the monotonicity of $f_k$ and the
fact that it is bounded by $1$. In addition, a
positive constant, which may vary from line to line,
will appear in the formulas below. Elementary computations and the very definitions in \eqref{e:uki} and
\eqref{e:vki} give, using $v^i_k\le v^A_k$,
\begin{multline}\label{e:i-1i-2}
F_k(u_k^i,v_k^i;B\cap(A_{i-1}\setminus {A}_{i-2}))\leq
\int_{B\cap(A_{i-1}\setminus {A}_{i-2})}f_k^2(v_k^A)|\nabla u_k^A|^2\,dx
+G_k(v_k^i;B\cap(A_{i-1}\setminus {A}_{i-2}))\\
\leq c\,\Big(F_k(u_k^A,v_k^A;B\cap(A_{i-1}\setminus {A}_{i-2}))
+F_k(u_k^B,v_k^B;B\cap(A_{i-1}\setminus {A}_{i-2}))\Big)\\
+\frac{c\,M^2\varepsilon_k}{\delta^2}
\int_{B\cap(A_{i-1}\setminus {A}_{i-2})}|v_k^A-v_k^B|^2\,dx,
\end{multline}
\begin{multline}\label{e:ii-1}
F_k(u_k^i,v_k^i;B\cap(A_{i}\setminus {A}_{i-1}))\\
\leq c\int_{B\cap(A_{i}\setminus {A}_{i-1})}f_k^2(v_k^A\wedge v_k^B)
\left(|\nabla u_k^A|^2+|\nabla u_k^B|^2+\frac{4M^2}{\delta^2}|u_k^A-u_k^B|^2\right)\,dx
+G_k(v_k^A\wedge v_k^B;B\cap(A_{i}\setminus {A}_{i-1}))\\
\leq c\Big(F_k(u_k^A,v_k^A;B\cap(A_{i}\setminus {A}_{i-1}))
+F_k(u_k^B,v_k^B;B\cap(A_{i}\setminus {A}_{i-1}))\Big)
+\frac{c\,M^2}{\delta^2}\int_{B\cap(A_{i}\setminus {A}_{i-1})}|u_k^A-u_k^B|^2\,dx,
\end{multline}
and
\begin{multline}\label{e:i+1i}
F_k(u_k^i,v_k^i;B\cap(A_{i+1}\setminus {A}_{i}))\leq
\int_{B\cap(A_{i+1}\setminus {A}_{i})}f_k^2(v_k^B)|\nabla u_k^B|^2\,dx
+G_k(v_k^i;B\cap(A_{i+1}\setminus {A}_{i}))\\
\leq c\,\Big(F_k(u_k^A,v_k^A;B\cap(A_{i+1}\setminus {A}_{i}))
+F_k(u_k^B,v_k^B;B\cap(A_{i+1}\setminus {A}_{i}))\Big)+\frac{c\,M^2\varepsilon_k}{\delta^2}
\int_{B\cap(A_{i+1}\setminus {A}_{i})}|v_k^A-v_k^B|^2\,dx.
\end{multline}
By adding \eqref{e:split}-\eqref{e:i+1i}, we deduce that
\begin{multline*}
F_k(u_k^i,v_k^i;A'\cup B)\leq F_k(u_k^A,v_k^A;A)+F_k(u_k^B,v_k^B;B)\\
+c\,\Big(F_k(u_k^A,v_k^A;B\cap(A_{i+1}\setminus {A}_{i-2}))
+F_k(u_k^B,v_k^B;B\cap(A_{i+1}\setminus {A}_{i-2}))\Big)\\
+\frac{c\,M^2}{\delta^2}\int_{B\cap(A_{i+1}\setminus {A}_{i-2})}|u_k^A-u_k^B|^2\,dx
+\frac{c\,M^2\varepsilon_k}{\delta^2}\int_{B\cap(A_{i+1}\setminus {A}_{i-2})}|v_k^A-v_k^B|^2\,dx.
\end{multline*}
Hence, by summing up on $i\in\{2,\ldots,M-1\}$ and taking the average, for each $k$ we may find an index
$i_k$ in that range such that
\begin{multline*}
F_k(u_k^{i_k},v_k^{i_k};A'\cup B)\leq F_k(u_k^A,v_k^A;A)+F_k(u_k^B,v_k^B;B)\\
+\frac{c}{M}\,\Big(F_k(u_k^A,v_k^A;B\cap(A\setminus {A^\prime}))
+F_k(u_k^B,v_k^B;B\cap(A\setminus {A^\prime}))\Big)\\
+\frac{c\,M}{\delta^2}\int_{B\cap(A\setminus {A^\prime})}|u_k^A-u_k^B|^2\,dx
+\frac{c\,M\varepsilon_k}{\delta^2}\int_{B\cap(A\setminus {A^\prime})}|v_k^A-v_k^B|^2\,dx.
\end{multline*}
By \eqref{e:rec1} we deduce that $(u_k^{i_k},v_k^{i_k})\to (u,1)$ in $L^1(\Omega){\times} L^1(\Omega)$,
and actually in $L^q(\Omega){\times} L^q(\Omega)$ for all $q\in[1,+\infty)$ thanks to the uniform boundedness
assumption in \eqref{e:rec3}. Therefore, in view of \eqref{e:rec2} and the definition of $\Gamma$-limsup
we infer that
\[
F''(u,1;A'\cup B)\leq \left(1+\frac{c}{M}\right)\Big(F''(u,1;A)+F''(u,1;B)\Big).
\]
The conclusion then follows by passing to the limit on $M\uparrow\infty$.
\end{proof}
We next prove that $F''(u,1;\cdot)$ is controlled in terms of the Mumford-Shah functional ${M\!S}$, whose
definition is given in \eqref{e:MS}. This result gives a first rough estimate for the upper bound inequality.
We shall improve on the jump part in Proposition~\ref{p:limsupndim} below and finally we shall conclude the proof of the $\Gamma$-limsup
inequality using a relaxation argument.
\begin{lemma}\label{l:boundglimsup}
For all $u\in L^1(\Omega)$ and $A\in\mathcal{A}(\Omega)$ it holds
\begin{equation}\label{e:FH}
F''(u,1;A)\leq {M\!S}(u;A).
\end{equation}
\end{lemma}
\begin{proof}
Denote by $\psi:[0,1]\to[0,1]$ any nondecreasing
lower-semicontinuous function such that $\psi^{-1}(0)=0$, $\psi(1)=1$ and
\[
\sup_{k}f_k(s)\leq\psi(s)\qquad\text{for all $s\in[0,1]$},
\]
for instance $\psi=\chi_{(0,1]}$ satisfies all the conditions written above.
Consider the corresponding functionals $AT_k^\psi:L^1(\Omega)\times L^1(\Omega)\to[0,+\infty]$
defined in \eqref{e:ATk},
and note that $F_k\leq AT_k^\psi$ for every $k$.
The upper bound inequality for $(F_k)$ then follows at once from the classical
results by Ambrosio and Tortorelli (cp. \cite{amb-tort2}, and see also \cite{focardi}).
\end{proof}
We are now ready to prove the upper bound inequality.
\begin{proposition}\label{p:limsupndim}
For every $(u,v)\in L^1(\Omega){\times}L^1(\Omega)$ it holds
\[
F''(u,v)\leq F(u,v).
\]
\end{proposition}
\begin{proof}
Since $L^1$ is separable, given any subsequence $(F_{k_j})$ of $(F_k)$ we may extract a further
subsequence, not relabeled for convenience, $\overline{\Gamma}$-converging to some
$\widehat{F}$ (see \cite[Theorem~16.9]{dalmaso}).
The functional $\widehat{F}(u,v;\cdot)$ is by definition increasing and inner regular. Since $F_k(u,v;\cdot)$ is
additive, one easily deduces that $F'$ is superadditive
and from this that its inner regular envelope $\widehat F=(F')_-$ is superadditive (see \cite[Proposition~14.18 or Proposition~16.12]{dalmaso}).
Using Lemma~\ref{l:wsub} one can show that
$\widehat F=(F'')_-$ is subadditive (see \cite[Lemma~14.20 and the proof of Proposition~18.4]{dalmaso}).
Therefore $\widehat F$ is the restriction to open sets of the Borel measure
\begin{equation*}
F_*(u,v;E)=\inf \{ \widehat F(u,v;A): A\in \mathcal{A}(\Omega); E\subset A\}\,,
\end{equation*}
see \cite[Theorem~14.23]{dalmaso}, in the following we identify $\widehat F$ and $F_*$.
If $u\in L^1(\Omega)$ is such that ${M\!S}(u;\Omega)<+\infty$, then by Lemma~\ref{l:boundglimsup} we obtain
$F''(u,1;\cdot)\le {M\!S}(u;\cdot)<+\infty$ on all open sets, and by the regularity
properties of Radon measures $F''$ coincides with its inner envelope.
Indeed, for a given open set $A$ and $\varepsilon>0$, choose open sets $A'$, $A''$ and $C$ with
$A'\subset\subset A''\subset\subset A$
and $A\setminus A'\subset C$ such that ${M\!S}(u;C)\le\varepsilon$.
Then use Lemmas~\ref{l:wsub} and \ref{l:boundglimsup} to estimate
$F''(u,1;A)\le F''(u,1;A'\cup C)\le F''(u,1;A'')+{M\!S}(u;C)\leq F''(u,1;A'')+\varepsilon$.
In other words, $\widehat{F}(u,1)$ is the
$\Gamma$-limit of $F_{k_j}$ for all $u$ such that ${M\!S}(u)<+\infty$.
For all $u\in SBV^2(\Omega)$ in particular the estimate in Lemma~\ref{l:boundglimsup} implies that
\begin{equation}\label{e:upbvolume}
\widehat{F}(u,1;\Omega\setminus J_u)\leq\int_\Omega |\nabla u|^2dx.
\end{equation}
We provide below for the same $u$ the estimate
\begin{equation}\label{e:upbvsalto}
\widehat{F}(u,1;J_u)\leq\int_{J_u}g(|[u]|)\,d{\mathcal H}^{n-1}.
\end{equation}
Given this for granted we conclude as follows: we consider the functional $F_\infty:BV(\Omega)\to[0,+\infty]$
\[
F_\infty(u):=\begin{cases}
\displaystyle{
\int_\Omega |\nabla u|^2dx+\int_{J_u}g(|[u]|)\,d{\mathcal H}^{n-1}} & \textrm{if $u\in SBV^2(\Omega)$}
\cr
+\infty & \textrm{otherwise on $BV(\Omega)$.}
\end{cases}
\]
Further, note that by \cite[Theorem~3.1 and Propositions~3.3-3.5]{bou-bra-but} its relaxation w.r.to the $w\ast\hbox{-}BV$
topology is given on $BV(\Omega)$ by $F(\cdot,1)$.
Since, by \eqref{e:upbvolume} and \eqref{e:upbvsalto} we have that $\widehat{F}\leq F_\infty$,
and $\widehat{F}(\cdot,1)$ is $L^1$-lower semicontinuous, we infer that
\[
\widehat{F}(u,1)\leq F(u,1)\quad\textrm{for all $u\in BV(\Omega)$.}
\]
We conclude that the same inequality is true for all $u\in GBV\cap L^1(\Omega)$ by the usual truncation
argument.
Finally, by combining the latter estimate with the lower estimate of Proposition~\ref{p:liminfndim}
allows us to deduce that the $\Gamma\hbox{-}$limit does not depend on the chosen subsequence
and it is equal to $F$. Hence, by Urysohn's property the whole family $(F_k)$ $\Gamma\hbox{-}$converges
to $F$ (cp. \cite[Proposition~8.3]{dalmaso}).
Let us now prove formula \eqref{e:upbvsalto}. To this aim, fixed $\lambda>0$ we introduce the perturbed
functional
\[
\widehat{F_{\lambda}}(u,1):=\widehat{F}(u,1)+\lambda\Big(\int_{\Omega}|\nabla u|^2dx
+\int_{J_u}(1+|[u]|)d{\mathcal H}^{n-1}\Big)
\]
for all $u\in SBV^2(\Omega)$. We may apply to $\widehat{F_{\lambda}}$ the integral representation result
\cite[Theorem~1]{bou-fon-leo-masc} to infer that for ${\mathcal H}^{n-1}$-a.e. $x\in J_u$
\bm{\label{e:flambda}
\frac{d\widehat{F_{\lambda}}(u,1;\cdot)}{d({\mathcal H}^{n-1}\res J_u)}(x)=\limsup_{\delta\downarrow 0}\frac{1}{\delta^{n-1}}
\inf\left\{\widehat{F_{\lambda}}(w,1;x+\delta\,Q_{\nu_u(x)}):\,w\in SBV^2\big(x+\delta\,Q_{\nu_u(x)}\big),\right.
\\ \left.w=u_x \text{ on a neighborhood of }x+\delta\,\partial Q_{\nu_u(x)}\right\},
}
where
\[
u_x(y):=
\begin{cases}
u^+(x) & \text { if }\langle y-x,\nu_u(x)\rangle>0\cr
u^-(x) & \text { if }\langle y-x,\nu_u(x)\rangle<0
\end{cases}
\]
and $Q_{\nu_u(x)}$ denotes any cube of side $1$ centered in the origin and with a face
orthogonal to $\nu_u(x)$.
Hence, it is enough to show that for ${\mathcal H}^{n-1}$-a.e. $x\in J_u$
\begin{equation}\label{e:minest}
\limsup_{\delta\downarrow 0}\frac 1{\delta^{n-1}}\widehat{F}(u_x,1;x+\delta\, Q_{\nu_u(x)})\leq g(|[u](x)|),
\end{equation}
since by taking $u_x$ itself as test function in \eqref{e:flambda} we get
$$
\frac{d\widehat{F_{\lambda}}(u,1;\cdot)}{d({\mathcal H}^{n-1}\res J_u)}(x)\leq \limsup_{\delta\downarrow 0}\frac{1}{\delta^{n-1}}
\widehat{F}(u_x,1;x+\delta\, Q_{\nu_u(x)})+\lambda(1+|[u](x)|),
$$
in turn implying
\[
\widehat{F}(u,1;J_u)\leq\widehat{F_{\lambda}}(u,1;J_u)\leq
\int_{J_u}\big(g(|[u](x)|)+\lambda+\lambda|[u](x)|\big)\,d{\mathcal H}^{n-1}.
\]
Finally, \eqref{e:upbvsalto} follows at once by letting $\lambda\downarrow 0$.
Formula \eqref{e:minest} easily follows by repeating the one-dimensional construction
of Proposition~\ref{p:limsupunidim}. More precisely, assume $x=0$ and $\nu_u(x)=e_n$ for simplicity.
With fixed $\eta>0$, let $T_{\eta}>0$ and $\alpha_{\eta},\beta_{\eta}\in H^1\big((0,T_{\eta})\big)$ be such that
$\alpha_{\eta}(0)=u^-(0),$ $\alpha_{\eta}(T_{\eta})=u^+(0)$, $\beta_{\eta}(0)=\beta_{\eta}(T_\eta)=1$,
$u^-(0)\leq\alpha_{\eta}\leq u^+(0)$, $0\leq\beta_{\eta}\leq 1$, and
\[
\int_0^{T_{\eta}} \left(f^2(\beta_{\eta})|\alpha_{\eta}'|^2+
\frac{|1-\beta_{\eta}|^2}{4}+|\beta_{\eta}'|^2\right)dt\leq g\big(|[u](0)|\big)+\eta.
\]
Let $A_j:=(-\frac{\varepsilon_{k_j}T_\eta}{2},\frac{\varepsilon_{k_j}T_\eta}{2})$, and set
\ba
{u_j(y):=\left\{ \begin{array}{ll}
\ {\displaystyle \alpha_{\eta}\left(\frac{y_n}{\varepsilon_{k_j}}+\frac{T_\eta}{2}\right)}
& \textrm{if $y_n\in A_j$}\\
u_0 & \text{otherwise},
\end{array} \right.\nonumber\\
v_j(y):=\left\{ \begin{array}{ll}
\ {\displaystyle \beta_{\eta}\left(\frac{y_n}{\varepsilon_{k_j}}+\frac{T_\eta}{2}\right)}
& \textrm{if $y_n\in A_j$}\\
1 & \text{otherwise}.
\end{array} \right.
}
Clearly, $(u_j,v_j)\to(u_0,1)$ in $L^1(Q_{e_n})\times L^1(Q_{e_n})$, and if
$Q'_{e_n}=Q_{e_n}\cap({\mathbb R}^{n-1}\times\{0\})$, a change of variable yields
\begin{multline*}
F_{k_j}(u_j,v_j;\delta\,Q_{e_n})=F_{k_j}\big(u_j,v_j,\delta\,Q'_{e_n}\times A_j\big)\\
\leq\delta^{n-1}\int_0^{T_{\eta}} \left(f^2(\beta_{\eta})|\alpha_{\eta}'|^2+
\frac{|1-\beta_{\eta}|^2}{4}+|\beta_{\eta}'|^2\right)dt\leq \delta^{n-1}(g\big(|[u](0)|\big)+\eta).
\end{multline*}
Therefore, by the very definition of $\widehat{F}$ we infer that
\[
\widehat{F}(u_0,1;\delta\,Q_{e_n})\leq \delta^{n-1}(g\big(|[u](0)|\big)+\eta),
\]
and estimate \eqref{e:minest} follows at once dividing by $\delta^{n-1}$ and taking the superior
limit as $\delta\downarrow 0$, and finally by letting $\eta\downarrow 0$ in the formula above.
\end{proof}
The proof of the compactness result Theorem~\ref{t:comp} follows the lines of \cite[Theorem 7.4]{dm-iur},
so we just sketch the relevant arguments and refer to \cite{dm-iur} for more details.
\begin{proof}[Proof of Theorem~\ref{t:comp}]
One first proves the thesis in the one-dimensional case under the hypothesis that $u_k$ is bounded in
$L^\infty(\Omega)$. Then one extends the proof to the $n$-dimensional case and finally removes the
boundedness assumption.
Let us start assuming that $n=1$ and that $\sup_k ||u_k||_{L^\infty(\Omega)}<+\infty$.
Up to a diagonalization argument, one reduces to study the case $\Omega=(0,1)$. Repeating the proof of
Theorem~\ref{p:liminfunidim} one finds that $v_k\to1$ in $L^1(\Omega)$ and that for every $\delta>0$ there
exists a finite subset $S\subset\Omega$ for which
$$(1-\omega(\delta))\int_{\Omega\setminus S_{\eta}}h(|u'_k|)dx\leq F_k(u_k,v_k,\Omega\setminus S_{\eta})$$
holds for $\eta>0$ small (dependently on $\delta$) and for $k$ large (dependently on $\eta$),
where $\omega$ is a modulus of continuity provided by $f$ and $S_{\eta}:=\bigcup_{i=1}^L(t_i-\eta,t_i+\eta)$.
This implies by assumption that $u_k$ is bounded in $BV(\Omega\setminus S_{\eta})$ uniformly with respect
to $k$ and $\eta$.
Hence up to subsequences $u_k$ converges to a function $u\in BV(\Omega\setminus S_{\eta})$ ${\mathcal L}^1$-a.e.\ in
$\Omega\setminus S_{\eta}$.
The boundedness hypothesis and a diagonalization argument yield that $u$ in fact belongs to $BV(\Omega)$
and that $u_k\to u$ $L^1(\Omega)$.
\medskip
In order to generalize the previous result to the case $n>1$, one applies a compactness result by
Alberti, Bouchitt\'{e}, and Seppecher \cite[Theorem 6.6]{alberti}. Indeed, fixed $\xi\in\mathbb{S}^{n-1}$ and $\delta>0$,
one can introduce the sequence $w_k$ whose slices satisfy
\bes{
& (w_k)^\xi_y:=\begin{cases}(u_k)^{\xi}_y & \textrm{if $y\in A_k$,}\\
0 & \textrm{otherwise,}
\end{cases}\\
& A_k:=\{y\in \Omega^{\xi}:F^1_k((u_k)^{\xi}_y,(v_k)^\xi_y)\leq L\},}
where $F_k^1$ denotes the one-dimensional counterpart of the functional $F_k$ and $L$ is chosen properly and
depends on $\delta$.
An easy computation shows that $w_k$ is bounded in $L^\infty(\Omega)$, that $u_k$ is in a $\delta$-neighborhood
of $w_k$ in $L^1(\Omega)$, and that $(w_k)^\xi_y$ is pre-compact in $L^1(\Omega)$ (the last property follows
from the first part of the proof).
Then the pre-compactness of $u_k$ in $L^1(\Omega)$ is ensured by \cite[Theorem 6.6]{alberti} as $\xi$ varies
in a basis of ${\R}^n$.
\medskip
If $u_k$ is not bounded in $L^\infty(\Omega)$ the argument above applies to the truncations, so that up
to subsequences $u_k^M\to u_M$ in $L^1(\Omega)$ and $\mathcal{L}^n\text{-a.e.\ in } \Omega$, with $u_M\in BV(\Omega)$, for every $M\in {\mathbb N}$.
One can prove that the function
$$u:=\lim_{M\to+\infty}u_M$$
is well-defined, finite $\mathcal{L}^n\text{-a.e.\ in } \Omega$, and its truncation $u^M$ coincides with $u_M$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$. This straightforwardly
implies that $u_k\to u$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$ and that $u\in GBV\cap L^1(\Omega)$.
\end{proof}
\section{Further results}\label{s:further}
In this section we build upon the results in Sections~\ref{s:stat}-\ref{s:ndim} to obtain in the
limit different models by slightly changing the approximating energies $F_k$'s.
More precisely, we shall approximate a cohesive model with the Dugdale's surface density, a
cohesive model with power-law growth at small openings, and a model in Griffith's brittle fracture.
This task will be accomplished by letting the function $f$ vary as in item (ii)
of Proposition~\ref{p:gl} in the first instance, as in item (iii) in the third, and suitably
in the second (cp. (iii) of Proposition~\ref{p:tp} below), respectively.
More precisely, we consider a sequence of functions $(\f{j})$ satisfying \eqref{f0} and \eqref{f1}
and for all $j,k\in{\mathbb N}$ introduce the energies
\begin{equation}\label{e:Fj}
\Fk jk(u,v):=\begin{cases}
\displaystyle{\int_\Omega{\Big((\fk jk)^2(v)|\nabla u|^2+\frac{(1-v)^2}{4\varepsilon_k}
+\varepsilon_k|\nabla v|^2\Big) dx}}
& \textrm{if $(u,v)\in H^1(\Omega){\times}H^1(\Omega)$}\cr
& \text{and $0\leq v\leq1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$},\cr
+ \infty & \text{otherwise},
\end{cases}
\end{equation}
where $\fk jk(s):=1\wedge\varepsilon_k^{1/2} \f{j}(s)$.
In each of Theorems~\ref{t:Dugdale}, \ref{t:sublin}, and \ref{t:MS} below we shall further
specify the nature of the sequence $(\f{j})$.
\subsection{Dugdale's cohesive model}
\label{subsecdugdale}
In order to approximate the Dugdale's model, i.e., to get in the limit
$\mathscr{D}:L^1(\Omega)\to[0,+\infty]$
\begin{equation}\label{e:Phit}
\mathscr{D}(u):=
\begin{cases}
\displaystyle{\int_\Omega h(|\nabla u|)dx+\int_{J_{u}}\big(\gD |[u]|\big)d{\mathcal H}^{n-1}+\ell|D^cu|(\Omega)}
& \textrm{if $u\in GBV(\Omega)$,}\cr\cr
+\infty & \textrm{otherwise},
\end{cases}
\end{equation}
with $h$ as in \eqref{h},
we shall consider the specific choice
\begin{equation}\label{e:fjPhit}
\f j(s):=(a_j\,s)\vee f(s)
\end{equation}
with $f$ satisfying \eqref{f0} and \eqref{f1}, and
\begin{equation}\label{e:alphaj}
\text{$(a_j)$ nondecreasing, $a_j\uparrow\infty$ and such that
$a_j\,\varepsilon_j^{1/2}\downarrow 0$.}
\end{equation}
\begin{theorem}\label{t:Dugdale}
Suppose that $(\f j)$ is as in \eqref{e:fjPhit} and \eqref{e:alphaj} above.
Then, the functionals $\Fk kk$ $\Gamma$-converge in $L^1(\Omega){\times}L^1(\Omega)$ to the functional $\widetilde{\D}$
defined as follows
\begin{equation}\label{e:F1}
\widetilde{\D}(u,v):=\begin{cases}
\mathscr{D}(u) & \textrm{if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$,}\cr
+\infty & \text{otherwise}.
\end{cases}
\end{equation}
\end{theorem}
\begin{proof}[Proof of Theorem~\ref{t:Dugdale}]
The very definitions in \eqref{Fk} and \eqref{e:Fj} give $\Fk jk\leq \Fk kk$ for $j\leq k$,
being $(\f j)$ nondecreasing by assumption. Hence, by Theorem~\ref{t:gamma-lim} we deduce
\begin{equation}\label{e:liminfFj0}
\Gamma\hbox{-}\liminf_k \Fk kk(u,v)\geq \F j(u,v),
\end{equation}
where $\F j$ is defined as $F$ in \eqref{F} with $f$ substituted by $\f j$ in formulas
\eqref{f1} and \eqref{h} defining the volume density, and \eqref{g} defining the surface density.
In particular, being $\ell_j=\ell$ for all $j$, the corresponding volume density $h_j$
equals the function $h$ in \eqref{h}.
Moreover, the surface energy densities $g_j$ are dominated by the constant $1$,
and by item (ii) in Proposition~\ref{p:gl} we have $\lim_jg_j(s)=\gD s$
for all $s\in[0,+\infty)$.
In conclusion, if $\Gamma\hbox{-}\liminf_k \Fk kk(u,v)<+\infty$, we infer that
$v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$, $u\in GBV(\Omega)$, and
\[
\Gamma\hbox{-}\liminf_k\Fk kk(u,1)\geq \mathscr{D}(u),
\]
by the Dominated Convergence theorem, as $j\uparrow+\infty$ in \eqref{e:liminfFj0}.
The upper bound inequality follows by arguing as in Proposition~\ref{p:limsupndim}.
Indeed, we first note that by a careful inspection of the proofs, Lemmas~\ref{l:wsub}
and \ref{l:boundglimsup} are still valid in this generalized framework.
More precisely, Lemma~\ref{l:wsub} continue to hold true as there we have only used that each
function $f_k=1\wedge\varepsilon_k^{1/2} f$ in \eqref{fk} is nondecreasing and bounded by $\chi_{(0,1]}$,
properties enjoyed by $\fk kk$ as well.
In conclusion, as a first step we establish the estimate
\begin{equation}\label{e:minest2}
\limsup_{\delta\downarrow 0}\frac 1{\delta^{n-1}}\widehat{F}(u_x,1;x+\delta\, Q_{\nu_u(x)})\leq
\gD{|[u](x)|},
\end{equation}
for $u\in SBV^2(\Omega)$ and for ${\mathcal H}^{n-1}$-a.e. $x\in J_u$, where $\widehat{F}$ is the
$\overline{\Gamma}$-limit of a properly chosen subsequence $(\Fk{k_j}{k_j})$ of $(\Fk kk)$
(cp. Proposition~\ref{p:limsupndim}).
Given \eqref{e:minest2}, the derivation of the upper bound inequality in general follows
exactly as in Proposition~\ref{p:limsupndim}.
Let us now prove \eqref{e:minest2} by means of a one-dimensional construction.
For the sake of simplicity we assume $x=0$ and $\nu_u(x)=e_n$.
Actually, in view of the estimate in Lemma~\ref{l:boundglimsup} we need only to discuss
the case $|[u](0)|<\ell^{-1}$. To this aim, set
\[
s_j:=\sup\{s\in[0,1):\,a_{k_j}s=f(s)\},
\]
it is easy to check that $s_j$ it is actually a maximum, i.e., $a_{k_j}s_j=f(s_j)$, and that
$s_j\leq s_{j+1}<1$ with $s_j\uparrow 1$.
Let now
\[
T_{j}:=|[u](0)|\frac{f(s_j)}{1-s_j},
\]
then $T_j\uparrow\infty$. Define $\alpha_j(t):=u^-(0)$ on $[-T_j-1,-T_j]$,
$\alpha_j(t):=[u](0)\cdot(\frac{t}{2T_j}+\frac12)+u^-(0)$ on $[-T_j,T_j]$,
$\alpha_j(t):=u^+(0)$ on $[T_j,T_j+1]$, and $\beta_j(t):=s_j$ on $[-T_j,T_j]$,
$\beta_j(t):=(1-s_j)(|t|-T_j)+s_j$ otherwise in $[-T_j-1,T_j+1]$.
Setting $A_j:=(-{\varepsilon_{k_j}(T_j+1)},{\varepsilon_{k_j}(T_j}+1))$, we have that ${\mathcal L}^1(A_j)\to0$ as $j\uparrow\infty$
by \eqref{e:alphaj}. Indeed, in view of \eqref{f1} and the definition of $s_j$ it is easy to deduce
that $(1-s_j)a_{k_j}\to\ell$ as $j\uparrow\infty$, so that
$\varepsilon_{k_j}T_j\sim\varepsilon_{k_j}a_{k_j}^2\to 0$
as $j\uparrow\infty$ thanks to \eqref{e:alphaj}. Therefore, if
\ba
{u_j(y):=\left\{ \begin{array}{ll}
{\displaystyle \alpha_j\left(\frac{y_n}{\varepsilon_{k_j}}\right)}
& \textrm{if $y_n\in A_j$}\\
u_0 & \text{otherwise},
\end{array} \right.\nonumber\\
v_j(y):=\left\{ \begin{array}{ll}
{\displaystyle \beta_j\left(\frac{y_n}{\varepsilon_{k_j}}\right)}
& \textrm{if $y_n\in A_j$}\\
1 & \text{otherwise},
\end{array} \right.
}
then $(u_j,v_j)\to (u_0,1)$ on $L^1(Q_{e_n})\times L^1(Q_{e_n})$,
where $u_0=u^-(0)\chi_{\{y_n\leq 0\}}+u^+(0)\chi_{\{y_n> 0\}}$.
Moreover, if $Q'_{e_n}=Q_{e_n}\cap({\mathbb R}^{n-1}\times\{0\})$, then a change of variable yields
\begin{multline*}
\Fk{k_j}{k_j}(u_j,v_j;\delta\,Q_{e_n})=\Fk {k_j}{k_j}\big(u_j,v_j;\delta\,Q'_{e_n}\times A_j\big)\\
\leq\delta^{n-1}\left(\int_{-T_j}^{T_j}\bigg(f^2(\beta_j)|\nabla\alpha_j|^2+\frac{(1-\beta_j)^2}4\bigg)dt
+2\int_{T_j}^{T_j+1} \left(\frac{|1-\beta_j|^2}{4}+|\beta'_j|^2\right)dt\right)\\
=\delta^{n-1}\left(
\bigg(f^2(s_j)\frac{|[u](0)|^2}{2T_j}+2T_j\frac{(1-s_j)^2}4\bigg)
+2(1-s_j)^2\int_{T_j}^{T_j+1}\frac{(t-(T_j+1))^2}{4}dt+2(1-s_j)^2\right)\\
=\delta^{n-1}\left((1-s_j)f(s_j)|[u](0)|+\frac{13}6(1-s_j)^2\right)
=\delta^{n-1}\big(\ell|[u](0)|+o(1)\big)\qquad\textrm{as $j\uparrow\infty$}.
\end{multline*}
Therefore, being $|[u](0)|<\ell^{-1}$, by the very definition of $\widehat{F}$ we infer that
\[
\widehat{F}(u_0,1;\delta\,Q_{e_n})\leq \delta^{n-1}(\gD{|[u](0)|}),
\]
and estimate \eqref{e:minest2} follows at once dividing by $\delta^{n-1}$ and taking the superior
limit as $\delta\downarrow 0$ in the formula above.
\end{proof}
\begin{remark}\label{r:regimes}
The analysis in the general case of a diverging sequence $f^{(k)}$ is much more intricate because
of the combination of several effects: the speed of divergence of the $f^{(k)}$'s compared with the
scaling $\varepsilon_k^{1/2}$ in the definition of $f^{(k)}_k$, and even more the behavior of each $f^{(k)}$
close to $1$.
In this remark we limit ourselves to consider those families of functions $f^{(k)}$ satisfying item
(ii) in Proposition~\ref{p:gl}, another instance shall be discussed in Remark~\ref{r:regimes2} below.
Therefore, assume for example that $f(s)=\frac{\ell s}{1-s}$, and that
$\f k$ is defined as in \eqref{e:fjPhit} above but with $a_k=\varepsilon_k^{-1/2}$, thus violating the
last condition in \eqref{e:alphaj}.
Then, one can show that the $\Gamma$-limit is given by the Mumford-Shah energy introduced
in \eqref{e:MS}. This claim follows easily by noting that with this choice
\[
\fk kk(s)=\begin{cases}
s & 0\leq s\leq 1-\ell\,\varepsilon_k^{1/2} \cr
\displaystyle{\varepsilon_k^{1/2}\frac{\ell s}{1-s}}
& 1-\ell\,\varepsilon_k^{1/2}\leq s\leq (1+\ell\,\varepsilon_k^{1/2})^{-1} \cr
1 & (1+\ell\,\varepsilon_k^{1/2})^{-1}\leq s\leq 1,
\end{cases}
\]
so that $\fk kk(s)\geq s$ for all $s\in[0,1]$, and
actually $(\fk kk)$ converges uniformly to the identity on $[0,1]$.
Therefore, $AT_k^{Id}\leq\Fk kk\leq AT_k^{\psi}$, with $\psi(s)=\chi_{(0,1]}(s)$
(cp. with \eqref{e:ATk} for the definition of $AT_k^\psi$),
and the result follows at once from Ambrosio and Tortorelli classical results
(cp. \cite{amb-tort2}, see also \cite{focardi}).
A similar argument works also in the regime $a_k\varepsilon_k^{1/2}\uparrow\infty$, in which
\[
\fk kk(s)=\begin{cases}
a_k\varepsilon_k^{1/2}s & 0\leq s\leq a_k^{-1}\varepsilon_k^{-1/2} \cr
1 & a_k^{-1}\varepsilon_k^{-1/2}\leq s\leq 1,
\end{cases}
\]
for $k$ sufficiently large, so that $\fk kk(s)\to\chi_{(0,1]}(s)$ for all $s\in[0,1]$, and
again we get the Mumford-Shah energy in the $\Gamma$-limit arguing as above.
Finally, note that for $a_k$ as in \eqref{e:alphaj}, we have
\[
\fk kk(s)=\begin{cases}
a_k\varepsilon_k^{1/2}s & 0\leq s\leq 1-\ella_k^{-1} \cr
\varepsilon_k^{1/2}\frac{\ell s}{1-s} & 1-\ella_k^{-1}\leq s\leq (1+\ell\varepsilon_k^{1/2})^{-1}\cr
1 & (1+\ell\varepsilon_k^{1/2})^{-1}\leq s\leq 1
\end{cases}
\]
so that $\fk kk(s)\to\chi_{\{1\}}(s)$ in $[0,1]$.
\end{remark}
\medskip
\subsection{A model with power-law growth at small openings}
\label{subsectpowerlaw}
In Theorem~\ref{t:sublin} below we approximate a model with sublinear surface density
in the origin and quadratic growth for the volume term. To this aim, let $p>1$ and
consider a function $\psi_p$ satisfying condition \eqref{f0} and
\be{\label{e:f1p}
\displaystyle\lim_{s\to1}(1-s)^p\psi_p(s)=\kappa, \quad \kappa\in(0,+\infty).
}
Clearly, one can take $\psi_p(s):=\frac{s}{(1-s)^p}$ as prototype.
The surface energy density $\vartheta_p:[0,+\infty)\to[0,+\infty)$ is defined as $g$ in \eqref{g}
by
\be{\label{e:tp}
\vartheta_p(s):=\inf_{(\alpha,\beta)\in\mathcal{U}_s}
\int_0^1|1-\beta|\sqrt{\psi_p^2(\beta)|\alpha'|^2+|\beta'|^2}\,dt,
}
where $\mathcal{U}_s$ has been introduced in \eqref{e:admfnctns}. In this case the integral is
finite only if $\beta<1$ almost everywhere on the set $\{\alpha'\ne0\}$.
We next prove some properties of $\vartheta_p$ in analogy to
Propositions~\ref{11}, \ref{p2} and \ref{p:gl}. In what follows, we keep the same notations
introduced there. We also note that given any curve $(\alpha,\beta)$, the integral to be minimized
in the definition of $\vartheta_p$ is invariant under reparametrizations of $(\alpha,\beta)$.
\begin{proposition}\label{p:tp}
Let $\psi_p$ satisfy \eqref{f0} and \eqref{e:f1p}, let $\vartheta_p:[0,+\infty)\to[0,+\infty)$ be
the corresponding surface energy in \eqref{e:tp}. Then,
\begin{itemize}
\item[(i)] $\vartheta_p(0)=0$, $\vartheta_p$ is nondecreasing, subadditive, and
\be{\label{e:tp4}
0\leq\vartheta_p(s)\leq 1\wedge c\, s^{\frac2{p+1}},\quad\textrm{for all $s\geq 0$},
}
where $c=c(\psi_p)>0$.
Moreover, $\vartheta_p\in C^{0,\frac 2{p+1}}\big([0,+\infty)\big)$ and
\begin{equation}\label{e:tp3}
\kappa^{\frac 2{p+1}}\le \lim_{s\downarrow 0}\frac{\vartheta_p(s)}{s^{\frac2{p+1}}}\le
\frac{p+1}{2^{\frac 2{p+1}}(p-1)^{\frac{p-1}{p+1}}}\,{\kappa}^{\frac{2}{p+1}};
\end{equation}
\item[(ii)] $\vartheta_p=\hat{\vartheta}_p$, where
\be{\label{e:tp2}
\hat{\vartheta}_p(s):=\lim_{T\uparrow\infty}\inf_{(\alpha,\beta)\in\mathcal{U}_s(0,T)}
\int_0^T\left(\psi_p^2(\beta)|\alpha^\prime|^2+\frac{(1-\beta)^2}{4}+|\beta^\prime|^2\right)dt;
}
\item[(iii)] the functions
\begin{equation}\label{e:fjPhitt}
f^{(j)}(s):=\frac{j\,s}{1-s}\wedge \psi_p(s),
\end{equation}
satisfy \eqref{f0} and \eqref{f1}. If $g_j$ denotes the corresponding surface energy in
\eqref{g}, then $g_j\leq g_{j+1}$ and
\be{\label{e:gjtp}
\lim_{j\to\infty}g_j(s)=\vartheta_p(s)\quad\textrm{for all $s\geq 0$}.
}
\end{itemize}
\end{proposition}
\begin{proof} We prove (i). The facts that $\vartheta_p(0)=0$ and that $\vartheta_p$ is nondecreasing
follow easily from the definition. The subadditivity follows as in
Proposition~\ref{11}\ref{e21}.
Moreover, $0\leq\vartheta_p\leq 1$ arguing as in \ref{e19} of Proposition~\ref{11}.
To show \eqref{e:tp4} and the upper bound in \eqref{e:tp3}, let $s,\,\lambda>0$ and consider
$\alpha:=0$ in $[0,1/3]$, $\alpha:=s$ in $[2/3,1]$ and set $\alpha$ to be the linear interpolation of
the values $0$ and $s$ on $[1/3,2/3]$; $\beta_\lambda:=1-(\lambda\,s)^{\frac 1{p+1}}$ in $[1/3,2/3]$
and set $\beta_\lambda$ to be the linear interpolation of that value to $1$ on $[0,1/3]\cup[2/3,1]$.
Then, clearly $(\alpha,\beta_\lambda)\in\mathcal{U}_s$ and a simple computation shows that
\begin{equation}\label{e:thetapub}
\vartheta_p(s)\leq\int_0^1|1-\beta_\lambda|\sqrt{\psi_p^2(\beta_\lambda)|\alpha'|^2
+|\beta_\lambda^{\prime}|^2}\,dt
=(\lambda\,s)^{\frac{1}{p+1}}\,\psi_p\big(1-(\lambda\,s)^{\frac 1{p+1}}\big)\,s
+(\lambda\,s)^{\frac 2{p+1}}.
\end{equation}
By taking $\lambda=1$, since $(1-t)^p\psi_p(t)\leq c$ for some constant $c=c(\psi_p)>0$ and
for all $t\in[0,1]$, we deduce that
\[
\vartheta_p(s)\leq (c+1)\,s^{\frac 2{p+1}},
\]
from which inequality \eqref{e:tp4} follows as $0\leq\vartheta_p\leq 1$.
Note that the H\"older continuity of $\vartheta_p$ then follows easily from \eqref{e:tp4} and
its subadditivity and monotonicity.
Further, by \eqref{e:thetapub} we infer
\[
\limsup_{s\downarrow 0}\frac{\vartheta_p(s)}{s^{\frac2{p+1}}}\le
\kappa\,\lambda^{-\frac{p-1}{p+1}}+\lambda^{\frac{2}{p+1}},
\]
minimizing the latter inequality over $\lambda\in(0,\infty)$ yields the upper bound in (\ref{e:tp3}).
\smallskip
We now prove the lower bound in \eqref{e:tp3}. Let $s_k\to 0$, $s_k>0$, and up to subsequences let the liminf in \eqref{e:tp3} be a limit. Let $\alpha_k,\beta_k$ be competitors for $\vartheta_p(s_k)$ such that
$$\int_{0}^{1} |1-\beta_k| \, \sqrt{\psi_p^2(\beta_k)|\alpha_k'|^2 + |\beta_k'|^2}\, dt\le \vartheta_p(s_k)+s_k.$$
If, after taking a subsequence, there is a sequence $x_j\in[0,1]$ such that
$$\frac{1-\beta_j(x_j)}{s_j^{\frac{1}{p+1}}}\ge\kappa^{1/(p+1)} \text{ for all $j$},$$
then
\be{\label{e:MM}\vartheta_p(s_j)+s_j\ge
(1-\beta_j(x_j))^2\ge\kappa^{2/(p+1)}\,s_j^{\frac{2}{p+1}}.}
Otherwise, for all $k$ large enough
$$\frac{1-\beta_k}{s_k^{\frac{1}{p+1}}}\leq \kappa^{1/(p+1)}$$
must hold uniformly, so that $\beta_k\to1$ uniformly and by (\ref{e:f1p}) for any $\varepsilon>0$
\begin{equation*}
(1-\beta_k)^p \psi_p(\beta_k) \ge \kappa-\varepsilon \text{ uniformly, for $k$ large enough.}
\end{equation*}
Therefore
\begin{equation}\label{e:the}
\vartheta_p(s_k)+s_k>\int_0^1\psi_p(\beta_k)(1-\beta_k)|\alpha'_k|dt\geq \int_0^1 \frac{\psi_p(
\beta_k)(1-\beta_k)^p}{(1-\beta_k)^{p-1}}|\alpha'_k|dt
\ge \frac{\kappa-\varepsilon}{\kappa^{(p-1)/(p+1)}} s_k^{2/(p+1)}\,.
\end{equation}
Since $\varepsilon$ was arbitrary this and (\ref{e:MM}) give
the lower bound in (\ref{e:tp3}).
Finally we prove that the limit in (\ref{e:tp3}) exists. We fix a sequence $s_j\downarrow 0$ and choose
$\alpha_j,\beta_j \in\mathcal{U}_{s_j}$ such that
\begin{equation*}
\int_0^1|1-\beta_j|\sqrt{\psi_p^2(\beta_j)|\alpha'_j|^2 +|\beta_j^{\prime}|^2}\,dt
\le \vartheta_p(s_j) + \frac1j s_j^{2/(p+1)}\,.
\end{equation*}
By the computation above we obtain $\beta_j\to1$ uniformly.
For $k\ge j$ we define $\alpha_k,\beta_k\in\mathcal{U}_{s_k}$ by
\begin{equation*}
\overline \alpha_k =\frac{s_k}{s_j}\alpha_j \text{ and }
\overline \beta_k =1- \Bigl(\frac{s_k}{s_j}\Bigr)^{1/(p+1)}(1-\beta_j)\,.
\end{equation*}
After a straightforward computation, using these test functions in
the definition of $\vartheta_p(s_k)$ leads to
\begin{equation*}
\vartheta_p(s_k)\le \Bigl(\frac{s_k}{s_j}\Bigr)^{2/(p+1)} \left[
\int_0^1|1-\beta_j|\sqrt{\psi_p^2(\beta_j)|\alpha'_j|^2 +|\beta_j^{\prime}|^2}\,dt\right]
\sup\bigl\{\frac{\psi_p(t)(1-t)^p}{\psi_p(t')(1-t')^p}: \min \beta_j\le t,t'<1\bigr\}\,.
\end{equation*}
Since $\beta_j\to1$ uniformly as $j\to\infty$, and $\psi_p(t)(1-t)^p$ has a finite limit as $t\to1$, the
$\sup$ converges to $1$ as $j\to\infty$. Therefore we obtain that for every $\varepsilon>0$ if $j$ is sufficiently large, then
\begin{equation*}
\frac{\vartheta_p(s_k)}{s_k^{2/(p+1)}} \le
(1+\varepsilon) \frac{\vartheta_p(s_j)}{s_j^{2/(p+1)}} +\frac1j\hskip5mm \text{ for all } k\ge j\,.
\end{equation*}
This implies that the sequence converges. Since the decreasing
sequence $s_j$ was arbitrary, the limit in (\ref{e:tp3}) exists.
\smallskip
To establish (ii), we note first that by Cauchy inequality $\vartheta_p\leq\hat{\vartheta}_p$.
For the sake of proving the converse inequality, we first claim that
$\alpha$ and $\beta$ in the infimum problem defining $\vartheta_p$ can be taken in $W^{1,\infty}\big((0,1)\big)$. Let $\eta>0$ small and let $\alpha,\beta\in H^1\big((1/3,2/3)\big)$
be competitors for $\vartheta_p(s)$ such that
\be{\label{e:quasimin2}\int_{1/3}^{2/3} |1-\beta| \, \sqrt{\psi_p^2(\beta)|\alpha'|^2 + |\beta'|^2}\, dt\le \vartheta_p(s)+\eta.}
We define $\beta^\eta(t):=\beta(t)\wedge (1-\eta)$ in $[1/3,2/3]$.
Since $(1-s)^p\psi_p(s)$ has a finite nonzero limit
at $1$, there is a function $\omega$, with $\omega(\eta)\to0$ as $\eta\to0$, such that
\begin{alignat}1\label{al:com}
(1-s')^p \psi_p(s') &\le (1+ \omega(\eta)) (1-s)^p \psi_p(s) \text{ for all } s,s'\in [1-\eta,1)\,.
\end{alignat}
In particular, if $1-\eta< \beta(t)< 1$, then
\begin{equation}\label{e:cont}
\eta \psi_p(1-\eta) \le \eta^{1-p} (1+\omega(\eta)) (1-\beta(t))^p \psi_p(\beta(t)) \le
(1+\omega(\eta)) (1-\beta(t)) \psi_p(\beta(t)) \,.
\end{equation}
We observe that $\beta^\eta=1-\eta$ and $(\beta^\eta)'=0$ almost everywhere on the set
$\{\beta\ne \beta^\eta\}$ and compute
\begin{alignat}1
\int_{\{\beta\ne \beta^\eta\}}
(1-\beta^\eta) \sqrt{\psi_p^2(\beta^\eta)\, |\alpha'|^2 + |(\beta^\eta)'|^2}dt &=
\int_{\{\beta\ne \beta^\eta\}}\eta \psi_p(1-\eta)\, |\alpha'| dt \nonumber\\
&\le (1+\omega(\eta))\int_{\{\beta\ne \beta^\eta\}}(1-\beta) \psi_p(\beta) |\alpha'|\,dt,\label{e:mini}
\end{alignat}
so that by \eqref{e:quasimin2} it follows
\be{\label{e:betaeta}\int_{1/3}^{2/3} |1-\beta^\eta| \, \sqrt{\psi_p^2(\beta^\eta)|\alpha'|^2 + |(\beta^\eta)'|^2}\, dt\le \vartheta_p(s)+\eta+\omega(\eta)+\eta\omega(\eta).}
By density we are able to find two sequences $\alpha_j,\beta_j^\eta\in W^{1,\infty}\big((1/3,2/3)\big)$ (actually in $C^\infty([1/3,2/3])$) such that
$\alpha_j(1/3)=0$, $\alpha_j(2/3)=s$, $\beta_j^\eta(1/3)=\beta_j^\eta(2/3)=1-\eta$, $0\leq\beta\leq 1-\eta$, and converging respectively
to $\alpha$ and $\beta^\eta$ in $H^1\big((1/3,2/3)\big)$.
Since the function $(1-s)^p\psi_p(s)$ is uniformly continuous in $[0,1-\eta]$ and since $\beta^\eta_j\to\beta^\eta$ also uniformly, we deduce that
for $j$ large it holds
$$\int_{1/3}^{2/3} |1-\beta_j^\eta| \, \sqrt{\psi_p^2(\beta_j^\eta)|\alpha_j'|^2 + |(\beta_j^\eta)'|^2}\, dt\le \vartheta_p(s)+2\eta+\omega(\eta)+\eta\omega(\eta).$$
Finally we extend $\alpha_j$ and $\beta_j^\eta$ in $[0,1]$ defining $\alpha_j:=0$ in $[0,1/3]$, $\alpha_j:=s$ in $[2/3,1]$, and $\beta_j^\eta$
as a linear interpolation of the values $1-\eta$ and $1$. Now $\alpha_j$ and $\beta_j^\eta$ are competitors for $\vartheta_p(s)$ and for $j$ large
they satisfy
$$\int_{0}^{1} |1-\beta_j^\eta| \, \sqrt{\psi_p^2(\beta_j^\eta)|\alpha_j'|^2 + |(\beta_j^\eta)'|^2}\, dt\le \vartheta_p(s)+2\eta+\omega(\eta)+\eta\omega(\eta)+\eta^2$$
and this concludes the proof of the claim.
Let us prove now that $\hat{\vartheta}_p(s)\leq\vartheta_p(s)$. We argue exactly as in Proposition~\ref{p2} until estimate
\eqref{al:p1}. In doing this we point out that $f$, $g$ and $\hat g$ have to be substituted by
$\psi_p$, $\vartheta_p$ and $\hat\vartheta_p$, respectively.
By keeping the same notation introduced there, we repeat the computations in \eqref{al:com}-\eqref{e:mini}
and we conclude that
\begin{alignat*}1
\hat \vartheta_p(s)
&\le \sqrt \eta + 3\eta^2
+(1+\omega(\eta)) \int_{0}^1
(1-\beta) \sqrt{\psi_p^2(\beta)\, |\alpha'|^2 + |\beta'|^2}dt \,.
\end{alignat*}
Since the last integral is less than $\vartheta_p(s)+\eta$ and $\eta$ can be made arbitrarily
small the inequality $\hat{\vartheta}_p\leq\vartheta_p$ follows at once.
\medskip
We finally prove (iii).
It is easy to check that $f^{(j)}\leq f^{(j+1)}$, and that $f^{(j)}(s)\to \psi_p(s)$ for all $s\in[0,1)$.
Hence, the sequence $(g_j)$ is nondecreasing and $g_j(s)\leq \vartheta_p(s)$ for all $s\geq 0$.
To prove \eqref{e:gjtp}, with fixed $s\in(0,+\infty)$, consider the functionals
$\mathscr{G}_j,\,\mathscr{G}_\infty:L^1\big((0,1)\big)\times L^1\big((0,1)\big)\to[0,+\infty]$ defined for
$(\alpha,\beta)\in\mathcal{U}_s$ by
\[
\mathscr{G}_j(\alpha,\beta):=
\int_0^1|1-\beta|\sqrt{(f^{(j)})^2(\beta)|\alpha^\prime|^2+|\beta^\prime|^2}\,dt
\]
and
\[
\mathscr{G}_\infty(\alpha,\beta):=
\int_0^1|1-\beta|\sqrt{\psi_p^2(\beta)|\alpha^\prime|^2+|\beta^\prime|^2}\,dt
\]
respectively, and set equal to $+\infty$ otherwise on $L^1\big((0,1)\big)\times L^1\big((0,1\big))$.
Note that $\mathscr{G}_j\leq\mathscr{G}_{j+1}$ and that $\mathscr{G}_j$ pointwise converge
to $\mathscr{G}_\infty$ by Beppo-Levi's theorem. Therefore, $(\mathscr{G}_j)$ $\Gamma$-converges
to $\overline{\mathscr{G}_\infty}$, the relaxation of $\mathscr{G}_\infty$ w.r.to the $L^1\times L^1$
topology.
Being $g_j(s)=\inf\mathscr{G}_j$ and $\vartheta_p(s)=\inf\mathscr{G}_\infty$ to conclude we
need only to discuss the compactness properties of the minimizing sequences of $\mathscr{G}_j$'s.
To this aim let $\alpha_j$, $\beta_j\in W^{1,\infty}\big((0,1)\big)$ be such that
$\alpha_j(0)=0$, $\alpha_j(1)=s$, $\beta_j(0)=\beta_j(1)=1$, and
\[
\mathscr{G}_j(\alpha_j,\beta_j)\leq g_j(s)+\frac1j.
\]
Hence, either there exists $\delta>0$ and a subsequence $j_k$ such
that $\inf_{[0,1]}\beta_{j_k}\geq\delta$, or $\inf_{[0,1]}\beta_j\to0$.
In the former case, we note that
\[
\big(\min_{t\in[\delta,1)}(1-t)f^{(j_k)}(t)\big)\|\alpha_{j_k}^\prime\|_{L^1}\leq g_{j_k}(s)+\frac1{j_k},
\]
and since for $k$ sufficiently large (depending on $\delta$)
\[
\min_{t\in[\delta,1)}(1-t)f^{(j_k)}(t)=\min_{t\in[\delta,1)}(1-t)\psi_p(t)>0\,,
\]
we conclude that $\sup_k\|\alpha_{j_k}^\prime\|_{L^1}<+\infty$. Moreover, as
\[
\frac12\sup_j\|\big((1-\beta_j)^2\big)^\prime\|_{L^1}\leq \sup_j g_j(s)\leq\vartheta_p(s),
\]
the sequence $(\alpha_{j_k},\beta_{j_k})$ is pre-compact in $L^1\times L^1$, so that
by standard properties of $\Gamma$-convergence we may conclude that
\begin{equation}\label{e:tpcasoa}
\lim_j g_j(s)=\lim_j\inf\mathscr{G}_j=\min\overline{\mathscr{G}_\infty}
=\inf\mathscr{G}_\infty=\vartheta_p(s).
\end{equation}
Instead, in case $\inf_{[0,1]}\beta_j\to0$, set $s_j=\mathrm{argmin}_{[0,1]}\beta_j$ and deduce that
\begin{equation}\label{e:tpcasob}
\lim_jg_j(s)\geq \liminf_j\left(\int_0^{s_j}|1-\beta_j||\beta_j^\prime|\,dt
+\int_{s_j}^1|1-\beta_j||\beta_j^\prime|\,dt\right)\geq1.
\end{equation}
Together with inequality $\vartheta_p(s)\leq 1$, the latter formula provides the conclusion.
\end{proof}
The functionals $F_k^{(k)}$ corresponding to the sequence $(f^{(j)})$ in \eqref{e:fjPhitt} of
Proposition~\ref{p:tp} provide an approximation of $\Phi_p:L^1(\Omega)\to[0,+\infty]$ defined by
\begin{equation}\label{e:Phitt}
\Phi_p(u):=
\begin{cases}
\displaystyle{\int_\Omega |\nabla u|^2dx+\int_{J_{u}}\vartheta_p(|[u]|)d{\mathcal H}^{n-1}}
& \textrm{if $u\in GSBV(\Omega)$,}\cr\cr
+\infty & \textrm{otherwise},
\end{cases}
\end{equation}
with $\vartheta_p$ is defined in formula \eqref{e:tp}.
\begin{theorem}\label{t:sublin}
Suppose that $(\f j)$ is as in \eqref{e:fjPhitt} above.
Then, the functionals $\Fk kk$ $\Gamma$-converge in $L^1(\Omega){\times}L^1(\Omega)$ to $\widetilde{\Phi}_p$, where
\begin{equation}\label{e:p}
\widetilde{\Phi}_p(u,v):=\begin{cases}
\Phi_p(u) & \textrm{if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$,}\cr
+\infty & \text{otherwise}.
\end{cases}
\end{equation}
\end{theorem}
\begin{proof}
By monotonicity of the sequence $(f^{(j)}_k)$ we have that $F_k^{(k)}\geq F_k^{(j)}$ for $k\geq j$,
so that by Theorem~\ref{t:gamma-lim} if $\Gamma\hbox{-}\liminf_kF_k^{(k)}(u,v)<+\infty$ then
$u\in GBV(\Omega)$, $v=1$ ${\mathcal L}^n$-q.o. on $\Omega$ and for all $j\in{\mathbb N}$
\[
\Gamma\hbox{-}\liminf_kF_k^{(k)}(u,1)\geq \Gamma\hbox{-}\lim_kF_k^{(j)}(u,1
=\int_\Omega h_j(|\nabla u|)dx+\int_{J_u}g_j(|[u]|)d{\mathcal H}^{n-1}+j|D^cu|(\Omega),
\]
where $h_j$ and $g_j$ are defined, respectively, by \eqref{h} and \eqref{g} with $f^{(j)}$ in place
of $f$. By letting $j\uparrow\infty$, we get that
\[
h_j(s)\uparrow s^2,\quad\text{ and}\quad g_j(s)\uparrow\vartheta_p(s)\quad \textrm{for all $s\geq 0$.}
\]
Indeed, the former convergence follows from the explicit formula $h_j(s)=s^2$ for $s\in[0,j/2]$ and
$h_j(s)=js-j^2/4$ for $s\in[j/2,+\infty)$, while the latter in view of (iii) in Proposition~\ref{p:tp}.
Therefore, by Beppo-Levi's theorem we conclude that $u\in GSBV(\Omega)$ with
\[
\Gamma\hbox{-}\liminf_kF_k^{(k)}(u,1)\geq\widetilde{\Phi}_p(u,1).
\]
To prove the upper bound inequality we note that Lemma~\ref{l:wsub} and \ref{l:boundglimsup}
still hold true in this setting as there we have only used that each function $f_k=1\wedge\varepsilon_k^{1/2} f$
in \eqref{fk} is nondecreasing and bounded by $1$ from above, properties enjoyed by $\fk kk$ as well
(cp. also Theorem~\ref{t:Dugdale}).
Hence, we may argue again as in Proposition~\ref{p:limsupndim} and reduce ourselves to prove
the estimate
\begin{equation}\label{e:minestp}
\limsup_{\delta\downarrow 0}\frac 1{\delta^{n-1}}\widehat{F}(u_x,1;x+\delta\, Q_{\nu_u(x)})\leq
\vartheta_p(|[u](x)|),
\end{equation}
for $u\in SBV^2(\Omega)$ and for ${\mathcal H}^{n-1}$-a.e. $x\in J_u$, where $\widehat{F}$ is the
$\bar{\Gamma}$-limit of a properly chosen subsequence $(\Fk{k_j}{k_j})$ of $(\Fk kk)$.
Given \eqref{e:minestp}, we deduce the upper bound estimate as follows: we employ first
\cite[Propositions~3.3-3.5]{bou-bra-but} to get the estimate
$\widehat{F}(\cdot,1)\leq \widetilde{\Phi}_p(\cdot,1)$
on the full $SBV$ space, by relaxing the functional $\Phi_\infty:BV(\Omega)\to[0,+\infty]$
\[
\Phi_\infty(u):=\begin{cases}
\displaystyle{
\int_\Omega |\nabla u|^2dx+\int_{J_u}\vartheta_p(|[u](x)|)\,d{\mathcal H}^{n-1}} & \textrm{if $u\in SBV^2(\Omega)$}
\cr
+\infty & \textrm{otherwise on $BV(\Omega)$,}
\end{cases}
\]
w.r.to the $w\ast\hbox{-}BV$ topology on $BV(\Omega)$. This implies $\widehat{F}(\cdot,1)\leq \Phi_p$
on $BV(\Omega)$. We get the required estimate on the whole $GSBV\cap L^1(\Omega)$ by the usual
truncation argument. We then argue as in Proposition~\ref{p:limsupndim} to show that the whole family
$(F_k^{(k)})$ $\Gamma\hbox{-}$converges to $\widetilde{\Phi}_p$.
The proof of
\eqref{e:minestp} is identical to the proof of \eqref{e:minest}
in Proposition~\ref{p:limsupndim} and therefore not repeated.
\end{proof}
\medskip
\subsection{Griffith's brittle fracture}
\label{subsecgriffith}
Finally, we show how to approximate the Mumford-Shah functional by means of any
sequence $\big(f^{(j)}\big)$ satisfying item (iii) in Proposition~\ref{p:gl}.
Thus, we recover the original approximation scheme of Ambrosio and Tortorelli
\cite{amb-tort1}, \cite{amb-tort2} (see also \cite{focardi}).
\begin{theorem}\label{t:MS}
Suppose that $(\f j)$ satisfies
$\f{j}\leq \f{j+1}$, $\ell_j\uparrow\infty$ and $\f{j}(s)\uparrow \infty$ pointwise in $(0,1).$
Then, the functionals $\Fk kk$ $\Gamma$-converge in $L^1(\Omega){\times}L^1(\Omega)$ to the functional $\widetilde{{M\!S}}$
defined as follows
\begin{equation}\label{e:Ft}
\widetilde{{M\!S}}(u,v):=\begin{cases}
{M\!S}(u) & \textrm{if $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$,}\cr
+\infty & \text{otherwise}.
\end{cases}
\end{equation}
\end{theorem}
\begin{proof}[Proof of Theorem~\ref{t:MS}]
We start off as in the proof of Theorem~\ref{t:Dugdale} and note that
by the very definitions $\Fk jk\leq \Fk kk$ for $j\leq k$ being $(\f j)$ nondecreasing
by assumption. Thus, by Theorem~\ref{t:gamma-lim} we deduce
\begin{equation}\label{e:liminfFj}
\Gamma\hbox{-}\liminf_k \Fk kk(u,v)\geq \F j(u,v),
\end{equation}
where $\F j$ is defined as $F$ in \eqref{F} with $f$ substitued by $\f j$ in formulas
\eqref{h} defining $h_j$, and \eqref{g} defining $g_j$. In particular, the corresponding
volume density is given by
\[
h_j(t)=\begin{cases}
t^2 & t\leq\frac{\ell_j}{2} \cr
\ell_j\,t-\frac{\ell_j^2}{4} & t\geq\frac{\ell_j}{2},
\end{cases}
\]
where $\ell_j$ is the value of the limit in \eqref{f1} and it satisfies $\ell_j\uparrow\infty$.
Thus $h_j(s)\leq s^2$ and $\lim_jh_j(s)=s^2$ for all $s\in[0,+\infty)$.
Moreover, the surface energy densities $g_j$ are dominated by the constant $1$,
and by item (iii) in Proposition~\ref{p:gl} we have $\lim_jg_j(s)=\chi_{(0,+\infty)}(s)$ for all
$s\in[0,+\infty)$.
In conclusion, if $\Gamma\hbox{-}\liminf_k\Fk kk(u,v)<+\infty$, by letting $j\uparrow\infty$ in
\eqref{e:liminfFj} we infer that $v=1$ $\mathcal{L}^n\text{-a.e.\ in } \Omega$, $u\in GSBV(\Omega)$ and by the Beppo-Levi's
theorem we get
\[
\Gamma\hbox{-}\liminf_k\Fk kk(u,v)\geq \widetilde{{M\!S}}(u).
\]
Eventually, we establish the limsup inequality. Set $\psi:=\chi_{(0,1]}$, we observe once more
that $\Fk kk\leq AT_k^\psi$ for every $k$, where $AT_k^\psi$ has been defined in \eqref{e:ATk}.
Therefore the conclusion follows by the Ambrosio and Tortorelli result \cite{amb-tort2} (see
also \cite{focardi}).
\end{proof}
\begin{remark}\label{r:regimes2}
In Remark~\ref{r:regimes} we have shown that both the divergence of the $f_k$'s and the scaling
with $\varepsilon_k^{1/2}$ in the definition of $f^{(k)}_k$ are influencing the asymptotic behavior of
the related sequence $(F_k^{(k)})$.
Here, we show that also the sequence of values of the limits in $1$ of the functions $(1-s)f^{(k)}(s)$
is playing a role.
In particular, we highlight that the pointwise limit of $(f_k^{(k)})$ is not determining
the asymptotics of $(F^{(k)}_k)$.
Indeed, suppose that $f^{(k)}(s):=a_k\frac{s}{1-s}$, where $a_k\uparrow\infty$, then
\[
f^{(k)}_k(s)=\begin{cases}
a_k\varepsilon_k^{1/2}\frac{s}{1-s} & 0\leq s\leq(1+a_k\varepsilon_k^{1/2})^{-1} \cr
1 & (1+a_k\varepsilon_k^{1/2})^{-1}\leq s\leq 1,
\end{cases}
\]
and by letting $k\uparrow\infty$ we infer that
\[
f^{(k)}_k(s)\to \begin{cases}
\chi_{\{1\}}(s) & \textrm{if $a_k\varepsilon_k^{1/2}\downarrow 0$} \cr
\gamma\,\frac{s}{1-s} \wedge 1 & \textrm{if $a_k\varepsilon_k^{1/2}\to\gamma\in(0,+\infty)$}\cr
\chi_{(0,1]}(s) & \textrm{if $a_k\varepsilon_k^{1/2}\uparrow \infty$}.
\end{cases}
\]
Hence, by taking also into account the examples in Remark~\ref{r:regimes}, we have
built two sequences of functions $f_k^{(k)}$ both converging to $\chi_{\{1\}}$ but giving
rise in the $\Gamma$-limit on one hand to the Dugdale's cohesive energy and on the other
hand to a Griffith's type energy.
\end{remark}
\section*{Acknowledgments}
Part of this work was conceived when M.~Focardi was visiting the University of Bonn in winter 2014.
He would like to thank the Institute for Applied Mathematics for the hospitality and the stimulating scientific
atmosphere provided during his stay.
M.~Focardi and F.~Iurlano are grateful to Gianni Dal Maso for stimulating discussions and for many
insightful remarks. They are members of the Gruppo Nazionale per
l'Analisi Matematica, la Probabilit\`a e le loro Applicazioni (GNAMPA)
of the Istituto Nazionale di Alta Matematica (INdAM).
F.~Iurlano was funded under a postdoctoral fellowship by the Hausdorff Center for Mathematics.
|
\section{Introduction}\label{sI}
In recent years there has been a great deal of interest in spectral singularities in non-Hermitian systems, especially in $\mathcal{PT}$(parity-time) symmetric systems \cite{muga,ali-prl-09,zafar-ss,ali-th,ali-wg,longhi-ss}. Such singularities with diverging scattered amplitudes (in reflection and transmission) were initially studied because of fundamental interest. A great deal of research has been devoted to understand the origin and the nature of these singularities and also the mechanism to regularize the infinities, since real systems cannot support such divergence. Indeed, no realistic system can ever support the infinite growth and such growing amplitudes will eventually render the response of the medium nonlinear. It was further understood that a dispersive Kerr nonlinearity is inadequate to regularize the singularity \cite{ali-kerr}. It was shown recently that regularization can be achieved by an all-order saturation mechanism in a balanced gain-loss system \cite{liu}. In addition to the fundamental interest in $\mathcal{PT}$-symmetric systems, there have been few attempts to find some basic applications in photonics \cite{ap1,ap2,ap3,ap4,ap5,ap6}. But any application would require a thorough control of the singularities. In this paper we show that such a control can be achieved if the gain and the loss media are kept in separate cavities with an overall feed-back provided by a third cavity. A similar ring cavity configuration with saturable absorbers was studied earlier in the context of demonstrating remarkable flexibility in the multi-stable response \cite{gsa}. Here we exploit the same flexibility albeit with a replacement of the saturable absorber with a saturable gain medium in one of the cavities. The balanced loss and gain cavities with identical parameters form the core of the $\mathcal{PT}$-symmetric system. We first investigate the linear response which clearly shows the singularity for a perfectly tuned system as reported by others \cite{ali-prl-09}. We show how $\mathcal{PT}$-symmetry is violated and these are regularized by introduction of the atomic and the cavity detunings leading to poles in scattered amplitudes away from the real axis. We then look at the nonlinear response which reiterates the earlier findings that saturable nonlinearity is able to limit the infinite growth. We further report on the bistable response and the non-reciprocity in transmission \cite{liu} which can lead to switching and optical isolation or diode action \cite{isolator}.
\begin{center}
\begin{figure}[h]
\includegraphics[width=8cm]{fig1.eps}
\caption{Schematic view of the $\mathcal{PT}$-symmetric ring cavity system with coupled cavities with balanced gain and loss medium.}%
\label{fig1}
\end{figure}
\end{center}
Consider the system shown in Fig.~\ref{fig1} comprising of two coupled ring cavities of length $L_1$ and $L_2$ and mirror transmission $T_1$ and $T_2$, respectively, enclosed in a third cavity with mirror transmission given by $T_3$. All the cavity mirrors are assumed to be lossless (i.e.,$T+R=1$, $R$- reflection coefficient). As mentioned earlier, optical multi-stability was investigated in a similar system with both the cavities containing lossy media \cite{gsa}. Unlike the general cavity configuration as in Fig.~\ref{fig1}, we consider a special case comprising of two identical coupled ring cavities of equal length ($L_1=L_2=L$) and mirror transmission ($T_1=T_2=T$). In order to realize $\mathcal{PT}$ symmetry, we further assume that the internal cavities are balanced with the same gain ($\alpha_1<0$) and loss ($\alpha_2>0$) coefficient ie., $|\alpha_1|=\alpha_2=\alpha$. This symmetry additionally implies that $\delta_1=\delta_2=\delta$ and $\theta_1=\theta_2=\theta$, where $\delta$ and $\theta$ give the atomic and the cavity detunings, respectively. Note that the exterior cavity can also be detuned from its resonance leading to a finite $\theta_3$. For such a case the input $y$ vs. output $x$ relation is given by
\begin{eqnarray}
y & = & x\left \lbrace \vphantom{\frac{ [ \tau + \zeta \sqrt{T_3/T} ] } { 1 + \delta^2 + \left| x [ \tau + \zeta \sqrt{T_3/T} ] \right|^2 } } \sqrt{\frac{T}{T_3}} \left[ \tau^2 -R_3 (1-i T_3 \theta_3) \right]+ \zeta \tau \right. \nonumber \\ \label{eq1}
& - & \left.\frac{\zeta (1+\delta^2+|x|^2)\left[ \tau + \zeta \sqrt{T_3/T} \right] } { 1 + \delta^2 + \left| x \left[ \tau + \zeta \sqrt{T_3/T} \right] \right|^2 } \right \rbrace, \\ \hspace{-2cm} \text{with} \nonumber \\
\zeta & = &\frac{\alpha L (1-i\delta)}{\sqrt{TT_3} (1+\delta^2+|x|^2)},~~~ \tau=1+iR\theta \label{eq2}.
\end{eqnarray}
The parameters in Eq.(\ref{eq1}) are given by \cite{lugiato}
\begin{eqnarray}
\label{eq3}
x & = & \frac{\mu_2}{\hbar} \frac{E_t}{\sqrt{TT_3\gamma_{\|}\gamma_{\bot}}}, ~~~~~~~~y= \frac{\mu_2}{\hbar} \frac{E_i}{\sqrt{\gamma_{\|}\gamma_{\bot}}} ,\\ \label{eq4}
\delta_i & = & \omega_i-\omega_0 ,~~\theta_i=(\omega_{ci}-\omega_0)/\kappa, ~~ \kappa=cT/L.
\end{eqnarray}
$\omega_i$ is the atomic transition frequency in the $i^{th}$ cavity, $\omega_0$ and $\omega_{ci}$ are the incident light and the $i^{th}$ cavity frequency. $R_i$ and $T_i$ give the intensity reflection and transmission coefficients of the lossless mirrors of the $i^{th}$ cavity. It is clear from Eq.~(\ref{eq1}) that in the linear regime ($|x|\rightarrow 0$) the occurrence of the the spectral singularity requires the vanishing of the expression in the curly brackets. This implies, for example, for perfectly tuned system the following simple relation for $\alpha L$
\begin{equation}
\label{eq5}
\alpha L=\sqrt{T_3T^2}.
\end{equation}
It is evident from Eq.~(\ref{eq5}) that the better the cavities the easier it is to obtain the spectral singularity at lower gain levels or smaller systems. In what follows we present a detailed study of the linear response of the system and show the occurrence of the singularity and its dependence on the cavity parameters.
We consider three specific cases:
\subsection*{Effect of the exterior cavity, no internal cavities}
We assume the mirrors of the internal cavities to be completely transparent ($T=1$) and we look at the influence of the exterior cavity. The results for $\log_{10}(|x/y|)$ as a function of $\alpha L$ for three specific cases, namely, (a) $\delta=0.0$, $\theta_3=0.0$ (b) $\delta=0.0$, $\theta_3=0.05$, (c) $\delta=0.001$, $\theta_3=0$ are shown in Figs.~\ref{fig2}(a),~\ref{fig2}(b) and \ref{fig2}(c), respectively. Curves from right to left are for $T_3=1.0,~0.49,~0.09$.
\begin{figure}[h]
\center{\includegraphics[width=8cm]{fig2.eps}}
\caption{Effect of the external cavity on the spectral singularity in absence of internal cavities ($T=1.0)$: (a) $\delta=0.0$, $\theta_3=0.0$ (b) $\delta=0.0$, $\theta_3=0.05$, (c) $\delta=0.001$, $\theta_3=0$. Curves from right to left are for $T_3=1.0,~0.49,~0.09$.}
\label{fig2}
\end{figure}
As can be seen from the solid curve in Fig.~\ref{fig2}(a) that the absence of the third cavity leads to the spectral singularity at $\alpha L=\sqrt{T_3}=1$. A decrease in $T_3$ leads to the same at lower values of $\alpha L$ and the singularity (being an inherent property of the $\mathcal{PT}$-symmetric component) survives. It is not surprising that a finite value of the detuning of the exterior cavity introduces the necessary dispersion to move the pole away from the real axis and the singularity is regularized (see Fig.~\ref{fig2}(b)). Recall that the true $\mathcal{PT}$-symmetry holds only for null detuning ($\delta=0$). A finite atomic detuning can kill the singularity even in a perfectly tuned exterior cavity (Fig.~\ref{fig2}(c)).
\subsection*{Effect of internal cavities in absence of the exterior cavity}
Results for the transmission in absence of the exterior cavity ($T_3=1$) are shown in Fig.~\ref{fig3} for (a) $\delta=0,~\theta=0,$ (b) $\delta=0,~\theta=0.05$ and (c) $\delta=0.001,~\theta=0$. The curves from right to left in each panel are for $T=1.0,~0.5,~0.01$, respectively. As before, the perfectly tuned system displays the singularities at $\alpha L=T$ (Fig.~\ref{fig3}(a)).
\begin{figure}[h]
\center{\includegraphics[width=8cm]{fig3.eps}}
\caption{Effect of the internal cavities on the spectral singularity in absence of the external cavity ($T_3=1.0)$: (a) $\delta=0.0$, $\theta=0.0$ (b) $\delta=0.0$, $\theta=0.05$, (c) $\delta=0.001$, $\theta=0$. Curves from right to left are for $T=1.0,~0.5,~0.1$.}
\label{fig3}
\end{figure}
A comparison of Fig.~\ref{fig2} and Fig.~\ref{fig3} reveals that the internal cavities play a more dominant role because of $T$ scaling of the location of the singularity as compared to the $\sqrt{T}$ dependence in the previous case. It is also clear that atomic detuning plays the more significant role in limiting the infinite growth. A word of caution regarding the parameter values used in Fig.~\ref{fig2}(b) and Fig.~\ref{fig3}(b) would be in place since finite cavity detuning can be arbitrary for unity cavity transmission (cavity not present).
\begin{figure}
\center{\includegraphics[width=8cm]{fig4.eps}}
\caption{Removal of the spectral singularity in detuned systems for $T=T_3=0.01$ for (a) $\delta= 0.0,~ 0.01,~0.05$ with $\theta_3=\theta=0.0$, (b) $\theta_3= 0.0,~0.05,~0.1$ with $\theta=\delta=0.0$, and (c) $\theta= 0.0,~0.0001,~0.0005$ with $\theta_3=\delta=0.0$.}
\label{fig4}
\end{figure}
\begin{figure}[b]
\center{\includegraphics[width=6cm]{fig5.eps}}
\caption{Regularization of the spectral singularity by nonlinearity in a cavity configuration ($T_3=T=0.01$,~ $\theta_3=0.005$ and $\delta=\theta=0$): $\log_{10}(|x/y|)$ as a function of $\alpha L$ for different values of $|y|$, namely, $|y|=10^{-6},~10^{-4}~,10^{-3}$. Curves from right to left are for increasing values of $|y|$.}
\label{fig5}
\end{figure}
\subsection*{All three cavities present}
Consider the case when all the three cavities are present, say, with $T_3=T=0.01$. The results for this case are shown in Fig.~\ref{fig4} for different combinations of the cavity and atomic detuning parameters. Fig.~\ref{fig4}(a) shows how increasing atomic detuning can move the system away from the singularity with other detunings $\theta=\theta_3=0$. Similar behavior is displayed in the Figs.~\ref{fig4}(b) and \ref{fig4}(c) when $\theta_3$ and $\theta,$ are changed, respectively. It is evident from Figs.~\ref{fig4}(b) and \ref{fig4}(c) that the spectral singularity can be regularized with much lower values of $\theta$ than $\theta_3$. Fig.~\ref{fig4} thus demonstrates clearly the role of detuning parameters in regularizing the singularity in linear systems.
\begin{figure}[h]
\center{\includegraphics[width=8cm]{fig6.eps}}
\caption{Regularization due to atomic detuning: $\log_{10}(|x/y|)$ as a function of $\log_{10}(|y|)$ (a) in a cavityless system ($T_3=T=1.0$) with $\alpha L=0.75$ for $\delta=0.0,~0.05,~0.2$ and (b) in presence of all the cavities $T_3=T=0.01$ with $\theta_3=\theta=0.0$ and $\alpha L=0.0002$ for $\delta=0.0,~0.15,~0.5$. Curves labeled $1-3$ are for the increasing values of $\delta$. In Fig.~\ref{fig6}(b) the point marked by a cross on curve 2 corresponds to very high transmission at low input, and the thick solid curve shows the diode action corresponding to curve $2$ in Fig.~\ref{fig6}(b).}
\label{fig6}
\end{figure}
\begin{figure}[t]
\center{\includegraphics[width=8cm]{fig7.eps}}
\caption{Effect of cavity detuning on spectral singularity in presence of all the cavities $T=T_3=0.01$ with $\delta=0.0,$ and $\alpha L=0.0002$ for (a) $\theta=0$, $\theta_3=0.0,~0.2,~0.5$ and (b)$\theta_3=0$, $\theta=0.0,~0.001,~0.003$. Curves labeled by $1-3$ are for increasing values of parameter of interest in each panel.}
\label{fig7}
\end{figure}
\subsection*{Nonlinear response}
Till now we have analyzed the linear properties of the system when $|x|\rightarrow 0.$ We now look at the system response at larger power levels. As mentioned earlier, nonlinearity starts to play an important role and a saturation type nonlinearity as in our system can limit the infinite growth of the amplitudes. We show this regularization for larger values of $|y|$ in Fig.~\ref{fig5}. For a given input the input-output relation (Eq.~(\ref{eq1})) can yield multiple roots as in many other nonlinear systems \cite{sdgreview}. In order to avoid the infinite build-up near the singularity at low power levels, we consider a system slightly detuned from the spectral singularity and investigate the effects of nonlinearity on it. The system parameters were chosen as follows: $T_3=T=0.01$, $\theta_3=0.005$ with $\delta=\theta=0$. The transmission as a function $\alpha L$ for three different values of $|y|$ is shown in Fig.~\ref{fig5}. It is clear from Fig.~\ref{fig5} that nonlinearity can introduce dramatic changes to the response leading to different possible bistable scenarios. For example, for $|y|=10^{-4}$ the resonance is bent towards left, while further increase in $|y|$ leads to an upward twist of the resonance. Effectively one now has the peak in transmission at a much lower value of $\alpha L$ at ($\alpha L=9.942\times 10^{-5}$). Thus cavity-designable nonlinear response can be handy for novel kind of switching phenomena in optical circuitry.
\par
The nonlinear response of the system is extremely sensitive to the detunings used. We have plotted transmission as a function incident field (both in $\log_{10}$ scale) in Fig.~\ref{fig6}(a) and Fig.~\ref{fig6}(b), where the two panels correspond to the case with no ($T=T_3=1$) and all cavities ($T=T_3=0.01$). In each panel we have presented the results for three distinct values of $\delta$. It is clear from figures that the ring cavity configuration can lead to bistable response for much lower values of $\alpha L$ ($=2.0\times10^{-4}$) as compared to $\alpha L$ (=0.75) for the cavityless situation. Moreover, there can be a multi-stable response (see curve 2) for suitable choice of $\delta$. The regularization feature for increasing values of $\delta$ can be easily seen from both the panels.%
\par%
Analogous response is also seen in transmission when cavity detuning is used as a parameter. This is shown in Fig.~\ref{fig7}, albeit with no multi-stable response. Fig.~\ref{fig7}(a) (Fig.~\ref{fig7}(b)) shows the transmission for three distinct detunings of the external (internal) cavity. As shown in Fig.~\ref{fig7} the diverging transmission can be regularized with much lower detuning values of the internal cavities than the external cavity. Note that this is consistent with the linear results shown in Figs.~\ref{fig4}(b) and \ref{fig4}(c). The control over the location and the height of the bent peaks in Fig.~\ref{fig6} and Fig.~\ref{fig7} can offer a true handle on the optical switching action and its contrast.
\par
Last but not the least, there have been ample discussions on nonreciprocity in linear and nonlinear systems. Nonreciprocity in reflection is a well known fact, while that in transmission is missing in linear and also in some Kerr nonlinear systems \cite{zafar-nrecp,sdg-nrecp,np-nrecp}. Nonreciprocity in transmission leading to optical isolation (diode action) was reported recently \cite{liu}. We now show that analogous nonreciprocity is exhibited by our system when the light is incident from the right or from the left. In the context of Eq.~(\ref{eq1}) it means that now $\alpha_1>0,$ and $\alpha_2<0,$ or the loss and the gain media are interchanged. The response for such a case is shown in Fig.~\ref{fig6}(b) by the thick solid line. It is clear that, the response is completely different for illumination from the left and the right. Evidently this can find many applications in optical circuitry. Another interesting point suggested by the nonlinear response, of say, Fig.~\ref{fig6}(b) is the following. Very large transmission can be realized at low powers. For example, the point marked by a cross on curve 2 of Fig.~\ref{fig6}(b) refers to such a situation when $|x/y|=6.4\times 10^{3}$ and $|y|=1.4\times 10^{-3}$. Of course one can achieve such a situation only after crossing the threshold (at $|y|=2.8 \times 10^{-3}$). The presence of cavity lowers the threshold value in comparison to that of cavityless system, for example, for curve 2 in Fig.~\ref{fig6}, the threshold value in presence of cavity is $|y|=2.8 \times 10^{-3}$, whereas for cavityless system $|y|=0.144$.\par
In conclusion, we have investigated a balanced gain-loss ring cavity system to show the presence of a singularity, which can be efficiently controlled by the cavity parameters and the atom-laser detuning. We compare various cases with or without the cavities to demonstrate this control on both linear and nonlinear response. We identified the atomic and cavity detuning parameters to be capable of removing the singularities. The regularization is also shown to be possible by changing the incident laser power, leading to skewed resonances and, in general multi-stability. Further, we demonstrated how to use it for the purpose of a controlled switch and optical diode action. Our results may find interesting applications in photonics devices.
\par
One of the authors (SDG) is thankful to Girish S Agarwal for fruitful discussions. KNR is thankful to Department of Science and Technology for an INSPIRE Fellowship.
|
\section{Dimensional Reduction of Brans-Dicke Theory in Arbitrary Dimensions}
\label{sec:1}
The original motivation of the induced-matter
theory (IMT)~\cite{ponce-wesson-92}
was to achieve the unification of matter
and geometry. Recently,
the idea of the IMT has been employed for generalizing
BD theory, as a fundamental underlying theory, in
four~\cite{Ponce2-10}
and arbitrary dimensions~\cite{RFMinprogress}.
In the following, we present only a brief review of the latter.
The variation of the $(N+1)D$ BD action in vacuum with
respect to metric and BD scalar field, $\phi$, give the equations
\begin{equation}\label{(D+1)-equation-1}
G^{^{(N+1)}}_{ab}=\frac{\omega}{\phi^{2}}
\left[(\nabla_a\phi)(\nabla_b\phi)-\frac{1}{2}\gamma_{ab}(\nabla^c\phi)(\nabla_c\phi)\right]
+\frac{1}{\phi}\Big(\nabla_a\nabla_b\phi-\gamma_{ab}\nabla^2\phi\Big),
\end{equation}
\begin{equation}\label{(D+1)-equation-2}
\nabla^2\phi=0,
\end{equation}
where the Latin indices run
from $0$ to $N$; $\gamma_{ab}$ is the metric associated to the $(N+1)D$ space-time,
$\nabla^2\equiv\nabla_a\nabla^a$ and $\omega$ is a
dimensionless parameter.
Here, we have chosen $c=1$.
In the following, we only employ the equations of the MBDT
in arbitrary dimensions, which convey relations between the $(N+1)D$ field
equations to the corresponding ones with sources in $ND$ space-time in the
context of BD theory~\cite{RFMinprogress}.
We can find the reduced field equations on the $ND$
hypersurface by employing the BD field Eqs.~(\ref{(D+1)-equation-1}), (\ref{(D+1)-equation-2})
and a $(N+1)D$ space-time with a line element
\begin{equation}\label{global-metric}
dS^{2}=\gamma_{ab}(x^c)dx^{a}dx^{b}=
g_{\mu\nu}(x^\alpha,l)dx^{\mu}dx^{\nu}+
\epsilon\psi^2\left(x^\alpha,l\right)dl^{2}\,,
\end{equation}
where the Greek indices run from zero to $(N-1)$, $l$ is a
non-compact coordinate associated to $(N + 1)$th dimension,
the parameter $\epsilon=\pm1$ allows we to choose the
extra dimension to be either time-like or space-like, and $\psi$
is the another scalar field taken as a function of all the coordinates.
Here, we only present some of the reduced equations on
the $ND$ hypersurface. These equations will be described in
two separated parts with a short interpretation.
\begin{enumerate}
\item We
can construct the Einstein tensor on the hypersurface as
\begin{eqnarray}\label{BD-Eq-DD}
G_{\mu\nu}^{^{(N)}}&=&\frac{8\pi}{\phi}T_{\mu\nu}^{^{(\rm BD)}}+
\frac{\omega}{\phi^2}\left[({\cal D}_\mu\phi)({\cal D}_\nu\phi)-
\frac{1}{2}g_{\mu\nu}({\cal D}_\alpha\phi)({\cal D}^\alpha\phi)\right]\cr
&+&\frac{1}{\phi}\left[{\cal D}_\mu{\cal D}_\nu\phi-
g_{\mu\nu}{\cal D}^2\phi\right]-g_{\mu\nu}\frac{V(\phi)}{2\phi},
\end{eqnarray}
where ${\cal D}_\alpha$ is the covariant derivative on $ND$ hypersurface,
which is calculated with $g_{\alpha\beta}$ , and ${\cal D}^2\equiv{\cal D}^\alpha{\cal D}_\alpha$.
The above equations correspond to the BD equations, obtained from
the standard BD action containing a scalar potential, but here
there are some differences which we clarify them in the following
\begin{itemize}
\item
The quantity introduced by $V(\phi)$ is actually the effective induced scalar potential
on the hypersurface which will be determined by a relation in part 2.
\item
The quantity $T_{\mu\nu}^{^{(\rm BD)}}$, can be interpreted as
an induced energy-momentum tensor (EMT)
for a BD theory in $N$-dimensions and it, in turn, contains three components as
\begin{eqnarray}\label{matt.def}
T_{\mu\nu}^{^{(\rm BD)}}\equiv T_{\mu\nu}^{^{(\rm I)}}+T_{\mu\nu}^{^{(\rm \phi)}}
+\frac{1}{16\pi}g_{\mu\nu}V(\phi),
\end{eqnarray}
where $8\pi/\phi T_{\mu\nu}^{^{(\rm I)}}$ is the same as the induced EMT appearing in IMT, while
\begin{eqnarray}\label{T-phi}
\frac{8\pi}{\phi}T_{\mu\nu}^{^{(\rm \phi)}}\equiv
\frac{\epsilon\phi_{_{,N}}}{2\psi^2\phi}\left[g_{\mu\nu,}{}_{_{N}}
+g_{\mu\nu}\left(\frac{\omega\phi_{_{,N}}}{\phi}-g^{\alpha\beta}g_{\alpha\beta,}
{}_{_{N}}\right)\right],
\end{eqnarray}
where $A{_{,N}}$ is the partial derivative of the quantity $A$ with respect to
$l$.
\end{itemize}
\item
The wave equation on the hypersurface is given by
\begin{eqnarray}\label{D2-phi}
{\cal D}^2\phi=\frac{8\pi T^{^{(\rm BD)}}}{(N-2)\omega+(N-1)}+
\frac{1}{(N-2)\omega+(N-1)}\left[\phi\frac{dV(\phi)}{d\phi}-\frac{N}{2}V(\phi)\right],
\end{eqnarray}
where
\begin{eqnarray}\label{v-def}
&\phi&\frac{dV(\phi)}{d\phi}\equiv-(N-2)(\omega+1)
\left[\frac{({\cal D}_\alpha\psi)({\cal D}^\alpha\phi)}{\psi}
+\frac{\epsilon}{\psi^2}\left(\phi_{_{,NN}}-
\frac{\psi_{_{,N}}\phi_{_{,N}}}{\psi}\right)\right]\\\nonumber
&-&\!\!\!\frac{(N-2)\epsilon\omega\phi_{_{,N}}}{2\psi^2}
\left[\frac{\phi_{_{,N}}}{\phi}+g^{\mu\nu}g_{\mu\nu,}{}_{_{N}}\right]
+\frac{(N-2)\epsilon\phi}{8\psi^2}
\left[g^{\alpha\beta}{}_{_{,N}}g_{\alpha\beta,}{}_{_{N}}
+(g^{\alpha\beta}g_{\alpha\beta,}{}_{_{N}})^2\right].
\end{eqnarray}
\end{enumerate}
Actually in this approach, the $(N+1)D$ field equations
(\ref{(D+1)-equation-1}) and (\ref{(D+1)-equation-2})
split naturally into four sets of equations on every $ND$ hypersurface,
in which we only have introduced the two sets (\ref{BD-Eq-DD}) and (\ref{D2-phi}).
Regarding the geometrical
interpretation of the other two sets, we will not discuss
them and leave them for next paper in this series.
In the following, we investigate the Kasner solution in BD theory in a $5D$ space-time;
then, as an application of the MBDT in cosmology, we present
the properties of a the reduced cosmology on the hypersurface.
\section{Kasner Solution in Brans-Dicke Theory and its Corresponding Reduced Cosmology}
\label{K. Solution}
We start with the generalized Bianchi type~I anisotropic model
in a $5D$ space-time as
\begin{equation}\label{bulk-metric}
dS^{2}=-dt^{2}+\sum^{3}_{i=1}a_i^{2}(t, l)dx_i^{2}+h^{2}(t, l)dl^{2}\,,
\end{equation}
where $t$ is the cosmic time, $(x_1, x_2, x_3)$ are the Cartesian
coordinates, $l$ is the non-compactified extra dimension, and $a_i(t, l)$, $h(t, l)$ are
different cosmological scale factors in each of the four
directions. We assume that there is no matter in $5D$
space-time and $\phi=\phi(t, l)$. In addition,
based on the usual spatial homogeneity, we solve the field
equations (\ref{(D+1)-equation-1}) and (\ref{(D+1)-equation-2})
by assuming separation of
variables as
\begin{equation}\label{sep.eq}
\phi(t,l)=\phi_0 t^{p_{0}}l^{s_{0}},\hspace{6mm}
a_i(t,l)\propto t^{p_{i}}l^{s_{i}},\hspace{6mm} h(t,l)=h_o
t^{p_{4}}l^{s_{4}},\hspace{6mm}
\end{equation}
where $h_0$ and $\phi_0$ are constants, and the $p_{_a}$'s and
$s_{_a}$'s ($a=0,2,3,4$ ) are parameters satisfying field equations. By replacing the {\it
ansatz} (\ref{sep.eq}) into Eq.~(\ref{(D+1)-equation-2}) and using~(\ref{bulk-metric}), we get
five classes of solutions.
In the following, we are interested to investigate just
the solutions that leads to a generalized Kasner relations in five
dimensions. Also, we then apply the MBDT to obtain the corresponding
reduced cosmology on a $4D$ hypersurface.
In order to have consistency and ignoring the trivial solutions we set $p_4\neq 1$ and
$s_4\neq -1$. Also for simplicity, we assume $h_o=1$ in {\it ansatz} (\ref{sep.eq}).
Hence after a little manipulation, we can obtain the following relations among the
generalized Kasner parameters~\cite{RFS11}
\begin{eqnarray}\label{kas-rel1}
\sum^{4}_{a=0}p_{_a}&=&1, \hspace{5mm}
(\omega+1)p_{0}^{2}+\sum^{4}_{m=1}p_m^{2}=1,\hspace{5mm}
\sum^{3}_{\mu=0}s_{\mu}=1+s_{4},\\\nonumber
(\omega+1)s_{0}^{2}&+&\sum^{3}_{i=1}s_i^{2}=(1+s_{4})^{2},\hspace{5mm}
(\omega+1)p_{0}s_{0}+\sum^{3}_{i=1}p_is_i=p_{4}(1+s_{4})\,.
\end{eqnarray}
Eqs.~(\ref{kas-rel1}) lead to a few constrains, so that there are only five
independent relations among the
Kasner parameters, designated as the generalized
Kasner relations in $5D$ BD theory.
In what follows, for the sake of brevity,
we would like to present a brief review of the
results of the reduced Kasner cosmology on the $4D$ hypersurface~\cite{RFS11}:
the pressure and energy density of the specified induced matter on any $4D$
hypersurface can be derived from (\ref{matt.def}).
These results show that, in general, we cannot consider it as a perfect fluid.
By applying (\ref{v-def}), (\ref{sep.eq}) and (\ref{kas-rel1}), the induced
scalar potential is obtained to be either in the power law or in the logarithmic form, in which we only
investigate the properties of the former.
The properties of a few cosmological quantities as well as
physical quantities such as the average scale factor, the mean
Hubble parameter, the expansion scalar, the shear scalar and the
deceleration parameter have been studied. First,
these quantities have been derived in terms of the
generalized Kasner parameters. Then, we find that
the induced EMT satisfies the barotropic equation of state, where the equation
of state parameter, $w$, is a function
of the Kasner parameters. And thus, the evolution of all the
quantities has been represented with respect to $w$, $\omega$ and the deceleration parameter, $q$.
We then probe the quantities, in the general case, versus $q$, $t$ and $\omega$ for the stiff fluid and the
radiation-dominated universe. We have shown that, for both of the fluids, there is an expanding
universe commenced with a big bang, and there is a horizon for each of them. Also, we have shown that the rate of
expansion slows down by time. By employing the weak energy condition, the allowed (or the well-behaved)
ranges of the deceleration and the BD coupling parameters have been obtained for each of
the fluids. The behavior of the quantities,
in the very early universe and the very large time show that the models yield
empty universes when the cosmic time tends to infinity.
However, both of the models, in general, do not approach isotropy for large values
of the cosmic time.
\begin{acknowledgement}
I sincerely appreciate to Prof. Paulo Vargas Moniz for a critical reading of the letter.
I am supported by the Portuguese Agency Funda\c{c}\~{a}o para a
Ci\^{e}ncia e Tecnologia through the fellowship SFRH/BPD/82479/2011.
\end{acknowledgement}
\input{referenc}
\end{document}
|
\section{Introduction}
The problem of neutrino propagation and interactions in dense external environments
is one of the most important issues
of the present-day particle physics. Neutrinos play an important role in very complex and physically diverse processes as of core-collapse supernovae that provides fascinating playground for new or exotic phenomena.
A detailed review on supernova mechanisms is given in \cite{Bethe:1990mw}, a discussion on some of the progress that has been made recently in the simulations of stellar core-collapse and supernova explosions can be also found in \cite{Janka:2006fh}.
The physical processes that lead from stellar core-collapse to supernova explosion provide
an extremely important and interesting astrophysical application of the discussed problem of neutrino interactions in external environments.
Under the influence of the extreme external conditions, such as dense matter and strong magnetic fields, neutrinos can manifest their yet hidden fundamental properties that might open a window to new physics.
Recently studies of neutrino electromagnetic properties (see \cite{Giunti:2014ixa, Broggini:2012df,Giunti:2008ve} for a review) have revealed a new mechanism of electromagnetic radiation that can be emitted by
a neutrino propagating in dense matter termed the spin
light of neutrino in matter ($SL\nu$) \cite{Lobanov:2002ur}. At first the $SL\nu$ was investigated within the quasi-classical treatment (see also \cite{Lobanov:2004um,Grigoriev:2004bm}) based on use of the generalized Bargmann-Michel-Telegdi equation for the neutrino spin evolution in the background environment \cite{Lobanov:2001ar,Dvornikov:2002rs}.
The quantum theory of the $SL\nu$ was first revealed in our studies
\cite{Studenikin:2004dx,Grigorev:2005sw} ( see also \cite{Lobanov:2005zn}) within implication of the so called ``method of wave equations exact solutions" that implies use of exact solutions of modified Dirac equations that contain the corresponding effective potentials accounting for the matter influence on neutrinos \cite{Studenikin:2004dx,Grigorev:2005sw,Studenikin:2005bq,Studenikin:2007zza,Studenikin:2008qk,Balantsev:2010zw,
Balantsev:2013aya,Studenikin:2012vi}.
Different aspects related to the proposed the $SL\nu$ have been discussed and investigated recently. For instance, the importance of the plasma influence on the proposed new mechanism of electromagnetic radiation as well as corrections to the effective matter density due to nonlocality of neutrino interaction with particles of the environment have been considered in the subsequent series of papers dedicated to the $SL\nu$ (see \cite{Kuznetsov:2007ar,Grigoriev:2012pw,Kuznetsov:2013tta}). The $SL\nu$ mechanism was also considered for the case when in addition to matter a gravitational field is also present \cite{Grigoriev:2004bm}, or the model describing neutrino interactions with the environment permits of the Lorentz and CPT invariance violation \cite{Kruglov:2013oia}.
The main properties of the $SL\nu$ are summarized \cite{Studenikin:2008qk}
in the following way: 1) a neutrino with nonzero magnetic moment when
moving in dense matter can emit electromagnetic waves; 2) the
$SL\nu$ radiation rate and power depend on the neutrino magnetic
moment and energy, and also on the matter density;
3) for a wide range of matter densities the radiation
is beamed along the neutrino momentum;
4) the
emitted photon energy is also essentially dependent on the
neutrino energy and matter density;
5) in the most
interesting for possible astrophysical and cosmology applications
case of ultra-high energy neutrinos, the average energy of the
$SL\nu$ photons equals a reasonable part of the initial neutrino energy so that the $SL\nu$ spectrum can span up to gamma-rays.
In spite of the listed above notable properties of the $SL\nu$ its possible role and impact in astrophysical processes is constraint due to the fact that the rate of the process is proportional to the neutrino magnetic moment squared that is in fact a very small quantity for the most of theories beyond the Standard Model (see, for instance, \cite{Giunti:2014ixa}). Other constraints on possible visualization of the $SL\nu$ are imposed by the mentioned above effects of the background plasma.
To avoid the suppression of the radiation produced by the spin light mechanism we
considered the electromagnetic radiation by an electron moving in matter (the ``spin light of electron in matter", $SLe$). From the order-of-magnitude estimation \cite{Studenikin:2005bq}, we
predicted that the ratio of rates of the $SLe$ and the $SL\nu$ in
matter is
\begin{equation}
R=\frac {\Gamma_{SLe}}{\Gamma_{SL\nu}}\sim \frac {e^2}{\omega^2
\mu^2},
\end{equation}
that gives $R \sim 10^{20}\div 10^{14}$ for the value of the neutrino magnetic
moment $\mu \sim 10^{-11}\mu_0$ and the predicted wide range $SLe$ photon's energies $\omega \sim 5 \ MeV \div 5 \ GeV $.
This estimation was confirmed by the direct evaluation \cite{Grigoriev:2006rv} of the $SLe$ properties based on the exact solutions of the modified Dirac equation for the electron in matter. Thus, we expect that in certain
cases the $SLe_{\nu}$ in matter would be more effective than the $SL\nu$. However, the possibility of phenomenological consequences of the $SLe$ in astrophysics is quite not obvious.
In this Letter we continue studies of the spin light mechanism of electromagnetic radiation in a dense environment and consider a new possible realization of this mechanism. The predicted new mechanism implies the electromagnetic radiation emitted by an electron in a dense flux of ultra-relativistic neutrinos. We term this mechanism ``the spin light of electron in dense neutrino fluxes", $SLe_{\nu}$. This new realization of the spin light provides a possibility to avoid two suppression factors in the radiation rate and power, the discussed above suppression due to smallness of a neutrino magnetic moment and one due to the plasma effects. We predict that this new realization of the spin light mechanism can have visible consequences for different astrophysical settings, for stellar core-collapse and supernova explosion phenomenology in particular.
\section{Modified Dirac equation}
Consider a neutrino flux propagating through medium composed of different particles like electrons, protons and neutrons. This situation can be found in various astrophysical settings, for example, in supernovae. In our approach we suppose that the neutrino flux is described by a set of macroscopic parameters
(namely, the particles average values of speeds and polarizations and number densities). We also assume that this kind of astrophysical medium is composed of three neutrino flavors ($\nu_e$, $\nu_{\mu}$ and $\nu_{\tau}$) forming three independent fluxes moving in the same direction. The generalization of our consideration
for the case when the antineutrino fluxes are also present is straightforward.
Let us now investigate the behavior of an electron in a dense medium composed of neutrinos
of different flavors. Within a set of assumptions listed above an effective Lagrangian
describing interaction of the electron with fluxes of different flavour neutrinos is \cite{Studenikin:2004dx,Grigorev:2005sw}
\be
\mathcal{L}=\sum\limits_{i=e,\mu,\tau}\bar{e}\gamma_{\nu}\frac{\epsilon_i (1+\gamma^5)-4\sin^2\theta_W}{2}ef^{\nu}_i,
\label{effective lagrangian_L}
\ee
where $\epsilon_i = -1$ for $i=e$ and $\epsilon_i = +1$ for $i= \mu, \, \tau$. Each of the
flavour neutrino $\nu_{i}$ fluxes is characterized by the neutrino matter potential
\be\label{f}
f^{\mu}_i = G(n_i,n_i\bs{v}_i),
\ee
where the neutrino number densities $n_{i}$ in the rest frame of the electron are determined by the corresponding invariant number densities $n^{0}_{i}$ in the rest frame of the particular flavour neutrino flux,
\be\label{n}
n_{i}=\frac{n^{0}_{i}}{\sqrt{1-\bs{v_{i}}^2}},
\ee
$v_{i}$ is the average speed of the flavour neutrino in the flux. Here $G=\frac{G_F}{\sqrt{2}}$, and $G_F$ is the Fermi constant.
We suppose that all flavour neutrino fluxes have the same speed, $\bm {v}_i=\bm {v}$,
and introduce the average value $n$ of the neutrino number densities and the parameter $\delta_e$,
\be \label{n_delta}
n=\frac{n_e+n_\mu+n_\tau}{3},\qquad \delta_{e}=\frac{n_\mu+n_\tau-n_e}{n}.
\ee
Using these notations we rewrite the effective Lagrangian (\ref{effective lagrangian_L}) in the form,
\be\label{effective lagrangian}
\mathcal{L}=\bar{e}\gamma_{\mu}\frac{c+\delta_e\gamma^5}{2}ef^{\mu},
\ee
where $f^\mu = Gn(1,\bs{v})$ is the effective neutrino potential and $c=\delta_e-12\sin^2\theta_W$.
For the relativistic neutrinos $v\simeq 1$, and the effective neutrino potential is
\be\label{fv1}
f^{\mu}=G(n,0,0,n).
\ee
From here we suppose that the neutrino flux propagation is relating to the direction of an axis $z$.
From the Lagrangian (\ref{effective lagrangian}) one gets the following
modified Dirac equation for an electron in the relativistic neutrino flux,
\be\label{modified Dirac equation}
\{\gamma_{\mu}p^{\mu}+\gamma_{\mu}\frac{c+\delta_e\gamma^5}{2}f^{\mu}-m\}\Psi(x)=0,
\ee
here $m$ and $p^{\mu}$ are the electron mass and momentum.
\section{Exact solution and kinematics}
In the considered case of a relativistic neutrino flux ($v\sim 1$) Eq. (\ref{modified Dirac equation}) with the matter potential given by Eq. (\ref{fv1}) can be solved exactly.
Performing evaluations similarly to those described in \cite{Studenikin:2004dx}, we get
the energy spectrum of an electron in the presence of the relativistic neutrino
flux (\ref{modified Dirac equation}),
\begin{gather}\label{branch E_s}
E^{\varepsilon}_{s}({p})=\varepsilon\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c-s\delta\big)\Big)^2}
-\frac{Gn}{2}\big(c-s\delta\big),
\end{gather}
where $\delta$ is the absolute value of $\delta_e$, $p_3$ is the electron momentum in the direction of the neutrino flux propagation and $p=\sqrt{p_{3}^2+ {p}_\bot ^2}$ is the total electron momentum.
The value $\varepsilon=\pm 1$ corresponds to the positive and
negative frequency solutions (for the electron $\varepsilon = +1$).
The energy spectrum
(\ref{branch E_s}) also contains the second integer number $s=\pm 1$. Comparing Eq. (\ref{branch E_s}) with corresponding spectra of an electron \cite{Grigoriev:2006rv}
or neutrino \cite{Studenikin:2004dx,Grigorev:2005sw} in nonmoving matter
we conclude that the number $s=\pm 1$ distinguishes two possible electron spin states.
Thus we see that the obtained spectrum branches are classified by
both the frequency sign $\varepsilon=\pm 1$ and the spin sign $s=\pm1$
without explicit introduction and using of the spin operator.
Two particular electron energy branches $E^{\varepsilon}_{s}({p})_{|\varepsilon=+1}=
E_{s}({p})$ with $s=\pm1$ as functions of the momentum $p$ are plotted in Fig. \ref{figure spectrum}.
It is clearly seen that two corresponding curves have no common points that means
there are no energy states with ``undefined'' spin.
It follows from Eq.(\ref{branch E_s}) that for the electron ($\varepsilon=+1$)
there are two different spin states with the same momentum ${p}$. It is interesting to note that
the energy of the state characterized by $s=+1$ always exceeds the energy of the state
with $s=-1$ for the fixed value of the electron momentum $p$.
Indeed, from Eq.(\ref{A_3}) of Appendix we get
\begin{gather}\label{E-E}
E_+({p}) - E_-({p})=\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c-\delta\big)\Big)^2}-\\-
\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c+\delta\big)\Big)^2}+Gn\delta>0,
\end{gather}
thus it is always $E_{+}({p})> E_{-}({p})$.
\begin{figure}[h!]
\begin{center}
\psset{xunit=0.75 cm, yunit=0.75 cm}
\begin{pspicture*}(-7,-3)(7,7)
\psline[linewidth=0.5 pt]{->}(-7,0)(7,0)
\psline[linewidth=0.5 pt]{->}(0,-1)(0,7)
\psplot[showpoints=false, linewidth=1.5 pt, plotstyle=curve, algebraic=true]{-1}{6}{2.5+((x-2.5)^2+3)^(1/2)}
\psplot[showpoints=false, linewidth=1.5 pt, plotstyle=curve, algebraic=true]{-6}{6.4}{-1.5+((x+1.5)^2+3)^(1/2)}
\psplot[showpoints=false, linewidth=1 pt, linestyle=dotted, algebraic=true]{-1}{6.5}{x}
\psline[linewidth=1 pt, linestyle=dotted]{-}(-1.5,0)(-1.5,0.2)
\psline[linewidth=1 pt, linestyle=dotted]{-}(2.5,0)(2.5,4.2)
\rput[l](0.1,-0.2){$0$}
\rput[l](0.1,6.8){$E$}
\rput[l](6.5,-0.3){$p_3$}
\rput[l](-3.5,-0.4){$-\frac{Gn}{2}(c+\delta)$}
\rput[l](1.5,-0.4){$-\frac{Gn}{2}(c-\delta)$}
\rput[l](-6,2.5){$E_-$}
\rput[l](-1,5.7){$E_+$}
\rput[l](4,3.5){$E=p_3$}
\end{pspicture*}
\end{center}
\vspace{-1cm}
\caption{
The dependence of the electron energies in two different spin states, $E_+(\bs{p})$ and $E_-(\bs{p})$, on electron momentum component $p_{3}$.}
\label{figure spectrum}
\end{figure}
The modified Dirac equation (\ref{modified Dirac equation})
can be solved exactly. For the electron in $s=+1$ spin state the wave function can be found in the following form,
\begin{gather}\label{wave function_i}
\psi_i(\bs{r},t)=\frac{1}{L^{\frac{3}{2}}C_+}e^{i(-E_+t+ \bs{p}\bs{r})}
\begin{pmatrix}
0\\m\\p_\bot\rme^{-i\phi}\\E_{+} -p_3
\end{pmatrix},
\end{gather}
and in $s=-1$ spin state the wave function is
\begin{gather}\label{wave function_j}
\psi_f(\bs{r},t)=\frac{1}{L^{\frac{3}{2}}C_-}e^{i(-E_{-}t+ \bs{p}\bs{r})}
\begin{pmatrix}
E_{-}-p_3\\-p_\bot e^{i\phi}\\m\\0
\end{pmatrix},
\end{gather}
where $L$ is the normalization length, $C_+$ and $C_-$ are normalization coefficients given by
\be\label{C}
C_{\pm}=\sqrt{m^2+p_\bot^2+(E_{\pm}-p_3)^2}.
\ee
\section{Spin light of electron in dense neutrino flux}
Using the ``method of exact solutions" \cite{Studenikin:2005bq,Studenikin:2007zza,Studenikin:2008qk} we consider the radiative transition of an electron with emission of a photon (plasmon) in the presence of the relativistic flux of neutrinos. This process we term the {\it spin light of electron in dense neutrino flux} ($SLe_\nu$). In the first order of the perturbation theory the Feynman diagram of the process (see Fig. \ref{Feynman diagram SLe}) is the one-photon emission diagram with the bold electron lines denoting the exact account for interaction of electrons with the neutrino background .
\begin{figure}[h]
\begin{center}
\begin{fmffile}{eLightEmission}
\begin{fmfchar*}(180,90)
\fmfleft{e_in}
\fmflabel{$e$}{e_in}
\fmflabel{$e$}{e_out}
\fmflabel{$\gamma$}{gamma}
\fmf{heavy}{e_in,Int1}
\fmf{heavy}{Int1,e_out}
\fmf{photon}{Int1,gamma}
\fmfright{e_out,gamma}
\fmfdot{Int1}
\end{fmfchar*}
\end{fmffile}
\end{center}
\caption{The Feynman diagram of the radiation by the electron in the neutrino flux.}
\label{Feynman diagram SLe}
\end{figure}
The element of $S$-matrix defining the process amplitude
is given by (see also \cite{Grigoriev:2006rv}):
\be\label{amplitude basic}
S^{(\lambda)}_{fi} = -e\sqrt{4\pi}\int\,d^4x \bar{\psi}_f(x)({\bs\gamma} \bs{e}^{(\lambda)}{}^\ast)\frac{\rme^{\rmi kx}}{\sqrt{2\omega L^3}}\psi_i(x),
\ee
where $e$ is the electron charge, $\psi_i(x)$ and $\psi_f(x)$ are the wave functions of the initial and final
electrons in the background neutrino flux, and $k = (\omega, \bs{k})$ and $\bs{e}^{(\lambda)}$ ($\lambda=1,2$) are the momentum and polarization vectors of the emitted photon. The considered radiative process is a quantum transition between two electron spin states with emission of a photon. It can proceed because the
condition $E_+ > E_-$ is fulfilled.
Consider the case when the relativistic flux of neutrinos is propagating through a shell of electrons that are at rest. This can be used as a model of real situations peculiar for astrophysical settings.
The electron rest frame in moving background is defined as one where the electron energy $E_+$ gets its minimum.
This means that the electron de Broglie wave group velocity is vanishing, $\frac{\partial E_+}{\partial\bs{p}}=0$.
This condition, accounting for the initial electron energy given by Eq. (\ref{branch E_s}), is provided in case the following relations are fulfilled,
\be
p_3 = -\frac{Gn}{2}(c-\delta),\qquad \bs{p}_\bot = 0.\label{condition of nonmoving electron}
\ee
It can be seen that in the background matter the minimum of the electron energy corresponds to a non-zero value for its momentum \footnote{This phenomenon was discussed in \cite{Chang:1988yn,Pantaleone:1991zr, Pantaleone:1992xh,Studenikin:2004dx}.}. For the case of a supernova environment $\frac{Gn}{m}\sim 10^{-8}$ therefore most naturally electrons can be considered as nonrelativistic particles.
Performing integration over time and the spatial coordinates in Eq. (\ref{amplitude basic}) we recover the $\delta$-functions providing the law of energy-momentum conservation for the considered process and the following relations are straightforward
(the primed quantities describe the final electron state):
\be\label{e_m_concervation}
\begin{aligned}
E_+ (p)&= E'_- (p')+\omega,\\
-\frac{Gn}{2}(c-\delta)&= p_3'+\omega\cos\theta,\\
0 &= p_\bot' + \omega\sin\theta,
\end{aligned}
\ee
where $\theta$ is the angle between the direction of the $SLe_{\nu}$ radiation and neutrino flux propagation.
\begin{center}
\psset{xunit=1 cm, yunit=1 cm}
\begin{pspicture*}(-3,-2.5)(3,2.5)
\psline[linewidth=0.5 pt]{-}(-3,0)(3,0)
\psline[linewidth=0.5 pt]{-}(0,-2.5)(0,2.5)
\psline[linewidth=0.5 pt]{-}(-3,0)(-3,0.2)
\psline[linewidth=0.5 pt]{-}(-2.5,0)(-2.5,0.1)
\psline[linewidth=0.5 pt]{-}(-2,0)(-2,0.2)
\psline[linewidth=0.5 pt]{-}(-1.5,0)(-1.5,0.1)
\psline[linewidth=0.5 pt]{-}(-1,0)(-1,0.2)
\psline[linewidth=0.5 pt]{-}(-0.5,0)(-0.5,0.1)
\psline[linewidth=0.5 pt]{-}(0.5,0)(0.5,0.1)
\psline[linewidth=0.5 pt]{-}(1,0)(1,0.2)
\psline[linewidth=0.5 pt]{-}(1.5,0)(1.5,0.1)
\psline[linewidth=0.5 pt]{-}(2,0)(2,0.2)
\psline[linewidth=0.5 pt]{-}(2.5,0)(2.5,0.1)
\psline[linewidth=0.5 pt]{-}(3,0)(3,0.2)
\psline[linewidth=0.5 pt]{-}(0,-2.5)(0.1,-2.5)
\psline[linewidth=0.5 pt]{-}(0,-2)(0.2,-2)
\psline[linewidth=0.5 pt]{-}(0,-1.5)(0.1,-1.5)
\psline[linewidth=0.5 pt]{-}(0,-1)(0.2,-1)
\psline[linewidth=0.5 pt]{-}(0,-0.5)(0.1,-0.5)
\psline[linewidth=0.5 pt]{-}(0,0.5)(0.1,0.5)
\psline[linewidth=0.5 pt]{-}(0,1)(0.2,1)
\psline[linewidth=0.5 pt]{-}(0,1.5)(0.1,1.5)
\psline[linewidth=0.5 pt]{-}(0,2)(0.2,2)
\psline[linewidth=0.5 pt]{-}(0,2.5)(0.1,2.5)
\psline[linewidth=1 pt]{->}(2.2,0.4)(3,0.4)
\psplot[showpoints=false, linewidth=1 pt, linestyle = dotted, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{2*1/(1-cos(x)+1)}
\psplot[showpoints=false, linewidth=1 pt, linestyle = dashed, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{2*4/(1-cos(x)+4)}
\psplot[showpoints=false, linewidth=1 pt, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{2*10/(1-cos(x)+10)}
\rput[l](0.1,-0.2){$0$}
\rput[l](2.7,0.6){$\bs{v}$}
\end{pspicture*}
\end{center}
\begin{figure}[h]
\caption{The angular dependence of the photon energy for different neutrino flux densities $\frac{m}{Gn\delta}=1$ (dotted),
$\frac{m}{Gn\delta}=4$ (dashed) and $\frac{m}{Gn\delta} = 10$ (solid),
$\bs{v}$ indicates the direction of the neutrino flux.}
\label{figure angular dependence of frequency}
\end{figure}
The emitted photon energy can be obtained as the only solution of Eqs. (\ref{e_m_concervation}),
\be\label{frequency}
\omega = \frac{m}{1-\cos\theta+\frac{m}{Gn\delta}}.
\ee
From Eq. (\ref{frequency}) it follows that in general the emission is possible in all directions. The angular dependence of the photon energy for different neutrino flux densities is shown in
Fig.\ref{figure angular dependence of frequency} . For the most realistic case, when $Gn\delta\ll m$,
the photon energy simplifies to
\be\label{omega_2}
\omega = Gn\delta.
\ee
Thus, for an initial charged particle with rather large mass or
for the case of the background environment with enough small density the emitted photon energy
does not depend on the direction of radiation and is determined only by the density of the environment.
Using expressions for the amplitude (\ref{amplitude basic}), the wave functions of the initial (\ref{wave function_i}) and final (\ref{wave function_j}) electrons, and for the photon energy (\ref{frequency})
for the $SLe_{\nu}$ in the neutrino flux rate and power we get,
\begin{align}
\Gamma&=\frac{e^2}{2}m\int\limits_0^\pi\frac{(1-\cos\theta)^2}
{(1-\cos\theta+a)^3}\sin\theta\,d\theta,\label{radiation rate diff}\\
I&=\frac{e^2}{2}m^2\int\limits_0^\pi\frac{(1-\cos\theta)^2}
{(1-\cos\theta+a)^4}\sin\theta\,d\theta.\label{power diff}
\end{align}
where the parameter
\be
a=\frac{m}{Gn\delta}
\ee
is used.
Further integration over the angle $\theta$
gives the closed expressions for the radiation rate
\be\label{radiation rate}
\Gamma=\frac{e^2}{2}m\Big[\frac{2}{a^2}-\frac{2}{a}+\ln(1+\frac{2}{a})\Big],
\ee
and power
\be\label{radiation power}
I=\frac{4}{3}e^2m^2\frac{1}{a^3(a+2)}.
\ee
Much simpler expressions can be obtained in the three limiting cases,
\begin{align*}
a &\gg 1, &\Gamma&=\frac{4}{3}e^2m\Big(\frac{Gn\delta}{m}\Big)^3, &I&=\frac{4}{3}e^2m^2\Big(\frac{Gn\delta}{m}\Big)^4,\\
a &\sim 1, &\Gamma&=\frac{\ln 3}{2}e^2m, &I&=\frac{4}{9}e^2m^2,\\
a &\ll 1, &\Gamma&=e^2m\Big(\frac{Gn\delta}{m}\Big)^2, &I&=\frac{2}{3}e^2m^2\Big(\frac{Gn\delta}{m}\Big)^3.
\end{align*}
For most of astrophysical environments $Gn\ll m$ and the case ($a\gg 1$) is realized.
From the angular distribution of the $SLe_{\nu}$ given by Eq.(\ref{power diff}) one can estimate the ratio of the radiation power emitted in the forward and backward directions in respect to the neutrino flux propagation,
\be
\frac{I_{forw}}{I_{back}}=\frac{\int\limits_0^{\frac{\pi}{2}}(1-\cos\theta)^2\sin\theta\,d\theta}
{\int\limits_{\frac{\pi}{2}}^{\pi}(1-\cos\theta)^2\sin\theta\,d\theta}=\frac{1}{7}.
\ee
Therefore, a fraction of about $\frac{1}{8}$ of the total $SLe_{\nu}$ radiation power is emitted in the forward direction (see also Fig. \ref{figure angular dependence of power}).
So that at the space around the supernova core an entire layer (characterized by an overall distance
$R=10 \ km$ from the star centre)
can emit electromagnetic waves with the photon energy decreasing with the increase of
the distance from the centre.
\begin{center}
\psset{xunit=1 cm, yunit=1 cm}
\begin{pspicture*}(-5,-2.5)(1,2.5)
\psline[linewidth=0.5 pt]{-}(-5,0)(1,0)
\psline[linewidth=0.5 pt]{-}(0,-2.5)(0,2.5)
\psline[linewidth=0.5 pt]{-}(-5,0)(-5,0.2)
\psline[linewidth=0.5 pt]{-}(-4.5,0)(-4.5,0.1)
\psline[linewidth=0.5 pt]{-}(-4,0)(-4,0.2)
\psline[linewidth=0.5 pt]{-}(-3.5,0)(-3.5,0.1)
\psline[linewidth=0.5 pt]{-}(-3,0)(-3,0.2)
\psline[linewidth=0.5 pt]{-}(-2.5,0)(-2.5,0.1)
\psline[linewidth=0.5 pt]{-}(-2,0)(-2,0.2)
\psline[linewidth=0.5 pt]{-}(-1.5,0)(-1.5,0.1)
\psline[linewidth=0.5 pt]{-}(-1,0)(-1,0.2)
\psline[linewidth=0.5 pt]{-}(-0.5,0)(-0.5,0.1)
\psline[linewidth=0.5 pt]{-}(0.5,0)(0.5,0.1)
\psline[linewidth=0.5 pt]{-}(1,0)(1,0.2)
\psline[linewidth=0.5 pt]{-}(0,-2.5)(0.1,-2.5)
\psline[linewidth=0.5 pt]{-}(0,-2)(0.2,-2)
\psline[linewidth=0.5 pt]{-}(0,-1.5)(0.1,-1.5)
\psline[linewidth=0.5 pt]{-}(0,-1)(0.2,-1)
\psline[linewidth=0.5 pt]{-}(0,-0.5)(0.1,-0.5)
\psline[linewidth=0.5 pt]{-}(0,0.5)(0.1,0.5)
\psline[linewidth=0.5 pt]{-}(0,1)(0.2,1)
\psline[linewidth=0.5 pt]{-}(0,1.5)(0.1,1.5)
\psline[linewidth=0.5 pt]{-}(0,2)(0.2,2)
\psline[linewidth=0.5 pt]{-}(0,2.5)(0.1,2.5)
\psline[linewidth=1 pt]{->}(0.2,0.4)(1,0.4)
\rput[l](-0.3,-0.2){$0$}
\rput[l](0.7,0.6){$\bs{v}$}
\psplot[showpoints=false, linewidth=1 pt, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{(1-cos(x))^2}
\end{pspicture*}
\end{center}
\begin{figure}[h]
\caption{The angular distribution of the $SLe_{\nu}$ radiation power given by Eq.(\ref{power diff}),
$\bs{v}$ indicates the direction of the neutrino flux.}
\label{figure angular dependence of power}
\end{figure}
\section{Polarization properties}
Consider the polarization properties of the discussed $SLe_{\nu}$ radiation. We define two linear polarization vectors $\bs{e}^{(1)}$ and $\bs{e}^{(2)}$ relative to the plane determined by two vectors
$\bs{\kappa} = (\sin\theta\cos\phi, \sin\theta\sin\phi, \cos\theta)$ and $\bs{e}_z=(0,0,1)$.
The vector $\bs{e}^{(1)}$ is perpendicular to this plane and $\bs{e}^{(2)}$ is parallel to it,
\begin{gather}
\bs{e}^{(1)}=\frac{[\bs{\kappa}\times \bs{e}_z]}{\sqrt{1-(\bs{\kappa}\bs{e}_z)^{2}}}= (\sin\phi,-\cos\phi,0),\\
\bs{e}^{(2)}=\frac{[\bs{\kappa}\times \bs{e}^{(1)}]}{\sqrt{1-(\bs{\kappa}\bs{e}_z)^{2}}}=(\cos\theta\cos\phi,\cos\theta\sin\phi,-\sin\theta).
\end{gather}
Decomposing the $SLe_{\nu}$ transition amplitude (\ref{amplitude basic}) in contributions from the photons of the two linear polarizations, determined by the vectors $\bs{e}^{(1)}$ and $\bs{e}^{(2)}$, we get
\be
|S^{(1)}_{fi}| = |S^{(2)}_{fi}|.
\ee
Thus, the $SLe_{\nu}$ radiation rate and power are equally distributed between the two emitted photon linear polarizations.
\section{Effect of initial electron motion}
The properties of the considered $SLe_{\nu}$ are influenced by possible motion of the initial electron. In particular, the emitted photon energy can be significantly shifted from the value given by Eq.(\ref{frequency}) if the initial electron is moving.
Consider the case when the initial electron moves with the momentum given by
\be\label{p3}
p_3 = -\frac{Gn}{2}(c-\delta) + \tilde{p}_3,\qquad {p}_\bot = 0,
\ee
here $\tilde{p}_3$ describes the deviation of the third component of the initial electron momentum
$p_3$ from its value (\ref{condition of nonmoving electron}) at rest. It is also supposed that the final electron is at rest. In case the initial electron moves along the neutrino flux propagation (${p}_\bot = 0$) the emitted photon energy is given by \be\label{omega_p_3}
\omega = \frac{\sqrt{m^2+\tilde{p}_3^2}-\tilde{p}_3}{1-(1+\frac{\tilde{p}_3}{Gn\delta})\cos\theta+\sqrt{(\frac{\tilde{p}_3}{Gn\delta})^2+(\frac{m}{Gn\delta})^2}}.
\ee
Note that the radiation photon energy in the forward direction ($\theta = 0$) does not depend on $\tilde{p}_3$ and again $\omega$ is given by Eq.(\ref{omega_2}).
On the contrary, the radiation photon energy in the backward direction (when $\theta = \pi$) is strongly dependent on $\tilde{p}_3$
(see Fig. \ref{figure angular dependence of frequency relativistic case}).
In the case the initial electron is moving against the neutrino flux ($\tilde{p}_3 < 0$) we get
\be\label{omega_-_p_3}
\omega = \frac{\sqrt{m^2+\tilde{p}_3^2}+|\tilde{p}_3|}{2-\frac{|\tilde{p}_3|}{Gn\delta}+
\sqrt{(\frac{\tilde{p}_3}{Gn\delta})^2+
(\frac{m}{Gn\delta})^2}}.
\ee
For big enough values of $\tilde{p}_3$, for instance when $|\tilde{p}_3|\gg m$ and also even
$|\tilde{p}_3Gn\delta|\gg m^2$,
there is a reasonable increase of the emitted photon energy in respect to Eq.(\ref{omega_2}),
\be\label{omega_big_p_3}
\omega = 4\Big(\frac{\tilde{p}_3}{m}\Big)^2Gn\delta.
\ee
For example, if $|\tilde{p}_3|\sim 100 \ m$ then the photon energy is $\omega\sim 4\cdot 10^4 \ Gn \delta$.
\begin{center}
\psset{xunit=1 cm, yunit=1 cm}
\begin{pspicture*}(-10,-3)(2,3)
\psline[linewidth=0.5 pt]{-}(-10,0)(1.5,0)
\psline[linewidth=0.5 pt]{-}(0,-2.5)(0,2.5)
\psline[linewidth=0.5 pt]{-}(-10,0)(-10,0.2)
\psline[linewidth=0.5 pt]{-}(-9.5,0)(-9.5,0.1)
\psline[linewidth=0.5 pt]{-}(-9,0)(-9,0.2)
\psline[linewidth=0.5 pt]{-}(-8.5,0)(-8.5,0.1)
\psline[linewidth=0.5 pt]{-}(-8,0)(-8,0.2)
\psline[linewidth=0.5 pt]{-}(-7.5,0)(-7.5,0.1)
\psline[linewidth=0.5 pt]{-}(-7,0)(-7,0.2)
\psline[linewidth=0.5 pt]{-}(-6.5,0)(-6.5,0.1)
\psline[linewidth=0.5 pt]{-}(-6,0)(-6,0.2)
\psline[linewidth=0.5 pt]{-}(-5.5,0)(-5.5,0.1)
\psline[linewidth=0.5 pt]{-}(-5,0)(-5,0.2)
\psline[linewidth=0.5 pt]{-}(-4.5,0)(-4.5,0.1)
\psline[linewidth=0.5 pt]{-}(-4,0)(-4,0.2)
\psline[linewidth=0.5 pt]{-}(-3.5,0)(-3.5,0.1)
\psline[linewidth=0.5 pt]{-}(-3,0)(-3,0.2)
\psline[linewidth=0.5 pt]{-}(-2.5,0)(-2.5,0.1)
\psline[linewidth=0.5 pt]{-}(-2,0)(-2,0.2)
\psline[linewidth=0.5 pt]{-}(-1.5,0)(-1.5,0.1)
\psline[linewidth=0.5 pt]{-}(-1,0)(-1,0.2)
\psline[linewidth=0.5 pt]{-}(-0.5,0)(-0.5,0.1)
\psline[linewidth=0.5 pt]{-}(0.5,0)(0.5,0.1)
\psline[linewidth=0.5 pt]{-}(1,0)(1,0.2)
\psline[linewidth=0.5 pt]{-}(1.5,0)(1.5,0.1)
\psline[linewidth=0.5 pt]{-}(0,-2.5)(0.1,-2.5)
\psline[linewidth=0.5 pt]{-}(0,-2)(0.2,-2)
\psline[linewidth=0.5 pt]{-}(0,-1.5)(0.1,-1.5)
\psline[linewidth=0.5 pt]{-}(0,-1)(0.2,-1)
\psline[linewidth=0.5 pt]{-}(0,-0.5)(0.1,-0.5)
\psline[linewidth=0.5 pt]{-}(0,0.5)(0.1,0.5)
\psline[linewidth=0.5 pt]{-}(0,1)(0.2,1)
\psline[linewidth=0.5 pt]{-}(0,1.5)(0.1,1.5)
\psline[linewidth=0.5 pt]{-}(0,2)(0.2,2)
\psline[linewidth=0.5 pt]{-}(0,2.5)(0.1,2.5)
\psline[linewidth=1 pt]{->}(0.7,0.4)(1.5,0.4)
\rput[l](-0.3,-0.2){$0$}
\rput[l](1.2,0.6){$\bs{v}$}
\psplot[showpoints=false, linewidth=1 pt, linestyle = dotted, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{5*((1+0.01)^(1/2)+0.1)/(1+101^(1/2))}
\psplot[showpoints=false, linewidth=1 pt, linestyle = dashed, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{5*(2^(1/2)+1)/(1+9*cos(x)+10*2^(1/2))}
\psplot[showpoints=false, linewidth=1 pt, plotstyle=curve, polarplot=true, algebraic=true]
{0}{TwoPi}{5*((1+9)^(1/2)+3)/(1+29*cos(x)+10*10^(1/2))}
\end{pspicture*}
\end{center}
\begin{figure}[h]
\caption{The photon energy angular dependence for a fixed neutrino flux density $\frac{m}{Gn\delta}=10$ and
different electron momenta:
$\frac{\tilde{p}_3}{m}=0.1$ (dotted line), $\frac{\tilde{p}_3}{m}=1$ (dashed line) and $\frac{\tilde{p}_3}{m}=3$ (solid line),
$\bs{v}$ indicates the direction of the neutrino flux.}
\label{figure angular dependence of frequency relativistic case}
\end{figure}
\section{Effect of plasma}
It is well know that in general the electromagnetic wave propagation in the background environment is influenced by the plasma effects. For the case of spin light of neutrino in matter these effects have been discussed in details in \cite{Grigorev:2005sw, Kuznetsov:2007ar,Grigoriev:2012pw}. In the above considerations of the $SLe_{\nu}$ an effect of possible nonzero emitted photon mass has been neglected.
Now consider the effect of nonzero emitted photon mass (the plasmon mass $m_\gamma$). The kinematic condition for the $SLe_{\nu}$ in this case is as follows,
\be\label{op_cond_1}
-\frac{Gn}{2}(c-\delta) + \sqrt{\tilde{p}_3^2+m^2} > -\frac{Gn}{2}(c+\delta) + m + m_\gamma.
\ee
After simplification this conditions reads
\be\label{op_cond_2}
\tilde{p}_3^2 > 2m(m_\gamma - Gn\delta).
\ee
If $m_\gamma < Gn\delta$ the process is opened for arbitrary values $\tilde{p}_3$ of the initial electron including the case when the initial electron is at rest.
In the
nonrelativistic classical plasma $m_\gamma = \omega_p = \sqrt{\frac{4\pi e^2 N_e }{m_e}}$, where $N_{e}$ is the electron number density in the plasme.
Thus, if the electron number density $N_e$ is relatively small the kinematic condition
(\ref{op_cond_2}) is fulfield. In the case $\frac{m_\gamma}{Gn\delta}\ll 1$ the effect of the plasmon mass
is not important. The latter conditions can be realized at a distance of about $R>10\ km$ from a supernova centre where $N_e < 10^{19}\,cm^{-3}$.
The Debye screening of electromagnetic waves is another plasma effect that could be important for the $SLe_{\nu}$ radiation propagation in an environment. Only photons (plasmons) with energy exceeding the plasmon frequency
\be\label{condition}
\omega>\omega_p
\ee
can propagate in the plasma. This sets an upper bound for the electron number density in the cloud,
\be
N_e<\frac{\omega^2 m_e}{4\pi e^2},
\ee
providing possibility for the $SLe_{\nu}$ propagation.
Substituting the corresponding values of $\omega$ one gets $N_e<10^{21}\,cm^{-3}$. Thus, the electron matter with $N_e\sim 10^{19} \ cm^{-3}$ considered above is quite transparent for the $SLe_{\nu}$.
\section{Conclusion}
The supernova phenomenon, that is the most energetic event in the Universe, provides a very promising
environment for of the proposed new scheme of the spin light radiation, $SLe_{\nu}$, applications.
Although there are still opened issues in understanding of the supernova mechanism, the two well confirmed main components of it are the
following \cite{Bethe:1990mw, Janka:2006fh}: (i) the collapse of the core under gravity, (ii) the supernova explosion. The second is of particular interest in the context of the presented studies because it is accompanied by the neutrino burst.
According to the present understanding of the phenomenon (see, for instance, \cite{Kachelriess:2008ze}), practically all gravitational binging energy stored in the star (about 99\%) is emitted in the form of a dense neutrino flux that penetrates surrounding interstellar medium.
Under these conditions the neutrino flux can be considered as a continuum
moving at a speed close to the speed of light.
Consider this flux of the ultra-relativistic neutrinos propagating through medium that contains a reasonable amount of electrons. In this case the exact solution for the electron energy spectrum is given by Eq.(\ref{branch E_s})
and the spin light of electrons can be emitted in the quantum transition of an electron from the state with the energy $E_+$ to the state characterized by $E'_-$. Note that the transition from the electron state characterized by $s=-1$ to the state with $s=+1$ is not possible for any values of the initial $p$ and final $p'$ momenta because in this case the energy-momentum conservation equations system, similar to Eq. (\ref{e_m_concervation}), has no nonzero solution.
The emitted photon energy, Eq.(\ref{frequency}), is determined by the effective neutrino number density $n$ and neutrino flux composition given by Eq.(\ref{n_delta}). In order to get the $SLe_{\nu}$ rate and power let us estimate the effective neutrino number density $n$ for the ultra-relativistic neutrino flux from a supernova.
The neutrino luminosity at the level of $L_{\nu}\sim 10^{53}\ \frac{erg}{s}$ have been predicted by the recent supernova simulations \cite{Liebendorfer:2004aj,Pastor:2002prl}. The same order of magnitude estimations, $L_{\nu}\sim 10^{52} - 10^{53}\ erg \ s^{-1}$, have been obtained \cite{Lunardini:2006ap} for the $\bar{\nu}_e$ luminosity with the average energy $\langle E_{\bar{\nu}_e}\rangle \sim 12 - 14\ MeV$ from the K2 and IBM experimental data.
The neutrino luminosity at a distance $R$ from the center of a supernova is determined by
\be\label{luminosity}
L\sim \langle E\rangle JR^2,
\ee
where $\langle E\rangle$ is the average neutrino energy
in the flux, $J$ is the neutrino flux density. For $\langle E\rangle\sim 10^7 \ eV$ and
$R\sim 10^6 \ cm$ we find $J\sim 10^{45} \ cm^{-2}s^{-1}$.
The neutrino flux density is proportional to the neutrino number density in the flux,
$J=nv$, where $v\sim 1$ for the ultra-relativistic neutrino flux. From this we obtain the neutrino number density in the rest frame of the electron,
\be\label{n_35}
n\sim 10^{35}\,cm^{-3}.
\ee
The same value for $n$ is used in \cite{Dasgupta:2008prd} for the $\bar{\nu}_e$ number density near
the neutrinosphere of the supernova.
Taking into account (\ref{n_35}), we get that $\frac{Gn}{m}\sim 10^{-8}$. Thus,
from Eq. (\ref{omega_2}) it follows for he $SLe_{\nu}$ photon energy
\be
\omega\sim 1 \ eV.
\ee
From (\ref{radiation rate}) and (\ref{radiation power})
we also find for the $SLe_{\nu}$ rate and power
\be\label{rate_power}
\Gamma\sim 10^{-19}\,eV\sim 10^{-4}\,s^{-1},\qquad I\sim 10^{-7} {eV} {s^{-1}}.
\ee
The corresponding characteristic time of the $SLe_{\nu}$ process is rather big, $\tau\sim 10^4 s$.
It means that $SLe_{\nu}$ from a single electron is hardly observable.
It is interesting to consider a collective effect of the $SLe_{\nu}$ radiation from a cloud of electrons
in space around a supernova core.
In \cite{Janka:2006fh,Liebendorfer:2004aj} it is shown that there could be
regions with reasonably high electron density. Thus, if the neutrino flux propagates through such
electron-rich clouds the $SLe_{\nu}$ effect can be increased.
At the distance $R=10 \ km$ from the star center the electron number density can be of order
$n_e\sim 10^{19} \ cm^{-3}$.
In this case the amount of $SLe_{\nu}$ flashes per second from $1 \ cm^3$ of the electron matter under the influence of a dense neutrino flux is $N\sim 10^{15} \ cm^{-3} \ s^{-1} $.
For the energy release per one second
of $1 \ cm^3$ of the considered electron-reach matter under the influence of a dense ultra-relativistic neutrino flux we get
\be\label{delta_E}
\frac{\delta E}{\delta t \delta V}\sim N \omega \sim I n_e \sim 10^{13}\ eV \ cm^{-3} \ s^{-1}.
\ee
It is possible to estimate efficiency of the energy transfer from the total neutrino flux to the electromagnetic radiation due to the proposed $SLe_{\nu}$ mechanism. The ``energy content" of the neutrino flux can be characterized by the product of the mean value of the neutrino energy and neutrino number density $\langle E\rangle n$. The corresponding energy characteristics of the $SLe_{\nu}$ radiation
is given by $\omega N_{e}$. Thus, for the $SLe_{\nu}$ luminosity we get the following estimation
\be\label{Lum}
L_{SLe}\sim \frac{\omega N_{e}}{\langle E\rangle n}L_{\nu}\sim 10^{31} \ erg \ s^{-1}
\ee
for the following choice of values $L_{\nu}\sim 10^{53}\ \frac{erg}{s}$, $\omega\sim 1 \ eV$,
$N_e \sim 10^{19}\,cm^{-3}$, $n\sim 10^{35}\,cm^{-3}$ and $\langle E\rangle\sim 10^7 \ eV$.
It is interesting to note that the $SLe_{\nu}$ luminosity can be drastically increased in the case when the emitting electrons are moving with relativistic speed against the neutrino flux propagation. As it follows from
Eq.(\ref{omega_big_p_3}), the energy of the $SLe_{\nu}$ photons also increases in this case. We predict that this should have important consequences in astrophysics and for the supernova process in particular.
\section{Acknowledgments}
We are thankful to Alexander Grigoriev, Alexey Lokhov and Alexei Ternov for many fruitful discussions on the considered phenomenon.
\section{Appendix}
Here we show that $E_{+}({p})> E_{-}({p})$ where the energy of the electron in the relativistic neutrino flux ${E_{s}({p})=E^\varepsilon}_{s}({p})_{|\varepsilon=+1}$ is given by Eq. (\ref{branch E_s}).
From Eq. (\ref{branch E_s}) we get the estimation
\begin{gather}\label{est}
\Biggl|\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c+\delta\big)\Big)^2}-
\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c-\delta\big)\Big)^2}\Biggr|=\nonumber\\
=\frac{2Gn\delta\big|p_3+\frac{Gn}{2}c\big|}{\sqrt{\Big(m^2+{p}_\bot^2+p_3+
\frac{Gn}{2}\big(c+\delta\big)\Big)^2}+\sqrt{m^2+{p}_\bot^2+
\Big(p_3+\frac{Gn}{2}\big(c-\delta\big)\Big)^2}}<\nonumber\\
<\frac{2Gn\delta\big|p_3+\frac{Gn}{2}c\big|}{\big|p_3+\frac{Gn}{2}c+
\frac{Gn}{2}\delta\big|+\big|p_3+\frac{Gn}{2}c-\frac{Gn}{2}\delta\big|}.
\end{gather}
Taking into account the following simple inequality
\be\label{A_3}
\frac{|x|}{\big|x+|y|\big|+\big|x-|y|\big|}=
\left\{
\begin{array}{rcl}
\frac{1}{2},& x\geq|y|\\
\frac{1}{2}\frac{|x|}{|y|},& -|y|<x<|y|\\
\frac{1}{2},& x\leq-|y|
\end{array}
\right\}\quad\leq\quad\frac{1}{2},
\ee
finally we find that
\be\label{A_4}
\Bigl|\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c+\delta\big)\Big)^2}-
\sqrt{m^2+{p}_\bot^2+\Big(p_3+\frac{Gn}{2}\big(c-\delta\big)\Big)^2}\Bigr|<Gn,
\ee
and validity of Eq. (\ref{E-E}) is just straightforward.
\bibliographystyle{model1a-num-names}
|
\section{Introduction}
A set of integers, denoted by $S$, is called a \emph{sum-free set} if $%
S\cap(S+S)=\emptyset$, where $S+S$ denotes the set of pairwise sums, i.e., $%
S+S=\{x+y:x,y\in S\}$. Thus, for any sum-free set $S$, there do not exist $%
x,y,z\in S$ for which $x+y=z$. It is natural to arrange the elements of a
sum-free set $S$ in ascending order, so $S=(S_{n})_{n\geq 0} $ can also
be treated as an integer sequence.
Cameron and Erd\H{o}s \cite{Ca,ca} studied the number of sum-free sets which
are contained in the first $n$ integers. They showed that the number of
sum-free subsets of $\{\frac{n}{3},\frac{n}{3}+1,\cdots,n\}$ is $%
O(2^{\frac{n}{2}})$. Calkin \cite{cs} showed that
the Hausdorff dimension of the sum-free sets is at most $0.599$.
Calkin \cite{number} proved that the number of the sum-free subsets of
$\{1,2,\cdots,n\}$ is $\emph{o}(2^{n(1/2+\epsilon)})$ for every $%
\epsilon>0$. {\L}uczak and Schoen \cite{TT} studied the properties of
$k$-sum-free sets with precise upper density.
There are several methods to construct infinite sum-free sets. One of them
is to construct such a set directly using its definition; that is, to construct it from numbers which are not the sum of two earlier numbers. Another way is to
construct the sum-free set from an infinite zero-one sequence. This method
was introduced by Cameron. Cameron \cite{Pj} defined a
natural bijection between $\Sigma$ and $\mathfrak{S}$, where $\Sigma$ and $%
\mathfrak{S}$ denote the set of all zero-one sequences and the set of all
sum-free sets respectively.
\subsection{The bijection between $\Sigma$ and $\mathfrak{S}$.}
Let $S\in\mathfrak{S}$ be a sum-free set and $\mathbf{v}$ be a ternary
sequence defined as follow:
\begin{equation} \label{vn}
v_n=\left\{
\begin{array}{ll}
1, & \hbox{if $n\in S$;} \\
\ast, & \hbox{if $n\in S+S$;} \\
0, & \hbox{otherwise.}%
\end{array}
\right.
\end{equation}
By deleting all the $\ast$'s in $\mathbf{v}$, we obtain a unique zero-one
sequence $\mathbf{v}^{\prime}$. This process introduces a mapping from $%
\mathfrak{S}$ to $\Sigma$. Calkin \cite{Nc} showed that this mapping is a
bijection and denoted its inverse by $\theta$. Thus $\theta:\mathbf{v}%
^{\prime}\mapsto S$ is a bijection from $\Sigma$ to $\mathfrak{S}$.
Since there is a one-to-one correspondence between the set of zero-one sequences and the set of sum-free sets (of positive integers), one would like to know the connection between these two.
\subsection{Periodicity of $S$}
A sum-free set $S$ is said to be \emph{ultimately periodic} if there exist
positive integers $m,n_{0}$ such that for all $n>n_{0}$, $n\in S$ if and
only if $n+m\in S$. If $n_{0}=0$, then $S$ is \emph{periodic}. It is not
easy to show whether a given sum-free set is periodic or not.
In \cite{Pc}, Cameron observed that if a sum-free set is (ultimately)
periodic, the corresponding zero-one sequence is also (ultimately) periodic.
This was proved by Calkin and Finch in \cite{Nc}.
Conversely, Cameron \cite{Nc} also asked whether sum-free sets
corresponding to (ultimately) periodic zero-one sequences are (ultimately)
periodic or not. This question is still open. With the help of a computer,
Calkin and Finch \cite{Nc} presented some sum-free sets, which correspond to
periodic zero-one sequences, and appear to be aperiodic (aperiodicity checked up
to $10^7$). There is no proof to show whether such sum-free sets are periodic or aperiodic.
Calkin and Erd\H{o}s \cite{aperiodic} showed that a class of aperiodic
sum-free sets $S$ is incomplete, i.e., $\mathbb{N}\backslash (S+S)$ is an
infinite set. Later, Calkin, Finch and Flowers \cite{difference} introduced
the concept of difference density which can be used to test whether specific sets
are periodic or not. These tests produced further evidence that certain sets are not ultimately periodic. Payne \cite{GP}
studied the properties of some sum-free sets over an additive group.
\subsection{Main results.}
Motivated by Cameron's question, in this paper, we will investigate the
regularity of $S$. We would like to know what kind of zero-one sequences
will lead to $k$-regular sum-free sets?
\begin{definition}
A sum-free set $S$ is called a \emph{$k$-regular sum-free set} if the
sequence $(S_n)_{n\geq 0}$ is $k$-regular\footnote{%
For the definition of $k$-regular sequences, refer to Definition \ref{def:regular} in
Section \ref{reg}.} for some $k\in\mathbb{N}$.
\end{definition}
Let $\mathbf{c}=c_{0}c_{1}c_{2}\cdots $ be a zero-one sequence with $c_0=1$.
Let $\mu_{n}$ denote the number of zeros between the $n$-th and the $(n+1)$-th ones in $%
\mathbf{c}$. An equivalent definition of $\mu_n$ is
in Definition \ref{deff}. The sum-free set corresponding to $%
\mathbf{c}$ is denoted by $S$. Our results are stated as follows:
\bigskip
\noindent \textbf{Main results I.} \emph{Suppose the integer sequence
$(\mu _{n})_{n\geq 1}$ is $%
2 $-regular and satisfies the conditions
\begin{equation}
\left\{
\begin{array}{l}
\mu _{2^{m}}>\sum_{i=1}^{2^{m}-1}\mu _{i}+2^{m}+\frac{3^{m}-1}{2}, \\
\mu _{2^{m}+k}=\mu _{k},\quad \forall ~0<k<2^{m},%
\end{array}%
\right. \label{hm}
\end{equation}%
} \emph{for all $m\geq 1$, then $S$ is a $2$-regular sum-free set.}
\bigskip
For any fixed $l_{1}\geq 0,~l_{2}\geq 0$ and $l_{3}\geq 3$, let $\sigma
(l_{1},l_{2},l_{3})$ be a substitution over the alphabet $\{0,1\}$ given by
\begin{equation}
1\mapsto 1\overbrace{0\cdots 0}^{l_{1}}1\overbrace{0\cdots 0}^{l_{2}},\quad
\ \ \ 0\mapsto \overbrace{0\cdots 0}^{l_{3}}. \label{sigma}
\end{equation}%
The sequence
\begin{eqnarray}
\mathbf{c}_{l_{1},l_{2},l_{3}} &=&\sigma (l_{1},l_{2},l_{3})^{\infty }(1)
\nonumber \\
&=&10\cdots 010\cdots 010\cdots 01\cdots \label{cm}
\end{eqnarray}%
is called a \emph{Cantor-like sequence.} In particular, $\mathbf{c}_{1,0,3}$ is the Cantor sequence. In the rest of this paper, since $%
l_{1},l_{2},l_{3}$ are fixed in the context, we will simply use $\mathbf{%
c}$ to denote such Cantor-like sequences. It is worth pointing out that
Cantor-like sequences contain a class of automatic sequences and a class of
non-automatic substitution sequences.
\medskip
\noindent \textbf{Main results II.} \emph{Let }$\mathbf{c}$\emph{\ be a
Cantor-like sequence satisfying the assumption in Main results $I$.
Then the corresponding sum-free set $S$ modulo $2$ is
the Thue-Morse sequence\footnote{The Thue-Morse sequence beginning by $1$
is the infinite sequence $\mathbf{t}=t_0t_1\cdots\in\{0,1\}^{\mathbb{N}}$
satisfying the following relations: $t_0=1,t_{2n}=t_n,t_{2n+1}=1-t_n$.}
beginning by $1$
up to a coding.}
\medskip
We also investigate the zero-one sequences corresponding to a class of
sum-free sets generated from the set $(bn+1)_{n\geq 0}$ through base
changing where $b\in\mathbb{N}$ and $b\geq 2$. For example, let $S=\{1,3,5,7,\cdots \}$ where $%
S_{n}=(2n+1)_{n\geq 0}$. Suppose $n=\sum_{i=1}^{k}n_{i}\cdot 2^{i-1}$, with$%
~n_{i}\in \{0,1\}$ for any $1\leq i\leq k$, denote $S_{n}^{\prime
}:=\sum_{i=1}^{k}n_{i}\cdot 3^{i}+1$. Then $S^{\prime
}=\{1,4,10,13,28,31,37,40,\cdots \}$ is also sum-free. Moreover, the
corresponding zero-one sequence of $S^{\prime }$ is the Cantor sequence (see Remark \ref{rem:cantor}).
\medskip
\noindent\textbf{Main results III.} \emph{Suppose $b\geq2$, let $S$ be the
sum-free set given by $S_n=\sum_{i=1}^kn_i(2b-1)^i+1$ where $%
n=\sum_{i=1}^kn_i2^{i-1}$ with $n_i\in\{0,1\}$ , then $
\theta^{-1}(S)$ is an automatic sequence.}
\medskip
This paper is organized as follows. In Section 2, we introduce some
definitions and auxiliary
lemmas which are useful
in the proof of Theorem \ref{mainthm}. In Section 3, we give the explicit value of the sum-free
set $S$ corresponding to $\mathbf{c}$ and discuss the regularity of $S$.
A class of sum-free sets $S$ corresponding to Cantor-like sequences
is investigated in Section 4. In the last section, we discuss the automaticity of zero-one
sequences corresponding to certain sum-free sets.
\section{Preliminary}
\if
It is of interest to know whether the corresponding sum-free sets of
automatic binary sequence is automatic, this work is intended as an attempt
to motivate that is true. Cobham \cite{Ch} presented an important connection
between automatic sequences and fixed points of substitution of constant
length. We introduce the notation of substitution sequence. \fi
\subsection{$m$-complement of nonnegative integers.}
For $b\in \mathbb{N}$, let $\Sigma_b :=\{0,1,\cdots,b-1\}$. Let $\Sigma_b^k$ be
the set of words of length $k$ over $\Sigma_b$ and $\Sigma_b^{\ast}=\bigcup_{k%
\geq0}\Sigma_b^k$. The \emph{$b$-ary expansion} of $n$ is denoted by
\[
(n)_b:=\epsilon_k\epsilon_{k-1}\cdots\epsilon_1\in\Sigma_b^{k},
\]
where $\epsilon_k\neq 0$. The \emph{$b$-ary number} of $\mathbf{w}%
=w_kw_{k-1}\cdots w_1\in\Sigma_b^k$ is given by
\[
[\mathbf{w}]_b:=\sum_{i=1}^{k}w_i\cdot b^{i-1}.
\]
For $\mathbf{w}=w_kw_{k-1}\cdots w_1\in\Sigma_2^k$, the \emph{complement} of $%
\mathbf{w}$ is the word
\[
\bar{\mathbf{w}}=\bar{w}_k\bar{w}_{k-1}\cdots \bar{w}_1,
\]
where $\bar{\epsilon}=1-\epsilon$ for any $\epsilon\in\Sigma_2$.
Let $m\in\mathbb{N}$ be fixed and $n=\epsilon_m2^{m-1}+%
\cdots+\epsilon_12^0<2^m$ with $\epsilon_i\in\Sigma_2$. In this case, $%
\epsilon_m\epsilon_{m-1}\cdots\epsilon_1$ is a $2$-ary representation of $n$
of length $m$ which allows leading zeros, denoted by $(n)_{2,m}$.
The \emph{$m$-complement} of $n$,
denoted by $\bar{n}$, is given by
\[
\bar{n}=\left[\overline{(n)}_{2,m}\right]_2=\bar{\epsilon}_m2^{m-1}+\cdots+\bar{\epsilon}_12^0.
\]
For example, when $m=3$, then $\bar{1}=1\cdot 2^2+1\cdot 2^1+0\cdot 2^0=6$, $%
\bar{3}=2^2=4$. We also use notation `$0n$' and `$1n$' in form, which
denote integers of $2$-ary representation `$0\epsilon_m\epsilon_{m-1}\cdots%
\epsilon_1$' and `$1\epsilon_m\epsilon_{m-1}\cdots\epsilon_1$' respectively.
In fact,
\[
0n=n,\quad 1n=2^m+n.
\]
\subsection{Two auxiliary bijections}
Let $(h(i))_{i\geq 1}$ be a strictly increasing positive integer sequence. Define
the mapping $f: \Sigma_{2}^m\times\Sigma_{2}^m\rightarrow\mathbb{N}$ by
\[
(x,y)\mapsto f(x,y)=\sum_{i=1}^m(x_i+y_i)h(i)+2.
\]
Let $\sim$ be the equivalence relation on $\Sigma_{2}^m\times\Sigma_{2}^m$
induced by $f$, i.e.,
\[
(x,y)\sim(x^{\prime},y^{\prime}) \textrm{ if and only if } f(x,y)=f(x^{\prime},y^{%
\prime}).
\]
Denote $\widetilde{\Sigma}^m:=(\Sigma_2^m\times\Sigma_2^m)/{\sim}$ for
simplicity. Hence, we have defined an injection
\[
f:\widetilde{\Sigma}^m\rightarrow\mathbb{N}.
\]
For any $m\geq 0$, define two mappings $\varphi_m,~\psi_m$ as follows:
for any $(u,v)\in%
\widetilde{\Sigma}^m$,
\begin{eqnarray*}
(u,v) &\mapsto& \varphi_m(u,v)=(\bar{u},\bar{v}), \\
(u,v) &\mapsto& \psi_m(u,v)=(0u,1v).
\end{eqnarray*}
Notice that
\begin{eqnarray}
f\circ\varphi_m(u,v)&=&f(\bar{u},\bar{v}) \nonumber \\
&=&\sum_{i=1}^m (\bar{u}_i+\bar{v}_i)h(i)+2 \nonumber \\
&=&2\sum_{i=1}^mh(i)-f(u,v)+4, \label{eqn:phi}
\end{eqnarray}
and
\begin{equation}
f\circ\psi_m(u,v)=f(0u,1v)=f(u,v)+h(m+1). \label{eqn:psi}
\end{equation}
By (\ref{eqn:phi}) and (\ref{eqn:psi}),
\[
(\bar{u},\bar{v})\sim(\bar{u}^{\prime},\bar{v}^{\prime})%
\Longleftrightarrow(u,v)\sim (u^{\prime},v^{\prime}) \Longleftrightarrow
(0u,1v)\sim (0u^{\prime},1v^{\prime}).
\]
Thus $\varphi_m$ and $\psi_m$ are well-defined injections.
But these two mappings are not always bijections. Additional conditions are
needed to ensure that. Before this, we need the following notation: for any
$0\leq k<2^m$ with $k=\epsilon_m2^{m-1}+\cdots+\epsilon_12^0$ and $%
\epsilon_i\in\Sigma_2$, denote
\begin{equation}
M_k:=\sum_{i=1}^m\epsilon_ih(i)+1, ~~M_{2^m+k}:=M_k+h(m+1).
\label{lem1eqn31}
\end{equation}
Note that $M_{2^m-1}=\sum_{i=1}^mh(i)+1$. Moreover, for any $m\geq0, 0\leq k<2^m$%
, let
\begin{eqnarray*}
L_m &:=& \{(u,v)\in\widetilde{\Sigma}^m:~f(u,v)\leq M_{2^m-1}\}, \\
R_m &:=& \{(u,v)\in\widetilde{\Sigma}^m:M_{2^m-1}+1<f(u,v)\leq 2M_{2^m-1}\},
\\
L(k)&:=& \{(u,v)\in\widetilde{\Sigma}^m:~M_k<f(u,v)<M_k+K\}, \\
R(k)&:=& \{(0u,1v)\in\widetilde{\Sigma}%
^{m+1}:~M_{2^m+k}<f(0u,1v)<M_{2^m+k}+K\},
\end{eqnarray*}
where $K=\mu_k+\alpha_k+1$, and $(\mu_i)_{i\geq 0},~(\alpha_i)_{i\geq 0}$
are positive integer sequences.
\begin{lemma}
\label{lm4} $\varphi_m|_{L_m}$ is a bijection from $L_m$ to $R_m$.
\end{lemma}
\begin{proof}
By (\ref{eqn:phi}), $f(\bar{u},\bar{v})=2M_{2^m-1}-f(u,v)+2.$ Hence, if $%
(u,v)\in L_m$, then
\[
M_{2^m-1}+2\leq f(\bar{u},\bar{v})\leq 2M_{2^m-1},
\]
which implies $\varphi_m(u,v)=(\bar{u},\bar{v})\in R_m$.
Note that $\varphi_m(\bar{u},\bar{v})=(u,v)$. To show $\varphi_m$ is a
surjection, we only need to show $(\bar{u},\bar{v})\in L_m$ for any $%
(u,v)\in R_m$. In fact, if $(u,v)\in R_m$, then
\[
f(\bar{u},\bar{v})=2M_{2^m-1}-f(u,v)+2<M_{2^m-1}+1.
\]
\end{proof}
\begin{lemma}
\label{lm5} For any $0\leq k<2^m$, $\psi_m|_{L(k)}$ is a bijection from $%
L(k) $ to $R(k)$.
\end{lemma}
\begin{proof}
Note that $M_{2^m+k}=h(m+1)+M_k$. By (\ref{eqn:psi}), $(u,v)\in L(k)$ if and
only if
\[
M_{2^m+k}<f(0u,1v)=f(u,v)+h(m+1)<M_{2^m+k}+K.
\]
Thus $\psi_m|_{L(k)}$ is a bijection.
\end{proof}
\begin{remark}
By Lemma \ref{lm4} and Lemma \ref{lm5}, for $0\leq k<2^m$,
\[
\textrm{Card}~L_m=\textrm{Card}~R_m,~\textrm{Card}~L(k)=\textrm{Card}~R(k).
\]
\end{remark}
\section{Regularity of $S$}
Let $\mathbf{w}$ be a zero-one sequence beginning by $1$, and $S=\theta (\mathbf{w})$ be the
corresponding sum-free set. In this section, we will characterize $S$ using the properties of $S+S$.
\subsection{Elementary observations of $S$}
Now assume that $\mathbf{v}=(v_{n})_{n\geq 0}$ is the sequence generated from $%
S $ by (\ref{vn}). In fact, $\mathbf{v}$ labels all the natural numbers by $0
,1$ or $*$. And only those integers belonging to $S$ are labeled by $1$. When
we just care about the number of integers labeled by $0$ or $*$, we will use
``the number of $0$'s (or $*$'s)'' rather than ``the number of integers
labeled by $0$'s (or $*$'s)''. This abuse of language does not cause any
ambiguity in the paper.
\begin{definition}\label{deff}
For any $n\geq 1$, denote by $\mu_n$ (resp. $\alpha_n$), the number of $0$'s
(resp. $*$'s) between $S_{n-1}$ and $S_{n}$. In detail,
\begin{eqnarray*}
&\mu_n := \textrm{Card}~\{i\in\mathbb{N}: v_i=0, S_{n-1}<i<S_{n}\}; \\
&\alpha_n := \textrm{Card}~\{i\in\mathbb{N}: v_i=*, S_{n-1}<i<S_{n}\}.
\end{eqnarray*}
\end{definition}
\begin{remark}
$(1)$ For any $n\geq 1$,
\[
S_{n}-S_{n-1}=\mu_n+\alpha_n+1.
\]
Thus to study $S$, we need to study the properties of $(\mu_n)_{n \geq 1}$ and $%
(\alpha_n)_{n\geq 1}$.
$(2)$ Since $\mathbf{v}$ converts to $\mathbf{w}$ by deleting all the $*$'s, $%
\mu_n$ represents the number of $0$'s between the $n$-th and $(n+1)$-th `$1$%
' in $\mathbf{w}$, i.e.,
\[
\mathbf{w}=1\overbrace{0\cdots 0}^{\mu_{1}}1\overbrace{0\cdots 0} ^{\mu_{2}}1%
\overbrace{0\cdots 0}^{\mu_{3}}1\cdots.
\]
\end{remark}
\label{reg} During the proof of Theorem \ref{mainthm}, we need the following
two notation: for any $n\geq 1$,
\begin{eqnarray}
g(n)&:=&\mu_n+\alpha_n, \label{gn} \\
h(n)&:=&\left(\sum_{i=1}^{2^{n-1}}g(i)\right)+2^{n-1}. \label{hn}
\end{eqnarray}
Thus $S_{n}=S_{n-1}+g(n)+1$ and $h(n)$ is strictly increasing.
\begin{lemma}
\label{lem:hn} Suppose $g(2^k+i)=g(i)$ for $0\leq k<m, ~0<i<2^k$. Then
\[
h(m+1)=\sum_{i=1}^{m}h(i)+g(2^m)+1.
\]
\end{lemma}
\begin{proof}
By (\ref{hn}),
\begin{eqnarray*}
&&h(k+1)-h(k) \\
&=&\left(\sum_{i=1}^{2^k}g(i)+2^k\right)-h(k) \\
&=& \left(\sum_{i=1}^{2^{k-1}-1}\big(g(i)+g(2^{k-1}+i)\big)%
+g(2^{k-1})+g(2^k)+2^k\right)-h(k) \\
&=& h(k)+g(2^k)-g(2^{k-1}).
\end{eqnarray*}
Add the last equation up from $k=1$ to $m$. The result follows.
\end{proof}
\begin{theorem}
\label{thm1} \label{mainthm} Let $\mathbf{c}:=1\overbrace{0\cdots0}^{\mu_1}1%
\overbrace{0\cdots0}^{\mu_2}1\cdots$. Denote its corresponding sum-free set by $%
S=(S_n)_{n\geq 0}$. If the sequence $(\mu_n)_{n\geq1}$ satisfies
$(\ref{hm})$, then for every integer $n\geq 1$, we have
\begin{enumerate}
\item if $n=2^k(2j+1)$ for some $k,j\geq 0$, then
\begin{equation} \label{alphan}
\alpha_n=\frac{3^{k}+1}{2};
\end{equation}
\item if $(n)_2=\epsilon_m\epsilon_{m-1}\cdots\epsilon_1$, then
\[
S_n=1+\sum_{i=1}^m\epsilon_ih(i).
\]
\end{enumerate}
\end{theorem}
\begin{proof}[Proof of Theorem \ref{thm1}]
We prove this theorem by induction on $n$. \newline
\textbf{Step1:} Since $\alpha_1=1,~S_1=\mu_1+3$, the conclusion is true for $%
n=1$. \newline
\textbf{Step2:} Assume the result is true for $n<2^m$; that is, if $%
n=2^k(2j+1)$, then $\alpha_n=\frac{3^k+1}{2}$. Thus $\alpha_{2^p+q}=%
\alpha_{q}, \forall~0<q<2^p, 0<p<m$. Moreover, if $(n)_2=\epsilon_m%
\epsilon_{m-1}\cdots\epsilon_1$, then $S_n=1+\sum_{i=1}^m\epsilon_ih(i)$.
Hence, it suffices to show that the result is also true for ~$2^m\leq
n<2^{m+1}$.
\textbf{Step2.1} Let $n=2^m$. By the induction hypothesis and (\ref{alphan}), we have
\begin{eqnarray}
\sum_{i=1}^{2^m-1}\alpha_i
&=& \sum_{i=1}^{2^{m-1}-1}\alpha_i+\alpha_{2^{m-1}}+\sum_{i=1}^{2^{m-1}-1}\alpha_{2^{m-1}+i}
=2\sum_{i=1}^{2^{m-1}-1}\alpha_i+\alpha_{2^{m-1}}\nonumber\\
&=&2^{2}\sum_{i=1}^{2^{m-2}-1}\alpha_i+2\alpha_{2^{m-2}}+\alpha_{2^{m-1}}
\cdots
=\sum_{i=1}^{m}2^{i-1}\alpha_{2^{m-i}}\nonumber\\
&=&\frac{3^{m}-1}{2}.\label{sum:alpha}
\end{eqnarray}
Now we will evaluate $\alpha_{2^m}$. By (\ref{sum:alpha}) and (\ref%
{hm}),
\begin{eqnarray*}
S_{2^m-1} &=& S_0+\sum_{i=1}^{2^m-1}(\mu_i+\alpha_i)+2^m-1 \\
&=& \sum_{i=1}^{2^m-1}\mu_i+2^m+\frac{3^m-1}{2}< \mu_{2^m}.
\end{eqnarray*}
Thus for $i,j\leq 2^m-1$,
\begin{equation} \label{ss}
S_i+S_j\leq 2S_{2^m-1}<S_{2^m-1}+\mu_{2^m}.
\end{equation}
Thus to evaluate $\alpha_{2^m}$, we only need to count the number of $\ast$'s
between $S_{2^m-1}$ and $2S_{2^m-1}$. It is easy to see that $%
S_0+S_{2^m-1}=S_{2^m-1}+1$ is one of such $\ast$'s. Denote
\begin{eqnarray*}
\widetilde{L}_m &:=& \{x\in S+S:S_0<x<S_{2^m-1}\}, \\
\widetilde{R}_m &:=& \{x\in S+S:S_{2^m-1}+1<x<2S_{2^m-1}\}.
\end{eqnarray*}
Then
\[
\alpha_{2^m}=\textrm{Card}~\widetilde{R}_m+1.
\]
Since $f$ is an injection and $\widetilde{L}_{m}=f(L_m), ~\widetilde{R}%
_{m}=f(R_m)$, by Lemma \ref{lm4}, we have
\[
\textrm{Card}~\widetilde{L}_{m}=\textrm{Card}~\widetilde{R}_{m}.
\]
Hence,
\begin{eqnarray*}
\alpha_{2^m}&=&\textrm{Card}~\widetilde{L}_m+1 \\
&=&\sum_{i=1}^m2^{m-i}\alpha_{2^{i-1}}+1 \\
&=&\frac{3^m+1}{2}.
\end{eqnarray*}
Now we will give the expression of $S_{2^{m}}$. By (\ref{hm}) and (\ref%
{alphan}), for $0<j<m$ and $0<k<2^{j}$,
\[
g(2^{j}+k)=\alpha _{2^{j}+k}+\mu _{2^{j}+k}=\alpha _{k}+\mu _{k}=g(k).
\]%
Let $(2^{m})_{2}=1\overbrace{0\cdots 0}^{m}=:\epsilon _{m+1}(2^{m})\epsilon
_{m}(2^{m})\cdots \epsilon _{1}(2^{m})$. By Lemma \ref{lem:hn} and the
induction assumption, we have
\begin{eqnarray*}
S_{2^{m}} &=&S_{2^{m}-1}+g(2^{m})+1 \\
&=&\sum_{k=1}^{m}h(k)+g(2^{m})+2 \\
&=&h(m+1)+1 \\
&=&1+\sum_{k=1}^{m+1}\epsilon _{k}(2^{m})h(k).
\end{eqnarray*}%
\textbf{Step2.2:} Now we prove the results for $2^{m}<n<2^{m+1}$. Let $%
n=2^{m}+k$ where $1\leq k<2^{m}$, and we prove the results by induction on $k$.
For $1\leq k<2^{m}$, assume the result is true for $n<2^{m}+k$. We will
prove the results for $n=2^{m}+k$.
Firstly, we need to determine the value of $S_{2^{m}+k}$. Denote $\Xi%
_{2^{m}+k-1}:=\{x\in S:x\leq S_{2^{m}+k-1}\}$. Then
\[
\Xi_{2^{m}+k-1}+\Xi_{2^{m}+k-1}=\mathcal{S}_{1}\cup
\mathcal{S}_{2}\cup \mathcal{S}_{3},
\]%
where
\begin{eqnarray*}
\mathcal{S}_{1}:= &\{S_{i}+S_{j}:&i,j<2^{m}\}, \\
\mathcal{S}_{2}:= &\{S_{i}+S_{j}:&i<2^{m},2^m\leq j\leq 2^{m}+k-1\}, \\
\mathcal{S}_{3}:= &\{S_{i}+S_{j}:&2^m\leq i,j\leq 2^{m}+k-1\}.
\end{eqnarray*}
In fact, if $i,j<2^m$, then by (\ref{ss}) we have $S_{i}+S_{j}<S_{2^{m}}$.
Thus $\max \mathcal{S}%
_{1}<S_{2^{m}}$. Note that by the induction hypothesis, $S_{2^{m}+k-1}=h(m+1)+S_{k-1}$ and
$S_{2^{m}}=h(m+1)+1$. Since $(S_{n})_{n\geq 1}$ is strictly increasing, we have for $i,j\geq 2^{m}$,
\begin{eqnarray*}
S_{i}+S_{j} &> &S_{k}+S_{2^{m}}-1 \\
&=&(S_{k-1}+\mu _{k}+\alpha _{k}+1)+(h(m+1)+1)-1 \\
&=&S_{2^{m}+k-1}+\mu _{k}+\alpha _{k}+1=:s.
\end{eqnarray*}%
Therefore $\min \mathcal{S}_{3}> s$. Note that $s=h(m+1)+S_k$. Set $K=g(k)+1$,
then $S_k=S_{k-1}+K$
and $s=S_{2^m+k-1}+K$. Let
\begin{eqnarray*}
& &\widetilde{L}(k):=\{x\in S+S:S_{k-1}<x<S_{k}\}, \\
& &\widetilde{R}(k):=\{x\in S+S:S_{2^{m}+k-1}<x<s\}.
\end{eqnarray*}%
Clearly $f(L(k))=\widetilde{L}(k)$ and $f(R(k))\subset \widetilde{R}(k)$.
Now we will show that $f(R(k))=\widetilde{R}(k)$. Since $f$ is an injection,
by Lemma \ref{lm5}, we have
\[
\textrm{Card}~\widetilde{R}(k)\geq \textrm{Card}~\widetilde{L}(k)=\alpha _{k}.
\]
Suppose $S_{2^m+k}<s$. The way that we construct $S$ from a zero-one sequence implies that there are $\mu_{k}$ $0$'s, at least $\alpha_{k}$ $\ast$'s and at least one $1$ between $S_{2^m+k-1}$ and $s$. However there are only $g(k)$ integers between $S_{2^m+k-1}$ and $s$, which is a contradiction. Therefore $S_{2^m+k}\geq s$. In this case, $f(R(k))= \widetilde{R}(k)$. Hence
$$\textrm{Card}~\widetilde{R}(k)=\alpha _{k}.$$
And the number $s$ must be labeled by $\ast$ or $1$, i.e., $v_s=\ast \textrm{ or } 1$. Since $\max\mathcal{S}_1<S_{2^m}$ and $\min\mathcal{S}_3>s$, then $s\in\mathcal{S}_2$ if $v_s=\ast$. That is,
$$\exists~ i<2^m,j\geq2^m, \textrm{ s.t., } s=S_i+S_j.$$
Then there exist $i^{\prime},j^{\prime}<2^m$ satisfying $i=0i^{\prime},j=1j^{\prime}$, and
\begin{eqnarray*}
s=S_k+h(m+1)=S_i+S_j=S_{i^{\prime}}+S_{j^{\prime}}+h(m+1).
\end{eqnarray*}
Thus $S_k=S_{i^{\prime}}+S_{j^{\prime}}$ which contradicts the construction of $S_k$.
Therefore $v_s=1$ and $s=S_{2^m+k}$. Combining this fact with (\ref{hm}), $\alpha_k=\alpha_{2^m+k}$. The proof is completed.
\end{proof}
\iffalse
Note that, for $i<2^{m},j>2^{m}$, there exist $i^{\prime },j^{\prime }$,
satisfying $i=0i^{\prime },j=1j^{\prime }$. By Lemma \ref{lm5}, we know, if $%
S_{i^{\prime }}+S_{j^{\prime }}\in \widetilde{L}(k)$, then $S_{i}+S_{j}\in
\widetilde{R}(k)$; if $S_{i^{\prime }}+S_{j^{\prime }}<S_{k-1}$, then $%
S_{i}+S_{j}<S_{2^{m}+k-1}$; if $S_{i^{\prime }}+S_{j^{\prime }}>S_{k}$, then
\begin{eqnarray*}
S_{i}+S_{j} &=&\sum_{q=1}^{m}(\epsilon _{q}(i^{\prime })+\epsilon
_{q}(j^{\prime }))h(q)+h(m+1)+2 \\
&>&S_{k-1}+\mu _{k}+\alpha _{k}+1+h(m+1)+1 \\
&=&s.
\end{eqnarray*}%
For any $m\geq 0$, let
\[
\mathfrak{S}_{2}^{(1)}:=\{x\in \mathfrak{S}_{2}:x-h(m+1)\notin \widetilde{%
L(k)}\},
\]%
\[
\mathfrak{S}_{2}^{(2)}:=\{x\in \mathfrak{S}_{2}:x-h(m+1)\in \widetilde{L(k)}%
\}.
\]
Then we have $\mathfrak{S}_2^{(1)}\cap (S_{2^m+k},s)=\emptyset$, and $\textrm{Card}~\{%
\mathfrak{S}_2^{(2)}\cap \widetilde{R(k)}\}=\alpha_{k+1}$, so $\textrm{Card}~%
\widetilde{R(k)}\geq\alpha_{k+1}$. And $\textrm{Card}~\widetilde{R(k)}=\textrm{Card}~%
\widetilde{L(k)}\leq\alpha_{k+1}$, hence
\[
\textrm{Card}~\widetilde{L(k)}=\textrm{Card}~\widetilde{R(k)}=\alpha_{k+1}.
\]
From above and (\ref{vn}), we know $v_{s}$ can not be $\ast$, since
$s$ is not the sum of any two elements in the sum-free set $S$, and $v_{s}$
can not be $0$, since $s>S_{2^m+k}+\mu_{2^m+k+1}$, so $v_{s}$ must be $1$,
which implies that $s\in S$, and $s=S_{2^m+k+1}$, i.e.,
\begin{eqnarray*}
S_{2^m+k+1} &=& S_{2^m+k}+\mu_{2^m+k+1}+\alpha_{k+1}+1 \\
&=& 1+h(m+1)+S_{k}+\mu_{k+1}+\alpha_{k+1}+1 \\
&=& 1+h(m+1)+S_{k+1}=1+\sum_{i=1}^{m+1}\epsilon_i^{\prime}h(i),
\end{eqnarray*}
where $(2^m+k+1)_2:=\epsilon_{m+1}^{\prime}\cdots\epsilon_1^{\prime}$. Then
from the definition of $S_n$, we have $S_{2^m+k+1}=S_{2^m+k}+\mu_{2^m+k+1}+%
\alpha_{2^m+k+1}+1$, so
\[
\alpha_{2^m+k+1}=\alpha_{k+1}.
\]
\fi
\begin{remark}
$(h(n))_{n\geq 1}$ is a numeration system (for detail, refer to \cite[Chapter 3]{jps}. In fact, by Lemma \ref{lem:hn},
for any $j\geq1$, $\sum_{i=1}^{j-1}h(i)<h(j)$.
\end{remark}
\begin{definition}[Allouche and Shallit \protect\cite{Jp}]
\label{def:regular} A sequence $(t(n))_{n\geq 0}$ is called \emph{$k$-regular} if
there exist $m$ subsequences of $n$, say $\{n_l^{(j)}\}_{l\geq0}$ $(0\leq j\leq
m-1)$, which satisfy for any $i\geq0$ and $0\leq b<k^i$, the subsequence $%
(t(k^in+b))_{n\geq 0}$ is a $\mathbb{Z}$-linear combination of $t(n_l^{(j)})$.
\end{definition}
\begin{lemma}[Allouche and Shallit \protect\cite{Jp}]
\label{lemmm} If $(u_{n})_{n\geq 1}$ and $(v_{n})_{n\geq 1}$ are both $k$%
-regular, then $(u_{n}+v_{n})_{n\geq 1}$ and $(\sum_{i=1}^{n}u_{i})_{n\geq
1} $ are also $k$-regular.
\end{lemma}
\begin{theorem}
\label{thm:regular}If $(\mu _{n})_{n\geq 1}$ is $2$-regular and
(\ref{hm}) holds, then the sequence $(S_{n})_{n\geq 0}$ is $2$%
-regular.
\end{theorem}
\begin{proof}
Since $(\mu _{n})_{n\geq 1}$ is $2$-regular, by Lemma \ref{lemmm}, so is $%
(\sum_{i=1}^{n}\mu _{i})_{n\geq 1}$. And (\ref{alphan}) in Theorem %
\ref{thm1} implies that
\[
\left\{
\begin{array}{l}
\alpha _{2n}=3\alpha _{n}-1, \\
\alpha _{2n+1}=1.%
\end{array}%
\right.
\]%
Thus $(\alpha _{n})_{n\geq 1}$ is $2$-regular. By Lemma \ref{lemmm}, $(\sum_{i=1}^{n}\alpha _{i})_{n\geq 1}$ is $2$-regular. Note that $(n+1)_{n\geq 1}$ is
also $2$-regular, so by Lemma \ref{lemmm},
\[
S_{n}=\sum_{i=1}^{n}\mu _{i}+\sum_{i=1}^{n}\alpha _{i}+(n+1)
\]%
is $2$-regular.
\end{proof}
\section{Examples}
Let $\Sigma_2^{\ast }$ be the set of finite words over alphabet $\{0,1\}$.
For any $w\in \Sigma_2^{\ast }$, the length of $w$ is denoted by $|w|$. Denote by $|w|_{0}$ and $|w|_{1}$ the number of $0$'s and $1$'s in $w$ respectively.
The following lemma completely characterizes the gap between two adjacent $1$%
's in the Cantor-like sequence $\mathbf{c}$.
\begin{lemma}
\label{lemma1} For any $l_{1}\geq 0,~l_{2}\geq 0,~l_{3}\geq 3$ and all $%
n\geq 1$,
\begin{equation}
\mu _{n}=\frac{l_{2}(l_{3}^{k}-1)}{l_{3}-1}+l_{1}l_{3}^{k} \label{mun}
\end{equation}%
where $n=2^{k}(2j+1)$ for some $k,j\geq 0$.
\end{lemma}
\begin{proof}
Let $\sigma:=\sigma(l_1,l_2,l_3)$. It is easy to see that $%
|\sigma^m(1)|_1=2^m $ and $|\sigma^m(0)|=|\sigma^m(0)|_0=l_3^m$ for any $%
m\geq 0$.
Now, we prove this result by induction on $n$. It is clear that $\mu_1=l_1$.
Assume that the result is true for all $n<2^m$; we prove it for $2^m\leq
n<2^{m+1} $. Since $\sigma^{\infty}(1)$ begins with $\sigma^{m+1}(1)$ and $%
|\sigma^{m+1}(1)|_1=2^{m+1}$, in order to evaluate $\mu_{n}$ for $2^m\leq
n<2^{m+1}$, we only need to investigate $\sigma^{m+1}(1)$. Note that $%
\sigma^{m}(1)$ begins with `$1$' and
\begin{eqnarray}
\sigma^{m+1}(1) &=& \sigma^m(\sigma(1))=\sigma^m(1\overbrace{0\cdots 0}%
^{l_1}1\overbrace{0\cdots 0}^{l_2}) \nonumber \\
&=&\sigma^m(1)\overbrace{\sigma^m(0)\cdots \sigma^m(0)}^{l_1}\sigma^m(1)%
\overbrace{\sigma^m(0)\cdots \sigma^m(0)}^{l_2} \nonumber \\
&=&\sigma^{m-1}(1\overbrace{0\cdots 0}^{l_1}1\overbrace{0\cdots 0}^{l_2})%
\overbrace{\sigma^m(0)\cdots \sigma^m(0)}^{l_1}\sigma^m(1)\overbrace{%
\sigma^m(0)\cdots \sigma^m(0)}^{l_2} \nonumber \\
&=&\sigma^{m-1}(1)\overbrace{\sigma^{m-1}(0)\cdots\sigma^{m-1}(0)}%
^{l_1}\sigma^{m-1}(1) \overbrace{\sigma^{m-1}(0)\cdots\sigma^{m-1}(0)}^{l_2}
\nonumber \\
& &\overbrace{\sigma^m(0)\cdots \sigma^m(0)}^{l_1}\sigma^m(1)\overbrace{%
\sigma^m(0)\cdots \sigma^m(0)}^{l_2}; \label{iteration_m}
\end{eqnarray}
then we have
\begin{equation}
\mu_{2^m}=\sum_{i=0}^{m-1}l_2l_3^i+l_1l_3^m=\frac{l_2(l_3^m-1)}{l_3-1}%
+l_1l_3^m. \label{lem1eqn1}
\end{equation}
From (\ref{iteration_m}) and the fact $|\sigma^{m+1}(1)|_1=2^{m+1}$%
, we know that the block of consecutive zeros between $n$-th and $(n+1)$-th
`1' will appear in the second $\sigma^m(1)$ of $\sigma^{m+1}(1)$ and for all
$0<i<2^m$,
\begin{equation}
\mu_{2^m+i}=\mu_i. \label{lem1eqn3}
\end{equation}
Since for any $i=2^k(2j+1)$ where $0<k<m$, we have $2^m+i=2^k(2j^{\prime}+1)$%
. Equation (\ref{lem1eqn3}) implies that for all $0<i<2^m$,
\begin{equation}
\mu_{2^m+i}=\mu_i=\frac{l_2(l_3^k-1)}{l_3-1}+l_1l_3^k. \label{lem1eqn2}
\end{equation}
Therefore the result follows from (\ref{lem1eqn1}) and (%
\ref{lem1eqn2}).
\end{proof}
\begin{theorem}
\label{thm2} Let $\mathbf{c}$ be the Cantor-like
sequence satisfying
$7l_{3}\geq 4(l_{1}+l_{2})+17$ and $l_{1}(l_{3}-1)+l_{2}>3$
, and $S$ be the sum-free set corresponding to $\mathbf{c}$. Then the sequence $(S_{n})_{n\geq 0}$ is $2$%
-regular.
\end{theorem}
\begin{proof}
According to Theorem \ref{thm:regular}, we only need to show that the gap
sequence $(\mu _{n})_{n\geq 1}$ of $\mathbf{c}$ is $2$-regular and
satisfies (\ref{hm}). By Lemma \ref{lemma1}, we have
\begin{equation}
\left\{
\begin{array}{l}
\mu _{2n}=l_{3}\mu _{n}+l_{2}, \\
\mu _{2n+1}=l_{1},%
\end{array}%
\right.\label{mu:rec}
\end{equation}%
which implies $(\mu _{n})_{n\geq 1}$ is $2$-regular.
Now we will show that $(\mu _{n})_{n\geq 1}$ satisfies (\ref{hm}). By (\ref{mu:rec}), for all $m\geq 1$,
\begin{eqnarray}
\sum_{i=1}^{2^m-1}\mu_i
&=& \sum_{i=0}^{2^{m-1}-1}\mu_{2i+1}+\sum_{i=1}^{2^{m-1}-1}\mu_{2i}\nonumber\\
&=&2^{m-1}l_{1}+l_{3}\sum_{i=1}^{2^{m-1}-1}\mu_{i}+(2^{m-1}-1)l_{2}.\nonumber
\end{eqnarray}
Thus
\begin{equation}
l_{3}\sum_{i=1}^{2^{m-1}-1}\mu_{i}=\sum_{i=1}^{2^m-1}\mu_i-2^{m-1}(l_{1}+l_{2})+l_{2}.\label{sum:mu}
\end{equation}
By (\ref{lem1eqn3}), we know that $(\mu _{n})_{n\geq 1}$ satisfies the second equation of (\ref{hm}). We will show the first equation of (\ref{hm}) by induction. Since $\mu_{2}=l_{3}l_{1}+l_{2}>l_{1}+3=\mu_{1}+2+\frac{3-1}{2}$, the first equation of (\ref{hm}) holds. Now suppose $\mu _{2^{m}}>\sum_{i=1}^{2^{m}-1}\mu _{i}+2^{m}+\frac{3^{m}-1}{2}$. Then by (\ref{mu:rec}) and (\ref{sum:mu}), we have
\begin{eqnarray}
\mu_{2^{m+1}} &=& l_{3}\mu_{2^{m}}+l_{2}\nonumber\\
&\geq& l_{3}(\sum_{i=1}^{2^{m}-1}\mu _{i}+2^{m}+\frac{3^{m}-1}{2}+1)+l_{2}\nonumber\\
&=& \sum_{i=1}^{2^{m+1}-1}\mu_{i}-2^{m}(l_{1}+l_{2})+l_{3}(2^{m}+\frac{3^{m}+1}{2})+2l_{2}.\label{eqn:thm3:1}
\end{eqnarray}
Since $7l_{3}\geq 4(l_{1}+l_{2})+17$, we have $l_{3}-l_{1}-l_{2}\geq 2-\frac{3(l_{3}-3)}{4}$. Then
\begin{eqnarray}
& & 2^{m}(l_{3}-l_{1}-l_{2})+l_{3}\frac{3^{m}+1}{2}\nonumber\\
&>& 2^{m}\left(2-\frac{3(l_{3}-3)}{4}\right)+l_{3}\frac{3^{m}}{2}\nonumber\\
&\geq& 2^{m+1}+3^{m}\left(\frac{l_{3}}{2}-\frac{3}{4}\left(\frac{2}{3}\right)^{m}(l_{3}-3)\right)\nonumber\\
&\geq& 2^{m+1}+3^{m}\left(\frac{l_{3}}{2}-\frac{1}{2}(l_{3}-3)\right)=2^{m+1}+\frac{3^{m+1}}{2}.\label{eqn:thm3:2}
\end{eqnarray}
By (\ref{eqn:thm3:1}) and (\ref{eqn:thm3:2}), we know that $(\mu _{n})_{n\geq 1}$ satisfies (\ref{hm}).
\end{proof}
In the following, we give two examples to illustrate Theorem \ref{thm2}.
Example \ref{eg:const} gives a sum-free set
corresponding to a zero-one sequence generated by a substitution of constant length.
Example \ref{eg:nonconst} gives a sum-free set corresponding to a sequence generated
by a substitution of non-constant length.
\begin{example}\label{eg:const}
Take $\sigma:=\sigma(3,0,5)$, that is, $\sigma:~1\rightarrow 10001,~0\rightarrow
00000$. The sum-free set corresponding to
$\sigma^{\infty}(1)=100010000000000000001000100000\cdots$
is
$$S=\{1,6,24,29,110,115,133,138,528,533,551,556,637,642,660,665,\cdots\}.$$
According to Theorem \ref{thm2}, $S$ is $2$-regular.
\end{example}
\begin{example}\label{eg:nonconst}
Take $\sigma:=\sigma(1,1,5)$, that is $\sigma:~1\rightarrow 1010,~0\rightarrow
00000$. The sum-free set corresponding to
$$\sigma^{\infty}(1)=10100000010100000000000000000000000000000001010101000000\cdots$$
is
$$S=\{1, 3, 15, 17, 69, 71, 83, 85, 333, 337, 349, 353, 415, 417, 431, 435,\cdots\}.$$
According to Theorem \ref{thm2},
$S$ is $2$-regular.
\end{example}
From Theorem \ref{thm2}, we have the following two results.
In the following, the symbol $\equiv $ means equality modulo $2$.
\begin{cor}\label{cor:thue}
Let $(S_{n})_{n\geq 0}$ be the sum-free set in Theorem \ref{thm2}. For $%
0\leq j\leq 3$, the subsequence $(S_{4n+j})_{n\geq 0}$ is either periodic or
the Thue-Morse sequence up to a coding.
\end{cor}
\begin{proof}
Recall that by (\ref{sum:alpha}), we have
$$\sum_{i=1}^{2^n-1}\alpha_i=\frac{3^{n}-1}{2}.$$ Using (\ref{sum:mu}) several times, we have
\begin{eqnarray*}
\sum_{i=1}^{2^{n}-1}\mu_{i}&=&l_{3}\sum_{i=1}^{2^{n-1}-1}\mu_{i}+[2^{n-1}(l_{1}+l_{2})-l_{2}]\\
&=&l_{3}^{2}\sum_{i=1}^{2^{n-2}-1}\mu_{i}+l_{3}[2^{n-2}(l_{1}+l_{2})-l_{2}]+[2^{n-1}(l_{1}+l_{2})-l_{2}]\\
&=&\cdots \\
&=&\sum_{i=1}^{n}l_{3}^{i-1}[2^{n-i}(l_{1}+l_{2})-l_{2}]\\
&=&\frac{l_{1}+l_{2}}{l_{3}-2}(l_{3}^{n}-2^{n})-l_{2}\frac{l_{3}^{n}-1}{l_{3}-1}.
\end{eqnarray*}
By (\ref{alphan}), (\ref{mun}) and the previous two equations, we have
\begin{eqnarray*}
\sum_{i=1}^{2^{n-1}}(\mu _{i}+\alpha _{i})&=&\sum_{i=1}^{2^{n-1}-1}(\mu _{i}+\alpha _{i})+\mu_{2^{n-1}}+\alpha_{2^{n-1}}\\
&=&\frac{l_{1}+l_{2}}{l_{3}-2}(l_{3}^{n-1}-2^{n-1})-l_{2}\frac{l_{3}^{n-1}-1}{l_{3}-1}+\frac{3^{n-1}-1}{2}\\
&&+(\mu_{2^{n-1}}+\alpha_{2^{n-1}})\\
&=&\frac{(l_{1}+l_{2})(l_{3}^{n-1}-2^{n-1})}{l_{3}-2}%
+l_{1}l_{3}^{n-1}+3^{n-1}.
\end{eqnarray*}
Now by (\ref{hn}), we obtain
\begin{eqnarray}
h(n) &=&\left( \sum_{i=1}^{2^{n-1}}(\mu _{i}+\alpha _{i})\right) +2^{n-1}\nonumber\\
&=&\frac{(l_{1}+l_{2})(l_{3}^{n-1}-2^{n-1})}{l_{3}-2}%
+l_{1}l_{3}^{n-1}+3^{n-1}+2^{n-1}.\label{eqn:hn}
\end{eqnarray}%
Since $\frac{l_{3}^{n-1}-2^{n-1}}{l_{3}-2}=\sum_{i=0}^{n-2}l_{3}^{i}2^{n-2-i}\equiv l_{3}$ ($n\geq 3$), then
\begin{eqnarray*}
h(n) &\equiv &\frac{(l_{1}+l_{2})(l_{3}^{n-1}-2^{n-1})}{l_{3}-2}+l_{1}l_{3}+1
\\
&\equiv &(l_{1}+l_{2})l_{3}+l_{1}l_{3}+1 \\
&\equiv &1+l_{2}l_{3}.
\end{eqnarray*}
Let $(j)_{2}:=j_{2}j_{1}$ where $0\leq j\leq 3$. Then
\begin{eqnarray}
S_{4n+j} &=&1+\sum_{i=1}^{m}\epsilon _{i}h(i+2)+j_{2}h(2)+j_{1}h(1)\nonumber\\
&\equiv &1+(1+l_{2}l_{3})t_{n}+j_{2}h(2)+j_{1}h(1),\label{eqn:thue}
\end{eqnarray}%
where $t_n$ is the $n$-th term of the Thue-Morse sequence.
By (\ref{eqn:thue}), when $1+l_{2}l_{3}\equiv 0$, $(S_{4n+j})_{n\geq 0}$ modulo $2$ is a constant
sequence. When $1+l_{2}l_{3}\equiv 1$, $(S_{4n+j})_{n\geq 0}$ modulo $2$ is
the Thue-Morse sequence up
to a coding.
\end{proof}
\begin{example}
According to Corollary \ref{cor:thue}, the sum-free set $S$ (modulo $2$) in Example
\ref{eg:const} is the Thue-Morse sequence beginning by $1$. For
the sum-free set $S$ in Example \ref{eg:nonconst}, the subsequences
$(S_{2n})_{n\geq0}$ and $(S_{2n+1})_{n\geq0}$ modulo $2$ are both the Thue-Morse sequence beginning
by $1$.
\end{example}
\begin{cor}
Let $S$ be the sum-free set corresponding to the sequences of Cantor type (i.e.,
the fixed point of $\sigma (l,0,l+2)\ (l\geq 2)$ beginning by $1$).
\begin{enumerate}
\item When $l$ is odd, then $(S_{n})_{n\geq 0}$ modulo $2$ is the Thue-Morse
sequence $(1-t_{n})_{n\geq 0}$.
\item When $l$ is even, then $(S_{2n})_{n\geq 0}\equiv(S_{2n+1})_{n\geq 0}$
modulo $2$, and they are both the Thue-Morse sequence $(1-t_{n})_{n\geq 0}$.
\end{enumerate}
\end{cor}
\begin{proof}
In this case, $l_{1}=l,l_{2}=0$ and $l_{3}=l+2$. Thus by (\ref{eqn:hn}), $%
h(1)=l+2\equiv l$ and
\[
h(n)=l^{2}+3l+3^{n-1}\equiv 1,\ \ \ \forall~ n>1.
\]
When $l\equiv 1$, by (\ref{eqn:thue}),
\begin{eqnarray*}
S_{4n+j} &\equiv &1+t_{n}+j_{2}+j_{1} \\
&\equiv &1+(\sum_{i=1}^{m}\epsilon _{i}+j_{2}+j_{1}) \\
&\equiv &1+t_{4n+j}.
\end{eqnarray*}%
Thus for any $n\geq 0$, $S_{n}\equiv 1+t_{n}.$
When $l\equiv 0$, for $n\geq 0$ and $j=0,1$,
\begin{eqnarray*}
S_{2n+j} &=&1+\sum_{i=1}^{m}\epsilon _{i}h(i+1)+j h(1) \\
&\equiv &1+\sum_{i=1}^{m}\epsilon _{i}\equiv 1+t_{n}.
\end{eqnarray*}%
Thus for any $n\geq 0$, $S_{2n}\equiv S_{2n+1}\equiv 1+t_{n}.$
\end{proof}
While $\mu_n$ increases fast, the corresponding sum-free set is not complicated. In fact, we have
\begin{proposition}
Suppose the sequence $(\mu_n)_{n\geq1}$ is increasing and
\begin{equation} \label{munn}
\mu_{n+1}>2\sum_{i=1}^n\mu_i.
\end{equation}
Let $S_0=1$. Then for any $n\geq1$,
\[
S_n=\sum_{i=1}^n\mu_i+\frac{(n+1)(n+2)}{2}.
\]
\end{proposition}
\begin{proof}
By Lemma \ref{lm4}, for every $n\geq1$, we have $\alpha_{n+1}=\alpha_n+1$
and
\[
\alpha_n=\alpha_{n-1}+1=\alpha_{n-2}+1+1=\cdots=\alpha_1+(n-1)=n,
\]
since $\alpha_1=1$. Hence
\[
S_n=S_0+\sum_{i=1}^n\mu_i+\sum_{i=1}^n\alpha_i+n=\sum_{i=1}^n\mu_i+\frac{%
(n+1)(n+2)}{2}.
\]
\end{proof}
\begin{remark}
$(1)$ The growth order of $\mu _{n}$ is larger than $O(3^{n}).$ In fact, if $%
\mu _{n}=2\sum_{i=1}^{n-1}\mu _{i}$ for any $n\geq 3$, then $\mu _{n}=3\mu
_{n-1}$, and $\mu_2=2\mu_1$. Hence $\mu _{n}=2\times 3^{n-2}\mu _{1}.$
$(2)$ The coefficient $2$ in $(\ref{munn})$ is not crucial.
\end{remark}
\begin{remark}
Assume $(S_{n})_{n\geq 0}$ is a sequence given by $S_{n}=1+\sum_{i=1}^{m}%
\epsilon _{i}h(i)$ where $(n)_{2}:=\epsilon _{m}\cdots \epsilon_1 $ and $%
(h(i))_{i\geq 1}$ is a positive integer sequence. If $h(i)\equiv 1$ for $%
i\geq 0$, then $(S_{n})_{n\geq 0}$ modulo $2$ is the Thue-Morse sequence.
If $h(i)\equiv 0 $ for $i\geq 0$, then $S_{n}\equiv 1$.
\end{remark}
\begin{example}
Let $h(i)=2^{i}$, then $S$ is sum-free. Moreover, $S=\{1,3,5,7,9,\cdots \}$
and it is periodic of period $2$.
\end{example}
\begin{example}
Let $h(i)=(i+1)!$, then $S$ is sum-free. Moreover,
\[
S=\{1,3,7,9,25,27,31,33,\cdots \}
\]%
and it is periodic of period $4$.
\end{example}
\bigskip
\section{Base changing}
The mapping $\theta :\Sigma \rightarrow \mathfrak{S}$ is bijective, so it is
natural to study the properties of the corresponding zero-one sequences
of some sum-free sets. In this section, we will show that the corresponding
zero-one sequences of the sum-free sets defined below are automatic.
\begin{definition}
\label{def1} For any $b\geq 2,~n\geq 0$, let $S=(S_{n})_{n\geq 0}$ be the
sum-free set given by $S_{n}=[(n)_{b}1]_{2b-1}$.
\end{definition}
\begin{definition}[Allouche and Shallit \protect\cite{jps}]
Let $\mathcal{A}$ be a set of non-negative integers. Then we say that $%
\mathcal{A}$ is a $k$\emph{-automatic set} if its characteristic sequence
\[
a_{n}=\left\{
\begin{array}{ll}
1, & \hbox{if $n\in \mathcal{A}$;} \\
0, & \hbox{otherwise,}%
\end{array}%
\right.
\]%
defines a $k$-automatic sequence.
\end{definition}
\begin{lemma}
\label{lem2} The set $S$ in Definition \ref{def1} is sum-free and $(2b-1)$%
-automatic.
\end{lemma}
\begin{proof}
By the definition of $S_n$, it is clear that for any integer $n\geq0$, $%
(S_n)_{2b-1}\in\Sigma_b^{\ast}1$. Hence, for any integers $m,~n\geq0$, $%
(S_m+S_n)_{2b-1}\in\Sigma_{2b-1}^{\ast}2$, which implies that $%
S_m+S_n\not\in S$. Thus, the set $S$ is sum-free.
Let $(a_{n})_{n\geq 0}$ be the characteristic sequence of $S$. Then
\[
a_{n}=\left\{
\begin{array}{ll}
1, & \hbox{if $(n)_{2b-1}\in\Sigma_b^{\ast}1$;} \\
0, & \hbox{otherwise.}%
\end{array}%
\right.
\]%
Thus the sequence $(a_{n})_{n\geq 0}$ can be generated by the $(2b-1)$%
-automaton in Figure 1, which implies that the sequence $(a_{n})_{n\geq 0}$
is $(2b-1)$-automatic.
\begin{figure}[tbp]
\centering
\begin{tikzpicture}[shorten >=1pt,node distance=2cm,on grid,>=stealth,every state/.style={draw=blue!50,very thick,fill=blue!20}]
\node[state,initial] (q_0) {$q_0/0$};
\node[state] (q_1) [below left=of q_0,yshift=-1.5cm,xshift=-2cm] {$q_1/1$};
\node[state] (q_2) [below right=of q_0,yshift=-1.5cm,xshift=2cm] {$q_2/0$};
\path[->]
(q_0) edge [bend right=15] node [above left] {1} (q_1)
edge node [above right] {$b,\cdots,2b-2$} (q_2)
(q_1) edge [bend right=15] node [below right] {$0,2,\cdots,b-1$} (q_0)
edge node [below ] {$b,\cdots,2b-2$} (q_2)
(q_0) edge [loop above] node {$0,2,\cdots,b-1$} ()
(q_1) edge [loop below] node {1} ()
(q_2) edge [loop below] node {$0,1,\cdots,2b-2$} ();
\end{tikzpicture}
\label{AUT}
\caption{Automaton generating the set $(a_n)_{n\geq 0}$ in Lemma \protect\ref%
{lem2}.}
\end{figure}
\end{proof}
\begin{theorem}
\label{the1} The zero-one sequence $\mathbf{c}$ corresponding to the
sum-free set $S$ in Definition \ref{def1} is $(2b-1)$-automatic.
\end{theorem}
\begin{proof}
Recall that the zero-one sequence $\mathbf{c}$ is obtained by deleting
all the $\ast $'s of the sequence $\mathbf{v}=(v_{n})_{n\geq 1}$. For any $%
n\geq 1$, we claim that
\begin{enumerate}
\item $v_n=1 \Leftrightarrow (n)_{2b-1}\in\Sigma_b^{\ast}1,$
\item $v_n=\ast \Leftrightarrow (n)_{2b-1}\in\Sigma_{2b-1}^{\ast}2,$
\item $v_n=0 \Leftrightarrow
(n)_{2b-1}\in\Sigma_{2b-1}^{\ast}\backslash(\Sigma_b^{\ast}1\cup%
\Sigma_{2b-1}^{\ast}2).$
\end{enumerate}
The third assertion is an immediate consequence of the first one and
the second one, and the first assertion follows directly from the
definitions of $S$ and $v_{n}$ in (\ref{vn}). Therefore it suffices to prove
the second assertion.
Since $v_n=\ast \Leftrightarrow n\in S+S$, we need to show that
\begin{equation} \label{star}
S+S=(2b-1)\mathbb{N}+2.
\end{equation}
For any $S_n\in S$, then $S_n=(2b-1)([(n)_b]_{2b-1})+1.$ Hence, for any
integer $m,n\geq1$, we have
\[
S_m+S_n=(2b-1)([(m)_b]_{2b-1}+[(n)_b]_{2b-1})+2\in(2b-1)\mathbb{N}+2.
\]
Conversely, for any $n\geq 0$, assume $(n)_{2b-1}:=a_{k}a_{k-1}\cdots a_{1}$%
. Hence, for any $1\leq i\leq k$, there exist $b_{i},d_{i}\in \Sigma _{b}$
such that $a_{i}=b_{i}+d_{i}$. Thus, there exist two integers
\begin{eqnarray*}
n_{1} &=&[b_{k}b_{k-1}\cdots b_{1}]_{2b-1}, \\
n_{2} &=&[d_{k}d_{k-1}\cdots d_{1}]_{2b-1},
\end{eqnarray*}%
such that $n=n_{1}+n_{2}$ and $(n_{1})_{2b-1},(n_{2})_{2b-1}\in \Sigma
_{b}^{\ast }.$ Moreover,
\begin{eqnarray*}
(2b-1)n+2 &=&((2b-1)n_{1}+1)+((2b-1)n_{2}+1) \\
&=&S_{[b_{k}b_{k-1}\cdots b_{1}]_{b}}+S_{[d_{k}d_{k-1}\cdots d_{1}]_{b}}\in
S+S.
\end{eqnarray*}%
This implies (\ref{star}) holds.
Now, we will show that $\mathbf{c}$ is $(2b-1)$-automatic. By (\ref{star}%
), $v_{(2b-1)n+2}=\ast $ for $n\geq 0$. Hence, $\mathbf{c}%
=c_{0}c_{1}\cdots $ satisfies
\begin{equation}
\left\{
\begin{array}{lll}
c_{(2b-2)n} & = & v_{(2b-1)n+1}; \\
c_{(2b-2)n+i} & = & v_{(2b-1)n+i+2},%
\end{array}%
\right. \label{77}
\end{equation}%
where $1\leq i\leq2b-3.$ By Lemma \ref{lem2}, $S$ is $(2b-1)$-automatic. Its
characteristic sequence $(a_{n})_{n\geq 0}$ is a $(2b-1)$-automatic sequence.
By Theorem 6.8.1 in \cite{jps}, $(a_{(2b-1)n+i})_{n\geq 0}$ is $(2b-1)$%
-automatic for $0\leq i\leq 2b-2$. Note that $v_{n}=a_{n}$ for any $n\notin $
$(2b-1)\mathbb{N}+2$, then $(v_{(2b-1)n+i})_{n\geq 0}$ is $(2b-1)$-automatic
for $0\leq i\leq 2b-2$ and $i\neq 2$. Thus $(c_{(2b-2)n+i})_{n\geq 0}$ is $%
(2b-1)$-automatic for $0\leq i\leq 2b-3$. By Theorem 6.8.2 in \cite{jps}, $%
\mathbf{c}$ is $(2b-1)$-automatic.
\end{proof}
\begin{remark}\label{rem:cantor}
$(1)$ $(2b-1)$ in Theorem \ref{the1} cannot be replaced by $p$, where $p>2b-1$%
, since there do not exist $x,y\in \mathbb{N}$ such that $S+S=x\mathbb{N}%
+y$. \newline
$(2)$ From Theorem \ref{the1}, if $b=2$, then the sequence $\mathbf{c}$ is the
Cantor sequence
\[
101000101000000000101000101\cdots
\]%
$(3)$ If we replace $S_{n}=[(n)_{b}1]_{2b-1}$ by $S_{n}=[(n)_{b}w]_{2b-1}$
where $w\in \Sigma _{b}^{\ast }$, then the corresponding zero-one sequence
is also automatic.
\end{remark}
\medskip
\noindent\textbf{Acknowledgements }
We would like to thank Professor Jacques Peyri\`{e}re for introducing us to this topic, and we gratefully acknowledge
of his many helpful suggestions. We would also like to thank the anonymous
referee for many helpful comments.
|
\section{Conclusion and Discussion}
\vspace*{0.3cm}
\noindent
{\it{\textbf{Discussion and Conclusions.}}}
As pointed out in Ref.~\cite{BDFG2003}, the string theory seems unlikely to realize a large effective decay constant for the natural inflation model. A more general argument has been provided from the weak gravity conjecture~\cite{AHMNV2006} motivated by arguments involving holography, the absence of Planck scale remnants and the incompatibility of global symmetries with quantum theories of gravity. It states that in an effective 4d theory with gravity and a $U(1)$ gauge field with a small gauge coupling or only mill-charged matters, there exist a ``hidden cutoff", $\Lambda_{\rm hidden} = q_{\rm eff} M_{\rm pl} \sim 10^{-3}\,M_{\rm pl}$, in the notation of our model. As can be seen from Fig.~\ref{fig:charge-r}, the current experimental data prefer to have the compactification scale $1/R \gtrsim 0.06\,M_{\rm pl}$ and above $\Lambda_{\rm hidden}$, so the weak gravity conjecture is violated from the 4d point of view. For the 5D theory, no tiny charges have been introduced in the original basis of $U(1)$'s. However, unless there are some other properties to distinguish those $U(1)$'s, one still has milli-charge from the beginning in the rotated basis. One may argue that gauge boson masses at lower energy scales can distinguish those $U(1)$'s without disturbing the inflation part. Unfortunately, the hierarchic scale separations between gauge boson masses and the compactification scale regenerate fine-tuning problems for the Higgs fields charged under $U(1)$'s, similar to the ``hierarchy problem" of the electroweak sector in the Standard Model.
In summary, we have constructed and analyzed a natural inflation model with multiple PNGB's as fifth components of 5D Abelian gauge bosons. A seesaw structure of the charge matrix is introduced to obtain a milli-charge for one linear combination of gauge fields. A trans-Planckian effective decay constant has been achieved with all scales kept sub-Planckian. The tensor-to-scalar ratio is preferred to be in the range of $0.033< r < 0.125$ $(0.066 < r < 0.146)$ for 60(50) $e$-folds, while the reheating temperature is a few $10^{11}$~GeV.
\vspace{3mm}
{\it{\textbf{Acknowledgements.}}}
We thank Haipeng An, Hsin-Chia Cheng, Aki Hashimoto, Josh Ruderman, Henry Tye and Xin-min Zhang for useful discussion. This work is supported by the U. S. Department of Energy under the contract DE-FG-02-95ER40896. YB thanks the Center for Future High Energy Physics, where this work is finished.
\providecommand{\href}[2]{#2}\begingroup\raggedright |
\section{Introduction}
For modeling the kinetic energy of a non-relativistic
spin-$\tfrac{1}{2}$ particle in the plane, moving
under a magnetic field $B$ in the perpendicular direction
to the plane, one uses the two-dimensional Pauli operator
$$ H_{\bf A} :=
\big[{\boldsymbol \sigma}\cdot
\big(- \ri \nabla- {\bf A}\big) \big]^2 =
\big(- \ri \nabla- {\bf A}\big)^2 - \sigma_3 B
\quad \mbox{on}
\ \hilbert ,
$$
where ${\bf A}$ is a vector potential associated to $B$,
i.e. $B=\curla :=\partial_1A_2 -\partial_1 A_1$.
Here, ${\boldsymbol \sigma} =(\sigma_1,\sigma_2)$ and
$\sigma_3$ are the Pauli matrices
\begin{align*}
\sigma_1 =
\begin{pmatrix}
0 & 1 \\ 1 & 0
\end{pmatrix} , \quad
\sigma_2 =
\begin{pmatrix}
0 & -\ri \\ \ri & 0
\end{pmatrix}, \quad
\sigma_3 =
\begin{pmatrix}
1 & 0 \\ 0 & -1
\end{pmatrix}.
\end{align*}
To study the behaviour of such spin-$\tfrac{1}{2}$ particles
(e.g. electrons)
in presence of an additional electric potential $V$, we
investigate the spectrum of the operator
$$
H := H_{\bf A} + V =
\big[{\boldsymbol \sigma}\cdot
\big(- \ri \nabla- {\bf A}\big) \big]^2 + V
\quad \mbox{on}
\ \hilbert.
$$
If $V$ is non-negative or decays at infinity
(e.g. potentials with Coulomb singularities), spectral
properties of the magnetic Schr\"odinger operator
$\big(- \ri \nabla- {\bf A}\big)^2 + V$,
as well as of the Pauli operator $\HA + V$
(in dimension d = 2 or 3)
have been widely studied over the last decades
(see, e.g. \cite{Avron_Herbst_Simon_1}, \cite{Kirsch_simon} or
\cite{Erdoes_overview} for a latest overview).
In this article, instead, we want to point out some
interesting features of the spectrum of $H$ for potentials $V$
tending to $-\infty$ as $|\bx| \to \infty$. Since such
scalar potentials result in an operator $H$, unbounded from below,
it is necessary to discuss questions related to the self-adjointness
of $H$. We emphasize that, since we also consider unbounded
magnetic fields $B$, the self-adjointness of $H$ cannot simply
be reduced to the one of the magnetic Schr\"odinger operator.
One motivation for the following considerations is an observation
made in \cite{MehringStock} for the two-dimensional massless
magnetic Dirac operator coupled to an electric potential $V$;
There, an accumulation process of spectral points has been observed,
governed by the ratio $|V^2/B|$ at infinity. This phenomenon can be
ascribed to the non-confining effect of $V$ in the case of the Dirac
operator. Regarding the Pauli operator, the influence of an additional
scalar potential $V$ on the spectrum $\sigma(H)$ depends crucially
on the sign of $V$. For simplicity, we outline this dependence in the
case of a constant magnetic field $B(x) = B_0$:
a positive potential $V$ growing at infinity
always leads to discrete spectrum of the operator $H$,
independently of the field strength $B_0$
(see e.g. \cite{kondratiev2005}). Such trapping potentials
only enhance a localization effect caused by $B$ that
generates eigenvalues and spectral gaps (proportional to $B_0$).
If we instead consider negative potentials $V$, the situation is
quite different since the particle lowers its energy by moving
in regions where $V$ is small.
A scalar potential $V$ converging to $- \infty$ as $|\bx| \to \infty$
has therefore a delocalizing effect, i.e. the particle tends to escape
any compact region of the plane. Our results show that such negative
potentials $V$ (describing for example constant radial fields)
counteract the localizing effect of ``hard'' magnetic fields $B$, as
they close spectral gaps induced by $B$:
\begin{itemize}
\item If $V$ converges to $-\infty$, but remains small compared
to $B$, the spectrum $\sigma(H)$ is discrete, i.e. it
consists only of eigenvalues of finite multiplicity.
\item If $V$ is comparable to $B$, more precisely $|V | \approx 2B$ at
infinity, points in the essential
spectrum occur.
\item If $V$ overtakes $B$, more explicitly $|V/B| \to \infty$ as
$|\bx| \to \infty $ (at least along a path), the spectrum
$\sigma(H)$ covers the whole real line.
\end{itemize}
One may compare the third claim with the result in \cite{SimonMiller80}
on $\HA$ for decaying magnetic fields.
The precise statements of the claims above are contained in
Theorems \ref{thm1}$-$\ref{thm4} of Sect. \ref{mainresults}.
We remark that the case $|V/B| \to \infty$ as
$|\bx| \to \infty$ is treated by Theorems \ref{thm3} and \ref{thm4}.
Unlike Theorem \ref{thm4}, which is only valid for constant
magnetic fields, Theorem \ref{thm3} covers also non-constant fields $B$,
but requires stronger constraints on the growth of $V$. Thus,
the important case $B =B_0$ is adressed by two theorems.
The ideas of the proofs of Theorem \ref{thm1}$-$\ref{thm3} originate
from those used to prove the results in \cite{MehringStock}. However,
since we work with a second-order operator, the proofs are technically
more laborious. Theorem \ref{thm4} is based on a further
construction of a Weyl sequence, obtained by treating $V$
locally as a potential of a constant electric field. This is a refined
ansatz compared to the method used
for the proof of Theorem \ref{thm3}.\\
{\it The organization of this article is as follows:} In the next
section some known facts about the Pauli operator are recapitulated.
We present our precise results in Section \ref{mainresults},
provided with some remarks and important applications.
In Section \ref{prooft1} we give the proof of Theorem \ref{thm1}.
The proofs of Theorems \ref{thm2} and \ref{thm3} are contained in
Section \ref{prooft2}, while the proof of Theorem \ref{thm4}
can be found in the last section.
In the appendix, attached to the main text, we give a proof
of the essential self-adjointness of the Pauli operator.
\section{Basic properties of the Pauli operator}\label{susy}
In this section we point out some basic facts about the Pauli operator
and the massless Dirac operator $\DA$, whose square equals $\HA$.
For a vector potential ${\bf A} \in C^1(\R^2,\R^2)$ generating the field
$B =\curla \in C(\R^2, \R)$,
the Hamiltonian $\DA$ is defined as the closure of the operator
\begin{align}
\label{eq:5}
{\boldsymbol \sigma}\cdot \big(- \ri \nabla- {\bf
A}\big)=\left(
\begin{array}{cc}
0&d^*\\
d&0
\end{array} \right)
\quad
\mbox{on} \ \core,
\end{align}
which is essentially self-adjoint on the given core (see
\cite{Chernoff77}). In particular, $d$ and $d^*$ can be
seen as closed operators, i.e. we use the notation
\begin{align}\label{carolina}
d =\overline{-\ri \partial_1 -A_1+
\ri(-\ri \partial_2 -A_2)
\upharpoonright}_{C^\infty_0(\R^2,\C)}
\end{align}
and analogously for $d^*$.
One observes that $d, d^*$ satisfy the commutation relation
\begin{align}
\label{eq:1}
[d,d^*]\varphi:=(dd^*-d^*d)\varphi=2B\varphi
\quad {\rm for}\ \varphi \in C^\infty_0(\R^2,\C).
\end{align}
We can write
$$\HA = \DA ^2 = \left(
\begin{array}{cc}
d^*d&0\\
0&dd^*
\end{array}
\right)
\quad
\mbox{on} \ \core $$
and consider $\HA$ as a self-adjoint operator on
$\{ \psi \in {\mathcal D}(\DA) \, | \, \DA \psi \in {\mathcal D} (\DA) \}$
given by the Friedrichs extension.
The two components $dd^*$ and $d^*d$ of $\HA$ are unitarily
equivalent on the orthogonal complement of
${\rm ker}(\HA) ={\rm ker}(\DA)$.
To verify this we note first that, due to
the matrix structure of $\DA$, we have
\begin{align}
\label{eq:8}
{\rm sgn}({\DA}):=\frac{\DA}{|\DA|}=\left(\begin{array}{cc}
0&s^*\\
s&0
\end{array}\right)
\end{align}
on
${\rm ker} (\DA)^\perp=
{\rm ker}(d)^\perp\oplus {\rm ker}(d^*)^\perp$.
Since $ {\rm sgn}({\DA})^2 = {\rm Id}$ on
${\rm ker}(\DA)^\perp$, the maps
\begin{align}\label{super0}
&s:{\rm ker}(d)^\perp\to{\rm ker}(d^*)^\perp, \qquad
s^*: {\rm ker}(d^*)^\perp\to {\rm ker}(d)^\perp
\end{align}
are unitary and conjugated to each other. By the operator
identity $\HA = \DA^2 = \SA \DA^2 \SA $ one concludes that
\begin{align}
\label{super}
\left(
\begin{array}{cc}
d^*d&0\\
0&dd^*
\end{array}
\right)\varphi = \left(\begin{array}{cc}
s^*dd^*s&0\\
0&sd^*ds^*
\end{array}
\right) \varphi
\end{align}
for any
$\varphi=(\varphi_1,\varphi_2)^{\rm T}$
with $\varphi_1\in \mathcal{D}({d^*d})\cap {\rm ker}(d)^\perp$ and
$\varphi_2\in \mathcal{D}({dd^*})\cap {\rm ker}(d^*)^\perp$. Hence,
on $\ker (d)^\perp$ the operator $d^*d$ is unitarily
equivalent to $dd^*$ (considered as an operator
on $\ker (d^*)^\perp$). Let us denote the orthogonal
projection on ${\rm ker} (\DA)$ by $P_0$ and the orthogonal
projections on ${\rm ker} (d)$, ${\rm ker} (d^*)$ by $\pi, \pi_*$.
Further, we set
$$ P_0^\perp := \mathbbm{1}- P_0 , \quad
\pi^\perp := \mathbbm{1}- \pi , \quad
\pi_*^\perp := \mathbbm{1}- \pi_* . $$
To define our full Hamiltonian let
${\bf A} \in C^1(\R^2,\R^2)$ and $B,V \in C(\R^2,\R)$
be such that $B=\curla$, then $H$ is given by
$$H \varphi = \big[\DA ^2 + V\big] \varphi =
\big[ \big(-\ri \nabla -{\bf A}\big)^2 - \sigma_3 B + V \big]
\varphi \quad
\mbox{for} \ \varphi \in \core\,. $$
In general, the closure of this densely defined operator is not
self-adjoint without any restriction on the growth rate of $V$
at infinity. However, there are conditions, very similar to those
for the classical Schr\"odinger operator, to ensure essential
self-adjointness.
\begin{proposition}\label{EssSelf}
Let $B, V \in C^1(\R^2,\R)$
and ${\bf A}\in C^2(\R^2,\R^2)$
with $B=\curla$. In addition,
assume that $V$ fulfills the lower bound
\begin{align}\label{growthcond}
V(\bx) \ \ge \ -c|\bx|^2 + d\,, \quad \ \bx \in \R^2,
\end{align}
for some constants $c>0$, $d \in \R$.
Then $H$ is essentially self-adjoint on $\core$.
\end{proposition}
\begin{remark}\label{relaxregularity}
Following the lines of the proof given in Appendix \ref{selfadj},
we see that the regularity condition on
$B, V$ can be relaxed to $B,V \in C^\alpha_{loc}(\R^2,\R)$, i.e.
both only need to be locally $\alpha$-H\"older continuous.
By a perturbation argument one can also see that it suffices
to assume that $B,V$ are $C^\alpha_{loc}$ outside some compact set
$K\subset \R^2$, while inside $K$ they only need to be continuous.
\end{remark}
\begin{remark}
The self-adjoint operator given by Proposition \ref{EssSelf}
is locally compact, i.e. for any characteristic function
$\chi_{B_R(0)}$ on the ball $B_R(0)$ with radius $R$, the operator
$\chi_{B_R(0)} (H-\ri)^{-1} $
is compact.
\end{remark}
\begin{remark}
Considering the case $V=0$, we obtain that $\HA$,
$dd^*$ and $d^*d$ are essentially self-adjoint on
$\core$, respectively on $C_0^\infty(\R^2,\C)$.
\end{remark}
Note that \eqref{growthcond} is the same lower bound on $V$ as one needs
for the (magnetic) Schr\"odinger operator to ensure the essential
self-adjointness, whereas no restriction on the growth of $B$ is necessary.
The regularity conditions on $V$ and ${\bf A}$ are quite strong compared
to those for the magnetic Schr\"odinger operator (see \cite{Reed_Simon_2}).
The reason is that due to the lack of a diamagnetic inequality for $\HA$,
one uses a direct argument that requires more regularity on the potentials
${\bf A}$ and $V$. The interesting question remains: Could one
relax these conditions for the Pauli operator as in the case of the
magnetic Schr\"odinger operator?
\section{Main results}\label{mainresults}
In this section we assume that $B,V$ and ${\bf A}$ satisfy
the conditions of Proposition \ref{EssSelf}. It is easy to see
that in the following results
$B,V \in C^1(\R^2,\R)$ can be relaxed to hold
only outside some compact set $K \subset \R^2$
as in Remark \ref{relaxregularity}.
\begin{theorem}\label{thm1}
Assume that
\begin{eqnarray}
&&V(\bx)\longrightarrow
-\infty\quad\mbox{as}\quad|\bx|\to \infty,\label{t1}\\
&&\left|\frac{\nabla V (\bx)}{V(\bx)}\right| \,
\longrightarrow 0\quad\mbox{as}\quad|\bx| \to \infty,\label{t2}\\
&&\limsup_{|\bx|\to \infty}
\left| \frac{V(\bx)}{2B(\bx)} \right| <1. \label{t3}
\end{eqnarray}
Then $\sigma_{\rm ess}(H)=\emptyset$, i.e.
$H$ has purely discrete spectrum.
\end{theorem}
Condition \eqref{t2} is a restriction on the growth rate of $V$
and rather of technical necessity.
The interplay between $B$ and $V$ (as mentioned in the
introduction) is described by condition \eqref{t3}.
Thus, it is worthwhile to investigate further the dependence
of $\sigma(H)$ on this quotient:
One can easily observe that if the quotient of \eqref{t3} surpasses
the constant $1$, the spectrum of $H$ changes its character.
To see this pick $\Omega \in {\rm ker}(d^*d)$, then
$$(dd^*+V)\Omega = (d^*d+ 2B +V)\Omega \approx 0$$
if $2B \approx -V$. Therefore, if ${\rm ker}(d^*d)$ contains
enough functions
(which is the case for fields $B$ bounded form below by some
positive constant), we obtain points in the essential
spectrum of $H$.
One can even show that the condition $2B \approx -V$ (at infinity)
does not need to hold globally for obtaining
$\sigma_{\rm ess}(H)\neq \emptyset$. We demonstrate this for a certain
class of fields $B$ and potentials $V$.
\begin{definition}
A function $f:\R^2\to\R$ varies with rate $\nu \in [0,1]$ on a
set $X \subset \R^2$ if there is a constant $C>0$ such that
for all $\bx \in X$ it holds that
\begin{align*}
|f(\bx +\by)|\le C |f(\bx)|,
\end{align*}
whenever $\by \in \R^2$ satisfies $|\by| \le \tfrac{1}{2}|\bx|^\nu$
\end{definition}
Note that functions of the form $f_1(\bx) = c|\bx|^s$ and
$f_2(\bx)= c|x_1|^s$, with $c,s \in \R$, vary with any rate
$\nu \in [0,1]$ on $\R^2\backslash B_1(0)$ and
on $\R^2\backslash [-1,1]\times \R$ respectively.
\begin{theorem}\label{thm2}
Assume that there is a sequence
$(\bx_n)_{n\in\N}$ with $|\bx_n|\to\infty$ as $n\to\infty$
and constants $k\in \N$, $\varepsilon\in(0,1)$
such that $|\nabla V|, |\nabla B|$ vary with rate $0$
on $(\bx_n)_{n \in \N}$, as well as
\begin{eqnarray}
\label{con0a}
&&V(\bx_n)\longrightarrow - \infty,\\
\label{con1a}
&& \frac{|\nabla B(\bx_n)|^2}{|B(\bx_n)|^{1-\varepsilon}},
\,\frac{|\nabla V(\bx_n)|^2}{|V(\bx_n)|^{1-\varepsilon}}
\longrightarrow 0,
\\
\label{con2a}
&&V(\bx_n)+2k|B(\bx_n)|\longrightarrow 0
\end{eqnarray}
as $n \to\infty$. Then $0\in \sigma_{\rm ess}(H)$.
\end{theorem}
Let us now consider the case $V \gg B$ at infinity. The next
two theorems state that the accumulation of eigenvalues intensifies,
creating more points in the essential spectrum and closing spectral gaps.
\begin{theorem}\label{thm3}
Assume that there is a continuous path $\gamma: \R^+ \to \R^2$,
with $|\gamma(t)| \to \infty$ as $t \to \infty$, and constants
$\epsilon > 0$, $\nu \in [0,1]$ such that
$|\nabla V|, |\nabla B|$ vary with rate $\nu$ on
$\mathrm{Im}(\gamma)$, as well as
\begin{eqnarray}
\label{con3.1}
&& \frac{V(\gamma(t))}{2|B(\gamma(t))|} \longrightarrow -\infty \\
\label{con3.2}
&& \left(
\frac{|\nabla B(\gamma(t))|}{|B(\gamma(t))|} +
\frac{|\nabla V(\gamma(t))|}{|V(\gamma(t))|} \right)
\left(\frac{|V(\gamma(t))|^3}{B^2(\gamma(t))}\right)^{\frac{1+\epsilon}{2}}
\longrightarrow 0,
\\[0.1cm]
\label{con3.3}
&&\frac{1}{|\gamma(t)|^{2\nu}}
\left(\frac{|V(\gamma(t))|}{B^2(\gamma(t))}\right)^{1+\epsilon}
\longrightarrow 0
\end{eqnarray}
as $t \to \infty$. In addition, suppose that
for all $t \in (0,\infty)$ the inequality
\begin{eqnarray}
\label{con3.4}
&& B_0\le|B(\gamma(t))|\le
\alpha \exp \left(\kappa\left| \frac{V(\gamma(t))}{B(\gamma(t))}\right|\right)
\end{eqnarray}
holds with constants $\alpha, \kappa, B_0 >0$. Then $\sigma_{\rm ess}(H) = \R$.
\end{theorem}
For our main application, potentials of power-like growth
(see discussion after the next theorem), condition \eqref{con3.3}
imposes unsatisfying restrictions on the growth rate of $V/B$.
At least in the case of a constant magnetic field they can be
weakened.
\begin{theorem}\label{thm4}
Let $B = B_0 > 0$ and $V \in C^2(\R^2,\R)$. Assume that there is a
continuous path $\gamma: \R^+ \to \R^2$, with
$|\gamma(t)| \to \infty$ as $t \to \infty$, and constants
$\epsilon >0 $, $\nu \in [0,1]$ such that
the matrix norm of the Hessian matrix
$\|\mathrm{Hess}(V)\|_2 : \R^2 \to \R$
varies with rate $\nu$ on $\mathrm{Im}(\gamma)$, as well as
\begin{eqnarray}
\label{con4.1}
&& V(\gamma(t))\longrightarrow -\infty , \\[1.3 mm]
\label{con4.2}
&&\|\mathrm{Hess}(V)\|_2 (\gamma(t)) |V(\gamma(t))|^{1+\epsilon}
\longrightarrow 0,
\\
\label{con4.3}
&&\frac{1}{|\gamma(t)|^{2\nu}} \,
|V(\gamma(t))|^{1+\epsilon}
\longrightarrow 0
\end{eqnarray}
as $t \to \infty$. In addition, let
\begin{eqnarray}
\label{con4.4}
&&\limsup_{t \to \infty} \frac{|\nabla V(\gamma(t))|^2}{|V(\gamma(t))|}
< (2B_0)^2.
\end{eqnarray}
Then $\sigma_{\rm ess}(H) = \R$.
\end{theorem}
\begin{remark}
Note that a well-known, basic example for this last theorem
is the case of a constant electric field $ {\mathcal E}_0$ in
$x_1$-direction with the corresponding
potential $V(\bx) = {\mathcal E}_0 x_1$.
\end{remark}
\begin{remark}
Results similar to that of Theorems \ref{thm1}$-$\ref{thm4} can be
obtained for the magnetic Schr\"odinger operator with scalar
potentials $V$ by using the same techniques as in the proofs of
Theorems \ref{thm1}$-$\ref{thm4}.
\end{remark}
Finally, we want to discuss some consequences of our results,
in particular with respect to spherically symmetric fields $B$ and
potentials $V$, i.e. $B(\bx) = b(|\bx|)$, $V(\bx) = v(|\bx|)$
for $\bx \in \R$. Using the rotational gauge
\begin{equation*}\label{goischt}
{\bf A}({\bf x}) := \frac{A(r)}{r}
\begin{pmatrix} -x_2 \\ x_1 \end{pmatrix}\,, \quad
A(r) = \frac{1}{r} \int_0^r b(s) s \mbox{d}s,
\end{equation*}
with $r=|\bx|$, we decompose $H$ in a direct sum of
operators on the half-line. More explicitly, there is
a unitary map
$$U: L^2(\R^2,\C^2) \to \bigoplus_{j\in \Z}L^2(\R^+,\C^2; \mbox{d}r)$$
such that
$ UHU^* =\bigoplus_{j\in \Z} h_j$,
with
\begin{align*}
h_j :=
\begin{pmatrix}
-\partial_r^2 + \frac{j^2-1/4}{r^2} & 0 \\
0 & -\partial_r^2 + \frac{(j+1)^2-1/4}{r^2}
\end{pmatrix}
+ A^2(r) - \frac{m_j}{r}A(r) + \sigma_3 A'(r)+ v(r)
\end{align*}
on $L^2(\R^+,\C^2; \mbox{d}r)$, where $m_j = j +\tfrac{1}{2}$
(see e.g. \cite{Thaller}). It is easy to verify that if
\begin{eqnarray}
\label{condRot.1}
&& \liminf_{r \to \infty} b(r) > 0, \\
\label{condRot.2}
&& A' (r) /A^2(r) \longrightarrow 0 \quad \mbox{as} \ r \to \infty ,\\[1.5mm]
\label{condRot.3}
&& \limsup_{r \to \infty} |v(r)|/A^2(r) < 1,
\end{eqnarray}
then $h_j$ has purely discrete spectrum for every $j \in \Z$.
As a consequence, one can use the relations
\begin{align*}
\sigma_{\#}(H)=
\overline{\bigcup_{j\in \Z} \sigma _{\#}{(h_j)}}
\,, \quad \quad \quad
\# \in \{ \mathrm{ac},\ \mathrm{sc}, \ \mathrm{pp} \}
\end{align*}
to conclude that $\sigma (H)=\sigma_{\mathrm{pp}}(H)$,
$\sigma_{\mathrm{ac}}(H)= \sigma_{\mathrm{sc}}(H) = \emptyset$
if \eqref{condRot.1}$-$\eqref{condRot.3} are satisfied.
To get more information on $\sigma (H)=\sigma_{\mathrm{pp}}(H)$,
we employ Theorems \ref{thm1}$-$\ref{thm4} and obtain:
\begin{corollary}
Let $b(r) = b_0 r^s$, $v(r) =v_0r^t$ with $v_0 < 0 < b_0$
and exponents $0\le s$, $0\le t \le 2$. Then
\begin{itemize}
\item[a)] $\sigma (H)$ is purely discrete if $0< t < s$ or
$0< t = s$ and $|v_0| < 2b_0$,
\item[b)] $0 \in \sigma_{\rm ess} (H)$ if $0< t=s$ and $|v_0| = 2kB_0$
for some $k \in \N$,
\item[c)] $\sigma (H) = \R$ is dense pure point
if $3s<3t < 2(s+1)$,
\item[d)] $\sigma (H) = \R$ is dense pure point
if $s=0$ and $0< t < 1$.
\end{itemize}
\end{corollary}
The origins of the strong restrictions on $s,t$ in c), d)
can easily be tracked back to conditions \eqref{con3.2},
\eqref{con3.3} of Theorem
\ref{thm3} and \eqref{con4.2} of Theorem \ref{thm4}. Unfortunately,
even in the case of a constant magnetic field ($s = 0$) we cannot
cover the full range of potentials ($0 < t \le 2$) for which one
might expect $\sigma (H) = \R$. \\
\section{Proof of Theorem \ref{thm1}}\label{prooft1}
Note that the assumptions imply that either $B(\bx) \to \infty$
or $B(\bx) \to -\infty$. It suffices to consider the case $B(\bx)\to \infty$
as $|\bx|\to \infty$ since otherwise we only have to interchange the
roles of $d$ and $d^*$ in the proof. By modifying $B$ and $V$ on a
compact set and comparing the corresponding resolvents, we may
assume that $B$ and $V$ satisfy
\begin{align}
\label{c1} &V(\bx) \le - 1/\delta,\\
\label{c2} &|\nabla V(\bx)| \le |\delta V(\bx)|, \\
\label{c3} &|V(\bx)|\le 2(1-\eta)B(\bx),
\end{align}
where $\delta \in (0,\tfrac{1}{4})$ is fixed, but can be chosen arbitrarily
small, and $\eta \in (0,1)$ is a fixed ($\delta$-independent) constant
(c.f. \cite{MehringStock} Appendix B).
Using the commutator relation \eqref{eq:1}, we see that
\begin{align}\label{commuteddstar}
dd^* \ge 2B \ge (1- \eta)^{-1} |V| \ge (1-\eta)^{-1}\delta^{-1}
\end{align}
on $C^\infty_0(\R^2,\C)$ and therefore on $\mathcal{D}(dd^*)$.
Since $dd^*$ and $d^*d$ are isospectral away from $0$, we
obtain a spectral gap $(0,\beta) \subset \varrho(\HA)$,
with $\beta = (1-\eta)^{-1}\delta^{-1}$. Thus,
$0$ can be regarded as an isolated point of the spectrum,
which is used in the following commutator estimates.
\begin{lemma}\label{lemmakey}
Let $V\in C^1(\R^2,\R)$, $B\in C(\R^2,\R)$ and
${\bf A}\in C^1(\R^2,\R^2)$ with
$B=\curla$. Assume further that the
conditions \eqref{c1}$-$\eqref{c3} are fulfilled for
$\delta\in(0,\tfrac{1}{4})$ and $\eta\in(0,1)$. Then:
\begin{itemize}
\item[a)] The operators $\left[ P_0^\perp,V^{-1} \right] V$,
$V\left[ P_0^\perp,V^{-1}\right]$ are well-defined on $\core$ and
extend to bounded operators on $\hilbert$ with
\begin{align*}
\left \lVert V\left[P_0^\perp,V^{-1}\right]\right \rVert ,
\left \lVert \left[P_0^\perp,V^{-1}\right] V\right \rVert
\le 4 \delta^\frac{3}{2}.
\end{align*}
The same holds true if we replace $P_0^\perp$ above by $P_0$.
\item[b)] $P_0\mathcal{D}(V), P_0^\perp \mathcal{D}(V) \subset
\mathcal{D}(V) $.
\end{itemize}
\end{lemma}
\begin{lemma}\label{lemmakey2}
Let $V\in C^1(\R^2,\R)$, $B\in C(\R^2,\R)$ and ${\bf A}\in
C^1(\R^2,\R^2)$ with $B=\curla$. Assume
further that the conditions \eqref{c1}$-$\eqref{c3} are fulfilled for
$\delta\in(0,\tfrac{1}{4})$ and $\eta\in (0,1)$. Then
$\left[{\rm sgn}(\DA)P_0^\perp,V^{-1} \right]$ maps $\hilbert$ into
$\mathcal{D}(V)$ and
\begin{equation}\label{rrr}
\left \lVert V\left[{\rm sgn}(\DA)P_0^\perp,V^{-1} \right] \right
\rVert \leq 4 \delta^\frac{3}{2}.
\end{equation}
\end{lemma}
The proofs of these commutator estimates can be found in
\cite{MehringStock}. Since $\DA$ is a first-order operator,
it is much more convenient to commute $V$ with functions
of $\DA$ instead of with functions of $\HA$.
For proving Theorem \ref{thm1}, it suffices to find a constant
$c>0$ such that
\begin{align}
\label{eq:2}
\| H \varphi \|\ge c\|V\varphi\|,\quad \varphi\in \core
\end{align}
holds (see Lemma \ref{Gertrud} in the appendix).
\begin{proof}[Proof of Theorem \ref{thm1}]
Let $\varphi\in\core$. By Lemma \ref{lemmakey}
we can split $\|H\varphi\|$ as
\begin{equation*}
\begin{split}
\left \lVert (\HA+V)\varphi \right \rVert ^2
&= \left \lVert (\HA+V)(P_0+P_0^\perp)\varphi \right \rVert^2\\
&= \left \lVert \big(VP_0+(\HA+V)P_0^\perp \big)\varphi \right \rVert^2\\
&= \left \lVert(\HA+V)P_0^\perp\varphi \right \rVert^2+
2{\rm Re}
\sps{(\HA+V)P_0^\perp\varphi}{VP_0\varphi}+
\left \lVert VP_0\varphi \right \rVert^2\\
&=\left \|(\HA+V)P_0^\perp\varphi \right \|^2-\delta \left \lVert VP_0^\perp\varphi \right \rVert^2\\
&\hspace{1.37cm}+
2{\rm Re}\sps{VP_0\varphi}{\HA
P_0^\perp\varphi}+\|V\varphi\|^2-(1-\delta) \left \lVert VP_0^\perp\varphi \right \rVert^2.
\end{split}
\end{equation*}
For the cross-term, condition \eqref{c2} yields
\begin{equation}
\label{eq:17}
\begin{split}
|\sps{VP_0\varphi}{\HA
P_0^\perp\varphi}|&=|\sps{\DA VP_0\varphi}{\DA
P_0^\perp\varphi}|\\
&= | \sps{ \left[\DA,V \right]V^{-1} V P_0\varphi}{\DA P_0^\perp\varphi}|\\
&\le \tfrac{1}{2}\delta^{-\frac{1}{2}}
\| (-\ri{\boldsymbol\sigma} \nabla V) V^{-1} V P_0\varphi\|^2
+ \tfrac{1}{2} \delta^{\frac{1}{2}}
\| \DA P_0^\perp\varphi \|^2\\
&\le \tfrac{1}{2}\delta^{\frac{3}{2}}
\| V P_0\varphi\|^2
+ \tfrac{1}{2} \delta^{\frac{1}{2}}
\| \HA P_0^\perp\varphi \| \|\varphi\|\\
&\le \tfrac{1}{4}\delta^{\frac{3}{2}}
\| \HA P_0^\perp\varphi \|^2
+ \tfrac{1}{4} \delta^{\frac{3}{2}}\big( \|V \varphi\|^2+2\|V P_0\varphi\|^2\big).
\end{split}
\end{equation}
Applying Lemma \ref{lemmakey} a) results in
\begin{align*}
\big \lVert VP_0^\perp\varphi \big \rVert ,
\big \lVert VP_0\varphi \big \rVert \le
\big(1+ 4\delta^\frac{3}{2} \big)\|V\varphi\|,
\end{align*}
and therefore
\begin{align}\label{eq:19}
\begin{split}
\|V\varphi\|^2-(1-\delta)\left \| VP_0^\perp\varphi \right\|^2
- \tfrac{1}{4} \delta^{\frac{3}{2}} \|V \varphi\|^2
&- \tfrac{1}{2} \delta^{\frac{3}{2}} \|V P_0 \varphi\|^2 \\
&\ge \big(\delta- 14\delta^\frac{3}{2}\big) \|V\varphi\|^2.
\end{split}
\end{align}
Because
\begin{equation*}
\begin{split}
\left\|(\HA+V)P_0^\perp\varphi\right\|^2 &-
\delta^{\frac{3}{2}} \| \HA P_0^\perp\varphi \|^2 -
\delta\|V P_0 \varphi\|^2 \\
& \ge (1-\varepsilon-\delta^{\frac{3}{2}})
\|\HA P_0^\perp\varphi\|^2+(1-\varepsilon^{-1}-\delta)
\|V P_0^\perp\varphi\|^2
\end{split}
\end{equation*}
for any $\varepsilon\in (0,1)$, it suffices to show,
in view of \eqref{eq:17} and \eqref{eq:19},
that
\begin{align}
\label{eq:20}
\sps{\HA P_0^\perp\varphi}{\HA
P_0^\perp\varphi}+\tfrac{1-\varepsilon^{-1}-\delta}
{1-\epsilon-\delta^\frac{3}{2}}
\sps{V P_0^\perp\varphi}{V P_0^\perp\varphi}\ge 0
\end{align}
for $\delta>0$ small enough and some $\varepsilon\in (0,1)$.
We choose $\epsilon = 1 - \delta^\frac{1}{2}$, then
\begin{align*}
-\frac{1-\epsilon^{-1}-\delta}{1-\epsilon-\delta^\frac{3}{2}} =
\frac{1}{1-\delta}\left(\frac{1}{1-\delta^\frac{1}{2}} +\delta^\frac{1}{2}\right)
=: c_\delta >0.
\end{align*}
Since $dd^* \ge 2B$ and therefore ${\rm ker}(d^*) = \{0\}$, we have
\begin{align}
\label{eq:21}
P_0^\perp =
\begin{pmatrix}
\pi^\perp&0\\
0& \pi_*^\perp
\end{pmatrix} =
\begin{pmatrix}
\pi^\perp&0\\
0& \mathbbm{1}
\end{pmatrix}\!.
\end{align}
Setting
$\varphi=(\varphi_1,\varphi_2)^{\rm T}$, one can rewrite \eqref{eq:20} as
\begin{align*}
\left \lVert \HA P_0^\perp\varphi \right \rVert^2-
c_{\delta}\left \lVert V P_0^\perp\varphi \right \rVert^2=
\|d^*d\pi^\perp\varphi_1\|^2- c_{\delta}\|V\pi^\perp\varphi_1\|^2
+\|dd^*\varphi_2\|^2- c_{\delta}\|V\varphi_2\|^2.
\end{align*}
By using the isometries $s,s^*$ given in \eqref{super0},
relation \eqref{super} and estimate \eqref{commuteddstar},
one obtains
\begin{align}\nonumber
\|dd^*\varphi_2\|^2-
c_{\delta}\|V\varphi_2\|^2
&=\sps{d^*\varphi_2}{d^*dd^*\varphi_2} -
c_{\delta}\Sps{\sqrt{-V}\varphi_2}{|V|\sqrt{-V}\varphi_2}\\
\nonumber
&=\sps{sd^*\varphi_2}{dd^*sd^*\varphi_2} -
c_{\delta}\Sps{\sqrt{-V}\varphi_2}{|V|\sqrt{-V}\varphi_2}\\
\nonumber
&\ge\sps{sd^*\varphi_2}{2Bsd^*\varphi_2} -
c_{\delta}\Sps{d^*\sqrt{-V}\varphi_2}{d^*\sqrt{-V}\varphi_2}\\
\nonumber
&\ge\sps{sd^*\varphi_2}{2Bsd^*\varphi_2} -
c_{\delta}\Sps{\sqrt{-V}d^*\varphi_2}{\sqrt{-V}d^*\varphi_2}\\
\nonumber & \hspace{3.15cm}
- c_{\delta}\Sps{\big[d^*,\sqrt{-V}\,\big]\varphi_2}{\sqrt{-V}d^*\varphi_2}\\
\nonumber & \hspace{3.15cm}
- c_{\delta}\Sps{\sqrt{-V}d^*\varphi_2}{\big[d^*,\sqrt{-V}\,\big]\varphi_2}\\
\nonumber & \hspace{3.15cm}
- c_{\delta}\Sps{\big[d^*,\sqrt{-V}\,\big]\varphi_2}{\big[d^*,\sqrt{-V}\,\big]\varphi_2}\\
\nonumber
&\ge \big\|\sqrt{2B}sd^*\varphi_2 \big\|^2 -
c_{\delta}\left(\big\|\big[d^*,\sqrt{-V}\,\big]\varphi_2 \big\|
+ \big\|\sqrt{-V} d^*\varphi_2 \big\| \right)^2 \\
\nonumber
&\ge \big\|\sqrt{2B}sd^*\varphi_2 \big\|^2 -
c_{\delta}\left(\delta \big\|sd^*\varphi_2 \big\|
+ \big(1+4\delta^\frac{3}{2}\big)\big\|\sqrt{-V} sd^*\varphi_2 \big\| \right)^2 \\
\begin{split}\label{ddstar}
&\ge \big[ 1-c_{\delta}(1-\eta)\big(1+15\delta^\frac{3}{2}\big) \big]
\big\|\sqrt{2B}sd^*\varphi_2 \big\|^2,
\end{split}
\end{align}
where we applied the bound
$\big \lVert \sqrt{-V} \big[s\pi^\perp,\sqrt{-V}^{-1}\big] \big \rVert
\le 4\delta^2$.
For the latter we write
\begin{align*}
\sqrt{-V} \left[{\rm sgn}(\DA)P_0^\perp,\sqrt{-V}^{-1}\right] =
\left(\begin{array}{cc}
0&\sqrt{-V}\left[s^*,\sqrt{-V}^{-1}\right]\\
\sqrt{-V}\left[s\pi^\perp, \sqrt{-V}^{-1}\right]&0
\end{array}
\right)
\end{align*}
and therefore, by
Lemma \ref{lemmakey2} with $\sqrt{-V}$ instead of $V$, we get
\begin{align*}
\begin{split}
\label{eq:75} \left \lVert \sqrt{-V} \left[s\pi^\perp,\sqrt{-V}^{-1}\right] \right \rVert
&\le
\left \lVert \sqrt{-V} \left[{\rm sgn}(\DA)P_0^\perp,\sqrt{-V}^{-1}\right] \right
\rVert\le 4\delta^\frac{3}{2}.
\end{split}
\end{align*}
\\
Similarly, we obtain a lower bound for
$ \|d^*d\pi^\perp\varphi_1\|^2-c_{\delta}\|V\pi^\perp\varphi_1\|^2$
by using again the upper relation of Equation
\eqref{super}. More precisely,
\begin{equation*}\label{dddd}
\begin{split}
\|d^*d\pi^\perp\varphi_1\|^2- c_{\delta}\|V\pi^\perp\varphi_1\|^2&=
\|d^*ds^*s\pi^\perp\varphi_1\|^2- c_{\delta}\|Vs^*s\pi^\perp\varphi_1\|^2\\
&= \|dd^*s\pi^\perp\varphi_1\|^2-
c_{\delta}\|Vs^*V^{-1} Vs\pi^\perp\varphi_1\|^2\\
&=\|dd^*s\pi^\perp\varphi_1\|^2-
c_{\delta}\big\|\big(s^*+V\left[s^*,V^{-1}\right]\big)Vs\pi^\perp\varphi_1\big\|^2,
\end{split}
\end{equation*}
where
$V\left[s^*,V^{-1}\right]$
is one of the components of the operator
\begin{align*}
V \left[{\rm sgn}(\DA)P_0^\perp,V^{-1}\right]= \left(\begin{array}{cc}
0&V\left[s^*,V^{-1}\right]\\
V\left[s\pi^\perp, V^{-1}\right]&0
\end{array}
\right)\!,
\end{align*}
so Lemma \ref{lemmakey2} yields
$\left \lVert V \left[s^*,V^{-1}\right] \right \rVert \le 4\delta^\frac{3}{2}$.
Thus,
\begin{align*}
\|d^*d\pi^\perp\varphi_1\|^2- c_{\delta}\|V\pi^\perp \varphi_1\|^2&
\ge \|dd^*s\pi^\perp\varphi_1\|^2 -
c_{\delta} (1+ 10 \delta^\frac{3}{2}) \big\|Vs\pi^\perp\varphi_1\big\|^2.
\end{align*}
We note that
$s\pi^\perp\varphi_1 \subset \mathcal{D} (dd^*)
\subset \mathcal{D}(V)$, hence we can use \eqref{ddstar}
(by approximating $s\pi^\perp\varphi_1$ through
$C_0^\infty$-functions in the graph norm of $dd^*$)
to conclude that
\begin{align*}
\|d^*d\pi^\perp\varphi_1\|^2 &- c_{\delta}\|V\pi^\perp \varphi_1\|^2 \\
& \ge \big[ 1-c_{\delta}(1-\eta)\big(1+10\delta^\frac{3}{2}\big)
\big(1+15\delta^\frac{3}{2}\big) \big]
\big\|\sqrt{2B}sd^*s\pi^\perp \varphi_1 \big\|^2.
\end{align*}
Combining this inequality with \eqref{ddstar} leads to
\begin{align*}
\left \lVert \HA P_0^\perp\varphi \right \rVert^2 & -
c_{\delta}\left \lVert V P_0^\perp\varphi \right \rVert^2 \\
&\ge
\big[ 1-c_{\delta}(1-\eta)(1+50\delta^\frac{3}{2})\big]
\left( \big\|\sqrt{2B}sd^*\varphi_2 \big\|^2 +
\big\|\sqrt{2B}d\pi^\perp\varphi_1 \big\|^2 \right)\!,
\end{align*}
where the r.h.s is non-negativ for $\delta $ small enough.
\end{proof}
\section{Proofs of Theorem \ref{thm2} and \ref{thm3}}\label{prooft2}
The basic strategy of the proofs is to represent $B$ and $V$ locally through
constant values $V_n:=V(\bx_n)$ and $B_n:=B(\bx_n)$ along a sequence
$(\bx_n)_{n\in\N}\subset\R^2$. Since one also needs to compare
vector potentials associated to $B_n$ and $B$, we use the gauges
\begin{align*}
&{\mathbf A}_n(\bx):=\int_0^1 B_n\wedge (\bx-\bx_n) s \rd
s=\tfrac{1}{2} B_n\wedge (\bx-\bx_n) ,\\
&\widetilde{{\mathbf A}}_n(\bx):= \int_0^1
B(\bx_n+s(\bx-\bx_n))\wedge (\bx-\bx_n) s \rd s,
\end{align*}
where $a\wedge {\mathbf v}:=a(-v_2,v_1)$ for $a\in \R$ and
${\mathbf v}=(v_1,v_2)\in \R^2$. The two given vector potentials satisfy
${\rm curl}\,{\mathbf A}_n = {\rm curl}\,\widetilde{\mathbf A}_n =B $,
hence for every $n \in \N$ there exists a function
$g_n\in C^2(\R^2,\R)$ such that $\nabla
g_n={\mathbf A}-\widetilde{{\mathbf A}}_n$ .
In addition, for every vector potential ${\mathbf A}_n$,
representing the constant magnetic fields $B_n$,
we obtain operators $d_n$ and $d_n^*$, $n \in \N$,
defined as in \eqref{carolina}.
For a sequence of natural numbers $(\pn)_{n\in\N}$
we set
\begin{align*}
\psi_n(\bx):=
\begin{pmatrix}
(d_n^* )^\pn e^{-B_n|\bx-\bx_n|^2/4}\\
0
\end{pmatrix}\!.
\end{align*}
Iterating commutator relation \eqref{eq:1} for $d_n,d_n^*$ yields
\begin{align}
\label{eq:25}
d_n^* d_n \big[(d_n^* )^{k_n} e^{-B_n|\bx-\bx_n|^2/4}\big]
=2k_nB_n \big[(d_n^* )^{k_n} e^{-B_n|\bx-\bx_n|^2/4}\big],
\quad n \in \N,
\end{align}
i.e. $ \psi_n$ is an eigenfunction of $H_{{\mathbf A}_n}$ with
the corresponding eigenvalue $2\pn B_n$.
For the localization let $\chi \in C^\infty_0(\R^2,[0,1])$
be such that $\chi(\bx)=1$ for $|\bx|\le 1$
and $\chi(\bx)=0$ for $|\bx|\ge 2$. We set
\begin{align*}
\chi_n(\bx):=\chi\left(\frac{\bx-\bx_n}{r_n}\right)\!,
\end{align*}
where the $r_n>0$ will be chosen in the proofs.
For the Weyl sequence we define the functions
$\varphi_n$ through
\begin{align}
\label{eq:26}
\varphi_n(\bx):= e^{\ri g_n (\bx)}\chi_n(\bx) \psi_n(\bx),
\quad \bx\in\R^2,
\end{align}
with $n \in \N$. Bounds on the norm of $\varphi_n$ can be obtained
as in \cite{MehringStock}. They are given by:
\begin{lemma}\label{landau}
For all $n\in \N$ large enough we have
\begin{align}\label{eq:29a}
& \|\varphi_n\|^2 \le \|\psi_n\|^2
= 2\pi \int_0^\infty (B_n r)^{2\pn} e^{-\frac{B_n}{2} r^2} \rd r
= 2^{\pn+1} \pi B_n^{\pn-1}\pn! , \\
\label{eq:29}
&\|\varphi_n\|^2\ge
\|\psi_n\|^2\left(1-\frac{1}{\pn!}
\int_{\frac{1}{2}B_n r_n^2}^\infty s^{\pn} e^{-s} \rd s\right)\!.
\end{align}
\end{lemma}
Now $H\varphi_n$ can be written as
\begin{equation}\label{weyl}
\begin{split}
e^{-\ri g_n} (\HA+V)\varphi_n= \ &
(H_{\widetilde{\mathbf A}_n}\!+V) \chi_n \psi_n\\
= \ & (H_{{\mathbf A}_n}\!+V) \chi_n \psi_n+
2({\widetilde{\mathbf A}_n} - {{\mathbf A}_n})
( -\ri \nabla -{\mathbf A}_n) \chi_n \psi_n \,+\\
& ({\widetilde{\mathbf A}_n} - {{\mathbf A}_n})^2\chi_n \psi_n
-\ri \nabla \cdot( {\widetilde{\mathbf A}_n}
- {{\mathbf A}_n})\chi_n\psi_n \,+\\
& (B-B_n)\chi_n \psi_n \,,
\end{split}
\end{equation}
with the localization error
\begin{align}
\begin{split} \label{localierror}
(H_{{\mathbf A}_n}\!+V) \chi_n \psi_n -\chi_n(H_{{\mathbf A}_n}\!
&+V)\psi_n \\
&=-(\Delta\, \chi_n)\psi_n +
2(-\ri\nabla \chi_n)( -\ri \nabla -{\mathbf A}_n)\psi_n.
\end{split}
\end{align}
To prove Theorem \ref{thm2} and Theorem \ref{thm3},
we estimate each term of \eqref{weyl} separately. For
the proofs we use the notation
$K_n := \{ x \in \R^2 \,|\, r_n \le |\bx-\bx_n| \le 2r_n\}$
with $n \in \N$.
\begin{proof}[Proof of Theorem \ref{thm2}]
We set $k_n = k$ and choose the radii to be
$r_n^{-4}=B_n^{(2-\epsilon)}$. Then, for any $p \ge 0$,
$$
\frac{(B_n)^p}{k_n!}
\int_{\frac{1}{2}B_n r_n^2}^\infty s^{\pn} e^{-s} \rd s
=
\frac{(B_n)^p}{k!}
\int_{\frac{1}{2}B_n^{\epsilon/2} }^\infty s^{k} e^{-s} \rd s
\longrightarrow 0 \quad \mbox{as} \ n \to \infty.
$$
Further, we have
$\|\psi_n\|^2 \le 2\|\varphi_n\|^2$ for $n \in \N$ large enough.
For treating the terms on the r.h.s. of \eqref{weyl}, we estimate
\begin{align*}
\ \big\|({\widetilde{\mathbf A}_n}&
- {{\mathbf A}_n})(-\ri\nabla -{\mathbf A}_n) \chi_n \psi_n \big\|^2
\\ \le
& \ C_1 r_n^4 |\nabla B(\bx_n)|^2
\big\|( -\ri \nabla -{\mathbf A}_n) \chi_n \psi_n \big\|^2 \\ \le
& \ 2 C_1 r_n^4 |\nabla B(\bx_n)|^2 \bigg[(2k+1)B_n \| \psi_n \|^2
+ r_n^{-2} \|\nabla \chi \|_\infty^2
\int_{K_n} |\psi_n(\bx)|^2 \,\rd^2 x \bigg] \\ \le
& \ 16kC_1 B_n\frac{|\nabla B(\bx_n)|^2}{B_n^{2-\epsilon}} \| \psi_n \|^2
+ 4C_1\|\nabla \chi \|_\infty^2 \| \psi_n \|^2 B_n^{(2-\epsilon)/2}
\frac{1}{k!}\int_{\frac{1}{2}B_n^{\epsilon/2}}^\infty s^ke^{-s}\rd s .
\end{align*}
In addition,
\begin{align*}
\|({\widetilde{\mathbf A}_n} - {{\mathbf A}_n})^2 \chi_n \psi_n \|^2
&\le C_2 r_n^4|\nabla B(\bx_n)|^4 \|\psi_n \|^2
\le C_2\frac{|\nabla B(\bx_n)|^4}{B_n^{2(1-\epsilon)}}
B_n^{-\epsilon}\|\psi_n \|^2,\\
\| \mathrm{div\,} ( {\widetilde{\mathbf A}_n} - {{\mathbf A}_n}) \chi_n \psi_n \|^2
&= \| \mathrm{div\,}{\widetilde{\mathbf A}_n} \chi_n \psi_n \|^2
\le C_3r_n^2 |\nabla B(\bx_n)|^2 \|\psi_n \|^2 \\
&\hspace{2.85cm}\le C_3 \frac{|\nabla B(\bx_n)|^2}{B_n^{1-\epsilon}}
B_n^{-\epsilon/2} \|\psi_n \|^2,\\
\|(B-B_n)\chi_n \psi_n \|^2
&\le C_4r_n^2 |\nabla B(\bx_n)|^2 \|\psi_n \|^2
\le C_4\frac{|\nabla B(\bx_n)|^2}{B_n^{1-\epsilon}} B_n^{-\epsilon/2}\|\psi_n \|^2.
\end{align*}
For the first term of the r.h.s of \eqref{weyl} we get,
due to \eqref{localierror},
\begin{align*}
\|(H_{{\mathbf A}_n}\!+V) &\chi_n \psi_n \| \\
&\le \|\chi_n(H_{{\mathbf A}_n}\!+V) \psi_n \|
+ \|(\Delta\, \chi_n)\psi_n\| +
2\|(-\ri\nabla \chi_n)( -\ri\nabla -{\mathbf A}_n)\psi_n\|,
\end{align*}
with
\begin{align*}
\|(\Delta\, \chi_n)\psi_n\|^2
& \le r_n^{-4} \|\Delta \chi \|_\infty^2
\int_{K_n} |\psi_n(\bx)|^2 \,\rd^2 x \\
& \le 2\|\Delta \chi \|_\infty^2 \| \psi_n \|^2\frac{1}{k!}
B_n^{2-\epsilon}\int_{\frac{1}{2}B_n^{\epsilon/2}}^\infty s^ke^{-s}\rd s,
\end{align*}
and
\begin{align*}
\|(-\ri\nabla \chi_n)( -\ri\nabla -{\mathbf A}_n)\psi_n\|^2
&\le r_n^{-2} \|\nabla \chi \|_\infty^2
\int_{K_n} \big|(-\ri\nabla -{\mathbf A}_n)\psi_n(\bx)\big|^2 \rd^2 x \\
&\le \|\nabla \chi \|_\infty^2 B_n^{(2-\epsilon)/2}(2k+1)B_n
\int_{K_n} \big|\psi_n(\bx)\big|^2 \rd^2 x \\
&\le \|\nabla \chi \|_\infty^2\| \psi_n \|^2 (2k+1) B_n^{2-\epsilon/2}
\int_{\frac{1}{2}B_n^{\epsilon/2}}^\infty s^ke^{-s}\rd s.
\end{align*}
Because of \eqref{eq:25} and since $|\nabla V|$ vary with rate $0$,
we conclude by the mean value theorem that
\begin{align*}
\|\chi_n(H_{{\mathbf A}_n}\!+V) \psi_n \|^2
&\le \|\chi_n(V+2kB_n) \psi_n \|^2 \\[0.2cm]
&\le C_5 |\nabla V(\bx_n)|^2 r_n^2 \|\chi_n \psi_n \|^2 +
(2kB_n + V_n)^2 \|\chi_n \psi_n \|^2 \\
&\le C_5 (4k)^{1-\epsilon} \frac{|\nabla V(\bx_n)|^2}{|V_n|^{1-\epsilon}}\|\varphi_n \|^2\
+ (2kB_n + V_n)^2 \|\varphi_n \|^2\,.
\end{align*}
Hence, by \eqref{weyl} and conditions \eqref{con0a}$-$\eqref{con2a},
we see that $ \|(\HA+V)\varphi_n\|/ \|\varphi_n\| \to 0 $
as $n \to \infty$. In addition, note that $r_n \to 0$ as $n \to \infty$,
so we can assume that the $\varphi_n$ have mutually disjoint support,
i.e. $(\varphi_n)_{n \in \N}$ is a Weyl sequence for $0$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm3}]
We first note that it suffices to proof $0 \in \sigma_{\rm{ess}}(H)$ since for
$E \in \R$ we consider $V_E := V-E$ instead of $V$, which also fulfills
\eqref{con3.1}$-$\eqref{con3.4} along $\gamma$. Because
$\R^+ \ni t \mapsto V(\gamma(t))/B((\gamma(t))$ is continuous and
\eqref{con3.1} holds, we find points $(\bx_n)_{n \in \N} \subset \rm{Im}(\gamma)$,
with $|\bx_n| \to \infty$ as $n \to \infty$, such that $2nB(\bx_n)=-V(\bx_n)$.
We choose $\pn = n$ and set
\begin{align}
\label{eq:rad2} r_n := \sqrt{2n^{1+\epsilon}/B_n} .
\end{align}
Note that $r_n/|\bx_n|^\nu \to 0$ as $n \to \infty$ by \eqref{con3.3}.
In particular, we might assume the $\varphi_n$'s to have mutually
disjoint support.
Further, for any $\lambda \ge 0$,
$$\frac{e^{\lambda n}}{n!}\int_{n^{1+\epsilon}}^\infty s^n e^{-s} \rd s
\le e^{\lambda n} \exp(n\ln(2n)-n^{1+\epsilon}/2 +n)\longrightarrow 0
\quad \mbox{as} \ n \to \infty.$$
Hence, we can choose $N \in \N $ so large that
$\|\varphi_n\|^2 \le \|\psi_n\|^2 \le 2 \|\varphi_n\|^2$ for $n \ge N$. Proceeding as
in the proof of Theorem \ref{thm2}, we obtain
\begin{align*}
\|({\widetilde{\mathbf A}_n} &- {{\mathbf A}_n})(-\ri \nabla -{\mathbf A}_n)
\chi_n \psi_n \|^2 \\ \le
& \ C_6 r_n^4 |\nabla B(\bx_n)|^2 (2n+1)B_n \bigg[\|\psi_n \|^2
+ r_n^{-2} \|\nabla \chi \|_\infty^2
\int_{K_n} |\psi_n(\bx)|^2 \rd^2 x \bigg] \\ \le
& \ 16C_6n^{3+2\epsilon} B_n\frac{|\nabla B(\bx_n)|^2}{B_n^2} \| \psi_n \|^2
+ C_6\|\nabla \chi \|_\infty^2 \| \psi_n \|^2
B_n\frac{1}{n!}\int_{n^{1+\epsilon}}^\infty s^ne^{-s}\rd s \\ \le
& \ \widetilde C_6\left(\frac{|V_n|^3}{B_n^2}\right)^{1+\epsilon}
\frac{|\nabla B(\bx_n)|^2}{B_n^2} \| \psi_n \|^2
+ \alpha C_6\|\nabla \chi \|_\infty^2 \| \psi_n \|^2
\frac{e^{2\kappa n}}{n!}\int_{n^{1+\epsilon}}^\infty s^ne^{-s}\rd s.
\end{align*}
Using (as in the first inequality above) that $|\nabla B|$ vary with
rate $\nu$, we conclude
\begin{align*}
\|({\widetilde{\mathbf A}_n} - {{\mathbf A}_n})^2 \chi_n \psi_n \|^2
&\le C_7 r_n^4|\nabla B(\bx_n)|^4 \|\psi_n \|^2
\le \frac{C_7}{4}\frac{|V_n|^6}{B_n^4}\frac{|\nabla B(\bx_n)|^4}{B_n^{4}}
\|\psi_n \|^2, \\
\|{\rm div}\, ( {\widetilde{\mathbf A}_n} - {{\mathbf A}_n})
\chi_n \psi_n \|^2
&\le C_8r_n^2 |\nabla B(\bx_n)|^2 \|\psi_n \|^2
\le \frac{C_8}{2}\frac{|V_n|^3}{B_n^2}\frac{|\nabla B(\bx_n)|^2}{B_n^2} \|\psi_n \|^2 ,
\end{align*}
\begin{align*}
\hspace{1.17cm}\|(B-B_n)\chi_n \psi_n \|^2
\le C_9r_n^2 |\nabla B(\bx_n)|^2 \|\psi_n \|^2
\le \frac{C_9}{2}\frac{|V_n|^3}{B_n^2}\frac{|\nabla B(\bx_n)|^2}{B_n^2}
\|\psi_n \|^2. \hspace{0.71cm}
\end{align*}
We estimate, using equality \eqref{localierror},
the first term on the r.h.s. of \eqref{weyl} by
\begin{align*}
\|(H_{{\mathbf A}_n}\!+V) &\chi_n \psi_n \| \\
&\le \|(V-V_n)\chi_n \psi_n \| + \|(\Delta\, \chi_n)\psi_n\|
+ 2\|(-\ri\nabla \chi_n)( -\ri \nabla - {\mathbf A}_n)\psi_n\|,
\end{align*}
with
\begin{align*}
\|(\Delta\, \chi_n)\psi_n\|^2
& \le \|\Delta \chi \|_\infty^2 \| \psi_n \|^2 \frac{B_n^2}{n^{2+2\epsilon}}
\frac{1}{n!}\int_{n^{1+\epsilon}}^\infty s^ne^{-s}\rd s \\
& \le \alpha^2 \|\Delta \chi \|_\infty^2 \| \psi_n \|^2
\frac{e^{4\kappa n}}{n^{2+2\epsilon}}\frac{1}{n!}
\int_{n^{1+\epsilon}}^\infty s^n e^{-s}\rd s,
\end{align*}
and
\begin{align*}
\|(-\ri\nabla \chi_n)( -\ri\nabla -{\mathbf A}_n)\psi_n\|^2
&\le \|\nabla \chi \|_\infty^2 \frac{B_n}{n^{1+\epsilon}}
\int_{K_n} (2n+1)B_n|\psi_n(\bx)|^2 \rd^2 x \\
&\le \alpha\|\nabla \chi \|_\infty^2\| \psi_n \|^2
\frac{e^{2\kappa n}}{n^{1+\epsilon}}\frac{1}{n!}
\int_{n^{1+\epsilon}}^\infty s^ne^{-s}\rd s.
\end{align*}
Since $|\nabla V|$ vary with rate $\nu$, we have
\begin{align*}
\|\chi_n(H_{{\mathbf A}_n}\!+V) \psi_n \|^2
&\le \|\chi_n(V-V_n) \psi_n \|^2 \\
&\le C_{10} |\nabla V(\bx_n)|^2 r_n^2 \|\psi_n \|^2
\le \widetilde{C}_{10} \left(\frac{|V_n|^3}{B_n^2}\right)^{1+\epsilon}
\frac{|\nabla V(\bx_n)|^2}{|V_n|^2}\|\psi_n \|^2.
\end{align*}
We see that $\|(H_{{\mathbf A}_n}\!+V) \varphi_n \|/ \|\varphi_n \| \to 0 $
as $n \to \infty$ and therefore, by \eqref{weyl} and the estimates
above, that $\|(\HA+V) \varphi_n \|/ \|\varphi_n \| \to 0 $ as
$n \to \infty$.
\end{proof}
\section{Proof of Theorem \ref{thm4}}\label{prooft4}
Throughout this section we consider the case of a constant
magnetic field $B(\bx) = B_0$. In addition, we assume that
${\bf A}$ is in the rotational gauge, i.e.
$${\bf A}(\bx) = \frac{B_0}{2}
\begin{pmatrix}
-x_2 \\ x_1
\end{pmatrix}\!.
$$
Note that $\HA$ is invariant under rotations. More precisely,
for a special orthogonal matrix ${\mathcal R} \in SO(2,\R)$
define the unitary map
\begin{align*}
U_{\mathcal R} : \hilbert \to \hilbert, \quad \quad \quad
\psi(\,\cdot\,) \mapsto \psi({\mathcal R}^{-1} \,\cdot\,) ,
\end{align*}
then $U_{\mathcal R}^{-1}\HA U_{\mathcal R} = \HA$ and therefore
\begin{align*}
U_{\mathcal R}^{-1} (\HA+V) U_{\mathcal R} = \HA + V_{\mathcal R} \quad \mbox{with} \
V_{\mathcal R}(\,\cdot\,) = V({\mathcal R} \,\cdot\,) .
\end{align*}
To construct a Weyl sequence, consider a second gauge
$\widetilde{\bf A}(\bx) = B_0 x_1 \hat e_2 $, called the
Landau gauge. Then our Hamiltonian reads
\begin{align}\label{HwcE}
\begin{split}
H_{\widetilde{\bf A}} +V
&= -\partial_1^2 + (-\ri \partial_2 -B_0x_1)^2
- \sigma_3 B_0+V \\
&={\tilde d}^{\,*} {\tilde d} +B_0 - \sigma_3 B_0+V ,
\end{split}
\end{align}
with
\begin{align*}
{\tilde d} =
-\ri \partial_1 + \ri (-\ri\partial_2 -B_0x_1) , \quad
{\tilde d}^{\,*} =
-\ri \partial_1 - \ri (-\ri\partial_2 -B_0x_1) .
\end{align*}
For electric fields of the form $V(\bx) = V_0 + {\mathcal E}_0 (x_1-\zeta)$,
with constants $V_0, {\mathcal E}_0, \zeta \in \R$, we can write
\begin{align}
\begin{split}
H_{\widetilde{\bf A}} +V
&= -\partial_1^2 + (-\ri \partial_2 -B_0x_1)^2 -
\sigma_3 B_0+V_0 + {\mathcal E}_0 (x_1- \zeta) \\
& = -\partial_1^2 +
B_0^2\big(x_1 -\tfrac{1}{B_0}\big(-\ri \partial_2 -
\tfrac{ {\mathcal E}_0}{2B_0}\big)\big)^2 \\
&\hspace{1.11cm}
+ \tfrac{ {\mathcal E}_0}{B_0} \big( -\ri\partial_2
-\tfrac{ {\mathcal E}_0}{2B_0}\big)
- {\mathcal E}_0 \zeta - \sigma_3 B_0+V_0
+\big(\tfrac{ {\mathcal E}_0}{2B_0}\big)^2.
\end{split}
\end{align}
Performing a Fourier transform in $x_2$, we obtain the direct
integral representation
\begin{align*}
H_{\widetilde{\bf A}} +V
\cong \int_\R^\oplus h(\xi) \mathrm{d} \xi
\end{align*}
on $L^2(\R_\xi, L^2(\R,\C^2))$, with
\begin{align*}
h(\xi) &=
-\partial_1^2 +
B_0^2\big(x_1 -\tfrac{1}{B_0}\big( \xi -\tfrac{ {\mathcal E}_0}{2B_0}\big)\big)^2 +
\tfrac{ {\mathcal E}_0}{B_0} \big( \xi-\tfrac{ {\mathcal E}_0}{2B_0}\big)
- {\mathcal E}_0 \zeta - \sigma_3 B_0+V_0
+\big(\tfrac{ {\mathcal E}_0}{2B_0}\big)^2 \\
&=
-\partial_1^2 +
B_0^2\big(x_1 -\hat\zeta \big)^2 +
{\mathcal E}_0\hat \zeta - {\mathcal E}_0 \zeta - \sigma_3 B_0+V_0
+\big(\tfrac{ {\mathcal E}_0}{2B_0}\big)^2 .
\end{align*}
Here we set $\hat \zeta =
\tfrac{1}{B_0}\big( \xi -\tfrac{ {\mathcal E}_0}{2B_0}\big)$.
Note that $h(\xi)$ is the Hamiltonian of a shifted
harmonic oscillator. Thus, we define for $n \in \N_0$
$$\phi_n(x) =
\frac{1}{\sqrt{2^nn!\sqrt{\pi} }}\, \vartheta_n(x)e^{-x^2/2},
\quad x \in \R,$$
where $\vartheta_n$ denotes the $n-$th Hermite polynomial.
The normalized functions
\begin{align*}
{\widehat \psi}_{{\mathcal E}_0, n,\xi}(x_1) :=
\sqrt[4] {B_0}
\begin{pmatrix}
\phi_n\big(\sqrt{B_0}\big(x_1- \tfrac{1}{B_0}\big( \xi
-\tfrac{ {\mathcal E}_0}{2B_0}\big)\big) \\
0
\end{pmatrix}
\end{align*}
fulfill the equation
\begin{align*}
h(\xi) {\widehat \psi}_{{\mathcal E}_0, n, \xi}=
\Big(2nB_0 + {\mathcal E}_0 \big(
\tfrac{1}{B_0}\big( \xi-\tfrac{ {\mathcal E}_0}{2B_0}\big) -\zeta\big)
+ V_0+ \big(\tfrac{ {\mathcal E}_0}{2B_0}\big)^2 \Big)
{\widehat \psi}_{{\mathcal E}_0, n, \xi}\,.
\end{align*}
Hence,
\begin{align}\label{monika}
{\psi}_{{\mathcal E}_0, n,\xi}(x_1, x_2) :=
e^{\ri \xi x_2} {\widehat \psi}_{{\mathcal E}_0, n, \xi} (x_1)
\end{align}
satisfies
\begin{align}\label{dominika}
\begin{split}
\big[H_{\widetilde{\bf A}} +V\big]{\psi}_{{\mathcal E}_0, n, \xi} =
\Big(2nB_0 + {\mathcal E}_0 \big(
\tfrac{1}{B_0}\big( \xi-\tfrac{ {\mathcal E}_0}{2B_0}\big) -\zeta\big) +
V_0+ \big(\tfrac{ {\mathcal E}_0}{2B_0}\big)^2 \Big)
{ \psi}_{{\mathcal E}_0, n, \xi}
\end{split}
\end{align}
for $\xi \in \R$ and $n \in \N_0$, seen as a differential equation.
In addition, we have
\begin{align}\label{annihilation}
{\tilde d} {\psi}_{{\mathcal E}_0, n,\xi} & =
-\ri \sqrt{2nB_0} {\psi}_{{\mathcal E}_0, n-1,\xi}
+\ri\tfrac{ {\mathcal E}_0}{2B_0}{\psi}_{{\mathcal E}_0, n,\xi} \,,\\
\label{creation}
{\tilde d}^{\,*} {\psi}_{{\mathcal E}_0, n,\xi} & =
\ri \sqrt{2(n+1)B_0} {\psi}_{{\mathcal E}_0, n+1,\xi}
-\ri\tfrac{ {\mathcal E}_0}{2B_0}{\psi}_{{\mathcal E}_0, n,\xi} \,.
\end{align}
\begin{proof}[Proof of Theorem \ref{thm4}]
As argumented in the proof of Theorem \ref{thm3}, it
suffices to find a Weyl sequence for $E =0$. Because of
\eqref{con4.1} and \eqref{con4.4}, there exists a sequence
$\{ \by_n\}_{n \in \N} \subset \mbox{Im}(\gamma)$ such that
\begin{align}\label{crosscond}
V(\by_n) = -2nB_0 - \left(\tfrac{|\nabla V(\by_n)|}{2B_0}\right)^2.
\end{align}
Further, there are rotations ${\mathcal R}_n \in SO(2,\R)$ such that
$\nabla V_{{\mathcal R}_n}(\bx_n)
= |\nabla V_{{\mathcal R}_n}(\bx_n)|\hat e_1$, with
$\bx_n = {\mathcal R}_n^{-1}\by_n =(x_{n,1}, x_{n,2})^T$
for $n \in \N$. We set
\begin{align}\label{Vn}
\hspace{1.0cm}
&V_n := V(\by_n) = V_{{\mathcal R}_n} (\bx_n) ,\\
\label{En}
&{\mathcal E}_n := |\nabla V(\by_n)| =
|\nabla V_{{\mathcal R}_n} (\bx_n)| ,
\end{align}
\begin{align}\label{Xin}
\xi_n := B_0x_{n,1} + \tfrac{ {\mathcal E}_n}{2B_0} .
\hspace{0.71cm}
\end{align}
For the Weyl functions let $\chi \in C_0^\infty(\R, [0, 1])$ with
$\chi(x) =1$ for $|x| \le 1$ and $\chi(x) =0$ for $|x| \ge 2$. Define
\begin{align*}
\chi_{n,j}(x) := \chi \left( \tfrac{x -x_{n,j}}{r_n} \right)\!,
\end{align*}
for $j = 1,2 $, and
\begin{align}\nonumber
\begin{split}
\varphi_n(\bx) &:=
\chi_{n,1}(x_1)
\chi_{n,2}(x_2)
{\psi}_{{\mathcal E}_n, n,\xi_n}(x_1, x_2)
\\ & \ =
\chi\left( \tfrac{x_2 -x_{n,2}}{r_n}\right)
e^{-\ri \xi_n x_2}
\chi\left( \tfrac{x_1 -x_{n,1}}{r_n}\right)
\begin{pmatrix}
\sqrt[4]{B_0}\phi_n\big(\sqrt{B_0}(x_1-x_{n,1})\big)
\\ 0
\end{pmatrix}\!,
\end{split}
\end{align}
where the localization radii $r_n$ are chosen to be
$r_n := \sqrt{n^{1+\epsilon}/B_0}$.
Note that
\begin{align}\label{normthm4}
r_n \le
2 r_n \int_{-\sqrt{n^{1+\epsilon}}}^{\sqrt{n^{1+\epsilon}}} |\phi_n(x)|^2 \rd x
\le \ \|\varphi_n\|^2 \ \le
4 r_n \int_{-2\sqrt{n^{1+\epsilon}}}^{2\sqrt{n^{1+\epsilon}}} |\phi_n(x)|^2 \rd x
\le 4 r_n
\end{align}
for $n \in \N$ large enough
(see Lemma \ref{hermiteestimate} in the appendix).
By denoting $g(\bx) = \tfrac{B_0}{2}x_1x_2$ for $\bx \in \R^2$, we get,
due to \eqref{HwcE}, \eqref{dominika}, \eqref{crosscond} and\eqref{Xin},
that
\begin{align}\label{applyphi}
\begin{split}
H U_{{\mathcal R}_n} e^{-\ri g} \varphi_n = &
\ U_{{\mathcal R}_n} e^{-\ri g}
\big[H_{\widetilde{\bf A}} +V_{{\mathcal R}_n} \big] \varphi_n
\\ = &\
U_{{\mathcal R}_n} e^{-\ri g}\big(
{\tilde d}^{\,*}{\tilde d} \varphi_n -
\chi_{n,1}\chi_{n,2}{\tilde d}^{\,*}{\tilde d}
{\psi}_{{\mathcal E}_n, n,\xi_n}\big) \ + \\ & \
U_{{\mathcal R}_n} e^{-\ri g}
\big[V_{{\mathcal R}_n} - V_n -
{\mathcal E}_n(x_1-x_{1,n})\big] \varphi_n .
\end{split}
\end{align}
The localization error results in
\begin{align*
\begin{split}
{\tilde d}^{\,*}{\tilde d} \varphi_n -
\chi_{n,1}\chi_{n,2}{\tilde d}^{\,*}{\tilde d}
{\psi}_{{\mathcal E}_n, n,\xi_n} \ = \
&\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} +
\chi_{n,1}\partial_2\chi_{n,2}\big]
{\tilde d}^{\,*} {\psi}_{{\mathcal E}_n, n,\xi_n} \ + \\
&\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} -
\chi_{n,1}\partial_2\chi_{n,2}\big]
{\tilde d}{\psi}_{{\mathcal E}_n, n,\xi_n} \ + \\
&\big[-\chi_{n,2} \partial_1^2\chi_{n,1} -\chi_{n,1} \partial_2^2\chi_{n,2}\big]
{\psi}_{{\mathcal E}_n, n,\xi_n} ,
\end{split}
\end{align*}
with, using \eqref{annihilation} and \eqref{creation},
\begin{align*}
\big\|
\big[-\ri\chi_{n,2} \partial_1&\chi_{n,1} +
\chi_{n,1}\partial_2\chi_{n,2}\big]
{\tilde d}^{\,*} {\psi}_{{\mathcal E}_n, n,\xi_n} \big \| \\
&\le \sqrt{2(n+1)B_0}\big\|
\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} +
\chi_{n,1}\partial_2\chi_{n,2}\big] {\psi}_{{\mathcal E}_n, n+1,\xi_n}
\big \| \\ & \hspace{1.52cm}+
\tfrac{ {\mathcal E}_n}{2B_0}\big\|
\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} +
\chi_{n,1}\partial_2\chi_{n,2}\big]{\psi}_{{\mathcal E}_n, n,\xi_n}
\big \| \\ &\le
2\sqrt{2(n+1)B_0} r_n^{-1}\|\chi'\|_\infty \sqrt{2r_n}
\|\phi_{n+1}\| +
2\tfrac{ {\mathcal E}_n}{2B_0}
r_n^{-1}\|\chi'\|_\infty \sqrt{2r_n} \|\phi_n\| \\
&\le 2\sqrt{2} \|\chi'\|_\infty
\Big( B_0\sqrt{\tfrac{2n+2}{n^{1+\epsilon}}} +
\tfrac{{\mathcal E}_n}{2}
\sqrt{\tfrac{B_0}{n^{1+\epsilon}}} \Big) \sqrt{r_n} \,,
\end{align*}
\begin{align*}
\big\|
\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} -
\chi_{n,1}\partial_2&\chi_{n,2}\big]
{\tilde d} {\psi}_{{\mathcal E}_n, n,\xi_n} \big \| \\
&\le \sqrt{2nB_0}\big\|
\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} -
\chi_{n,1}\partial_2\chi_{n,2}\big] {\psi}_{{\mathcal E}_n, n-1,\xi_n}
\big \| \hspace{0.27cm}
\\ & \hspace{0.63cm}+
\tfrac{ {\mathcal E}_n}{2B_0}\big\|
\big[-\ri\chi_{n,2} \partial_1\chi_{n,1} -
\chi_{n,1}\partial_2\chi_{n,2}\big]{\psi}_{{\mathcal E}_n, n,\xi_n}
\big \| \\ &\le
2\sqrt{2} \|\chi'\|_\infty
\Big( B_0\sqrt{\tfrac{2n}{n^{1+\epsilon}}} +
\tfrac{{\mathcal E}_n}{2}
\sqrt{\tfrac{B_0}{n^{1+\epsilon}}} \Big) \sqrt{r_n}
\end{align*}
and
\begin{align*}
\big\|
\big[-\chi_{n,2} \partial_1^2\chi_{n,1} -\chi_{n,1} \partial_2^2\chi_{n,2}\big]
{\psi}_{{\mathcal E}_n, n,\xi_n}
\big \| \le
2 \sqrt{2}\|\chi''\|_\infty r_n^{-2} \sqrt{r_n} \,.
\end{align*}
Thus, in view of condition \eqref{con4.4} and estimate
\eqref{normthm4}, we get
\begin{align}\label{normlocalierror}
\big\|
{\tilde d}^{\,*}{\tilde d} \varphi_n -
\chi_{n,1}\chi_{n,2}{\tilde d}^{\,*}{\tilde d}
{\psi}_{{\mathcal E}_n, n,\xi_n} \big\| \big/
\|\varphi_n\| \longrightarrow 0
\quad \mathrm{as} \ n \to \infty .
\end{align}
For estimating the remaining term on the r.h.s of \eqref{applyphi},
we expand $V_{{\mathcal R}_n}$ up to second order and obtain,
by \eqref{Vn} and \eqref{En}, that
\begin{align*}
\big| \big[ V_{{\mathcal R}_n}(\bx) - V_n -
{\mathcal E}_n(x_1-x_{1,n})\big] \varphi_n(\bx) \big|
& \le
\big\| \mathrm{Hess} (V_{{\mathcal R}_n}) \big\|_2
(\boldsymbol{\eta_{x,x_n}})
|\bx-\bx_n|^2 | \varphi_n(\bx)|,
\end{align*}
with $\boldsymbol{\eta_{x,x_n}} \in [\bx, \bx_n]$.
Because ${\mathcal R}_n$ are rotations, we have that
$\| \mathrm{Hess} (V_{{\mathcal R}_n})\|_2(\,\cdot\,) =
\|\mathrm{Hess} (V)\|_2({\mathcal R}_n\,\cdot\,)$
for $n \in \N$. Since
$\|\mathrm{Hess} (V)\|_2$
varies with rate $\nu$ along $\mathrm{Im}(\gamma)$ and
since, by \eqref{con4.3} and \eqref{crosscond},
$r_n/|\bx_n|^\nu \to 0$ as $n \to \infty$,
we find a constant $C_{11}>0$ such that
for $n \in \N$ large enough
\begin{align*}
\| \mathrm{Hess} (V_{{\mathcal R}_n})\|_2(\boldsymbol{\eta})
\le C_{11} \| \mathrm{Hess} (V_{{\mathcal R}_n})\|_2(\bx_n) , \quad
\boldsymbol{\eta} \in B_{2r_n}(\bx_n)
\end{align*}
holds. As a consequence,
\begin{align*}
\big\| U_{{\mathcal R}_n} e^{-\ri g}
\big[V_{{\mathcal R}_n} - V_n -
{\mathcal E}_n(x_1-x_{1,n})\big] \varphi_n \big\|
&\le 4C_{11}r_n^2
\| \mathrm{Hess} (V)\|_2({\mathcal R}_n \bx_n)
\|\varphi_n \| \\ & \le
4C_{11} \| \mathrm{Hess} (V)\|_2(\by_n)
\left(\tfrac{|V_n|}{B_0} \right)^{1+\epsilon}
\|\varphi_n \|
\end{align*}
for $n \in \N$ large enough.
In view of \eqref{con4.2}, we conclude that
$(U_{{\mathcal R}_n} e^{-\ri g}\varphi_n)_{n \in \N}$
is a Weyl sequence for $0$.
\end{proof}
\noindent
{\bf Acknowledgments.}
This work has been supported by SFB-TR12
"Symmetries and Universality in Mesoscopic Systems" of the DFG.
The author also wants to thank Edgardo Stockmeyer for useful
discussions and remarks as well as the
{\it Faculdad de F\'isica de la Pontificia Universidad Cat\'olica de Chile}
for the great hospitality.
\begin{appendix}
\section{Essential self-adjointness of the Pauli operator}\label{selfadj}
In this section we recapitulate an argument, originally given
in \cite{Iwatsuka1990}, for proving the essential self-adjointness of
the Pauli operator. As we will see, this argumentation works also for
the relaxed regularity conditions on $B$ and $V$ of Proposition \ref{EssSelf}.
For the proof we first note that for $\varphi \in C_0^\infty(\R^2,\C)$
we can write
$$[(-\ri \nabla -{\bf A})^2 + B]\varphi =
\sum_{k,l = 1}^2 (-\ri \partial_k -A_k)\overline{C_{k,l}}(-\ri \partial_l -A_l)\varphi\,,$$
$$[(-\ri \nabla-{\bf A})^2 - B]\varphi =
\sum_{k,l = 1}^2 (-\ri \partial_k -A_k)C_{k,l}(-\ri \partial_l -A_l)\varphi\,,$$
where $C_{k,l}$ denote the coefficients of the symmetric non-negativ definite matrix
$$C = \mathbbm{1}-\sigma_2 =
\begin{pmatrix}
1&\ri\\
-\ri & 1
\end{pmatrix} = C^*.$$
Furthermore, along the proof we use the notation
$B_R:= \{\bx \in \R^2 |\ |\bx| \le R\}$ and
$S_R := \{\bx \in \R^2 |\ |\bx| = R\} $.
\begin{proof}[Proof of Proposition \ref{EssSelf}]
Since $H$ is a diagonal matrix operator, it suffices to show that
both operators on the diagonal,
$$Q_{\pm}:= [(-\ri \nabla -{\bf A})^2 \pm B +V],$$
are essentially self-adjoint on $C_0^\infty(\R^2,\C)$. Because
$Q_\pm$ are symmetric on $C_0^\infty(\R^2,\C) $,
we have to show $Q_+^*\varphi= \pm \ri \varphi $
implies $\varphi \equiv 0$ for $\varphi \in \mathcal{D}(Q_+^*)$,
respectively $Q_-^*\varphi= \pm \ri \varphi $
implies $\varphi \equiv 0$ for $\varphi \in \mathcal{D}(Q_-^*)$.
We only treat the case $Q_-^*\varphi= \ri \varphi $ since the
others are completely analogous.
Let $\varphi \in \mathcal{D}(Q_-^*)$ be such that
$Q_-^*\varphi= \ri \varphi$, then
\begin{equation} \label{eq:dis}
[(-\ri \nabla-{\bf A})^2 - B + V]\varphi = \ri \varphi
\end{equation}
holds in distributional sense. Due to elliptic regularity theory
(see e.g. \cite{Gilbarg_Trudinger}, \cite{Hellwig}), we obtain
that $\varphi \in C^2(\R^2,\C)$ and that \eqref{eq:dis} holds
strongly. Applying integration by parts results in
\begin{align} \label{eq:dis7}
\begin{split}
& \int_{B_R}\bigg[\sum_{k,l = 1}^2 (-\ri \partial_k -A_k)C_{k,l}
(-\ri \partial_l -A_l)\varphi\bigg] \bar \varphi\, \rd^2x
\\ & \hspace{2.63cm}
= \int_{B_R}\bigg[\sum_{k,l = 1}^2 (-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}\bigg] \, \rd^2x
\\ & \hspace{4.15cm}
+ \ri \int_{S_R}\bigg[\sum_{k,l = 1}^2 \bold \nu_k C_{k,l}
(-\ri \partial_l -A_l)\varphi\bigg] \bar \varphi\, \rd S,
\end{split}
\end{align}
with $R>0$, where $\bold \nu_k(\bx) = x_k/|\bx|$ for $k= 1,2$.
By taking the imaginary part of \eqref{eq:dis7},
we conclude with \eqref{eq:dis} that
$$\int_{B_R} |\varphi|^2 \, \rd^2 x =
\int_{S_R}\bigg[\sum_{k,l = 1}^2 \bold \nu_k C_{k,l}
(-\ri \partial_l -A_l)\varphi\bigg] \bar \varphi\, \mbox{d}S$$
for any $R>0$. The Cauchy-Schwarz inequality yields
$$\int_{B_R} |\varphi|^2 \, \rd^2 x \le \!
\bigg(\int_{S_R}\sum_{k,l = 1}^2 (-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}\,\mbox{d}S\bigg)^{1/2} \!
\left(\int_{S_R} |\varphi|^2 \mbox{d}S\right)^{1/2}\!. $$
Hence, it suffices to show that
\begin{equation} \label{eq:dis5}
\int_{\R^2}\sum_{k,l = 1}^2 \frac{(-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}}{|\bx|^2+1}\,\rd^2 x \ <\infty
\end{equation}
since this implies that
$(1,\infty) \ni r \mapsto r^{-1}\int_{B_r} |\varphi|^2 \, \mbox{d}\bx $
is an $L^1$-function, i.e. $\varphi \equiv 0$. \\ For \eqref{eq:dis5}
we consider the function
$$f(R) :=
\int_{B_R}\sum_{k,l = 1}^2 \frac{(-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}}{|\bx|^2+1} \,\rd^2 x\,, $$
with $R>0$. Using Equation \eqref{eq:dis} and integration by parts, we
obtain, with $\zeta(\bx) = ( |\bx|^2+1)^{-1}$ and $M \ge c+ |d|$, that
\begin{align*}
f(R) - M\|\varphi\|^2
&\le f(R) + \int_{B_R} \zeta V |\varphi|^2\, \rd^2 x \\
& = \int_{B_R} \zeta (Q_-^*\varphi) \bar{\varphi} \, \rd^2 x -
\ri \int_{B_R} \bigg[\sum_{k,l = 1}^2 (\partial_l\zeta)
C_{k,l}(-\ri \partial_k -A_k)\varphi \bigg] \bar \varphi\, \,\rd^2x \\
&\hspace{3.19cm} + \ri \int_{S_R} \zeta \bigg[\sum_{k,l = 1}^2 \bold\nu_l C_{k,l}
(-\ri \partial_k -A_k)\varphi
\bigg] \bar \varphi\, \,\mbox{d}S .
\end{align*}
By the estimates
\begin{align*}
\bigg| \int_{B_R} \bigg[\sum_{k,l = 1}^2 (\partial_l\zeta)C_{k,l}
& (-\ri \partial_k -A_k) \varphi \bigg] \bar \varphi\, \,\rd^2 x \bigg| \\
& \le \int_{B_R} 2\zeta^{1/2} \bigg|\sum_{k,l = 1}^2
\bold \nu_l C_{k,l}(-\ri \partial_k -A_k)\varphi \bigg| |\varphi|\,\rd^2 x \\
& \le 2\int_{B_R}\bigg[\sum_{k,l = 1}^2 \zeta(-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}\bigg]^{1/2} |\varphi| \,\rd^2 x \\
& \le 2[f(R)]^{1/2}\|\varphi\| \le \frac{1}{2}f(R) + 2 \|\varphi\|^2
\end{align*}
and
\begin{align*}
\bigg|\int_{S_R}&\zeta\bigg[\sum_{k,l = 1}^2\bold\nu_k C_{k,l}
(-\ri \partial_l - A_l)\varphi\bigg] \bar \varphi\, \rd S \bigg|
\\ & \le
\int_{S_R}\zeta\bigg[\sum_{k,l = 1}^2 (-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}\bigg]^{1/2} |\varphi|\, \rd S
\\ & \le
\bigg(\int_{S_R}\sum_{k,l = 1}^2 \zeta(-\ri \partial_k -A_k)\varphi C_{k,l}
\overline{(-\ri \partial_l -A_l)\varphi}\,\rd S\bigg)^{1/2} \!
\left(\int_{S_R} |\varphi|^2 \rd S\right)^{1/2} \\ & =
\left({f'(R)} \,\int_{S_R} |\varphi|^2 \rd S\right)^{1/2} ,
\end{align*}
we conclude that
$$f(R) \le 2(3+c) \|\varphi\|^2+
2\left({f'(R)} \,\int_{S_R} |\varphi|^2\rd S\right)^{1/2}\!.
$$
If $f(R)= 0$ for all $R>0$, then clearly \eqref{eq:dis5} holds. If
$f(R_0) > 0$ for some $R_0>0$, then $f(R) > 0$ for all $R>R_0$ and
$f'(R)/f^2(R) \in L^1((R_0, \infty))$, implying that
$$\left(\frac{f'(R)}{f^2(R)} \,
\int_{S_R} |\varphi|^2 \mbox{d}S\right)^{1/2} \ \in L^1((R_0, \infty)).$$
Hence, there exists a sequence $(R_n)_{n\in \N} \subset (0,\infty)$ such that
$R_n \to \infty$ as $n \to \infty$ and
$$\left({f'(R_n)}\,
\int_{S_{R_n}} |\varphi|^2 \mbox{d}S\right)^{1/2} \le \frac{1}{4} \, f(R_n).$$
Therefore, we have $f(R_n) \le 4(3+c) \|\varphi\|^2$ for all $n\in\N$,
which implies
\eqref{eq:dis5} since $f(R)$ is a monotonically increasing function.
\end{proof}
\section{Remarks on locally compact operators}
\begin{lemma}\label{Gertrud}
Let $A$ be a locally compact, self-adjoint operator on $L^2(\R^n,\C^m)$
with $n,m \ge 1$. Assume there is a function
$W \in L_{loc}^\infty(\R^n, [0,\infty))$,
with $W(\bx) \to \infty$ as $|\bx| \to\infty$,
such that
$$\| A \varphi\| \ge \|W\varphi\| \quad \mbox{for} \
\varphi \in \mathcal{D}(A). $$
Then $\sigma_{{\rm ess}}(A) = \emptyset$, i.e.
$A$ has only discrete spectrum.
\end{lemma}
\begin{proof}
Assume $\lambda \in \sigma_{{\rm ess}}(A) \subset \R$. Then there is
a normalized sequence $(\varphi_n)_{n \in\N} \subset \mathcal{D}(A)$
such that
$\varphi_n \rightharpoonup 0$
and $\|(A-\lambda)\varphi_n\| \to 0$
as $n \to \infty$. Let $R> 0$ be a fixed constant. We have
\begin{align*}
\| \chi_R \varphi_n \| &=
\| \chi_R (A-\ri - \lambda)^{-1} (A-\ri - \lambda) \varphi_n \|
\\ & \le
\| \chi_R (A-\ri - \lambda)^{-1}\|
\| (A- \lambda) \varphi_n \| +
\| \chi_R (A-\ri - \lambda)^{-1} \varphi_n \| \,,
\end{align*}
using the notation $\chi_R := \chi_{B_R(0)}$. Since
$ \chi_R (A-\ri -\lambda)^{-1}$ is compact, this inequality implies
that $\|\chi_R\varphi_n\|\to 0$ as $n \to \infty$. Let $ R>0$ be so
large that $ W(\bx) \ge 5|\lambda| + 1 $ if $|\bx| \ge R$.
Choosing $N \in \N$ large enough, we can estimate, for $n \ge N$,
\begin{align*}
\| (A-\lambda) \varphi_n\| & \ge
\|W\varphi_n\| - |\lambda| \\ & \ge
\|W(\mathbbm{1}-\chi_R) \varphi_n\| -
\|W\chi_R\|_\infty \|\chi_R\varphi_n\| -|\lambda| \\ & \ge
(5|\lambda|+1)\|(\mathbbm{1}-\chi_R)\varphi_n\|-
\|W\chi_R\|_\infty \|\chi_R\varphi_n\| -|\lambda| \\ & \ge
(|\lambda| +1/2) - \|W\chi_R\|_\infty \|\chi_R\varphi_n\|\,.
\end{align*}
Hence, $\|(A-\lambda)\varphi_n\| \nrightarrow 0$ as $n \to \infty$,
which is a contradiction.
\end{proof}
\section{Integral estimates}
\begin{lemma}\label{hermiteestimate}
Let $\vartheta_n$ be the $n-$th Hermite polynomial. Then,
for any $\epsilon >0$, it holds that
\begin{align*}
\frac{1}{2^n n!}\int_{\sqrt{n^{1+\epsilon}}}^\infty
|\vartheta_n (x)|^2 e^{-x^2} \rd x \ &\longrightarrow 0
\quad \mbox{as} \ n \to \infty,\\[0.2cm]
\frac{1}{2^n n!}\int_{-\infty}^{-\sqrt{n^{1+\epsilon}}}
|\vartheta_n (x)|^2 e^{-x^2} \rd x \ &\longrightarrow 0
\quad \mbox{as} \ n \to \infty.
\end{align*}
\end{lemma}
\begin{proof}
We only treat the first case since the second claim can
be deduced from the first one by a symmetry argument.
Due to the identity
$$\vartheta_n(x) = (-1)^n \sum_{k_1+2 k_2=n}
\frac{n!}{k_1!k_2!} (-1)^{k_1+k_2}(2x)^{k_1}$$
for the $n-$th Hermite polynomial (see e.g. \cite{Abramowitz}),
we obtain for $|x|\ge 1$ the estimate
\begin{align*}
|\vartheta_n(x)| \le \sum_{k_1+2 k_2=n}
\frac{n!}{k_1!k_2!} (2|x|)^{k_1}
\le \frac{n+1}{2}\, 2^n n! |x|^n.
\end{align*}
Thus, for $n \in \N$ large enough we have
\begin{align*}
\frac{1}{2^nn!}\int_{\sqrt{n^{1+\epsilon}}}^\infty
|\vartheta_n(x)|^2 e^{-x^2}\rd x
&\le \frac{(n+1)^2}{4}2^n n!\int_{\sqrt{n^{1+\epsilon}}}^\infty
x^{2n} e^{-x^2} \rd x \\[0.1cm]
&\le \frac{(n+1)^2}{4}2^n
n^{1+\epsilon} n! \, \exp{(-n^{1+\epsilon}/2)}
\int_{\sqrt{n^{1+\epsilon}}}^\infty e^{-x^2/2} \rd x,
\end{align*}
and the r.h.s. tends to $0$ as $n \to \infty$ by Stirling's
Formula.
\end{proof}
\end{appendix}
\bibliographystyle{plain}
|
\section{Introduction}
\label{sec:introduction}
A typical problem in extremal combinatorics has the following form: What is the largest size of a structure which does not contain any forbidden configurations? Once this extremal value is known, it is very natural to ask how many forbidden configurations one is guaranteed to find in every structure of a certain size which is larger than the extremal value. There are many results of this kind. Most notably, there is a very large body of work on the problem of determining the smallest number of $k$-vertex cliques in a graph with $n$ vertices and $m$ edges, attributed to Erd{\H{o}}s and Rademacher; see~\cite{Er62, Er69, ErSi83, LoSi83, Ni11, Ra08, Re12}. In extremal set theory there, is an extension of the celebrated Sperner's theorem, where one asks for the minimum number of chains in a family of subsets of $\{1, \ldots, n\}$ with more than $\binom{n}{\lfloor n/2 \rfloor}$ members; see~\cite{DaGaSu13, DoGrKaSe13, ErKl74, Kl68}. Another example is a recent work in \cite{DaGaSu14}, motivated by the classical theorem of Erd\H{o}s, Ko, and Rado. It studies how many disjoint pairs must appear in a $k$-uniform set system of certain size.
One can ask analogous questions in Ramsey theory. Once we know the maximum size of a~structure which does not contain some unavoidable pattern, we may ask how many such patterns are bound to appear in every structure whose size exceeds this maximum. This direction of research has also been explored in the past. For example, a well-known problem of Erd{\H{o}}s is to determine the minimum number of monochromatic $k$-vertex cliques in a $2$-coloring of the edges of $K_n$;
see, e.g.,~\cite{Co12, FrRo93, Th89}. This may be viewed as a natural extension of Ramsey's theorem.
In this paper, we consider a similar generalization of another classical result in Ramsey theory, the famous theorem of Erd{\H{o}}s and Szekeres~\cite{ErSz35}, which states that for every positive integer $k$, any sequence of more than $k^2$ numbers contains a monotone (that is, monotonically increasing or monotonically decreasing) subsequence of length $k+1$. To be more precise, we shall be interested in the following very natural problem.
\begin{prob}
\label{prob:main}
For every $k$ and $n$, determine the minimum number of monotone subsequences of length $k+1$ in a sequence of $n$ numbers.
\end{prob}
It is not clear when Problem~\ref{prob:main} was originally posed. It appears first in print in a paper of Myers~\cite{My02}, who attributes it to Albert, Atkinson, and Holton. It follows from the aforementioned theorem of Erd{\H{o}}s and Szeker{\'e}s that every sequence of $n$ numbers contains at least $n-k^2$ monotone subsequences of length $k+1$. When $n \le k^2+k$, this is easily seen to be sharp by considering a sequence built from $k$ increasing sequences of lengths $k$ or $k+1$ by concatenating them in decreasing order, such as the sequence $\tau_{k,n}$ defined below. Without loss of generality, we may restrict our attention to sequences that are permutations of the set $\{1, \ldots, n\}$, which we shall from now on abbreviate by~$[n]$.
Let us denote by $S_n$ the set of all permutations of $[n]$. Following~\cite{My02}, given a permutation $\sigma \in S_n$, we let $m_k(\sigma)$ denote the number of monotone subsequences of length $k+1$ in $\sigma$ and let
\[
m_k(n) = \min\{m_k(\sigma) \colon \sigma \in S_n \}.
\]
In order to give an upper bound on $m_k(n)$, consider the permutation $\tau_{k,n}$ described by
\[
\begin{array}{llll}
\lfloor (k-1)n/k \rfloor + 1, & \lfloor (k-1)n/k \rfloor + 2, & \ldots, & n \\
\lfloor (k-2)n/k \rfloor + 1, & \lfloor (k-2)n/k \rfloor + 2, & \ldots, & \lfloor (k-1)n/k \rfloor \\
\vdots & \vdots & & \vdots \\
1, & 2, & \ldots, & \lfloor n/k \rfloor.
\end{array}
\]
Let $r_{k,n}$ be the unique number $r \in \{0, \ldots, k-1\}$ satisfying $r \equiv n \pmod k$. Since $\tau_{k,n}$ contains no decreasing subsequences of length $k+1$, it is easy to see that
\begin{equation}
\label{eq:mktau}
m_k(\tau_{k,n}) = r_{k,n} \binom{\lceil n/k \rceil}{k+1} + (k-r_{k,n}) \binom{\lfloor n/k \rfloor}{k+1}.
\end{equation}
It seems quite natural to guess that $m_k(n) = m_k(\tau_{k,n})$ for all $k$ and $n$, that is, that $\tau_{k,n}$ contains the minimum number of monotone subsequences of length $k+1$ among all permutations of $[n]$. This was conjectured by Myers~\cite{My02}, who noticed that Goodman's formula~\cite{Go59}, indeed proves that $m_2(n) = m_2(\tau_{2,n})$ for all $n$ and yields a characterisation of all permutations achieving equality.
\begin{conj}[Myers~\cite{My02}]
\label{conj:main}
Let $n$ and $k$ be positive integers. In any permutation of $[n]$, there are at least $m_k(\tau_{k,n})$ monotone subsequences of length $k+1$.
\end{conj}
Very recently, Balogh et al.~\cite{BaHuLiPiUdVo} proved this conjecture for $k=3$ and sufficiently large $n$ and described all extremal permutations. Their proof uses computer assistance and is based on the framework of flag algebras.
\subsection{Our results}
In this paper, we provide first evidence supporting Conjecture~\ref{conj:main} for large $k$. Our main result is that the conjecture holds for all sufficiently large $k$,
as long as $n$ is not much larger than $k^2$. This provides a good understanding of how the minimum number of monotone subsequences of length $k+1$ grows in a short interval above the extremal threshold. Our results are similar in spirit to the ones of~\cite{Er62, LoSi83}, which determine the minimum number of cliques in graphs whose number of edges is slightly supercritical (larger than the Tur{\'a}n number for the clique).
\begin{thm}
\label{thm:main}
There exist an integer $k_0$ and a positive real $c$ such that $m_k(n) = m_k(\tau_{k,n})$ for all $k$ and $n$ satisfying $k \ge k_0$ and $n \le k^2 + c k^{3/2}/\log k$. Moreover, if $n \neq k^2+k+1$ and $m_k(\sigma) = m_k(n)$ for some $\sigma \in S_n$, then $\sigma$ contains monotone subsequences of length $k+1$ of only one type (increasing or decreasing).
\end{thm}
\begin{figure}[h]
\centering
\begin{tabular}{ccccc}
\begin{tikzpicture}
\def0.8{0.35}
\def\sigma{
{9, 10, 11, 12, 13, 5, 6, 7, 8, 1, 2, 3, 4}
}
\draw[step=0.8,black,very thin] (0,0) grid (12*0.8,12*0.8);
\foreach \x in {0, ..., 12}
{
\filldraw (\x*0.8, \sigma[\x]*0.8-0.8) circle[radius=2pt];
}
\end{tikzpicture}
&
\quad
&
\begin{tikzpicture}
\def0.8{0.35}
\def\sigma{
{10, 6, 11, 12, 13, 2, 7, 8, 9, 1, 3, 4, 5}
}
\draw[step=0.8,black,very thin] (0,0) grid (12*0.8,12*0.8);
\foreach \x in {0, ..., 12}
{
\filldraw (\x*0.8, \sigma[\x]*0.8-0.8) circle[radius=2pt];
}
\end{tikzpicture}
&
\quad
&
\begin{tikzpicture}
\def0.8{0.35}
\def\sigma{
{10, 6, 11, 12, 13, 3, 7, 8, 9, 1, 2, 4, 5}
}
\draw[step=0.8,black,very thin] (0,0) grid (12*0.8,12*0.8);
\foreach \x in {0, ..., 12}
{
\filldraw (\x*0.8, \sigma[\x]*0.8-0.8) circle[radius=2pt];
}
\end{tikzpicture}
\end{tabular}
\caption{Permutations $\tau_{3,13}$, $\sigma_3^1$, and $\sigma_3^2$.}
\label{fig:sigma-ik}
\end{figure}
Somewhat surprisingly, if $n = k^2+k+1$, then there are $\sigma \in S_n$ with $m_k(\sigma) = m_k(n) = 2k+1$ which contain both increasing and decreasing subsequences of length $k+1$. Two such permutations are $\sigma_k^1$ and $\sigma_k^2$, where $\sigma_k^i$ is described by
\[
\begin{array}{llllll}
k^2+1, & k^2-k, & k^2+2, & k^2+3, & \ldots, & k^2+k+1, \\
& k^2-2k-1, & k^2-k+1, & k^2-k+2, & \ldots, & k^2,\\
& \vdots & \vdots & \vdots & & \vdots \\
& k+3, & 2k+5, & 2k+6, & \ldots, & 3k+4, \\
& 1+i, & k+4, & k+5, & \ldots, & 2k+3,\\
& 1, & 4-i, & 4, & \ldots, & k+2.
\end{array}
\]
One can check that $\sigma_k^i$ contains $2k+1-i$ increasing subsequences of length $k+1$ and $i$ decreasing subsequences of length $k+1$, see Figure~\ref{fig:sigma-ik}. However, no permutation $\sigma$ with $m_k(\sigma) = m_k(n) = 2k+1$ can have more monotone subsequences of length $k+1$ of the `odd' type than $\sigma_k^2$. It will follow from our proof of Theorem~\ref{thm:main} that for each extremal permutation, at least $2k-1$ out of its $2k+1$ monotone subsequences of length $k+1$ are of the same type. For details, we refer the reader to Theorem~\ref{thm:posets} and Example~\ref{example:extremal-posets}.
\subsection{Chains and antichains in posets}
Every permutation admits a natural representation as a poset (partially ordered set) in which its increasing and decreasing subsequences are mapped to chains and antichains, respectively. Indeed, given a permutation $\sigma$ of $[n]$, one may define a binary relation $\le_\sigma$ on $[n]$ by letting $i \le_\sigma j$ if and only if $i \le j$ and $\sigma(i) \le \sigma(j)$. It is not hard to see that $P_\sigma = ([n], \le_\sigma)$ is a poset whose chains and antichains are in a one-to-one length-preserving\footnote{By `length' of a chain or an antichain we mean the number of its elements.} correspondence with increasing and decreasing subsequences in $\sigma$, respectively. Via this correspondence, one may easily deduce the theorem of Erd{\H{o}}s and Szekeres from the famous theorem of Dilworth~\cite{Di50}, which says that every poset containing no antichain with $k+1$ elements admits a partition into $k$ chains, or its much easier to prove dual version (due to Mirsky~\cite{Mi71}), which says that every poset containing no chain with $k+1$ elements can be partitioned into $k$ antichains, see Section~\ref{sec:decomposition}.
Let us call a set $A$ of elements of a poset \emph{homogenous} if $A$ is a chain or an antichain. A natural generalization of Problem~\ref{prob:main} to posets would be the following.
\begin{prob}
\label{prob:posets}
For every $k$ and $n$, determine the minimum number of homogenous $(k+1)$-element sets in a poset with $n$ elements.
\end{prob}
Given a poset $P$, we let $h_k(P)$ denote the number of homogenous sets of cardinality $k+1$ in $P$ and
\[
h_k(n) = \min\{h_k(P) \colon \text{$P$ is a poset with $n$ elements}\}.
\]
It follows from the above discussion that $h_k(n) \le m_k(n)$ and we think that it is natural to ask the following.
\begin{quest}
Is it true that $h_k(n) = m_k(n)$ for all $n$ and $k$?
\end{quest}
Clearly, not every poset is isomorphic to $P_\sigma$ for some permutation $\sigma$. In fact, this is the case precisely for posets of \emph{order dimension} at most two, that is, posets that are the intersection of two linear orders. Nevertheless, more for the sake of convenience rather than generality, we shall present our arguments using the language of posets. We remark here that several times in the proof, we will use the fact that we can `flip' our poset $P$, exchanging the roles of chains and antichains. That is, we will assume that there is a dual poset $P^*$ defined on the same set as $P$ such that every pair of elements is comparable in either $P$ or $P^*$ but not both of them. This is possible only for posets of order dimension at most two, that is, ones that represent permutations. Indeed, for every permutation $\sigma \in S_n$, we have $P_\sigma^* = P_{\sigma^*}$, where $\sigma^*(i) = n+1-\sigma(i)$. The converse statement was proved by Dushnik and Miller~\cite{DuMi41}. Let us now rephrase Theorem~\ref{thm:main} in the language of posets.
\begin{thm}
\label{thm:posets}
There exist an integer $k_0$ and a positive real $c$ such that the following is true. Let $k$ and $n$ be integers satisfying $k \ge k_0$ and $n \le k^2 + c k^{3/2} / \log k$. If $P$ is an $n$-element poset of order dimension at most two, then
\begin{equation}
\label{eq:hkP-lower}
h_k(P) \ge m_k(\tau_{k,n}).
\end{equation}
Moreover, if equality holds in~\eqref{eq:hkP-lower}, then $P$ can be decomposed into $k$ chains or $k$ antichains of length $\lfloor n/k \rfloor$ or $\lceil n/k \rceil$ each, unless $n = k^2 + k + 1$ and $P$ (or $P^*$) is one of the posets described in Example~\ref{example:extremal-posets} below.
\end{thm}
\begin{example}
\label{example:extremal-posets}
Suppose that $n = k^2+k+1$ and observe that $m_k(\tau_{k,n}) = 2k+1$. We describe two families of $n$-element posets with exactly $2k+1$ homogenous $(k+1)$-sets that contain both chains and antichains with $k+1$ elements. Each of these posets has precisely $k+1$ minimal elements; denote the set of these minimal elements by $A_1$. Moreover, $P \setminus A_1$ can be decomposed into $k$ chains as well as into $k$ antichains. In particular, each chain and each antichain in every such decomposition has precisely $k$ elements. Furthermore, $P \setminus A_1$ contains only $k$ chains of length $k$. Let $A_2$ be the set of minimal elements of $P \setminus A_1$ and note that $|A_2| = k$. The comparability graph of the subposet of $P$ induced by $A_1 \cup A_2$ is either (i) a path with $2k+1$ vertices or (ii) the disjoint union of a path with $2k-1$ vertices and an edge. Moreover, if (ii) holds, then one of the elements of $A_1$ belonging to the path of length $2k-1$ is smaller than the second smallest element of the unique $k$-element chain in $P \setminus A_1$ whose smallest element is the unique element of $A_2$ that does not belong to the path. One can check that if (i) holds, then $P$ has precisely $2k$ chains and one antichain with $k+1$ elements and that if (ii) holds, then $P$ has precisely $2k-1$ chains and $2$ antichains with $k+1$ elements. Finally, in both cases, there exist posets of order dimension at most two fitting the description. Two examples of such posets are $P_{\sigma_k^1}$ and $P_{\sigma_k^2}$, where $\sigma_k^1$ and $\sigma_k^2$ are the permutations defined below Theorem~\ref{thm:main}; see Figure~\ref{fig:posets-sigma-ik}.
\end{example}
\begin{figure}[h]
\centering
\begin{tabular}{ccc}
\begin{tikzpicture}
\def0.8{0.8}
\def1.4{1.4}
\def\p{
{10,6,2,1,11,7,3,12,8,4,13,9,5}
}
\foreach \x [evaluate=\x as \lab using ({\p[\x]})] in {0, ..., 3}
{
\node [label=right:$\lab$] at (2*\x*0.8, 0) {};
\filldraw (2*\x*0.8, 0) circle[radius=2pt];
}
\foreach \x in {0, ..., 2}
{
\draw [very thick] (2*\x*0.8, 0) -- (2*\x*0.8 + 0.8, 1.4) -- (2*\x*0.8 + 2*0.8, 0);
\draw (2*\x*0.8 + 0.8, 1.4) -- (2*\x*0.8 + 0.8, 3*1.4);
}
\foreach \y in {1, ..., 3}
{
\foreach \x [evaluate=\x as \lab using ({\p[3*\y+\x+1]})] in {0, ..., 2}
{
\node [label=right:$\lab$] at (2*\x*0.8+0.8, \y*1.4) {};
\filldraw (2*\x*0.8+0.8, \y*1.4) circle[radius=2pt];
}
}
\end{tikzpicture}
&
\qquad\qquad\qquad
&
\begin{tikzpicture}
\def0.8{0.8}
\def1.4{1.4}
\def\p{
{10,6,3,1,11,7,2,12,8,4,13,9,5}
}
\foreach \x [evaluate=\x as \lab using ({\p[\x]})] in {0, ..., 3}
{
\node [label=right:$\lab$] at (2*\x*0.8, 0) {};
\filldraw (2*\x*0.8, 0) circle[radius=2pt];
}
\foreach \x in {0, ..., 1}
{
\draw [very thick] (2*\x*0.8, 0) -- (2*\x*0.8 + 0.8, 1.4) -- (2*\x*0.8 + 2*0.8, 0);
}
\draw (4*0.8, 0) -- (5*0.8, 2*1.4);
\draw [very thick] (5*0.8, 1.4) -- (6*0.8, 0);
\foreach \x in {0, ..., 2}
{
\draw (2*\x*0.8 + 0.8, 1.4) -- (2*\x*0.8 + 0.8, 3*1.4);
}
\foreach \y in {1, ..., 3}
{
\foreach \x [evaluate=\x as \lab using ({\p[3*\y+\x+1]})] in {0, ..., 2}
{
\node [label=right:$\lab$] at (2*\x*0.8+0.8, \y*1.4) {};
\filldraw (2*\x*0.8+0.8, \y*1.4) circle[radius=2pt];
}
}
\end{tikzpicture}
\end{tabular}
\caption{Hasse diagrams of posets $P_{\sigma_3^1}$ and $P_{\sigma_3^2}$. The edges of the comparability graph induced by $A_1 \cup A_2$ are thickened.}
\label{fig:posets-sigma-ik}
\end{figure}
\subsection{Outline of the paper}
The remainder of the paper is organized as follows. In Section~\ref{sec:decomposition}, we describe a canonical decomposition of an arbitrary poset into antichains and introduce several pieces of notation used in the proof of Theorem~\ref{thm:posets} and in Section~\ref{sec:auxiliary-lemmas}, we collect several auxiliary lemmas. In Section~\ref{sec:outline-proof}, we present a brief outline of the proof of Theorem~\ref{thm:posets}. Section~\ref{sec:large-surplus} is devoted to the proof of one of our main lemmas, which provides a lower bound on the number of homogenous sets in posets with large `surplus' (this notion will be defined in Section~\ref{sec:outline-proof}). Finally, Section~\ref{sec:proof} contains the proof of Theorem~\ref{thm:posets}. We close the paper with several concluding remarks.
\section{Decomposition into antichains}
\label{sec:decomposition}
In our arguments, we shall rely on the following canonical decomposition of an arbitrary poset into antichains, cf.\ Mirsky's theorem~\cite{Mi71}. Fix an arbitrary poset $P$. Recall that the height and the width of $P$, which we shall denote by $h(P)$ and $w(P)$, are the cardinalities of the largest chain and the largest antichain in $P$, respectively. For each positive integer $i$, let
\[
A_i = \{x \in P \colon \text{the longest chain $L$ with $\max L = x$ has $i$ elements}\}.
\]
In other words, $A_1$ is the set of minimal elements of $P$ and for every $i \ge 1$, $A_{i+1}$ is the set of minimal elements in $P \setminus (A_1 \cup \ldots \cup A_i)$. For each $i$, the set $A_i$ is an antichain and thus $|A_i| \le w(P)$, and
\[
P = \bigcup_{i=1}^{h(P)} A_i.
\]
For each $i$ with $1 \le i < h(P)$, let $G_i$ be the bipartite graph on the vertex set $A_i \cup A_{i+1}$ whose edges are all pairs $xy$ with $x \in A_i$ and $y \in A_{i+1}$ such that $x \le y$. In other words, $G_i$ is the Hasse diagram of the subposet of $P$ induced by $A_i \cup A_{i+1}$. Observe that each vertex in $A_{i+1}$ has at least one $G_i$-neighbor in $A_i$ (as otherwise it would belong to $A_1 \cup \ldots \cup A_i$). On the other hand, it is possible that some vertices in $A_i$ have degree zero in $G_i$ (as clearly not all elements of $P$ have to belong to some chain of maximum length).
Let $h = h(P)$. For every $i \in [h]$ and $x \in A_i$, we let $u_i(x)$ be the number of chains $L \subseteq P$ of length $h - i + 1$ with $\min L = x$. Observe that $u_h(x) = 1$ for every $x \in A_h$ and that for all $i \in [h-1]$ and $x \in A_i$,
\[
u_i(x) = |\{(x_{i+1}, \ldots, x_h) \in A_{i+1} \times \ldots \times A_h \colon x \le x_{i+1} \le \ldots \le x_h\}|.
\]
Upon making this definition, one easily verifies that the number of chains of length $h$ in $P$ is $\sum_{x \in A_1} u_1(x)$ and that for each $i \in [h-1]$ and $x \in A_i$,
\begin{equation}
\label{eq:ui-uip}
u_i(x) = \sum_{xy \in G_i} u_{i+1}(y).
\end{equation}
Since no element $x \in A_i$ with $u_i(x) = 0$ will be contributing anything to the total count of chains of length $h$, we shall often be focusing our attention on the set $A_i' \subseteq A_i$ defined by
\[
A_i' = \{x \in A_i \colon u_i(x) \ge 1\}.
\]
Let us note here for future reference that~\eqref{eq:ui-uip} implies that there are no edges of $G_i$ between $A_{i+1}'$ and $A_i \setminus A_i'$. As we shall often estimate the sum of $u_i(x)$ over all $x \in A_i$, let us abbreviate it by $\Sigma_i$. That is, for each $i \in [h]$, let
\[
\Sigma_i = \sum_{x \in A_i} u_i(x).
\]
Since each $y \in A_{i+1}$ has at least one $G_i$-neighbor in $A_i$, it follows from~\eqref{eq:ui-uip} that
\begin{equation}
\label{eq:sum-ui-monotone}
\Sigma_{i} \ge \Sigma_{i+1}.
\end{equation}
Clearly, \eqref{eq:sum-ui-monotone} holds with equality if and only if each $y \in A_{i+1}'$ has exactly one $G_i$-neighbor in $A_i$. This naturally leads to the final definition of this section. Namely, for each $i \in [h-1]$, we let
\[
B_{i+1} = \{y \in A_{i+1}' \colon \deg_{G_i}(y) = 1\}.
\]
\section{Auxiliary lemmas}
\label{sec:auxiliary-lemmas}
In this section, we collect a few auxiliary lemmas that will be repeatedly used in the proof of Theorem~\ref{thm:posets}. Our first lemma is a straightforward corollary of the Kruskal--Katona theorem~\cite{Ka68, Kr63}.
\begin{lemma}
\label{lemma:KK}
Suppose that $a \ge b > 0$, let $\mathcal{F}$ be an arbitrary family of $a$-element sets, and define
\[
\partial_b \mathcal{F} = \{B \colon \text{$|B| = b$ and $B \subseteq A$ for some $A \in \mathcal{F}$}\}.
\]
Then
\[
|\partial_b \mathcal{F}| \ge \min \{ |\mathcal{F}|/2, 2^b \}.
\]
\end{lemma}
\begin{proof}
If $\mathcal{F} = \emptyset$, $a=b$, or $a \ge 2b$, then the assertion of the lemma is trivial as
\begin{equation}
\label{eq:a-choose-b}
\binom{2b}{b} \ge 2^b.
\end{equation}
Otherwise, let $m$ be the smallest integer such that $\binom{m}{a} > |\mathcal{F}|$. By the Kruskal--Katona theorem~\cite{Ka68,Kr63}, $|\partial_b \mathcal{F}| \ge \binom{m-1}{b}$. If $m-1 \ge 2b$, then the conclusion follows from~\eqref{eq:a-choose-b}. Otherwise, since $b < a \le m \le 2b$, then $\binom{m}{a} \le \binom{m}{b+1}$ and consequently,
\[
|\partial_b \mathcal{F}| \ge \binom{m-1}{b} > \binom{m-1}{b} \frac{|\mathcal{F}|}{\binom{m}{a}} \ge \frac{\binom{m-1}{b}}{\binom{m}{b+1}} |\mathcal{F}| = \frac{b+1}{m} |\mathcal{F}| \ge \frac{|\mathcal{F}|}{2}.\qedhere
\]
\end{proof}
Our second lemma will be essential in proving a lower bound on the number of chains of length $k+1$ in a poset of height larger than $k+1$ in terms of the number of chains of maximum length. The lemma is somewhat abstract, but it is immediately followed by a much more concrete corollary.
\begin{lemma}
\label{lemma:signatures}
Suppose that $M$ is a positive integer, $X$ and $Y$ are arbitrary sets, and $f_1, \ldots, f_M \colon X \to Y$ are pairwise different functions. There exist sets $X_1, \ldots, X_M \subseteq X$ with $|X_i| \le \log_2 M$ for all $i \in [M]$ such that
\begin{equation}
\label{eq:fifj}
f_i|_{X_i \cup X_j} \neq f_j|_{X_i \cup X_j} \quad \text{for all $i \neq j$}.
\end{equation}
\end{lemma}
\begin{proof}
We prove the lemma by induction on $M$. The statement is trivial if $M=1$ (one takes $I_1 = \emptyset$), so we may assume that $M \ge 2$. Since $f_1, \ldots, f_M$ are pairwise different, there is an $x \in X$ such that not all $f_i$ take the same value at $x$. For each $y \in Y$, let
\[
I(x, y) = \{i \in [M] \colon f_i(x) = y\}
\]
and let $y \in Y$ be a value that maximizes $|I(x, y)|$. Note that $|I(x, y)| < M$ by our choice of $x$ and that $|I(x, z)| \le M/2$ for each $z \in Y \setminus \{y\}$. We apply the inductive assumption separately to $\{f_i \colon i \in I(x, z)\}$ for each $z \in Y$ with $I(x,z) \neq \emptyset$ to obtain sets $X_1', \ldots, X_M' \subseteq X$ such that $|X_i'| < \log_2M$ for every~$i$, $|X_i'| \le \log_2 (M/2)$ for every $i \notin I(x,y)$, and~\eqref{eq:fifj} holds for each pair $\{i,j\}$ which is fully contained in one of the sets $I(x,z)$. It is straightforward to check that the sets $X_1, \ldots, X_M$ defined by
\[
X_i =
\begin{cases}
X_i' & \text{if $i \in I(x,y)$,} \\
X_i' \cup \{x\} & \text{if $i \notin I(x, y)$,}
\end{cases}
\]
satisfy the assertion of the lemma.
\end{proof}
\begin{cor}
\label{cor:signatures}
Let $k$, $\ell$, and $M$ be positive integers, let $P$ be a poset of height $k+\ell$, and suppose that $m := \log_2 M + 1 \le k/4$.
\begin{enumerate}[(i)]
\item
\label{item:signatures-global}
If $P$ contains at least $M$ chains of length $k+\ell$, then it contains at least
\[
\exp \left( - \frac{2(\ell-1)m}{k} \right) \cdot M \binom{k+\ell}{k+1}
\]
chains of length $k+1$.
\item
\label{item:signatures-local}
Given any $y \in P$, (\ref{item:signatures-global}) still holds if we replace `chains' with `chains containing $y$'.
\end{enumerate}
\end{cor}
\begin{proof}
We prove (\ref{item:signatures-global}) and (\ref{item:signatures-local}) simultaneously. Suppose that $L_1, \ldots, L_M$ are pairwise distinct chains of length $k+\ell$. For (\ref{item:signatures-local}), assume moreover that each $L_i$ contains $y$. Viewing each chain $L_i$ as a function from $[k+\ell]$ to $P$, we invoke Lemma~\ref{lemma:signatures} to obtain sets $X_1, \ldots, X_M \subseteq [k+\ell]$ such that $|X_i| \le \log_2M$ for each $i$ and $L_i(X_i \cup X_j) \neq L_j(X_i \cup X_j)$ whenever $i \neq j$. For (\ref{item:signatures-local}), add to each $X_i$ the unique index $x$ such that $L_i(x) = y$. By the definition of $m$, we have that $|X_i| \le m$ for each $i$.
Let $N$ denote the number of chains of length $k+1$ that are obtained by fixing an $i \in [M]$ and an arbitrary $(k+1)$-set $C_i \subseteq [k+\ell]$ such that $X_i \subseteq C_i$ and considering the set $L_i(C_i)$. Note crucially that for any $i \neq j$ and any choice of $C_i$ and $C_j$ as above, the sets $L_i(C_i)$ and $L_j(C_j)$ are \emph{different} chains (containing $y$). Indeed, since $L_i$ and $L_j$ are chains of maximum length, then for every $z \in L_i \cap L_j$, there is a unique $x \in [k+\ell]$ such that $L_i(x) = z = L_j(x)$. Hence,
\[
\begin{split}
N & = \sum_{i=1}^M \binom{|L_i| - |X_i|}{k+1 - |X_i|} = \sum_{i=1}^M \binom{k+\ell - |X_i|}{\ell - 1} \ge M \binom{k+\ell-m}{\ell-1} \\
& \ge \left(1 - \frac{1}{k-m}\right)^{(\ell-1) m} \cdot M\binom{k+\ell}{\ell-1} \ge \exp\left(-\frac{2(\ell - 1)m}{k}\right) \cdot M \binom{k+\ell}{k+1}.
\end{split}
\]
Above, we used the fact that $\binom{a}{b} \ge \left(1 - \frac{1}{a-b}\right)^b \binom{a+1}{b}$ and that $1-x \ge e^{-3x/2}$ if $x \le 1/3$.
\end{proof}
We close this section with a simple lower bound on the number of connected sets in trees. We shall use this bound in the analysis of one of the almost extremal cases in the proof of Theorem~\ref{thm:posets}.
\begin{lemma}
\label{lemma:connected-sets}
If $1 \le c \le t$, then every tree with $t$ vertices contains at least $t-c+1$ connected subsets with $c$ elements.
\end{lemma}
\begin{proof}
We prove the statement by induction on $t-c$. It is certainly true if $t = c$. Assume that $t \ge c+1$ and let $T$ be a tree with $t$ vertices. Let $v$ be an arbitrary leaf of $T$ and set $T' = T - v$. By the inductive assumption, $T'$ contains at least $t-c$ connected subsets with $c$ elements. On the other hand, it is easy to check that $v$ is contained in at least one connected subset of $T$ of any given size between $1$ and $t$.
\end{proof}
\section{Outline of the proof}
\label{sec:outline-proof}
Roughly speaking, our proof of Theorem~\ref{thm:posets} is a combination of a stability-type argument and an induction on $n$. More precisely, given an $n$-element poset $P$, we either find an element $x \in P$ which belongs to at least $m_k(\tau_{k,n}) - m_k(\tau_{k,n-1})$ homogenous $(k+1)$-sets, in which case we may simply appeal to the inductive assumption on $P \setminus \{x\}$, or we show more `directly' that $P$ contains at least $m_k(\tau_{k,n})$ homogenous $(k+1)$-sets. Some extra work is needed to deduce the claimed structural description of $P$ when $h_k(P) = m_k(\tau_{k,n})$. At all times, we rely heavily on the assumption that the order dimension of $P$ is at most two and hence $P$, as well as each of its induced subposets, has a dual poset $P^*$. This assumption allows us to focus on counting chains, since we may always replace $P$ with $P^*$, exchanging the roles of chains and antichains. We shall tacitly assume that $h(P) \ge w(P)$, as $h(P^*) = w(P)$, and that $h(P) < \lceil n/k \rceil$, as otherwise each element of a longest chain in $P$ belongs to at least $m_k(\tau_{k,n}) - m_k(\tau_{k,n-1})$ chains of length $k+1$.
We first show that if $P$ is `far' from being a union of $k$ chains (or $k$ antichains), then the number of homogenous $(k+1)$-sets is much greater than $m_k(\tau_{k,n})$. To this end, we define a simple parameter termed surplus which measures the distance between a poset and a union of $k$ chains. Let $P$ be an arbitrary poset of height $h$ and let $k$ be an integer. The \emph{$k$-surplus} of $P$, denoted by $s_k(P)$, is defined by $s_k(P) = n - hk$. Observe that
\begin{equation}
\label{eq:surplus}
s_k(P) = \sum_{i = 1}^h (|A_i| - k),
\end{equation}
where $(A_i)_{i=1}^h$ is the canonical decomposition of $P$ into antichains. In Section~\ref{sec:large-surplus}, we show that if $s_k(P) = \Omega(k)$, then $h_k(P) = 2^{\Omega(\sqrt{k})} \gg m_k(\tau_{k,n})$. On the other hand, $s_k(P) = o(k)$ together with $h(P) < \lceil n/k \rceil$ imply that $h(P) = \lceil n/k \rceil - 1$.
In the remainder of the proof, we prove a sequence of lower bounds on $\Sigma_1$, the number of chains of maximum length. The proof of each of these bounds relies on the analysis of the graphs $G_i$ for various indices $i$ such that $A_i \cup A_{i+1}$ contains an antichain with $k+1$ elements. Roughly speaking, we show that for each such $i$, either $\Sigma_i - \Sigma_{i+1}$ is large or $A_i \cup A_{i+1}$ contains many $(k+1)$-element antichains. Each of these bounds gives some sufficient conditions on $P$ which imply that $h_k(P) > m_k(\tau_{k,n})$; here, we use Corollary~\ref{cor:signatures} to translate a lower bound on the number of chains of length $h(P)$ into a lower bound on the number of chains of length $k+1$. If $P$ does not satisfy any of these conditions, then the canonical decomposition of $P$ into antichains becomes greatly restricted. In particular, there are very few indices $i$ with $|A_i| > k$ and $|A_i'| \approx |A_i|$ for all $i$. Finally, some careful case analysis, which involves counting both chains and antichains with $k+1$ elements, shows that $h_k(P) \ge m_k(\tau_{k,n})$ and this inequality is strict unless $n = k^2+k+1$ and $P$ (or $P^*$) is one of the posets described in Example~\ref{example:extremal-posets}.
\section{Posets with large surplus}
\label{sec:large-surplus}
In this section, we prove one of our key lemmas. It says that posets with large $k$-surplus and no `bottlenecks' (small sets whose deletion reduces the height) contain many chains of maximum length or many $(k+1)$-element antichains.
\begin{lemma}
\label{lemma:large-surplus}
Let $d$, $k$, and $s$ be integers satisfying $1 \le d \le k$ and suppose that $P$ is a poset such that $s_k(P) \ge s$ and deletion of no $s/2$ elements reduces the height of $P$. Then $P$ contains either at least $2^d$ antichains with $k+1$ elements or at least $2^{\lfloor s/(2d) \rfloor}$ chains of length $h(P)$.
\end{lemma}
\begin{proof}
We fix $d$ and $k$ with $1 \le d \le k$ and prove the statement by induction on $s$. If $0 \le s < 2d$, then the assertion of the lemma holds vacuously. Suppose now that $s \ge 2d$ and that $P$ satisfies the assumptions of the lemma. Let $(A_j)_{j=1}^{h(P)}$ be the canonical decomposition of $P$ into antichains and let $i$ be the smallest index such that $|A_i| > k$; such $i$ exists since $s_k(P) > 0$, see~\eqref{eq:surplus}. Since $A_i$ is an antichain, we may assume that $|A_i| \le k+d$ since otherwise the number $N$ of $(k+1)$-element antichains in $A_i$ alone satisfies
\[
N \ge \binom{k+d+1}{k+1} = \binom{k+d+1}{d} = \prod_{j=1}^d \frac{k+j+1}{j} \ge 2^d,
\]
where the last inequality follows from our assumption that $d \le k$. Recall the definition of $G_i$ and let $B_{i+1}$ be the set of all $y \in A_{i+1}$ with at most (exactly) one $G_i$-neighbor in $A_i$. For every $Y \subseteq B_{i+1}$, the set $(A_i \setminus N_{G_i}(Y)) \cup Y$ is an antichain with at least $|A_i|$ elements and therefore we may further assume that $|B_{i+1}| < d$ as otherwise $A_i \cup B_{i+1}$ contains at least $2^d$ antichains of size $k+1$. To see this, for every $Y \subseteq B_{i+1}$ with $|Y| \le d$, consider an arbitrary $(k+1)$-element set $L$ with $Y \subseteq L \subseteq (A_i \setminus N_{G_i}(Y)) \cup Y)$. By~\eqref{eq:ui-uip} and~\eqref{eq:sum-ui-monotone},
\begin{equation}
\label{eq:chains}
\sum_{x \in A_1} u_1(x) \ge \sum_{x \in A_i} u_i(x) = \sum_{y \in A_{i+1}} \deg_{G_i}(y) u_{i+1}(y) \ge 2 \sum_{y \in A_{i+1} \setminus B_{i+1}} u_{i+1}(y).
\end{equation}
Let $h = h(P)$ and observe that the sum in the right-hand side of~\eqref{eq:chains} is precisely the number of chains of length $h-i$ in the poset $P'$ obtained from $P$ by deleting $A_1 \cup \ldots \cup A_i \cup B_{i+1}$. In order to estimate this sum, we shall apply the inductive assumption to $P'$. First, note that $h(P') \le h-i$, as $A_{i+1} \setminus B_{i+1}, A_{i+1}, \ldots, A_h$ partition $P'$ into $h-i$ antichains. Moreover, if $h(P' \setminus X) < h-i$ for some $X \subseteq P'$, then $h(P \setminus (B_{i+1} \cup X)) < h$ as every chain in $P$ contains at most one element from each of $A_1, \ldots, A_i$. Therefore, not only $h(P') = h-i$, as $|B_{i+1}| < d \leq s/2$, but also the deletion of no $s/2-d$ elements reduces the height of $P'$. Furthermore,
\[
\begin{split}
s_k(P') & = |P'| - k(h-i) = |P| - \sum_{j=1}^i (|A_j| - k) - |B_{i+1}| - hk \\
& \ge s_k(P) - (|A_i| - k) - |B_{i+1}| > s-2d.
\end{split}
\]
The first inequality above follows from the minimality of $i$ and the second inequality from the assumptions that $|A_i| \le k+d$ and $|B_{i+1}| < d$. Hence, $P'$ satisfies the assumptions of the lemma with $s$ replaced by $s-2d$. This means that either $P'$ contains at least $2^d$ antichains of size $k+1$ or
\[
\sum_{y \in A_{i+1} \setminus B_{i+1}} u_{i+1}(y) \ge 2^{\lfloor s/(2d) \rfloor - 1}.
\]
By~\eqref{eq:chains}, this completes the proof of the lemma.
\end{proof}
\begin{cor}
\label{cor:large-surplus}
Let $k$ and $t$ be integers satisfying $0 < t \le k/2$ and suppose that $P$ is a poset of order dimension at most two such that $h(P) \ge w(P)$ and $s_k(P) \ge 3t$. Then $P$ contains at least $2^{\sqrt{t}-1}$ homogenous $(k+1)$-sets.
\end{cor}
\begin{proof}
We first `prune' $P$ by repeatedly performing the following two-step procedure:
\begin{enumerate}[(1)]
\item
\label{item:prune-1}
If $P$ contains a set $S$ of at most $t$ elements whose deletion reduces the height of $P$, then remove the smallest such $S$ from $P$.
\item
\label{item:prune-2}
If $h(P) < w(P)$, then replace $P$ with $P^*$, exchanging the roles of chains and antichains.
\end{enumerate}
Let us list several properties of the `pruning' procedure. First, performing (\ref{item:prune-1}) decreases the height of $P$ by exactly one at the cost of deleting at most $t$ elements. Thus, each time (\ref{item:prune-1}) is executed, the $k$-surplus of $P$ increases by at least $k-t$. Second, since in the beginning and after (\ref{item:prune-2}) is performed, $h(P) \ge w(P)$, step (\ref{item:prune-2}) can be executed only in conjunction with (\ref{item:prune-1}). Third, each time (\ref{item:prune-2}) is performed, it increases the height of $P$ by exactly one, thus reducing the $k$-surplus of $P$ by $k$. Moreover, this cannot happen in two consecutive rounds, since immediately after (\ref{item:prune-2}) is triggered, $h(P) > w(P)$.
Therefore, letting $P'$ denote the final outcome of the `pruning' procedure and $r$ the number of rounds, we have (recall that $k \ge 2t$)
\[
\begin{split}
s_k(P') & \ge s_k(P) + r(k-t) - \lceil r/2 \rceil k \ge 3t + \lfloor r / 2 \rfloor k - rt \\
& \ge 3t + (r-1)t - rt = 2t > 0.
\end{split}
\]
In particular,
\[
h(P') \ge w(P') \ge k + \frac{s_k(P')}{h(P')} > k.
\]
Since $P'$ clearly does not contain a set of $t$ elements whose deletion reduces the height of $P'$, Lemma~\ref{lemma:large-surplus} with $d = \lfloor \sqrt{t} \rfloor$ and $s = 2t$ implies that $P'$ contains either at least $2^{\sqrt{t}-1}$ antichains of size $k+1$ or at least $2^{\sqrt{t}}$ chains of length $h(P')$ and consequently, by Lemma~\ref{lemma:KK}, also at least $2^{\sqrt{t}-1}$ chains of length $k+1$. Finally, since $P$ contains either $P'$ or $(P')^*$, every homogenous set in $P'$ is also homogenous in $P$.
\end{proof}
\section{Proof of Theorem~\ref{thm:posets}}
\label{sec:proof}
Let $k$ be a sufficiently large integer. We prove the theorem by induction on $n$. The assertion is trivial when $n \le k^2$ as then $m_k(\tau_{k,n}) = 0$ and every poset with no chain of length $k+1$ can be covered by $k$ antichains, see Section~\ref{sec:decomposition}. Therefore, suppose that $k^2 < n \le k^2 c k^{3/2}/\log_2 k$, where $c = 1/300$, and let $P$ be an arbitrary $n$-element poset of order dimension at most two. Without loss of generality, we may assume that $h(P) \ge w(P)$ as otherwise we may replace $P$ by $P^*$, exchanging the roles of chains and antichains. Let $\ell$ and $q$ be the unique nonnegative integers satisfying $0 < q \le k$ and
\begin{equation}
\label{eq:n-q-k-ell}
n = q \cdot (k+\ell+1) + (k-q) \cdot (k+\ell).
\end{equation}
In other words, we let $\ell = \lceil n / k \rceil - k - 1$ and $q = n - k(k+\ell)$. Our upper bound on $n$ implies that $\ell \le c\sqrt{k}/\log_2 k$, which we note here for future reference.
Observe that $m_k(\tau_{k,n})$ is precisely the number of chains of length $k+1$ in the poset which is the disjoint union of $k$ pairwise incomparable chains: $q$ chains of length $k+\ell+1$ and $k-q$ chains of length $k + \ell$, cf.~\eqref{eq:mktau}. In particular,
\begin{equation}
\label{eq:mkn-mknm}
m_k(\tau_{k,n}) - m_k(\tau_{k,n-1}) = \binom{k+\ell}{k}.
\end{equation}
With the view of~\eqref{eq:mkn-mknm}, we may and shall assume that $P$ contains no element $x$ that belongs to more than $\binom{k+\ell}{k}$ homogenous $(k+1)$-sets. Indeed, otherwise we could apply the inductive assumption to the poset $P \setminus \{x\}$ and conclude that $h_k(P) > m_k(\tau_{k,n})$.
\subsection{Posets of height larger than $k+\ell$}
Our inductive assumption allows us to easily deal with the case $h(P) \ge k + \ell + 1$. Indeed, every element of each longest chain in $P$ lies in at least $\binom{h(P)-1}{k}$ chains of length $k+1$ and hence by~\eqref{eq:mkn-mknm}, for any such element $x$,
\begin{equation}
\label{eq:induction}
h_k(P) \ge h_k(P \setminus \{x\}) + \binom{h(P)-1}{k} \ge m_k(\tau_{k,n-1}) + \binom{k+\ell}{k} = m_k(\tau_{k,n}).
\end{equation}
Since we have promised to characterize all posets satisfying $h_k(P) = m_k(\tau_{k,n})$, we still need to analyze the case when all inequalities in~\eqref{eq:induction} are actually equalities. This means, in particular, that $h(P) = k + \ell + 1$ and that no longest chain intersects a $(k+1)$-element antichain. We claim that $P$ may be partitioned into $k$ chains.
Let $x$ be the top element of some longest chain $L$ in~$P$. By the inductive assumption, $P \setminus \{x\}$ can be partitioned into $k$ chains or $k$ antichains. Let us argue that $P \setminus \{x\}$ can actually be partitioned into $k$ chains. When $n-1 = k^2$, this follows from Dilworth's theorem (or its dual version applied to $(P \setminus \{x\})^*$) as $h_k(P \setminus \{x\}) = m_k(\tau_{k,n-1}) = 0$. Otherwise, when $n-1 > k^2$, if $P \setminus \{x\}$ could be partitioned into $k$ antichains, then each of them would have to intersect the chain $L \setminus \{x\}$, which has at least $k$ elements, and one of them would have at least $k+1$ elements, contradicting our assumption above.
Suppose that the $k$ chains decomposing $P \setminus \{x\}$ are $L_1, \ldots, L_k$. It suffices to show that $x \ge \max L_i$ for some $L_i$, since then $L_1, \ldots, L_{i-1}, L_i \cup \{x\}, L_{i+1}, \ldots, L_k$ form a partition of $P$ into $k$ chains. Let $y$ be the second largest element of $L$. Clearly, $y \in L_i$ for some $i$. If $y = \max L_i$, then there is nothing left to prove, so suppose that $z = \max L_i > y$ and consider the set $L \cup \{z\} \subseteq P$. It is easy to see that $y$ belongs to at least $\binom{k+\ell}{k} + \binom{k+\ell-1}{k-1}$ chains of length $k+1$, contradicting our assumption.
\subsection{Posets of height smaller than $k+\ell$}
The case $h(P) \le k+\ell-1$ can be easily resolved with the use of Corollary~\ref{cor:large-surplus}. Indeed, if $h(P) \le k+\ell-1$, then
\[
s_k(P) \ge n - k(k+\ell-1) = k+q > k.
\]
Since $h(P) \ge w(P)$, Corollary~\ref{cor:large-surplus} implies that $h_k(P) \ge 2^{\sqrt{k/3}-1}$. On the other hand, our assumption that $n \le k^2 + ck^{3/2}/\log_2 k$ and~\eqref{eq:mktau} yield
\begin{equation}
\label{eq:mk-exp-sqrt-k}
\begin{split}
m_k(\tau_{k,n}) & \le k \binom{\lceil n/k \rceil}{k+1} \le k \binom{k+c\sqrt{k}/\log_2 k+1}{k+1} \\
& \le k \binom{2k}{c\sqrt{k}/\log_2 k} \le \left(\frac{2e\sqrt{k}\log_2k}{c}\right)^{\frac{c \sqrt{k}}{\log_2k}} < 2^{\sqrt{k}/4}.
\end{split}
\end{equation}
The fourth inequality above is $\binom{a}{b} \le (ea/b)^b$ and the last inequality follows since $c < 1/2$ and $k$ is sufficiently large.
\subsection{Posets of height $k+\ell$}
In view of the above considerations, for the remainder of the proof, we may assume that
\begin{equation}
\label{eq:height-kpell}
h(P) = k + \ell \ge w(P).
\end{equation}
Since $n \le h(P)w(P)$, this means, in particular, that $n > k^2+k$, as otherwise $\ell = 0$ and we have assumed above that $n > k^2$. As $n > k(k+\ell)$, see~\eqref{eq:n-q-k-ell}, assumption~\eqref{eq:height-kpell} implies that $P$ cannot be decomposed into $k$ chains or $k$ antichains. Thus, we shall show that $h_k(P) > m_k(\tau_{k,n})$, unless $n = k^2+k+1$ and $P$ is one of the posets described in Example~\ref{example:extremal-posets}. Observe that
\[
m_k(\tau_{k,n}) = k \binom{k+\ell}{k+1} + q \binom{k+\ell}{k} = \left( q + \frac{k \ell}{k+1} \right) \binom{k+\ell}{k},
\]
cf.~\eqref{eq:mktau} and~\eqref{eq:n-q-k-ell}. In particular, since $\ell, q \le k$ and $k$ is sufficiently large,
\begin{equation}
\label{eq:mktau-k-to-ell}
m_k(\tau_{k,n}) \le (q + \ell) \binom{k+\ell}{\ell} \le (k + \ell) \binom{k+\ell}{\ell} < k^{2\ell}.
\end{equation}
\subsubsection{The key lemma}
Let $(A_i)_{i=1}^{k+\ell}$ be the canonical decomposition of $P$ into antichains and recall the definition of $G_i$ from Section~\ref{sec:decomposition}. We shall provide various lower bounds on the number of homogenous sets by analyzing the graphs $G_i$ for various indices $i$ such that $A_i \cup A_{i+1}$ contains an antichain with $k+1$ elements. First and foremost, we shall be looking at $i \in F$, where
\begin{equation}
\label{eq:F}
F = \{i \in [k+\ell] \colon |A_i| \ge k+1\}.
\end{equation}
We start by establishing a lower bound on the size of $F$.
\begin{obs}
\label{obs:F-size}
For every $I \subseteq [k+\ell]$,
\[
|F| \ge \frac{q + \sum_{i \in I} (k-|A_i|)}{\ell}.
\]
\end{obs}
\begin{proof}
Recall that the sets $(A_i)_{i = 1}^{k+\ell}$ form a partition of $P$ into antichains and that $|A_i| \le w(P) \le k+\ell$ for all $i$. Hence,
\[
\begin{split}
n - \sum_{i \in I} |A_i| & = \sum_{i \not\in I \cup F} |A_i| + \sum_{i \in F \setminus I} |A_i| \le (k + \ell - |I \cup F|) \cdot k + |F \setminus I| \cdot (k+\ell) \\
& = (k + \ell - |I|) \cdot k + |F \setminus I| \cdot \ell \le n - q - |I| \cdot k + |F| \cdot \ell.\qedhere
\end{split}
\]
\end{proof}
Recall the definitions of $u_i$, $\Sigma_i$, $A_i'$, and $B_i$ from Section~\ref{sec:decomposition}. The following lemma is key
\begin{lemma}
\label{lemma:F}
If $i \in F \cap [k+\ell-1]$, then $A_i \cup B_{i+1}$ contains at least $2^{\min\{k, |B_{i+1}|\}}$ antichains with $k+1$ elements and
\[
\Sigma_i \ge \Sigma_{i+1} + \sum_{y \in A_{i+1}' \setminus B_{i+1}} u_{i+1}(y) \ge \Sigma_{i+1} + |A_{i+1}'| - |B_{i+1}|.
\]
\end{lemma}
\begin{proof}
Note that for any $Y \subseteq B_{i+1}$, the set $Y \cup (A_i \setminus N_{G_i}(Y))$ is an antichain with at least $|A_i|$ elements. Since $|A_i| \ge k+1$, each of these antichains that additionally satisfies $|Y| \le k+1$ contains a $(k+1)$-element subset $L$ such that $L \cap B_{i+1} = Y$. This proves the first assertion of the lemma. The second assertion holds since
\[
\Sigma_i = \sum_{xy \in G_i} u_{i+1}(y) = \sum_{y \in A_{i+1}'} u_{i+1}(y) \deg_{G_i}(y) \ge \sum_{y \in A_{i+1}'} u_{i+1}(y) + \sum_{i \in A_{i+1}' \setminus B_{i+1}} u_{i+1}(y).\qedhere
\]
\end{proof}
\subsubsection{The first lower bound on $\min_i |A_i'|$}
In view of Lemma~\ref{lemma:F}, we shall aim at proving a lower bound on the minimum size of $A_{i+1}'$. We first derive a somewhat weak bound on $\min_i |A_i'|$ from Corollary~\ref{cor:large-surplus}.
\begin{claim}
Either $h_k(P) > m_k(\tau_{k,n})$ or $|A_i'| \ge k/3$ for all $i \in [k+\ell]$, possibly after substituting $P^*$ for $P$.
\end{claim}
\begin{proof}
Suppose that $|A_i'| < k/3$ for some $i$. If $w(P) < h(P)$, then we let $P' = P \setminus A_i'$ and note that $h(P') = h(P) - 1 = k+\ell-1$ as every chain of maximum length in $P$ contains one element of $A_i'$. Thus
\[
s_k(P') = |P'| - k \cdot h(P') \ge k(k+\ell) - |A_i'| - k(k+\ell-1) \ge 2k/3.
\]
Since $w(P') \le w(P) \le h(P) - 1 = h(P')$, we may apply Corollary~\ref{cor:large-surplus} with $t = 2k/9$ to $P'$ and conclude that, recalling~\eqref{eq:mk-exp-sqrt-k},
\[
h_k(P) \ge h_k(P') \ge 2^{\sqrt{2k}/3-1} > m_k(\tau_{k,n}).
\]
If $w(P) = h(P)$, then we let $(A_i^*)_{i=1}^{k+\ell}$ be the canonical decomposition of $P^*$ into antichains. Now, if $|(A_j^*)'| \ge k/3$ for all $j$, then we work with $P^*$ instead of $P$. Otherwise, if $|(A_j^*)'| < k/3$ for some $j$, then we let $P' = P \setminus (A_i' \cup (A_j^*)')$ and note that $h(P') < h(P)$ and $w(P') < w(P)$. Consequently, letting $P'' = P'$ or $P'' = (P')^*$ so that $w(P'') \le h(P'')$, we have
\[
s_k(P'') = |P''| - k \cdot h(P'') \ge k(k+\ell) - |A_i'| - |(A_j^*)'| - k(k+\ell-1) \ge k/3.
\]
Now, we may again apply Corollary~\ref{cor:large-surplus} with $t = k/9$ to $P''$ to conclude that, again recalling~\eqref{eq:mk-exp-sqrt-k},
\[
h_k(P) \ge h_k(P'') \ge 2^{\sqrt{k}/3 - 1} > m_k(\tau_{k,n}).\qedhere
\]
\end{proof}
\subsubsection{Posets with large $F$}
For the remainder of the proof, we may and shall assume that $|A_i'| \ge k/3$ for all $i \in [k+\ell]$. Together with Lemma~\ref{lemma:F}, this bound already allows us to deal with the case $|F| \ge 40 \log_2 k$. The~crucial observation here is the following.
\begin{obs}
\label{obs:chains-per-element}
If some element of $P$ belongs to more than $2k/\ell$ chains of length $k+\ell$, then $h_k(P) > m_k(\tau_{k,n})$.
\end{obs}
\begin{proof}
Let $M = 2k/\ell$ and suppose that some $y \in P$ is contained in $M$ different chains of length $k+\ell$. We shall show that this implies that $y$ belongs to more than $\binom{k+\ell}{k}$ chains of length $k+1$, which, by the inductive assumption, implies that $h_k(P) > m_k(\tau_{k,n})$, see~\eqref{eq:mkn-mknm}. This follows easily from Corollary~\ref{cor:signatures}~(\ref{item:signatures-local}) as, letting $m = \log_2 M + 1$, since $\ell, m \ll \sqrt{k}$, we have that
\[
\exp\left(-\frac{2 (\ell-1) m}{k}\right) M \binom{k+\ell}{k+1} \ge \frac{3k}{2\ell} \binom{k+\ell}{k+1} = \frac{3k}{2(k+1)} \binom{k+\ell}{k}.\qedhere
\]
\end{proof}
\begin{claim}
\label{claim:F-size}
If $|F| \ge 40\log_2 k$, then $h_k(P) > m_k(\tau_{k,n})$.
\end{claim}
\begin{proof}
We may assume that $|B_{i+1}| < 2\ell\log_2k$ for every $i \in F$ as otherwise Lemma~\ref{lemma:F} implies that $h_k(P) \ge k^{2\ell} > m_k(\tau_{k,n})$. We have also assumed that $|A_i'| \ge k/3$ for every $k$ and hence, again by Lemma~\ref{lemma:F},
\[
\Sigma_{i+1} - \Sigma_i \ge k/4 \quad \text{ for every $i \in F \cap [k+\ell-1]$}.
\]
Partition $F \cap [k+\ell-1]$ into $I_1$ and $I_2$ with $|I_1| \ge 32 \log_2 k$ and $|I_2| \ge 4 \log_2 k$ such that $\min I_1 > \max I_2$. By~\eqref{eq:sum-ui-monotone},
\[
\Sigma_{\min I_1} = \Sigma_{k+\ell} + \sum_{i = \min I_1}^{k+\ell-1} (\Sigma_i - \Sigma_{i+1}) \ge |I_1| \cdot k/4 \ge 8 k \log_2 k.
\]
Now, consider an arbitrary $i \in I_2$. In accordance with Observation~\ref{obs:chains-per-element}, we may assume that $u_{i+1}(y) \le 2k/\ell$ for every $y \in A_{i+1}$. As $i+1 \le \min I_1$ and $|B_{i+1}| \le 2\ell\log_2k$,
\[
\sum_{y \in B_{i+1}} u_{i+1}(y) \le |B_{i+1}| \cdot 2k/\ell \le 4k \log_2 k \le \Sigma_{\min I_1} / 2 \le \Sigma_{i+1}/2,
\]
where the final inequality follows from~\eqref{eq:sum-ui-monotone}. Hence, by Lemma~\ref{lemma:F},
\[
\Sigma_i \ge \Sigma_{i+1} + \sum_{y \in A_{i+1}' \setminus B_{i+1}} u_{i+1}(y) \ge 3\Sigma_{i+1}/2.
\]
It follows that
\[
\Sigma_1 \ge (3/2)^{|I_2|} \cdot \Sigma_{\min I_1} \ge k^3,
\]
implying that there is an $x \in A_1$ which belongs to more than $2 k / \ell$ (actually more than $k^2/2$) chains of length $k+\ell$. Consequently Observation~\ref{obs:chains-per-element} implies that $h_k(P) > m_k(\tau_{k,n})$.
\end{proof}
\subsubsection{A sufficient condition on $\Sigma_1$}
For the remainder of the proof, we shall therefore assume that $|F| \le 40\log_2k$ and, as a consequence of Observation~\ref{obs:F-size}, that $q \le 40\ell\log_2k$. In view of Corollary~\ref{cor:signatures}~(\ref{item:signatures-global}), in order to conclude that $h_k(P) > m_k(\tau_{k,n})$, it is enough to provide a sufficiently strong lower bound on $\Sigma_1$, the number of chains of length $k+\ell$ in $P$. To this end, define
\begin{equation}
\label{eq:S}
S = \left(1 + \frac{q}{\ell}\right)k + 50\sqrt{k}\log_2k
\end{equation}
and note that our assumption on $q$ implies that $S \le 42k\log_2k$.
\begin{obs}
\label{obs:Sigma-1}
If $\Sigma_1 \ge S$, then $h_k(P) > m_k(\tau_{k,n})$.
\end{obs}
\begin{proof}
Since we have assumed that $q \le 40\ell\log_2k$, then
\[
S \ge \left(1+\frac{1}{\sqrt{k}}\right) \left(1+\frac{q}{\ell}\right)(k+1).
\]
As $\binom{k+\ell}{k+1} = \frac{\ell}{k+1}\binom{k+\ell}{k}$, it now follows from Corollary~\ref{cor:signatures}~(\ref{item:signatures-global}) that
\[
\begin{split}
h_k(P) & \ge \exp\left(-\frac{3\ell\log_2k}{k}\right)\left(1+\frac{1}{\sqrt{k}}\right)\left(1+\frac{q}{\ell}\right)(k+1) \binom{k+\ell}{k+1} \\
& \ge \left(1 - \frac{1}{2\sqrt{k}}\right)\left(1 + \frac{1}{\sqrt{k}}\right) (\ell + q) \binom{k+\ell}{k} > (\ell + q) \binom{k+\ell}{k} \ge m_k(\tau_{k,n}),
\end{split}
\]
where the second inequality holds since $\ell \le \sqrt{k} / (300\log_2 k)$ and the last inequality is~\eqref{eq:mktau-k-to-ell}.
\end{proof}
\subsubsection{The second lower bound on $\min_i |A_i'|$}
We shall now focus on proving the following strong lower bound on $\min_i |A_i'|$.
\begin{claim}
\label{claim:Ap-2nd}
Either $h_k(P) > m_k(\tau_{k,n})$ or $|A_i'| \ge k - 240 \ell \log_2 k$ for all $i \in [k+\ell]$.
\end{claim}
Recall from Section~\ref{sec:decomposition} that for each $i \in [k+\ell-1]$, the set $(A_i \setminus A_i') \cup A_{i+1}'$ is an antichain and consequently,
\begin{equation}
\label{eq:Ai-Aip-Aipp}
|A_i| - |A_i'| + |A_{i+1}'| \le w(P) \le k+\ell.
\end{equation}
With foresight, define
\begin{equation}
\label{eq:Fp}
F' = \{i \in [k+\ell-1] \colon |A_i| - |A_i'| + |A_{i+1}'| \ge k+1\}.
\end{equation}
Let $J$ be an arbitrary subset of $[k+\ell]$. By~\eqref{eq:Ai-Aip-Aipp} and~\eqref{eq:Fp},
\begin{equation}
\label{eq:Ap-telescope}
\sum_{i \in J} (|A_{i+1}'| - |A_i'|) \le \sum_{i \in J \cap F'} (k+\ell-|A_i|) + \sum_{j \in J \setminus F'} (k - |A_i|) \le \sum_{i \in J} (k-|A_i|) + |F'| \cdot \ell.
\end{equation}
Now, fix a $j \in [k+\ell]$ and let $J = \{j, \ldots, k+\ell-1\}$. It follows from \eqref{eq:Ap-telescope} and Observation~\ref{obs:F-size} with $I = J \cup \{k+\ell\}$ that (recalling that $A_{k+\ell}' = A_{k+\ell}$)
\[
k - |A_j'| = k - |A_{k+\ell}'| + \sum_{i \in J} (|A_{i+1}'| - |A_i'|) \le \sum_{i \in J \cup \{k+\ell\}} (k - |A_i|) + |F'| \cdot \ell \le (|F| + |F'|) \cdot \ell.
\]
Therefore, in order to establish Claim~\ref{claim:Ap-2nd}, it suffices to prove that $|F'| > 200\log_2 k$ implies that $h_k(P) > m_k(\tau_{k,n})$. This fact is a fairly straightforward consequence of the following lemma, which one may consider as the `dual' version of Lemma~\ref{lemma:F}.
\begin{lemma}
\label{lemma:Fp}
Either $h_k(P) > m_k(\tau_{k,n})$ or $\Sigma_i \ge \Sigma_{i+1} + k/4$ for all $i \in F'$.
\end{lemma}
\begin{proof}
Fix some $i \in F'$ and let
\[
C_i = \{x \in A_i' \colon \deg_{G_i}(x, A_{i+1}') \le 1\}.
\]
Observe that for every $X \subseteq C_i$, the set $(A_i \setminus A_i') \cup X \cup (A_{i+1}' \setminus N_{G_i}(X))$ is an antichain with at least $|A_i| - |A_i'| + |A_{i+1}'|$ elements. Each of these antichains that additionally satisfies $|X| \le k+1$ contains a $(k+1)$-element subset $L$ such that $L \cap A_i' = X$. Therefore, if $|C_i| \ge 2\ell\log_2k$, then $h_k(P) \ge k^{2\ell} > m_k(\tau_{k,n})$. On the other hand, if $|C_i| < 2\ell\log_2k$, then
\[
\begin{split}
\Sigma_i & = \sum_{xy \in G_i} u_{i+1}(y) = \sum_{y \in A_{i+1}'} u_{i+1}(y)\deg_{G_i}(y) \ge \Sigma_{i+1} + e_{G_i}(A_i', A_{i+1}') - |A_{i+1}'| \\
& \ge \Sigma_{i+1} + 2|A_i'| - |C_i| - |A_{i+1}'| \ge \Sigma_{i+1} + |A_i'| - (|A_{i+1}'| - |A_i'|) - 2\ell\log_2k.
\end{split}
\]
By Observation~\ref{obs:F-size} with $I = \{i\}$, we have $k - |A_i| \le |F| \ell$. Therefore, by~\eqref{eq:Ai-Aip-Aipp} and our assumption on $|F|$,
\[
|A_{i+1}'| - |A_i'| \le k+\ell-|A_i| \le (|F| + 1) \cdot \ell \ll k.
\]
Consequently, as we have assumed that $|A_i'| \ge k/3$, we have $\Sigma_i \ge \Sigma_{i+1} + k/4$.
\end{proof}
\begin{proof}[Proof of Claim~\ref{claim:Ap-2nd}]
Observe first that Lemma~\ref{lemma:Fp} yields $\Sigma_1 \ge |F'| \cdot k/4$. Indeed, similarly as in the proof of Claim~\ref{claim:F-size},
\[
\Sigma_1 = \Sigma_{k+\ell} + \sum_{i=1}^{k+\ell-1} (\Sigma_i - \Sigma_{i+1}) \ge |F'| \cdot k/4.
\]
As $q \le 40\ell\log_2k$ by our assumption, if $|F'| > 200\log_2k$, then $\Sigma_1 \ge 50k\log_2k \ge S$, see~\eqref{eq:S}, and consequently $h_k(P) > m_k(\tau_{k,n})$ by Observation~\ref{obs:Sigma-1}.
\end{proof}
\subsubsection{Strengthening Lemmas~\ref{lemma:F} and~\ref{lemma:Fp}}
For the remainder of the proof, we shall assume that $|A_i'| \ge k-240\ell\log_2k$ for every $i$. This assumption will allow us to prove the following strengthening of Lemmas~\ref{lemma:F} and~\ref{lemma:Fp}. Recall the definitions of $F$ from~\eqref{eq:F} and $F'$ from~\eqref{eq:Fp}.
\begin{lemma}
\label{lemma:Fp-strong}
Either $h_k(P) > m_k(\tau_{k,n})$ or
\begin{enumerate}[(i)]
\item
\label{item:Fp-strong-one}
$\Sigma_i \ge \Sigma_{i+1} + k - \sqrt{k}$ for all $i \in F \cup F'$ and
\item
\label{item:Fp-strong-two}
$\Sigma_i \ge \Sigma_{i+1} + 2k - 3\sqrt{k}$ for all $i \in F$ such that $i < \max F'$.
\end{enumerate}
\end{lemma}
\begin{proof}
Suppose first that $i \in F$. We may assume that $|B_{i+1}| \le 2\ell\log_2k$, as otherwise $h_k(P) \ge k^{2\ell} > m_k(\tau_{k,n})$ by Lemma~\ref{lemma:F}. Consequently,
\[
\Sigma_i - \Sigma_{i+1} \ge |A_{i+1}'| - |B_{i+1}| \ge k - 242\ell\log_2k \ge k - \sqrt{k},
\]
as $\ell \le \sqrt{k}/(300 \log_2 k)$ and we have assumed that $|A_{i+1}'| \ge k - 240\ell\log_2k$.
Suppose now that $i \in F'$. As in the proof of Lemma~\ref{lemma:Fp}, let
\[
C_i = \{x \in A_i' \colon \deg_{G_i}(x, A_{i+1}') \le 1\}
\]
and recall that either $|C_i| < 2\ell\log_2k$ or $h_k(P) \ge k^{2\ell} > m_k(\tau_{k,n})$ and that
\begin{equation}
\label{eq:Fp-strong}
\Sigma_i \ge \Sigma_{i+1} + |A_i'| - (|F|+1) \cdot \ell - |C_i| \ge \Sigma_{i+1} +k - \sqrt{k},
\end{equation}
as $\ell \le \sqrt{k}/(300 \log_2 k)$ and we have assumed that $|F| \le 40\log_2k$ and that $|A_i'| \ge k - 240\ell\log_2k$.
We now turn to proving~(\ref{item:Fp-strong-two}). First, for each $i \in [k+\ell]$, define
\[
A_i'' = \{x \in A_i \colon u_i(x) \ge 2\}
\]
and observe that $A_i'' \supseteq A_i' \setminus C_i$. Indeed, if $x \in A_i' \setminus C_i$, then by~\eqref{eq:ui-uip},
\[
u_i(x) = \sum_{xy \in G_i} u_{i+1}(y) \ge \deg_{G_i}(x,A_{i+1}') \ge 2.
\]
Consequently, if $i \in F'$ and $h_k(P) \le m_k(\tau_{k,n})$, then $|A_i''| \ge |A_i'| - |C_i| \ge k - \sqrt{k}$, see~\eqref{eq:Fp-strong}.
Assume now that $F' \neq \emptyset$ and let $f' = \max F'$. We claim that either $h_k(P) > m_k(\tau_{k,n})$ or $|A_{i+1}''| \ge k - 2\sqrt{k}$ for each $i < f'$. To this end, observe first that for each $i \in [k+\ell-1]$, the set $(A_i \setminus A_i'') \cup A_{i+1}''$ is an antichain and therefore $|A_i| - |A_i''| + |A_{i+1}''| \le k + \ell$. Indeed, if $x \in A_i$ and $y \in A_{i+1}$ satisfy $u_i(x) < u_{i+1}(y)$, then~\eqref{eq:ui-uip} implies that $xy \not\in G_i$. With foresight, let
\[
F'' = \{i \in [f' - 1] \colon |A_{i+1}''| - |A_i''| + |A_i| \ge k+1\}.
\]
Assume that $\min_{i \le f'} |A_i''| < k - 2\sqrt{k}$ and let $i \in [f']$ be the largest index such that $|A_i''| < k - 2\sqrt{k}$. Since, as we have shown above, $|A_{f'}''| \ge k - \sqrt{k}$, then
\begin{equation}
\label{eq:Fpp}
\begin{split}
\sqrt{k} & < |A_{f'}''| - |A_i''| = \sum_{j = i}^{f'-1} (|A_{i+1}''| - |A_i''|) \le \sum_{j = i}^{f'-1} (k-|A_i|) + |F'' \cap \{i, \ldots, f'-1\}| \cdot \ell \\
& \le |F| \cdot \ell + |F'' \cap \{i+1, \ldots, f'-1\}| \cdot \ell + \ell,
\end{split}
\end{equation}
where the last inequality follows from Observation~\ref{obs:F-size}. Since $\ell \le \sqrt{k}/(300 \log_2 k)$ and we have assumed that $|F| \le 40\log_2k$, it follows from~\eqref{eq:Fpp} that $|F'' \cap \{i+1, \ldots, f'-1\}| \ge 50 \log_2k$. We claim that this implies that $h_k(P) > m_k(\tau_{k,n})$. To this end, consider some $j \in F''$, let
\[
C_j' = \{x \in A_j'' \colon \deg(x, A_{j+1}'') \le 1\},
\]
and note that for every $X \subseteq C_j'$, the set $(A_j \setminus A_j'') \cup X \cup (A_{j+1}'' \setminus N_{G_j}(X))$ is an antichain with at least $k+1$ elements. Therefore, if $|C_j''| \ge 2\ell\log_2k$, then $h_k(P) \ge k^{2\ell} > m_k(\tau_{k,n})$. On the other hand, if $|C_j| < 2\ell\log_2k$, then
\[
\Sigma_j - \Sigma_{j+1} \ge e_{G_j}(A_j'' + A_{j+1}'') - |A_{j+1}''| \ge 2|A_j''| - (k+\ell) - 2\ell\log_2k.
\]
Therefore, if $j > i$ and $j \in F''$, then $\Sigma_j - \Sigma_{j+1} \ge k - o(k)$ and consequently, recalling~\eqref{eq:S},
\[
\Sigma_1 \ge |F'' \cap \{i+1, \ldots, f'-1\}| \cdot (k-o(k)) \ge (1-o(1)) \cdot 50k\log_2k \ge S,
\]
which, by Observation~\ref{obs:Sigma-1}, yields $h_k(P) > m_k(\tau_{k,n})$.
Finally, suppose that $i \in F$ and $i < \max F'$. We may now assume that $|A_{i+1}''| \ge k - 2\sqrt{k}$ and therefore, by Lemma~\ref{lemma:F},
\[
\Sigma_i - \Sigma_{i+1} \ge \sum_{y \in A_{i+1}' \setminus B_{i+1}} u_{i+1}(y) \ge |A_{i+1}''| + |A_{i+1}'| - |B_{i+1}| \ge 2k - 3\sqrt{k}.\qedhere
\]
\end{proof}
\subsubsection{Narrowing down to almost extremal posets}
The following observation will further narrow down our search for $P$ with $h_k(P) \le m_k(\tau_{k,n})$.
\begin{obs}
\label{obs:FFp}
If $|(F \cup F') \cap [k+\ell-1]| \ge \frac{q+1}{\ell}$, then $h_k(P) > m_k(\tau_{k,n})$.
\end{obs}
\begin{proof}
Recall that $\Sigma_{k+\ell} = |A_{k+\ell}| = |A_{k+\ell}'| \ge k - 240\ell\log_2k \ge k - \sqrt{k}$. By Lemma~\ref{lemma:Fp-strong}~(\ref{item:Fp-strong-one}) and~\eqref{eq:sum-ui-monotone},
\[
\Sigma_1 \ge \Sigma_{k+\ell} + \sum_{i \in (F \cup F') \cap [k+\ell-1]} (\Sigma_i - \Sigma_{i+1}) \ge \left(1 + \frac{q+1}{\ell}\right)\left(k-\sqrt{k}\right).
\]
Now, since $q \le 40\ell\log_2k$ and $\ell \le \sqrt{k}/(300 \log_2 k)$, then $\Sigma_1 \ge S$ and the conclusion follows from Observation~\ref{obs:Sigma-1}.
\end{proof}
Recall the definition of $F$ from~\eqref{eq:F}. We shall now split into cases depending on whether or not $k + \ell \in F$, that is, whether or not $|A_{k+\ell}| \ge k+1$.
\begin{case}
\label{case:k-ell-not-F}
$k + \ell \notin F$.
\end{case}
The assumption that $k+\ell \not\in F$ and Observation~\ref{obs:F-size} imply that
\[
|F \cap [k+\ell-1]| = |F| \ge \left\lceil \frac{q + |\{i \in [k+\ell] \colon |A_i| < k\}|}{\ell} \right\rceil.
\]
By Observation~\ref{obs:FFp}, we may assume that $|A_i| \ge k$ for all $i$, $\ell$ divides $q$, and $|F| = q/\ell$, as otherwise $h_k(P) > m_k(\tau_{k,n})$. This implies that $|A_i| = k+\ell$ for each $i \in F$ and $|A_i| = k$ otherwise.
\begin{subcase}
$F' \neq \emptyset$ and $\max F' > \min F$.
\end{subcase}
Let $j = \max F'$ and let $i$ be the largest element of $F$ that is smaller than $j$. By Lemma~\ref{lemma:Fp-strong}~(\ref{item:Fp-strong-two}),
\[
\Sigma_i - \Sigma_{i+1} \ge 2k - 3\sqrt{k}.
\]
Consequently, using Lemma~\ref{lemma:Fp-strong} as in the proof of Observation~\ref{obs:FFp},
\[
\Sigma_1 \ge \Sigma_{k+\ell} + \sum_{j \in F \setminus \{i\}} (\Sigma_j - \Sigma_{j+1}) + \Sigma_i - \Sigma_{i+1} \ge \left(1 + \frac{q}{\ell}\right)\left(k - \sqrt{k}\right) + k - 2\sqrt{k} \ge S,
\]
which yields $h_k(P) > m_k(\tau_{k,n})$.
\begin{subcase}
$F' = \emptyset$ or $\max F' \le \min F$.
\end{subcase}
If $F' \neq \emptyset$, then we may also assume that $\min F' \ge \min F$. Indeed, otherwise
\[
|(F \cup F') \cap [k+\ell-1]| \ge |F| + 1 \ge \frac{q}{\ell}+1
\]
and hence $h_k(P) > m_k(\tau_{k,n})$ by Observation~\ref{obs:FFp}. Hence $\min F' = \max F' = \min F$.
We first claim that $|A_i'| \ge k$ for every $i$. Suppose not and let $i$ be the largest index for which $|A_i'| < k$ and note that $i < k+\ell$ as $A_{k+\ell}' = A_{k+\ell}$ and we have assumed that $|A_j| \ge k$ for each $j \in [k+\ell]$. Consequently,
\[
|A_i| + |A_{i+1}'| - |A_i'| \ge |A_i| + k - |A_i'| > |A_i| \ge k,
\]
implying that $i \in F'$ and thus $i = \min F$. But this is impossible as $|A_i| = k+\ell$ and hence $(A_i \setminus A_i') \cup A_{i+1}'$ would be an antichain with more than $k+\ell$ elements.
We now claim that $\max F \le \min F + 1$, that is, that $|F|=1$ or $F = \{f, f+1\}$ for some $f \in [k+\ell-1]$. To see this, note that if $\max F \in F'$, then our assumption implies that $\max F = \min F$, that is, $|F| = 1$. Otherwise, if $\max F \not\in F'$, then
\[
|A_{\max F}| - |A_{\max F}'| + |A_{\max F+1}'| \le k
\]
and hence, as $|A_{\max F+1}'| \ge k$, we have $|A_{\max F}'| = |A_{\max F}| = k+\ell$. Consequently, either $\max F = 1$ or $\max F - 1 \in F'$ and therefore $\max F - 1 = \min F$ by our assumption.
Finally, let $f = \min F$. If $f > 1$ and $|A_f'| > k$, then $f - 1 \in F'$, contradicting our assumption. We may thus assume that $f = 1$ or $|A_f'| = k$. Consequently, if $|F| > 1$, then $F = \{1, 2\}$. Indeed, if $|F| > 1$, then $F = \{f, f+1\}$ and $|A_{f+1}'| = k+\ell$. Moreover, since $A_{f+1}' \cup (A_f \setminus A_f')$ is an antichain and $w(P) \le k+\ell$, then $|A_f'| = |A_f| = k+\ell$. Finally, we have shown above that if $f > 1$, then $|A_f'| = k$.
\begin{case}
$k + \ell \in F$.
\end{case}
Consider the poset $\hat{P}$ obtained from $P$ by reversing the $\le$ relation, that is, by letting $x \le y$ in $\hat{P}$ if and only if $y \le x$ in $P$. Clearly, the same sets form chains and antichains in both $P$ and $\hat{P}$. Let $(\hat{A}_i)_{i=1}^{k+\ell}$ be the canonical decomposition of $\hat{P}$ into antichains and observe that $\hat{A}_{k+\ell} = A_1'$. Indeed,
\[
\begin{split}
\hat{A}_{k+\ell} & = \{x \in \hat{P} \colon \text{there is a chain $L \subseteq \hat{P}$ of length $k+\ell$ with $x = \max L$}\} \\
& = \{x \in P \colon \text{there is a chain $L \subseteq P$ of length $k+\ell$ with $x = \min L$}\} = A_1'.
\end{split}
\]
Thus, we may assume that $|A_1'| > k$ as otherwise $\hat{P}$ falls into Case~\ref{case:k-ell-not-F}. Thus $1 \in F$. We show that this implies that $\Sigma_1 \ge S$ and consequently, by Observation~\ref{obs:Sigma-1}, that $h_k(P) > m_k(\tau_{k,n})$.
Since $|A_{k+\ell}'| = |A_{k+\ell}| > k$, then $k+\ell-1 \in F'$ and hence, by Lemma~\ref{lemma:Fp-strong}, we may assume that $\Sigma_{k+\ell-1} \ge \Sigma_{k+\ell} + k - \sqrt{k} \ge 2k-\sqrt{k}$ and $\Sigma_i \ge \Sigma_{i+1} + 2k-3\sqrt{k}$ for all $i \in F \cap [k+\ell-2]$. This yields
\begin{equation}
\label{eq:Sigma-1-case-2}
\Sigma_1 \ge \big( 2 + 2|F \cap [k+\ell-2]| + |(F' \setminus F) \cap [k+\ell-2]|\big) (k - 2\sqrt{k}).
\end{equation}
By our assumption that $1 \in F$ and Observation~\ref{obs:F-size},
\begin{equation}
\label{eq:F-case-2}
|F \cap [k+\ell-2]| \ge \max\{1, \lceil q / \ell \rceil - 2\}.
\end{equation}
It follows that either $q = 3\ell$ and we have equality in~\eqref{eq:F-case-2} or
\[
\Sigma_1 \ge \left( 1 + \frac{q+1}{\ell} \right)(k - 2\sqrt{k}) = \left(1 + \frac{q}{\ell}\right)k + \frac{k}{\ell} - 2\sqrt{k}\left(1 + \frac{q+1}{\ell}\right) \ge S;
\]
to see the last inequality, recall that $\ell \le \sqrt{k}/(300 \log_2k)$. The former (i.e., $q = 3\ell$ and equality in~\eqref{eq:F-case-2}) implies that $F = \{1, k+\ell-1, k+\ell\}$, $A_i = k+\ell$ for all $i \in F$, and $|A_i| = k$ for all $i \not\in F$. Since $A_{k+\ell}' \cup (A_{k+\ell-1} \setminus A_{k+\ell-1}')$ is an antichain and $w(P) \le k+\ell = |A_{k+\ell}'|$, we must have $|A_{k+\ell-1}'| = |A_{k+\ell-1}| = k+\ell$ and consequently, $k+\ell-2 \in F'$. Now, \eqref{eq:Sigma-1-case-2} again yields $\Sigma_1 \ge 5(k-2\sqrt{k}) \ge S$.
\subsection{Almost extremal posets}
Summarizing the above discussion, if $h_k(P) \le m_k(\tau_{k,n})$, then $h(P) = k+\ell$ and either $P$ or $\hat{P}$ (the poset obtained from $P$ by reversing the $\le$ relation) satisfy one of the following two lists of conditions:
\begin{enumerate}[(1)]
\item
\label{item:extremal-1}
$F = \{1, 2\}$, $|A_1'| = |A_2'| = k+\ell$, and $|A_i'| = k$ for every $i \ge 3$,
\item
\label{item:extremal-2}
$F = \{f\}$ for some $f \in [k+\ell-1]$ and $|A_i'| \ge k$ for all $i$; if $f > 1$, then $|A_f'| = k$.
\end{enumerate}
From now on, we shall have to count homogenous sets somewhat more carefully, as there are posets of either of these two types that contain fewer than $m_k(\tau_{k,n})$ chains of length $k+1$ and fewer than $m_k(\tau_{k,n})$ antichains with $k+1$ elements. (So far, we have always managed to show that our poset contains more than $m_k(\tau_{k,n})$ homogenous sets of one of the two types.)
\subsubsection{Bounding the number of antichains}
We first derive a lower bound for the number of $(k+1)$-element antichains which we shall use in both~(\ref{item:extremal-1}) and~(\ref{item:extremal-2}). To this end, for each $i \in [k+\ell-1]$, let
\[
D_{i+1} = \{x \in A_{i+1}' \colon \deg_{G_i}(x) = 2\} \subseteq A_{i+1}' \setminus B_{i+1}.
\]
Note for future reference that it follows from (the proof of) Lemma~\ref{lemma:F} that for each $i \in F$,
\begin{equation}
\label{eq:lemma-F}
\Sigma_i - \Sigma_{i+1} = \sum_{y \in A_{i+1}'} u_{i+1}(y)(\deg_{G_i}(y)-1) \ge 2|A_{i+1}'| - |D_{i+1}| - 2|B_{i+1}|,
\end{equation}
which improves the lower bound of $|A_{i+1}'| - |B_{i+1}|$ for $\Sigma_i - \Sigma_{i+1}$ stated in Lemma~\ref{lemma:F} whenever $D_{i+1}$ is smaller than $A_{i+1}' \setminus B_{i+1}$. On the other hand, if $D_{i+1}$ is large, then there are many $(k+1)$-element antichains in $A_i \cup D_{i+1}$, as the following lemma shows.
\begin{lemma}
\label{lemma:C-antichains}
Let $i \in [k+\ell-1]$ and suppose that $|A_i| = k+\ell$ and $|A_{i+1}'| = k$. If $|D_{i+1}| \ge k-k^{2/3}$, then $A_i \cup D_{i+1}$ contains at least $20(\ell-1)\binom{k+\ell}{k+1}$ antichains with $k+1$ elements that intersect $D_{i+1}$.
\end{lemma}
\begin{proof}
For each $z \in A_i$, let $H_z$ be an arbitrary tree with vertex set $N_{G_i}(z) \cap D_{i+1}$ and let $H$ be the multigraph with vertex set $D_{i+1}$ which is the union of all $H_z$ as $z$ ranges over $A_i$. Clearly,
\[
e(H) \ge \sum_{z \in A_i} (\deg_{G_i}(z, D_{i+1}) - 1) \ge 2|D_{i+1}| - |A_i| \ge k-3k^{2/3},
\]
as $|A_i| = k+\ell$ and $\ell \ll k^{2/3}$. Note crucially that:
\begin{enumerate}[(i)]
\item
\label{item:HX}
For every $X \subseteq D_{i+1}$, the set $X \cup (A_i \setminus N_{G_i}(X))$ is an antichain.
\item
\label{item:eHX}
For every $X \subseteq D_{i+1}$, we have $|N_{G_i}(X)| \le 2|X| - e_H(X)$.
\end{enumerate}
To see~(\ref{item:eHX}), recall that each $H_z$ is a tree and therefore
\[
|N_{G_i}(X)| = 2|X| - \sum_{z \in A_i} \max\{\deg_{G_i}(z,X)-1, 0\} \le 2|X| - \sum_{z \in A_i} e_{H_z}(X) = 2|X| - e_H(X).
\]
We first show that we may assume that $H$ contains fewer than $\sqrt{k}$ cycles (we consider two parallel edges to be a cycle). Since $w(P) \le k+\ell \le |A_i|$, (\ref{item:HX}) implies that $|N_{G_i}(X)| \ge |X|$ for each $X \subseteq D_{i+1}$ and hence each component $T$ of $H$ has at most one cycle as otherwise $|N_{G_i}(T)| < |T|$ by~(\ref{item:eHX}). Suppose now that $X_1, \ldots, X_m$ are cycles in $H$. They belong to different components of $H$, so in particular they are vertex-disjoint. By~(\ref{item:eHX}), for any $J \subseteq [m]$, the set $X = \bigcup_{j \in J} X_j$ satisfies $|N_{G_i}(X)| = |X|$. By~(\ref{item:HX}), $A_i \cup D_{i+1}$ contains at least $2^m$ antichains with $k+\ell$ elements and thus, by Lemma~\ref{lemma:KK}, at least $2^{m-1}$ antichains with $k+1$ elements. Finally, as $\ell < \sqrt{k}/(300 \log_2 k)$, then $2^{\sqrt{k}} \gg \ell \binom{k+\ell}{k+1}$, cf.~\eqref{eq:mk-exp-sqrt-k}.
Assume that $H$ has fewer than $\sqrt{k}$ cycles and delete from $H$ one edge in each of its cycles to obtain a forest $H'$ with $e(H') \ge e(H) - \sqrt{k} \ge k - 4k^{2/3}$. Let $N'$ be the number of sets $X \subseteq D_{i+1}$ with $|X| \le 41$ which are connected in $H'$. As $H' \subseteq H$, it follows from~(\ref{item:eHX}) that $|N_{G_i}(X)| \le |X| + 1$ for each such $X$. Thus by~(\ref{item:HX}), the number $N$ of antichains in $A_i \cup D_{i+1}$ satisfies
\[
\begin{split}
N & \ge \sum_{X \subseteq D_{i+1}} \binom{k+\ell-|N_{G_i}(X)|}{k+1-|X|} \ge N' \binom{k+\ell-42}{k-40} \\
& \ge N' \cdot \left(\frac{k-40}{k+\ell}\right)^{41} \cdot \binom{k+\ell-1}{k+1} \ge \frac{N'}{2} \cdot\frac{\ell-1}{k} \binom{k+\ell}{k+1},
\end{split}
\]
where we have used the assumption that $\ell = o(k)$. Finally, let $t_1, \ldots, t_m$ be the orders of the trees constituting $H'$. Since $m = k - e(H') \le 4k^{2/3}$, it follows from Lemma~\ref{lemma:connected-sets} that
\[
N' \ge \sum_{c=1}^{41} \sum_{j=1}^m (t_j-c) \ge \sum_{c=1}^{41} (k - cm) \ge 41k - O(k^{2/3}) \ge 40k.\qedhere
\]
\end{proof}
\subsubsection{Almost extremal posets of type~(\ref{item:extremal-1})}
Recall that $P$ is of type~(\ref{item:extremal-1}) if and only if $|A_i| = |A_i'| = k+\ell$ for $i \in \{1, 2\}$ and $|A_i| = |A_i'| = k$ otherwise. In particular, $q = 2\ell$ and hence $S = 3k + 50\sqrt{k}\log_2k$. We first show that $h_k(P) > m_k(\tau_{k,n})$ for each such $P$. Let $b_2 = |B_2|$ and $b_3 = |B_3|$. By Lemma~\ref{lemma:F} and \eqref{eq:lemma-F},
\[
\Sigma_1 \ge \Sigma_3 + 2|A_3'| - |D_3| - 2b_3 + |A_2'| - b_2 \ge 4k + \ell - |D_3| - b_2 - 2b_3.
\]
We may assume that $b_2, b_3 \le 2\ell\log_2k$ as otherwise Lemma~\ref{lemma:F} implies that $h_k(P) > m_k(\tau_{k,n})$. If $|D_3| < k - k^{2/3}$, then $\Sigma_1 \ge S$ and, by Observation~\ref{obs:Sigma-1}, $h_k(P) > m_k(\tau_{k,n})$. Thus, we may assume that $|D_3| \ge k - k^{2/3}$.
Let us carefully count homogenous $(k+1)$-sets in $P$. First, consider the collection of all chains of length $k+1$ obtained by taking a triple $(x_1,x_2,x_3) \in A_1' \times A_2' \times A_3'$ with $x_1 \le x_2 \le x_3$ and an arbitrary set of $k-2$ elements from some chain of length $k+\ell$ that contains $\{x_1,x_2,x_3\}$. The number of such chains containing a fixed triple is
\[
\binom{k+\ell-3}{k-2} = \frac{(k+1)k(k-1)}{(k+\ell)(k+\ell-1)(k+\ell-2)} \binom{k+\ell}{k+1} \ge \left(1 - \frac{\ell-1}{k-1}\right)^3 \binom{k+\ell}{k+1}
\]
and all chains constructed in this way are distinct. Let $N_t$ be the number of such triples. Since neither $G_1$ nor $G_2$ contain any isolated vertices, $A_2 = A_2'$, $A_3 = A_3'$, $e(G_1) \ge 2|A_2'| - b_2$, and $e(G_2) \ge 2|A_3'| - b_3$, then
\[
N_t = \sum_{x \in A_2'} \deg_{G_1}(x) \deg_{G_2}(x) \ge \sum_{x \in A_2'} (\deg_{G_1}(x) + \deg_{G_2}(x) - 1) \ge 2|A_3'| + |A_2'| - b_2 -b_3.
\]
Therefore, the number $N_c$ of chains of length $k+1$ satisfies
\[
N_c \ge (2|A_3'| + |A_2'| - b_2 - b_3) \binom{k+\ell-3}{k+1} \ge \left(1 - \frac{\ell-1}{k-1}\right)^3 (3k+\ell-b_2-b_3) \binom{k+\ell}{k+1}.
\]
To estimate the number of antichains, note that for any $i \in \{1, 2\}$ and any $Y \subseteq B_{i+1}$, the set $Y \cup (A_i \setminus N_{G_i}(Y))$ is an antichain with $k+\ell$ elements. Hence, the number $N_{a,1}$ of $(k+1)$-element antichains that are contained in either $A_1 \cup B_2$ or $A_2 \cup B_3$ satisfies
\[
\begin{split}
N_a & \ge \sum_{i=2}^3 \sum_{Y \subseteq B_i} \binom{k+\ell-|Y|}{k+1-|Y|} \ge \sum_{i=2}^3 2^{b_i} \binom{k+\ell-b_i}{k+1-b_i} \ge \sum_{i=2}^3 2^{b_i} \left(1 - \frac{\ell-1}{k+\ell-b_i} \right)^{b_i} \binom{k+\ell}{k+1} \\
& \ge \left[\left(\frac{7}{4}\right)^{b_2} + \left(\frac{7}{4}\right)^{b_3}\right] \binom{k+\ell}{k+1},
\end{split}
\]
where the last inequality follows since $\ell, b_i \ll k$. Moreover, by Lemma~\ref{lemma:C-antichains}, there are additionally at least $20(\ell-1)\binom{k+\ell}{k+1}$ antichains with $k+1$ elements that contain an element of $D_3$. Thus, using the inequality $\big(1 - \frac{\ell-1}{k-1}\big)^3 \ge 1 - 3\frac{\ell-1}{k-1}$,
\begin{equation}
\label{eq:example-1}
\begin{split}
h_k(P) \binom{k+\ell}{k+1}^{-1} & \ge \left(1 - \frac{\ell-1}{k-1}\right)^3 (3k + \ell - b_2 - b_3) + \left(\frac{7}{4}\right)^{b_2} + \left(\frac{7}{4}\right)^{b_3} + 20(\ell-1) \\
& \ge 3k+\ell-b_2-b_3 + 10(\ell-1) + \left(\frac{7}{4}\right)^{b_2} + \left(\frac{7}{4}\right)^{b_3} \\
& \ge 3k + 1 + \left(\frac{7}{4}\right)^{b_2} - b_2 + \left(\frac{7}{4}\right)^{b_3} - b_3.
\end{split}
\end{equation}
Finally, using the fact that $(7/4)^b - b \ge 3/4$ for every integer $b$, we see that the right hand side of~\eqref{eq:example-1} is strictly greater than $3k+2$. This completes the analysis as $m_k(\tau_{k,n}) = (3k+2)\binom{k+\ell}{k+1}$.
\subsubsection{Almost extremal posets of type~(\ref{item:extremal-2})}
Recall that $P$ is of type~(\ref{item:extremal-2}) if and only if there is some $f \in [k+\ell-1]$ such that $|A_i| = |A_i'| = k$ for all $i \neq f$, $|A_f| = k+\ell$, and $|A_f'| \ge k$. Moreover, $|A_f'| = k$ if $f \neq 1$. In particular, $q = \ell$ and hence $S = 2k + 50\sqrt{k}\log_2k$. We now show that $h_k(P) > m_k(\tau_{k,n})$ for each such $P$, unless $\ell = 1$. Let $b = |B_{f+1}|$. By~\eqref{eq:lemma-F},
\[
\Sigma_f \ge \Sigma_{f+1} + 2|A_{f+1}'| - |D_{f+1}| - 2|B_{f+1}| \ge 3k-|D_{f+1}|-2b.
\]
As before, by Lemma~\ref{lemma:F}, we may assume that $b \le 2\ell\log_2k$ since otherwise $h_k(P) > m_k(\tau_{k,n})$. If $|D_{f+1}| < k - k^{2/3}$, then $\Sigma_1 \ge \Sigma_f \ge S$, and, by Observation~\ref{obs:Sigma-1}, $h_k(P) > m_k(\tau_{k,n})$. Thus, we may assume that $|D_{f+1}| \ge k-k^{2/3}$.
Let us now carefully count homogenous $(k+1)$-sets in $P$. First, consider the collection of all chains of length $k+1$ obtained by taking an edge $xy$ of $G_f$ and an arbitrary set of $k-1$ elements from some chain of length $k+\ell$ that contains both $x$ and $y$ (such a chain exists as $A_f' = A_f$. The number of such chains containing a fixed edge is
\[
\binom{k+\ell-2}{k-1} = \frac{(k+1)k}{(k+\ell)(k+\ell-1)} \binom{k+\ell}{k+1} \ge \left(1 - \frac{\ell-1}{k}\right)^2 \binom{k+\ell}{k+1}
\]
and all chains constructed this way are distinct. Therefore, the number $N_c$ of chains of length $k+1$ satisfies
\[
N_c \ge e(G_f) \cdot \binom{k+\ell-2}{k-1} \ge \left(1-\frac{\ell-1}{k}\right)^2(2k-b)\binom{k+\ell}{k+1}.
\]
To estimate the number of antichains, note that for any $Y \subseteq B_{f+1}$, the set $Y \cup (A_f \setminus N_{G_f}(Y))$ is an antichain with $k+\ell$ elements. Hence, the number $N_{a,1}$ of $(k+1)$-element antichains that are contained in $A_f \cup B_{f+1}$ satisfies
\[
N_{a,1} \ge 2^b \binom{k+\ell-b}{k+1-b} \ge 2^b \left(1 - \frac{\ell-1}{k+\ell-b} \right)^b \binom{k+\ell}{k+1} \ge \left(\frac{7}{4}\right)^b \binom{k+\ell}{k+1},
\]
where the last inequality holds since $\ell, b \ll k$. Moreover, by Lemma~\ref{lemma:C-antichains}, there are additionally at least $20(\ell-1)\binom{k+\ell}{k+1}$ antichains with $k+1$ elements that contain an element of $D_{f+1}$. It follows that, using the inequality $\big(1 - \frac{\ell-1}{k}\big)^2 \ge 1 - 2\frac{\ell-1}{k}$,
\begin{equation}
\label{eq:example-2}
h_k(P) \binom{k+\ell}{k+1}^{-1} \ge 2k-b - 4(\ell-1) + \left(\frac{7}{4}\right)^b + 20(\ell-1).
\end{equation}
As $m_k(\tau_{k,n}) = (2k+1)\binom{k+\ell}{k+1}$, we conclude that $h_k(P) > m_k(\tau_{k,n})$ unless $\ell=1$ and $b \le 1$.
\subsubsection{Almost extremal posets of type~(\ref{item:extremal-2}) when $\ell = 1$}
We finally show that $h_k(P) \ge m_k(\tau_{k,n})$ for each poset $P$ of type~(\ref{item:extremal-2}) and provide a rough structural characterization of such posets which contain precisely $m_k(\tau_{k,n})$ homogeneous $(k+1)$-sets. Since we may assume that $\ell = 1$, then $m_k(\tau_{k,n}) = 2k+1$ and $\Sigma_1$ is the number of chains of length $k+1$ in $P$. Let $b = |B_{f+1}|$ and recall that $b \le 1$. By~\eqref{eq:lemma-F},
\[
\Sigma_1 \ge \Sigma_{f+1} + 2k - |D_{f+1}| - 2b \ge 2k - b.
\]
Moreover, $A_f \cup B_{f+1}$ contains at least $2^b$ antichains with $k+1$ elements. It follows that $h_k(P) \ge 2k - b + 2^b \ge m_k(\tau_{k,n})$. Moreover, either $h_k(P) > m_k(\tau_{k,n})$ or $\Sigma_{f+1} = k$, $D_{f+1} = A_{f+1} \setminus B_{f+1}$ (i.e., each $x \in A_{f+1} \setminus B_{f+1}$ has degree two in $G_f$), and there are only $2^b$ antichains with $k+1$ elements (plus, they are both contained in $A_f \cup B_{f+1}$). This means that for every $X \subseteq A_{f+1}$ such that $X \neq \emptyset$ and $X \neq B_{f+1}$, we have $|N_{G_f}(X)| \ge |X|+1$, as otherwise $X \cup (A_f \setminus N_{G_f}(X))$ would be an additional antichain with $k+1$ elements (other than $A_f$ and $B_{f+1} \cup (A_f \setminus N_{G_f}(B_{f+1}))$, which were already counted above in the $2^b$ term). In particular, $f=1$ and $A_1' = A_1$, as otherwise $|A_f'|=|A_{f+1}'|=k$ and consequently $|N_{G_f}(A_{f+1}')| = |A_{f+1}'|$. We now show that these conditions uniquely determine the graph $G_1$.
\begin{claim}
Suppose that $h_k(P) = m_k(\tau_{k,n})$.
\begin{enumerate}[(i)]
\item
\label{item:b0}
If $b = 0$, then $G_1$ is a path with $2k+1$ vertices.
\item
\label{item:b1}
If $b = 1$, then $G_1$ is the disjoint union of a path with $2k-1$ vertices and en edge.
\end{enumerate}
\end{claim}
\begin{proof}
Let $H$ be the auxiliary multigraph on $A_1$ where the multiplicity of each pair $xy$ is the number of common $G_1$-neighbors of $x$ and $y$ in $A_2$. As we have assumed that $D_2 = A_2 \setminus B_2$, it follows that $e(H) = |A_2 \setminus B_2| = k - b$. Moreover, the condition $|N_{G_1}(X)| \ge |X|+1$ implies that $H$ is a forest. (Here again we consider two parallel edges to be a cycle.) We claim that each tree in $H$ is a path. If this is not the case, then $H$ would have a vertex of degree at least $3$ and thus $G_1$ would contain the $1$-subdivision of $K_{1,3}$ as an induced subgraph. But this is not possible as both $G_1$ and its complement are comparability graphs (since $P$ is a poset of order dimension at most two) and one can check that the $1$-subdivision of $K_{1,3}$ does not have this property.
Now, (\ref{item:b0}) follows since $b=0$ implies that $H$ is a path with $k+1$ vertices and hence $G_1$ is a path with $2k+1$ vertices. To see~(\ref{item:b1}), note first that $b=1$ implies that $H$ has two connected components. The unique vertex $z \in B_2$ has a $G_1$-neighbor in one of these components. Denote it by $C$ and let $X = N_{G_1}(C)$. One easily checks that $|X| = |N_{G_1}(X)| = |C|$ and hence $X = \{z\}$ and $|C|=1$, as we have assumed that $X = \emptyset$ and $X = B_2$ are the only subsets of $A_2$ with $|X| \le |N_{G_1}(X)|$. It follows that $H$ is the union of a path with $k$ vertices and an isolated vertex and hence $G_1$ is the union of a path with $2k-1$ vertices and an edge.
\end{proof}
Finally, the following lemma (where we take $i = 2$ and $j = k+1$) shows that $P \setminus A_1$ may be partitioned into $k$ chains, as otherwise $\Sigma_2 > k$.
\begin{lemma}
\label{lemma:disjoint-chains}
Suppose that $i, j \in [k+\ell]$ with $i \le j$ are such that $|A_i'| = \ldots = |A_j'| = k$. There is some $d \ge 0$ such that:
\begin{enumerate}[(i)]
\item
\label{item:disjoint-chains-one}
There are pairwise disjoint chains $L_1, \ldots, L_{k-d}$ of length $j - i + 1$ with $\min L_p \in A_i'$ and $\max L_p \in A_j'$ for each $p \in [k-d]$ and
\item
$\Sigma_i \ge \Sigma_j + d$.
\end{enumerate}
\end{lemma}
\begin{proof}
We prove the claim by reverse induction on $i$. The statement is vacuously true for $i = j$ as $|A_j'| = k$. Suppose now that $i < j$ and, appealing to the inductive assumption, let $L_1', \ldots, L_{k-d'}'$ be a collection of pairwise disjoint chains of length $j-i$ with $\min L_p \in A_{i+1}'$ and $\max L_p \in A_j'$ for each $p \in [k-d']$. Let
\[
X = \{\min L_p \colon p \in [k-d']\} \subseteq A_{i+1}'
\]
and let $k-d$ be the size of the largest $G_i$-matching between $X$ and $A_i'$. Since this matching naturally extends some $k-d$ chains in $\{L_1', \ldots, L_{k-d'}'\}$ to pairwise disjoint chains $L_1, \ldots, L_{k-d}$ satisfying assertion~(\ref{item:disjoint-chains-one}) of the lemma, it is enough to show that $\Sigma_i \ge \Sigma_{i+1} + d-d'$. By Hall's theorem, there is a $Y \subseteq X$ such that $|N_{G_i}(Y)| \le |Y| - (d-d')$. Since $G_i[A_i',A_{i+1}']$ has no isolated vertices, it follows that
\[
e_{G_i}(A_i', A_{i+1}') \ge e_{G_i}(A_i',Y) + e_{G_i}(A_i' \setminus N_{G_i}(Y), A_{i+1}') \ge |Y| + |A_i'| - |N_{G_i}(Y)| \ge k + d-d'.
\]
Consequently,
\[
\Sigma_i = \sum_{x \in A_{i+1}'} u_{i+1}(x) \ge \Sigma_{i+1} + e_{G_i}(A_i', A_{i+1}') - |A_{i+1}'| \ge \Sigma_{i+1} + d-d'.\qedhere
\]
\end{proof}
This completes the proof of Theorem~\ref{thm:posets}.
\section{Concluding remarks}
\label{sec:concluding-remarks}
In this paper, we have determined the minimum number of monotone subsequences of length $k+1$ in a sequence of $n \le k^2 + k^{3/2} / (300\log_2k)$ numbers for all sufficiently large $k$. This minimum, which we denoted by $m_k(n)$, is achieved by taking $k$ increasing (decreasing) sequences of lengths $\lfloor n/k \rfloor$ or $\lceil n/k \rceil$ in such a way that there is no decreasing (increasing) subsequence of length $k+1$. One such sequence is $\tau_{k,n}$, defined in Section~\ref{sec:introduction}. Moreover, we have shown that if $n \neq k^2+k+1$, then no extremal sequence contains both increasing and decreasing subsequences of length $k+1$. Our results provide strong evidence supporting Conjecture~\ref{conj:main}, which asserts that the above statements remain true for all pairs of $k$ and $n$. It is also worth mentioning that, although we have not stated it explicitly, our proof establishes a stability statement of the following form: If a sequence of $n$ numbers is not `close' to a union of $k$ increasing (decreasing) sequences of almost equal lengths, then it contains `many' more than $m_k(n)$ monotone subsequences of length $k+1$.
Since we are still far from proving Conjecture~\ref{conj:main}, one may ask to determine at least the asymptotic behavior of the function $m_k(n)$. Let $\mu_k(n) = m_k(n) \binom{n}{k+1}^{-1}$.
Then Conjecture~\ref{conj:main} suggests that $\mu_k(n) = (1+o(1))k^{-k}$. Standard averaging arguments can be used to show that $\mu_k(n)$ is
non-decreasing in $n$. Therefore, the theorem of Erd{\H{o}}s and Szekeres implies that
\begin{equation}
\label{eq:mukn-lower}
\mu_k(n) \ge \mu_k(k^2+1) = \binom{k^2+1}{k+1}^{-1} \sim \sqrt{\frac{2\pi e}{k}} \cdot (ek)^{-k}.
\end{equation}
Our main theorem yields an improvement of~\eqref{eq:mukn-lower} by a factor of (only) $2^{\Theta(\sqrt{k})}$ and it would be interesting to improve this lower bound further.
We find very promising the prospect of studying Erd{\H{o}}s--Rademacher-type problems in other settings. In principle, one can investigate such extensions for any extremal or Ramsey-type result. Some motivation for studying these problems
comes from the recently renewed interest in `supersaturation' results, which have been used in conjunction with the `transference' theorems of Conlon and Gowers~\cite{CoGo} and of Schacht~\cite{Sc} as well as the `hypergraph containers' theorems of
Balogh, Morris, and the first author~\cite{BaMoSa} and of Saxton and Thomason~\cite{SaTh} to prove numerous `sparse random analogues' of classical extremal and Ramsey-type results.
\bibliographystyle{amsplain}
|
\section{Introduction}
As a tool of the 21st century for solving theoretical, practical and mathematical problems in all scientific areas,
the Heun's functions are a universal method for treatment of a vast variety of phenomena
in complicated systems of different kinds:
in solid state physics, crystalline materials, graphene, in celestial mechanics, quantum mechanics,
quantum optics, quantum field theory, atomic and nuclear physics, heavy ion physics,
hydrodynamics, atmosphere physics, gravitational physics, black holes, compact stars,
and especially in extremely urgent and expensive search for gravitational waves,
astrophysics, cosmology, biophysics, studies of the genome structure, mathematical chemistry,
economic and financial problems, etc.
This wide area of application is a result of the general type of the Heun's
differential equation that properly describes processes in all scientific areas.
The general Heun's equation written in the Fuchsian form
\ben
\hskip -.truecm
H''+\left({\frac {\gamma_{{}_G}} {z}}+{\frac{\delta_{{}_G}}{z-1}}+{\frac{\epsilon_{{}_G}}{z-a_{{}_G}}}\right)H'+
{\frac{\alpha_{{}_G}\beta_{{}_G} z-\lambda}{z(z-1)(z-a_{{}_G})}}H = 0,\,\,\,\,\,
\gamma_{{}_G}\!+\!\delta_{{}_G}\!+\!\epsilon_{{}_G}\!=\!\alpha_{{}_G}\!+\!\beta_{{}_G}\!+\!1;
\hskip 1truecm
\la{dHeunG}
\een
was constructed by Karl Heun in \cite{Heun} as a generalization of the standard hypergeometric equation
by adding one more regular singular point in complex plane: $z=a_{{}_G}\in \mathbb{C}$.
\footnote{Everywhere in this paper the prime denotes a derivative with respect to the variable $z$.}
At present, this is the most popular form of the Heun's equation, see \cite{Ron,SL} and the literature therein.
It is not symmetric with respect to four regular singular points $0, 1, a_{{}_G}, \infty$ with the corresponding indices
\ben
\{ 0, 1-\gamma_{{}_G}\},\quad \{ 0, 1-\delta_{{}_G} \},\quad \{ 0, \gamma_{{}_G}+\delta_{{}_G}-\alpha_{{}_G}-\beta_{{}_G} \},\quad \{ \alpha_{{}_G}, \beta_{{}_G} \}.
\la{indexG}
\een
The Heun's general function
$\text{HeunG}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G}, z)$
is defined as the unique local regular solution around the regular singular point $z=0$ under normalization
$\text{HeunG}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G}, 0)=1$
\footnote{Here we are using the notations of the widespread computer package Maple.}.
The second linearly independent local solution can be obtained via a proper change of the parameters,
as described, for example, in \cite{Ron,SL}. Using proper Mobius transformations (See the Appendix \ref{ApA}.) one can also obtain
similar local solutions around other regular singular points implementing the Heun's general function
(see, for example, \cite{Ron,SL}). Thus, the problem of finding all local solutions of Eq. \eqref{dHeunG}
is reduced to the study of the Heun's general function defined in the vicinity of the point $z=0$ by the absolutely
convergent series
\ben
\text{HeunG}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G},z)=
\sum_{n=0}^\infty h_{n}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G})z^n.
\la{HeunG}
\een
Replacing the function $H(z)$ in Eq. \eqref{dHeunG} with the series \eqref{HeunG}
one easily obtains the simple three-term recurrence relation
\ben
h_{n}+R_{n-1}h_{n-1}+R_{n-2}h_{n-2}=0
\la{recurrenceG}
\een
with the coefficients
\ben
R_{n-1}= -1 -{\frac 1 {a_{{}_G}}} +
{\frac {\lambda-\gamma_{{}_G}(a_{{}_G}\delta_{{}_G}-a_{{}_G}+\alpha_{{}_G}+\beta_{{}_G}-\delta_{{}_G}-\gamma_{{}_G})} {a_{{}_G}(\gamma_{{}_G} +n-1)(\gamma_{{}_G}-1)}},\nonumber \\
R_{n-2}={\frac 1 {a_{{}_G}}} + {\frac {-\alpha_{{}_G}\beta_{{}_G}+\alpha_{{}_G}\gamma_{{}_G}+\beta_{{}_G}\gamma_{{}_G}-\gamma_{{}_G}^2+\alpha_{{}_G}+\beta_{{}_G}-2\gamma_{{}_G}-1} {a_{{}_G}(\gamma_{{}_G} +n-1)(\gamma_{{}_G}-1)}}.
\la{R}
\een
Using relations \eqref{recurrenceG}, \eqref{R} and the initial conditions $h_0=1, h_1=\lambda/a_{{}_G}\gamma_{{}_G}$
one can effectively calculate the values of the series \eqref{HeunG} in the circle around the point $z=0$ with the circle-radius $<1$,
i.e., before approaching the next regular singular point $z=1$.
Trying to continue the series \eqref{HeunG} outside this circle, one meets hard numerical problems,
as seen from the ten-year not very satisfactory attempts to improve the only existing computer code for work
with the Heun's functions -- Maple.
At present, this is a serious obstacle for numerous applications of these extremely useful functions.
The main idea of the present paper is to find a novel representation of the solutions of the general Heun's equation
which gives an equal treatment of all regular singular points and yields series expansions
which are valid simultaneously in the vicinities of all of them.
We succeeded in finding such an approach but, as one can expect,
it leads to new and more complicated series expansions
of solutions of the general Heun's functions defined by the nine-term recurrence relations.
Fortunately, such recurrence relations are not a problem for modern computers.
In the present paper, we introduce for the first time these new series and study their basic properties.
One can hope that the new series will be a
useful tool for solution of the basic open problems in the theory of the general Heun's functions, like
connection problem, study of the asymptotics, monodromy group, relations between the derivatives of the Heun's functions,
e.t.c.,
as well as for development of new more efficient computational techniques.
\section{Symmetric form of the Heun's equation}
\subsection{The symmetric form of the general Fuchsian equation}
The {\em symmetric} form of the general Fuchsian equation with $N\geq 4$ arbitrary
regular singular points $z_{j=1,...,N}\in \mathbb{C}$ was adopted
by Felix Klein as early as in \cite{Klein}:
\ben
\mathcal{W}''+\left(\sum_{j=1}^N{\frac{1-\alpha_j-\beta_j}{z-z_j}}\right)\mathcal{W}'+
{\frac 1 {P(z)}\left(\Lambda(z)+\sum_{j=1}^N{\frac{q_j}{z-z_j}}\right)}\mathcal{W} = 0,
\la{dWN}
\een
see also \cite{Forsyth}. Here
\ben
\hskip -1.truecm P(z)=\prod_{j=1}^N(z-z_j)=\sum_{n=0}^N(-1)^n\sigma_{{}_{N-n}} z^n,\quad \Lambda(z)=\sum_{l=0}^{N-4}\lambda_l z^l,\quad
\text{and}\,\,\,q_j=\alpha_j\beta_j P^\prime(z_j)\,\,\, \text{for}\,\,\,{j=1,...,N}.
\la{LP}
\een
Under the additional condition
\ben
\sum_{j=1}^N \left(\alpha_j+\beta_j\right)=N-2
\la{constr}
\een
the point $z=\infty$ is a regular one for Eq. \eqref{dWN}. Thus,
it remains with only $N$ finite regular singular points $z_{j=1,...,N}\in \mathbb{C}$
with arbitrary indices $\{\alpha_j,\beta_j\}_{j=1...N}\in \mathbb{C}$.
As seen, such equations are determined altogether by $4(N-1)$ arbitrary complex numbers:
singular points $z_{j=0,...,N}$, their indices $\{\alpha_j,\beta_j\}:{j=0,...,N}$ with constraint \eqref{constr},
and auxiliary parameters $\lambda_{l=0,...,N-4}\in \mathbb{C}$.
Now one can use the following transformation of the unknown function $\mathcal{W}(z)$ with the properly chosen parameters $\nu_{j=1,...,N}$:
\ben
\mathcal{W}(z) = \mathcal{F}(z)\prod_{j=1}^N \left(z-z_j\right)^{\nu_j}, \qquad \sum_{j=1}^N \nu_j=0,
\la{nu_transf}
\een
to fix $N-1$ of the parameters $\{\alpha_j,\beta_j\}:{j=0,...,N}$, or some $(N-1)$-in-number their combinations.
The second condition in Eq. \eqref{nu_transf} is necessary to preserve relation \eqref{constr}.
As a result of the last two constraints, we remain with altogether $N$ free parameters between the indices $\{\alpha_j,\beta_j\}:{j=0,...,N}$.
For example, an asymmetric choice $\beta_{j=1,...,N-1}=0$, similar to \eqref{indexG}, is possible.
Thus, we remain with altogether $3(N-1)$ free complex parameters.
Instead of the above asymmetric choice, which destroys the symmetric treatment of the regular singular points,
we prefer to use the following $N$-in-number {\it symmetric} constraints on the indices $\{\alpha_j,\beta_j\}:{j=0,...,N}$:
\ben
\alpha_j+\beta_j=1 - {\tfrac 2 N}, \quad \alpha_j\beta_j P^\prime(z_j)=q_j,\quad \text{for}\quad j=1,...,N,
\la{index}
\een
thus preserving equal treatment of all $N$ regular singular points and relation \eqref{constr}.
Introducing new $N$-in-number free uniformization parameters $\chi_{j=1,...,N} \in \mathbb{C}$
we obtain for all $j=1,...,N$:
\ben
\alpha_j=\left(1-{\tfrac 2 N}\right) \left(\cos\chi_j\right)^2,\quad \beta_j=\left(1-{\tfrac 2 N}\right) \left(\sin\chi_j\right)^2,\nonumber\\
\quad q_j=\left(\left({\tfrac 1 2}-{\tfrac 1 N}\right)\sin(2\chi_j)\right)^2 \prod_{k\neq j}^N\left(z_j-z_k\right).
\la{alphabeta}
\een
and Eq. \eqref{dWN} acquires its simplest $3(N-1)$-parameter final form:
\ben
\mathcal{F}''+{\frac 2 N}\left(\sum_{j=1}^N{\frac 1{z-z_j}}\right)\mathcal{F}'+
{\frac 1 {P(z)}\left(\Lambda(z)+\sum_{j=1}^N{\frac{q_j}{z-z_j}}\right)}\mathcal{F} = 0,
\la{dWFsym}
\een
Note that:
i) In Eq. \eqref{dWFsym} one can consider the parameters $q_{j=1,...,N}$ as independent ones,
instead of the uniformization parameters $\chi_{j=1,...,N}$.
The disadvantage of this approach is in the introduction of branching points
of the indices $\{\alpha_j,\beta_j\}:{j=1,...,N}$,
since indices of singular points of Eq. \eqref{dWFsym} are the roots $x_j^\pm$ of the corresponding quadratic equations
$x_j^2-2\left({\tfrac 1 2}-{\tfrac 1 N}\right)x_j +q_j/P^\prime(z_j)=0$, $j=1,...,N$.
The presence of such branching points is undesirable
since it requires special care during numerical calculations.
ii) We still have the freedom to lower the number of the free parameters of the problem,
moving some three different singular points
to any convenient different places in the complex plane $\mathbb{C}$,
for example to $0,1,a_{{}_G}$, as in the case of general Heun's functions.
This can be done by using proper Mobius transformation
without changing the number and the character of the singular points of Eq. \eqref{dWFsym},
see Appendix \ref{ApA}.
Thus, we will end with $3(N-2)$ essential free parameters of Eq. \eqref{dWFsym}, as illustrated
in the rest of the paper by the basic example $N=4$
\footnote{For another form of the general Fuchsian equation with $N$ singular points
and the corresponding count of the number of free parameters in it see \cite{NIST}.}.
iii) There exist two quite different cases of positions of the singular points $z_{j=0,...,N}$.
$\bullet$ The first one is the special case when all regular singular points $z_{j=0,...,N}$ of Eq. \eqref{dF} lie
on some circle $\mathfrak{C}\in \mathbb{C}$.
A special and natural case is the one when $z_{j=0,...,N}\in \mathbb{R}$, i.e.,
all singular points are real, as in the important Smirnov's Thesis \cite{Smirnov}.
We will call this case {\it the circular case}.
In the circular case, one is able to move all singular points $z_{j=0,...,N}$ on any other circle
$\mathfrak{\tilde C}\in \mathbb{C}$ using Mobius transformation, see Appendix \ref{ApA}.
$\bullet$ The opposite (general) case is the one in which the singular points $z_{j=0,...,N\geq 4}$ do not lie
on any circle in the complex plane. We call it {\it the non circular case}.
In the non circular case, the theory of solutions of Eq. \eqref{dF} is much more complicated.
It is not developed enough even for the general Heun's equation \eqref{dHeunG}.
Hence, the choice of the position of the regular singular points $z_{j=0,...,N}$ of Eq. \eqref{dF}
is an important component of the general theory. In the present paper, we investigate this problem only
for the circular case with $N=4$, since it corresponds to the general Heun's functions.
As we shall show, in this case the general theory is a relatively simple one.
Equation \eqref{dWFsym} can be written down also in the following self-adjoint form:
\ben
\left(P(z)\right)^{1-2/N}\left(P(z)^{2/N}\mathcal{F}'\right)^\prime+
\left(\Lambda(z)-\left({\tfrac 1 2}-{\tfrac 1 N}\right)^2{\tfrac 1 {P(z)}}
\sum_{j=1}^N \left(\sin(2\chi_j)\right)^2\partial_z P(z_j)\partial_{z_j}P(z) \right)\mathcal{F} = 0.
\la{dWFsymSA}
\een
{\bf Proposition 1:} Let $\Lambda_{1,2}(z)$ be two polynomials of degree $(N-4)$ with equidistant coefficients:
$\lambda_{2,l}-\lambda_{1,l}=\Delta\lambda\neq 0\quad \forall\quad l=0,...,(N-4)$. Hence,
\ben
\Lambda_2(z)-\Lambda_1(z)=\Delta\lambda\,{\frac{z^{N-3}-1}{z-1}}.
\la{Lambda_1-2}
\een
Then, the solutions $\mathcal{F}_{\Lambda_{1,2}}(z)$ of the corresponding Eqs. \eqref{dWFsym}, or \eqref{dWFsymSA}
are orthogonal with respect to the measure
\ben
d\mu(z)={\frac{z^{N-3}-1}{z-1}}\left(P(z)\right)^{{\frac 2 N}-1},\quad\text{i.e.,}
\la{measureN}
\een
\ben
\int_{z_j}^{z_j}\mathcal{F}_{\Lambda_{1}}(z)\mathcal{F}_{\Lambda_{2}}(z)d\mu(z)=0,
\la{orthogonalN}
\een
if the boundary conditions
\ben
\left(P(z)\right)^{{\frac 2 N}}
\left(\mathcal{F}_{\Lambda_{1}}(z)\mathcal{F}_{\Lambda_{2}}^\prime(z)-
\mathcal{F}_{\Lambda_{2}}(z)\mathcal{F}_{\Lambda_{1}}^\prime(z)\right)\,\upharpoonleft_{z_j,z_j}=0
\la{boundaryN}
\een
are satisfied.\,$\blacktriangleleft$ \footnote{The
sign $\blacktriangleleft$ denotes the end of the corresponding statement.}
Thus, we arrived at a quite unusual boundary-value problem for Eqs. \eqref{dWFsym} and \eqref{dWFsymSA}.
The proof is based on the integration of the general identity valid for any functions $\Lambda_{1,2}(z)$ in Eqs. \eqref{dWFsym} and \eqref{dWFsymSA}:
\ben
\left(\Lambda_2(z)-\Lambda_1(z)\right)\left(P(z)\right)^{{\frac 2 N}-1}\mathcal{F}_{\Lambda_{1}}(z)\mathcal{F}_{\Lambda_{2}}(z)\equiv
\left(\left(P(z)\right)^{{\frac 2 N}} \left(\mathcal{F}_{\Lambda_{1}}(z)\mathcal{F}_{\Lambda_{2}}^\prime(z)-
\mathcal{F}_{\Lambda_{2}}(z)\mathcal{F}_{\Lambda_{1}}^\prime(z)\right)\right)^\prime.
\la{identityN}
\een
{\bf Proposition 2:} Equations \eqref{dWFsym} and \eqref{dWFsymSA} are invariant under
the {\it extension} of the Mobius group $\widehat{\mathfrak{G}}_{Mobius}$
that acts on the functions of $(3N-2)$ variables
$\mathcal{F}\left(z;z_1,...,z_N;q_1,...,q_N;\lambda_0,...\lambda_{N-4}\right)$
and is produced by the following basic transformations:
\begin{description}
\item[(i)] Complex translations with arbitrary $\zeta \in \mathbb{C}$:
\ben
\hskip - 1.2truecm z\to z+\zeta; \quad z_j\to z_j+\zeta:\, j=1,...,N;\quad q_j\to q_j:\, j=1,...,N; \quad
\lambda_l\to \sum_{m=l}^{N-4} {\binom {m}{l}}\zeta^{m-l}\lambda_{m}:\, l=0,...,N-4.
\la{translationN}
\een
\item[(ii)] Complex dilatations with arbitrary $t\in \mathbb{C}$, $t\neq 0$:
\ben
z\to t\, z; \quad z_j\to t\, z_j:\, j=1,...,N; \quad q_j\to t^{N-1} q_j:\, j=1,...,N;\quad \lambda_l\to t^{N-l-2} \lambda_l:\, l=0,...,N-4.
\la{rescalingN}
\een
\item[(iii)] Inversion
\ben
z&\to& 1/ z; \quad z_j\to 1/ z_j:\, j=1,...,N; \quad q_j \to {\tfrac {(-1)^{N-1}}{\sigma_{\!{}_N}}} z_j^{1-N} q_j :\, j=1,...,N;\nonumber\\
\lambda_l &\to& {\tfrac {(-1)^{N-1}}{\sigma_{\!{}_N}}}\bigg(\Big(\sum_{j=1}^N z_j^{l+3-N}q_j\Big)-\lambda_{N-4-l}\bigg):\, l=0,...,N-4;
\la{inversionN}
\een
where $\sigma_{\!{}_N}=\prod_{j=1}^N z_j$.\nonumber\,$\blacktriangleleft$
\end{description}
Indeed, it is not hard to check directly the invariance of Eq. \eqref{dWFsym}
under transformations \eqref{translationN}, \eqref{rescalingN}, and \eqref{inversionN}.
Using proper compositions of these basic transformations (see Appendix \ref{ApA}) we are able to construct
a representation of the whole extended Mobius group $\widehat{\mathfrak{G}}_{Mobius}$ that acts on the solutions
$\mathcal{F}\left(z;z_1,...,z_N;q_1,...,q_N;\lambda_0,...\lambda_{N-4}\right)$ of Eq. \eqref{dWFsym}
without bringing us outside of the variety of these solutions. Hence, $\widehat{\mathfrak{G}}_{Mobius}$
is the group of invariance of the variety of solutions to Eq. \eqref{dWFsym}.
\subsection{Symmetric form of the general Heun's equation and its Taylor series solutions}
\subsubsection{Symmetric form of the general Heun's equation.}
The symmetric form \eqref{dWN} for the special case of the general Heun's Eq. \eqref{dHeunG}
(i.e., for $N=4$) was pointed out in \cite{Ron}.
For brevity, in this case, we denote by $\lambda$ the single auxiliary parameter.
Then, the symmetric form \eqref{dWFsym} of the general Heun's Eq. \eqref{dHeunG} reads
\ben
\mathcal{F}^{\prime\prime}+{\frac 1 2}\left(\sum_{j=1}^4{\frac 1{z-z_j}}\right)\mathcal{F}^\prime+
{\frac 1 {P(z)}\left(\lambda+\sum_{j=1}^4{\frac{q_j}{z-z_j}}\right)}\mathcal{F} = 0,
\la{dF}
\een
or in a self-adjoint form:
\ben
\left(P(z)\right)^{1/2}\left(\left(P(z)\right)^{1/2}\mathcal{F}^\prime\right)^\prime +
\left(\lambda+Q(z)\right)\mathcal{F} = 0,
\la{dFsa}
\een
where $Q(z)=\sum_{j=1}^4{\frac{q_j}{z-z_j}}=-{\tfrac 1 {16}}{\tfrac 1 {P(z)}}
\sum_{j=1}^4 \left(\sin(2\chi_j)\right)^2\partial_z P(z_j)\partial_{z_j}P(z)$.
\subsubsection{The orthogonality of solutions.}
The form \eqref{dFsa} shows that the auxiliary parameter $\lambda$ actually plays the role of eigenvalue of the problem.
This form is also convenient for discussing the orthogonality of the solutions
$\mathcal{F}$ on the contours $\mathcal{L}\in\mathbb{C}$ under the measure $d\mu(z)=\left(P(z)\right)^{-1/2}dz$ \footnote{For polynomial $P(z)$
of the fourth degree this measure is obviously related
with the elliptic integrals \cite{NIST},
thus giving the basis for the well-known relation of the general Heun's functions
with elliptical ones, see, for example, \cite{Fiziev10} and the references therein.}:
{\bf Proposition 3:}
For any two solutions $\mathcal{F}_{\lambda_{1,2}}(z)$ of the Eq. \eqref{dFsa} with $\lambda_1\neq\lambda_2$ we have
\ben
\hskip -.5truecm \int_{\mathcal{L}_{ij}}\mathcal{F}_{\lambda_1}(z)\mathcal{F}_{\lambda_2}(z)d\mu(z)=0.
\la{orthogonality}
\een
Here $\mathcal{L}_{ij}\in\mathbb{C}$ is any contour which starts at the singular point $z_j$ and ends at the singular point $z_j$
without going through the other singular points $z_{k\neq i,j}$. Besides, the singular boundary conditions
\ben
\left(P(z)\right)^{1/2}
\left(\mathcal{F}_{\lambda_1}(z)\mathcal{F}_{\lambda_2}^\prime(z)-
\mathcal{F}_{\lambda_2}(z)\mathcal{F}_{\lambda_1}^\prime(z)\right)\,\upharpoonleft_{z_j,z_j}=0
\la{boundary}
\een
\noindent are supposed to be satisfied. The same boundary conditions ensure the self-adjoint property of the differential operator
in Eq. \eqref{dFsa} with respect to the measure $d\mu(z)=\left(P(z)\right)^{-1/2}dz$.\,$\blacktriangleleft$.
Indeed, the orthogonality relation \eqref{orthogonality} is an immediate consequence of the identity
\ben
\hskip -1.3truecm \left(\lambda_2 -\lambda_1\right) \int_{\mathcal{L}_{ij}}\mathcal{F}_{\lambda_2}(z)\mathcal{F}_{\lambda_1}(z)d\mu(z)=
\left(P(z)\right)^{1/2}
\left(\mathcal{F}_{\lambda_2}(z)\mathcal{F}_{\lambda_1}^\prime(z)-
\mathcal{F}_{\lambda_1}(z)\mathcal{F}_{\lambda_2}^\prime(z)\right)\,\Big|_{z_j,z_j},
\la{identity}
\een
which follows from Eq. \eqref{dFsa} by applying the well-known procedure for two solutions $\mathcal{F}_{\lambda_{1,2}}(z)$
with $\lambda_1 \neq\lambda_2$, and yields the boundary conditions \eqref{boundary}.
It is not hard to justify Proposition 3.
Indeed, Eq. \eqref{dF} has the following two linearly independent local Frobenius
solutions in the vicinity of each regular singular point $z_j$ \cite{Golubev, CL}:
\ben
\hskip -.6truecm \mathcal{F}_{\alpha_j}(z)=\left(z-z_j\right)^{\alpha_j}\sum_{n=0}^\infty f_{\alpha_j, n}\left(z-z_j\right)^n,\quad
\mathcal{F}_{\beta_j}(z)=\left(z-z_j\right)^{\beta_j} \sum_{n=0}^\infty f_{\beta_j, n}\left(z-z_j\right)^n.
\la{Frobenius}
\een
Then, in the vicinity of the point $z_j$ the solutions $\mathcal{F}_{\lambda_{1,2}}(z)$ allow the representation
\ben
\mathcal{F}_{\lambda_{1,2}}(z)=C^{\alpha_j}_{\lambda_{1,2}}\mathcal{F}_{\alpha_j}(z)+C^{\beta_j}_{\lambda_{1,2}}\mathcal{F}_{\beta_j}(z)
\la{localF_z_j}
\een
with proper constants $C^{\alpha_j}_{\lambda_{1,2}},C^{\beta_j}_{\lambda_{1,2}}$. Taking into account that in the same vicinity
$P(z)=P^\prime(z_j)(z-z_j)+O_2(z-z_j)$, one obtains from Eq. \eqref{localF_z_j}
\ben
\left(P(z)\right)^{1/2}
\left(\mathcal{F}_{\lambda_2}(z)\mathcal{F}_{\lambda_1}^\prime(z)-
\mathcal{F}_{\lambda_1}(z)\mathcal{F}_{\lambda_2}^\prime(z)\right)=\\
=\left(\alpha_j-\beta_j\right)\left(P^\prime(z_j)\right)^{1/2}
\left(C^{\alpha_j}_{\lambda_{1}}C^{\beta_j}_{\lambda_{2}}-C^{\alpha_j}_{\lambda_{2}}C^{\beta_j}_{\lambda_{1}}\right) +\mathcal{O}(z-z_j).
\nonumber
\la{orth_cond}
\een
Hence, the boundary condition at the singular point $z_j$ can be satisfied either if
$\alpha_j=\beta_j$ which gives $\chi_j/\hskip -.2truecm \mod\!(2\pi)=\pm \pi/4, \pm 3\pi/4$
or if $C^{\alpha_j}_{\lambda_{1}}/C^{\beta_j}_{\lambda_{1}}= C^{\alpha_j}_{\lambda_{2}}/C^{\beta_j}_{\lambda_{2}}$
which leads to a special coherent choice of the solutions $\mathcal{F}_{\lambda_{1,2}}(z)$ \eqref{localF_z_j}
with the coefficient ratio being independent of the eigenvalues $\lambda_{1,2}$.
If one imposes the same boundary condition also at the second singular point $z_j\neq z_j$,
then one arrives at a specific two-singular-point boundary problem \cite{Ron,SL,KLS,LS}.
In this case the auxiliary parameter of the solution, i.e. the eigenvalue $\lambda$, can have only
some definite values which define the spectrum of the self-adjoint operator in Eq. \eqref{dFsa}, see \cite{Smirnov},
where the standard approach to the general Heun's functions was substantially elaborated.
This confirms once again our interpretation of the auxiliary parameter $\lambda$ as an eigenvalue parameter
in Eq. \eqref{dFsa}.
Note that in our approach it is possible to impose simultaneously
regular boundary conditions at two regular singular points
since $\mathcal{F}(z)$ is not the standard {\it local} solution like
$\text{HeunG}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G}, z)$\footnote{We
remind the reader that the term {\it general-Heun's-function} is in use for
the local solution $\text{HeunG}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G}, z)$,
and can not be applied to any other solution, like $\mathcal{F}(z)$, to the general Heun's equation.}.
In the last case, the regularity condition is already imposed at the point $z=0$ by definition.
Therefore, one is able to impose on the local regular solutions like
$\text{HeunG}(a_{{}_G}, \lambda, \alpha_{{}_G}, \beta_{{}_G}, \gamma_{{}_G}, \delta_{{}_G}, z)$
only one more regularity condition at some different regular singular point\footnote{The
author is grateful to Professor S. Yu. Slavyanov for this remark,
as well as for drawing the author's attention to reference \cite{Smirnov}.}.
Obviously, the last (widely accepted) approach is equivalent to ours.
Our treatment seems to be more natural and corresponds to the standard
boundary problem for an ordinary differential equation (See also \cite{Smirnov}.).
\subsubsection{Elementary symmetric functions related to the problem.}
Further on, we use the representation $P(z)=z^4-\sigma_1 z^3+\sigma_2 z^2-\sigma_3 z +\sigma_4$
of this fourth-degree-polynomial, thus introducing the standard elementary symmetric functions
\ben
\sigma_1=z_1+z_2+z_3+z_4, \quad
\sigma_2=z_1z_2+z_1z_3+z_1z_4+z_2z_3+z_2z_4+z_3z_4,\nonumber\\
\sigma_3=z_2z_3z_4+z_1z_3z_4+z_1z_2z_4+z_1z_2z_3,\quad
\sigma_4=z_1z_2z_3z_4,
\la{sigma}
\een
\noindent and the additional notation
\ben
\sigma_1^j=\sigma_1(z_j=0),\quad\text{for example,}\quad \sigma_1^1=z_2+z_3+z_4,\quad \text{etc},\nonumber\\
\sigma_2^j=\sigma_2(z_j=0),\quad\text{for example,}\quad \sigma_2^1=z_2z_3+z_2z_4+z_3z_4,\quad \text{etc},\\
\sigma_3^j=\sigma_3(z_j=0),\quad\text{for example,}\quad \sigma_3^1=z_2z_3z_4, \quad \text{etc} .\nonumber
\la{sigma_j}
\een
\subsubsection{Invariance of the symmetric form of the general Heun's equation under inversion.}
Now it is easy to check that the symmetric form of the general Heun's equation Eq. \eqref{dF} (as well as Eq. \eqref{dFsa})
is covariant under a proper extension of the Mobius group $\mathfrak{G}_{Mobius}$. Indeed, applying the results of Proposition 2,
in the case $N=4$ we obtain much simpler results.
{\bf Proposition 4:}
Equation \eqref{dF} is invariant under the {\it extension} of the Mobius group $\widehat{\mathfrak{G}}_{Mobius}$
that acts on the functions of $10$ variables $\mathcal{F}\left(z;z_1,...,z_4;q_1,...,q_4;\lambda\right)$
and is produced by the following basic transformations:
\begin{description}
\item[(i)] Complex translations with arbitrary $\zeta \in \mathbb{C}$:
\ben
\hskip - 1.2truecm z\to z+\zeta; \quad z_j\to z_j+\zeta:\, j=1,...,4;\quad q_j\to q_j:\, j=1,...,4; \quad
\lambda \to \lambda.
\la{translationN4}
\een
\item[(ii)] Complex dilatations with arbitrary $t\in \mathbb{C}$, $t\neq 0$:
\ben
z\to t\, z; \quad z_j\to t\, z_j:\, j=1,...,4; \quad q_j\to t^3 q_j:\, j=1,...,4;\quad \lambda \to t^2 \lambda.
\la{rescalingN4}
\een
\item[(iii)] Inversion
\ben
\hskip -1.4truecm z&\to& 1/ z; \quad z_j\to 1/ z_j:\, j=1,...,4; \quad q_j \to -q_j/\left( z_j^2 \sigma_4 \right) :\, j=1,...,4;\quad
\lambda \to \Big(\lambda -\sum_{j=1}^4 q_j/z_j\Big)/\sigma_4.\, \blacktriangleleft
\la{inversionN4}
\een
\end{description}
Using proper compositions of these basic transformations we are able to construct
a representation of the whole extended Mobius group $\widehat{\mathfrak{G}}_{Mobius}$ that acts on the solutions
$\mathcal{F}\left(z;z_1,...,z_4;q_1,...,q_4;\lambda\right)$ of Eq. \eqref{dF}
without bringing us outside of the variety of these solutions. The extended group $\widehat{\mathfrak{G}}_{Mobius}$
is the group of invariance of the variety of solutions to Eq. \eqref{dF}.
\subsubsection{The Taylor series expansion of the solutions.}
Our next step is to adopt the following basic assumption which is of crucial importance for further work:
\ben
z_{j=0,...,4}\neq 0.
\la{zero}
\een
Then the function $\mathcal{F}(z)$
is an analytical one in the vicinity of the point $z=0$ and has an absolutely convergent
Taylor series expansion
\ben
\mathcal{F}(z)\equiv\mathcal{F}(z;z_1,...z_4;q_1,...q_4;\lambda) = \sum_{n=0}^\infty f_n(z_1,...z_4;q_1,...q_4;\lambda) z^n
\la{taylorF}
\een
with the coefficients $f_n(q_1,...,q_4;\lambda)$ defined by the nine-term recurrence relation
\ben
f_n+\sum_{k=1}^8 r_{n-k}f_{n-k}=0,
\la{frecurrence}
\een
which can be obtained from Eqs. \eqref{dF} and \eqref{taylorF}.
After some lengthly but straightforward calculations one derives the following relations
for eight coefficients $r_{n-1},...,r_{n-8}$:
\begin{subequations}\label{r:abcdefgh}
\ben
(\sigma_4)^2r_{n-1}&=&-\left( 2-{\tfrac 7 2}{\tfrac 1 n}\right)\sigma_3\sigma_4,\label{r:a}\\
(\sigma_4)^2r_{n-2}&=&{\tfrac{\sigma_4}{n(n-1)}}\Big(\lambda- \sum_{j=1}^4 q_j/z_j\Big)-
\left(1-{\tfrac 5 n}+{\tfrac 3 2}{\tfrac 1 {n-1}}\right)\left((\sigma_3)^2+2\sigma_2\sigma_4\right),\label{r:b}\\
(\sigma_4)^2r_{n-3}&=&-{\tfrac 1 {n(n-1)}}\Big(\lambda\sigma_3- \sum_{j=1}^4 q_j\sigma_2^i\Big)-
\left(2-{\tfrac {39} 2}{\tfrac 1 n}+{\tfrac 9 {n-1}}\right)\left(\sigma_2\sigma_3+\sigma_1\sigma_4\right),\label{r:c}\\
(\sigma_4)^2r_{n-4}&=&{\tfrac 1 {n(n-1)}}\Big(\lambda\sigma_2- \sum_{j=1}^4 q_j\sigma_1^i\Big)+
\left(1-{\tfrac {16} n}+{\tfrac 9 {n-1}}\right)\left((\sigma_2)^2+2\sigma_1\sigma_3+2\sigma_4\right),\label{r:d}\\
(\sigma_4)^2r_{n-5}&=&-{\tfrac 1 {n(n-1)}}\Big(\lambda\sigma_1- \sum_{j=1}^4 q_j\Big)-
\left(2-{\tfrac {95} 2}{\tfrac 1 n}+{\tfrac {30} {n-1}}\right)\left(\sigma_1\sigma_2+\sigma_3\right),\label{r:e}\\
(\sigma_4)^2r_{n-6}&=&{\tfrac \lambda {n(n-1)}}+\left(1-{\tfrac {33} n}+{\tfrac {45} 2}{\tfrac 1 {n-1}}\right)\left((\sigma_1)^2+2\sigma_2\right),\label{r:f}\\
(\sigma_4)^2r_{n-7}&=&-\left(2-{\tfrac {175} 2}{\tfrac 1 n}+{\tfrac {63} {n-1}}\right)\sigma_1,\label{r:g}\\
(\sigma_4)^2r_{n-8}&=&1-{\tfrac {56} n}+{\tfrac {42} {n-1}}.\label{r:h}
\een
\end{subequations}
Using the two initial conditions for the recurrence relation \eqref{frecurrence}:
\begin{subequations}\label{inicond:1,2}
\ben
f_{-7}=0,...f_{-1}=0,\,f_0=1,\, f_1=0,\la{inicond:1}\\
f_{-7}=0,...f_{-1}=0,\,f_0=0,\, f_1=1,\la{inicond:2}
\een
\end{subequations}
we obtain two linearly independent solutions of Eq. \eqref{dF} $\mathcal{F}_{1,2}(z)$.
Both of them are analytical functions in some vicinity of $z=0$
and define the general solution: $\mathcal{F}(z)=C_1\mathcal{F}_{1}(z)+C_2\mathcal{F}_{2}(z)$ ($C_{1,2}=\text{const}$),
having standard properties of a fundamental basis of solutions of Eq. \eqref{dF}:
\begin{subequations}\label{F:1,2}
\ben
\mathcal{F}_{1}(0)=1,\quad \mathcal{F}_{2}(0)=0, \la{F:1}\\
\mathcal{F}_{1}^\prime(0)=0,\quad \mathcal{F}_{2}^\prime(0)=1.\la{F:2}
\een
\end{subequations}
In addition, these solutions obey the relation
\ben
\mathcal{F}_{1}(z)\mathcal{F}_{2}^\prime(z)-\mathcal{F}_{2}(z)\mathcal{F}_{1}^\prime(z)=\left(P(0)/P(z)\right)^{1/2}.
\la{WronF1F2}
\een
For example, for any $j=1,2,3,4$ one is able to represent the corresponding general Heun's functions in the novel form
\ben
\text{HeunG}(a_{{}_{G,j}}, \lambda, \alpha_{{}_{G,j}}, \beta_{{}_{G,j}}, \gamma_{{}_{G,j}}, \delta_{{}_{G,j}}, z-z_j)=
\Gamma_j^1\,\mathcal{F}_{1}(z;z_1,...,z_4;q_1,...,q_4;\lambda)+\Gamma_j^2\,\mathcal{F}_{2}(z;z_1,...,z_4;q_1,...,q_4;\lambda)
\la{HGF12}
\een
with some coefficients $\Gamma_{j,1}, \Gamma_{j,2}$ which play a fundamental role in our approach to these functions.
Further detailed study of relation \eqref{HGF12} is outside the scope of the present paper.
\subsection{The symmetric choice of the positions of singular points}
We shall take advantage of the freedom to put the singular points $z_{j=0,...,4}$ in the proper places in the complex plane $\mathbb{C}$
for simplifying, as much as possible, the coefficients \eqref{r:a}-\eqref{r:h} and thus, the very solutions $\mathcal{F}_{1,2}(z)$.
This can be done in a symmetric way by imposing additional conditions on the elementary symmetric functions $\sigma_{j=1,2,3,4}$.
Using the proper Mobius transformation one is able to impose three independent constraints on $z_{j=0,...,4}$
without changing the problem, see Appendix \ref{ApA}.
Obvious simple choice is to reduce the quartic equation $P(z)=0$ to the following biquadratic one: $z^4- 2\cos(2\phi)z^2+1=0$
with the roots
\ben
z_1=e^{i\phi},\,\,z_2=-e^{-i\phi},\,\,z_3=-e^{i\phi},\,\,z_4=e^{-i\phi},
\la{z_j}
\een
by imposing three symmetric constraints
\ben
\sigma_1=\sigma_3=0,\,\, \sigma_4=1,
\la{B_sigma_constraits}
\een
and replacing $\sigma_2= -2\cos(2\phi)$ with one more complex uniformization parameter $\phi\in \mathbb{C}$.
This time our goal is to avoid branching points of the roots of the above biquadratic equation.
The meaning of the new variable $\phi$ is revealed by the formula for the invariant $a(z_1,z_2,z_3,z_4)$
of the Mobius transformation -- the so called {\it cross-ratio}, see Appendix \ref{ApA}.
In our problem it acquires the form
\ben
a={\frac {(z_1-z_3)(z_2-z_4)}{(z_2-z_3)(z_1-z_4)}}={\frac 1{\left(\sin\phi\right)^2}}
\quad \Rightarrow\quad\sigma_2=-2\left(1-{\frac 2 a}\right).
\la{a}
\een
Now, it is not hard to obtain the relations
\begin{subequations}\label{rho:2,3,4,5}
\ben
\sum_{k=1}^4 q_j/z_j &=& +{\tfrac i 4}\,\sin(2\phi)\,\rho_2,\label{rho:2}\\
\sum_{k=1}^4 q_j\sigma_2^j &=&-{\tfrac i 4}\,\sin(2\phi)\,\rho_3,\label{rho:3}\\
\sum_{k=1}^4 q_j\sigma_1^j &=&-{\tfrac i 4}\,\sin(2\phi)\,\rho_4,\label{rho:4}\\
\sum_{k=1}^4 q_j &=&+{\tfrac i 4}\,\sin(2\phi)\,\rho_5, \label{rho:5}
\een
\end{subequations}
where we introduce the following four elementary functions of five variables $\{\phi,\chi_1,\chi_2\chi_3,\chi_4\}$:
\begin{subequations}\label{rh:2,3,4,5}
\ben
\rho_2=\hskip .8truecm\left(\left(\sin(2\chi_1)\right)^2+\left(\sin(2\chi_3)\right)^2\right)&-&\hskip .7truecm\left(\left(\sin(2\chi_2)\right)^2+\left(\sin(2\chi_4)\right)^2\right),\\
\rho_3=e^{-i\phi}\left(\left(\sin(2\chi_1)\right)^2-\left(\sin(2\chi_3)\right)^2\right)&+&\,
e^{i\phi}\,\left(\left(\sin(2\chi_2)\right)^2-\left(\sin(2\chi_4)\right)^2\right),\\
\rho_4=e^{2i\phi}\,\left(\left(\sin(2\chi_1)\right)^2+\left(\sin(2\chi_3)\right)^2\right)&-&
e^{-2i\phi}\left(\left(\sin(2\chi_2)\right)^2+\left(\sin(2\chi_4)\right)^2\right),\\
\rho_5=\,e^{i\phi}\,\,\left(\left(\sin(2\chi_1)\right)^2-\left(\sin(2\chi_3)\right)^2\right)&+&\,
e^{-i\phi}\left(\left(\sin(2\chi_2)\right)^2-\left(\sin(2\chi_4)\right)^2\right).
\een
\end{subequations}
As a result, one obtains much simpler formulas for the coefficients in recurrence \eqref{frecurrence}:
\begin{subequations}\label{r134:abcdefgh}
\ben
\hskip -.6truecm r_{n-1}&=&0,\label{r134:a}\\
\hskip -.6truecm r_{n-2}&=&{\tfrac 1 {n(n-1)}}\Big(\lambda- {\tfrac i 4}\,\sin(2\phi)\,\rho_2\Big)+
4\left(1-{\tfrac 5 n}+{\tfrac 3 2}{\tfrac 1 {n-1}}\right)\cos(2\phi), \label{r134:b}\\
\hskip -.6truecm r_{n-3}&=&-{\tfrac 1 {n(n-1)}}{\tfrac i 4}\,\sin(2\phi)\,\rho_3, \label{r134:c}\\
\hskip -.6truecm r_{n-4}&=&{\tfrac 1 {n(n-1)}}\Big(\!-2\lambda \cos(2\phi)+ {\tfrac i 4}\,\sin(2\phi)\,\rho_4\Big)+
2\left(1-{\tfrac {16} n}+{\tfrac 9 {n-1}}\right)\left((\cos(2\phi)^2+1\right),\label{r134:d}\\
\hskip -.6truecm r_{n-5}&=&{\tfrac 1 {n(n-1)}}{\tfrac i 4}\,\sin(2\phi)\,\rho_5, \label{r134:e}\\
\hskip -.6truecm r_{n-6}&=&{\tfrac \lambda {n(n-1)}}-4\left(1-{\tfrac {33} n}+{\tfrac {45} 2}{\tfrac 1 {n-1}}\right)\cos(2\phi),\label{r134:f}\\
\hskip -.6truecm r_{n-7}&=&0,\label{r134:g}\\
\hskip -.6truecm r_{n-8}&=&1-{\tfrac {56} n}+{\tfrac {42} {n-1}}.\label{r134:h}
\een
\end{subequations}
If in addition to constraints \eqref{B_sigma_constraits} one imposes one more symmetric constraint, namely:
\ben
\sigma_2=0,
\la{sigma_2_0}
\een
then $a=2$, $\phi / \hskip -.2truecm \mod\!(2\pi)=\pm\pi/4,\pm 3\pi/4$,
and one obtains the simplest possible coefficients in recurrence \eqref{frecurrence}:
\begin{subequations}\label{sr134:abcdefgh}
\ben
r_{n-1}&=&0,\label{sr134:a}\\
r_{n-2}&=&{\tfrac \lambda {n(n-1)}}, \label{sr134:b}\\
r_{n-3}&=&0, \label{sr134:c}\\
r_{n-4}&=&2\left(1-{\tfrac {16} n}+{\tfrac 9 {n-1}}\right),\label{sr134:d}\\
r_{n-5}&=&0, \label{sr134:e}\\
r_{n-6}&=&{\tfrac \lambda {n(n-1)}},\label{sr134:f}\\
r_{n-7}&=&0,\label{sr134:g}\\
r_{n-8}&=&1-{\tfrac {56} n}+{\tfrac {42} {n-1}}.\label{sr134:h}
\een
\end{subequations}
Note that the additional constraint \eqref{sigma_2_0} brings us not only to a simplification
of the coefficients in recurrence \eqref{frecurrence},
but also restricts the class of the solutions of Eq. \eqref{dF} under consideration.
\section{The circular case}
The case of the general Heun's functions when the singular points of Eq. \eqref{dHeunG}
are placed on the real axis $\mathbb{R}$ by construction \cite{Smirnov} is circular one,
since $\mathbb{R}$ can be considered as a circle with an infinite radius.
The Mobius transformation preserves the circular property, see Appendix \ref{ApA},
as well as \cite{Smirnov}, where the circular case for the general Heun's equation was substantially
elaborated using the standard Eq. \eqref{dHeunG} and without any relation with the choice \eqref{z_j} in Eq. \eqref{dF}.
The value of the invariant cross-ratio \eqref{a} $a\in\mathbb{R}$ is real
for any four complex points $z_{j=1,2,3,4}$ on a circle in $\mathbb{C}$.
Then, from relation \eqref{a} follows that in the circular case the angle $\phi\in\mathbb{R}$ is real and
the singular points \eqref{z_j} lie on the unit circle with the center $z=0$.
According to the basic results of the standard analytic theory of ordinary differential equations
\cite{Golubev,CL,KF}, we obtain our key result:
{\bf Proposition 5:} In the circular case, the series \eqref{taylorF} with coefficients \eqref{r134:a}-\eqref{r134:h}
and $\phi\in\mathbb{R}$ are absolutely convergent inside the unit circle, i.e., for any $z\in \mathbb{C}$ with $|z|<1$.\,$\blacktriangleleft$
{\bf Corollary:} In the circular case, the four regular singular points $z_{j=1,2,3,4}$
of the general Heun's equation \eqref{dF} can be treated equally from inside the unit circle
using the Taylor series \eqref{taylorF}.
Next important step is to restrict Proposition 4 (iii) to the circular case.
{\bf Proposition 6:}
In the circular case, equation \eqref{dF} preserves its form if one makes the following substitutions
\ben
z\to 1/z,\quad, z_j \to 1/z_j, \quad F(z)\to F(1/z)\quad P(z)\to P(1/z),\nonumber \\
\lambda \to \lambda -{\tfrac i 4}\sin(2\phi)\rho_2,\quad q_j \to -q_j/z_j^2.
\la{inverssion_circular}
\een
This way we obtain from solutions \eqref{taylorF} new solutions or Eq. \eqref{dF} in the form of the
Laurent series expansions which are absolutely convergent for any $z\in \mathbb{C}$ with $|z|>1$.\,$\blacktriangleleft$
{\bf Corollary:} In the circular case, the four regular singular points $z_{j=1,2,3,4}$
of the general Heun's equation \eqref{dF} are mapped under inversion on the same points,
removed to the initial positions $z_{j=4,3,2,1}$, respectively. Hence,
one can treat all singular points equally from outside the unit circle
using the corresponding Laurent series, described in Proposition 6.
As a final result, in the circular case we reach a totally symmetric treatment of the singular points
$z_{j=1,2,3,4}$ in the whole complex plane $\mathbb{\tilde C}$.
\section{Some comments and concluding remarks}
In the present paper, we introduced and studied a novel representation of the general Heun's functions.
It is based on the symmetric form of the Heun's differential equation
yielded by a further development of the Felix Klein
symmetric form of the Fuchsian equations with an
arbitrary number $N\geq 4$ of regular singular points.
We derived the symmetry group of these equations and their solutions.
It turns to be a proper extension of the Mobius group.
The basic relations for the general Heun's equation with $N=4$ are derived and discussed in detail.
Special attention was paid to the nine-term recurrence relation for the coefficients of the Taylor series solutions
of the novel symmetric form of the general Heun's equation.
We described in detail the simplification of these coefficients
using the proper Mobius transformation of the singular points.
We also showed that in the circular case,
when the four singular points of the symmetric form of the gemeral Heun's equation lie on the unite circle,
the novel Taylor series solutions are absolutely convergent inside it.
After the simple inversion of the independent variable $z\to 1/z$ one obtains also the Laurent series solutions
which are absolutely convergent outside the circle with unit radius.
Hence, in the circular case one can use the new solutions for a simultaneous equal treatment of all singular points.
A more detailed study of the basic relation \eqref{HGF12},
as well as consideration of the corresponding confluent cases
of the Heun's equation will be presented elsewhere.
One can hope that this new approach will simplify
the solution of the existing basic open problems in the theory of the general Heun's functions.
The novel representation will allow also development of new effective computational methods
for calculations with the Heun's functions
which at present are still a quite problematic issue.
\vskip .7truecm
\noindent{\bf \Large Acknowledgments}
\vskip .3truecm
The author is grateful to Professor Sergei Yu. Slavyanov for his helpful comments,
help with the literature, and his kind encouragement,
to Professors Alexander Kazakov and Artur Ishkhanyan, as well as
to other participants in the Section {\it Heun's equations and their applications}
of the Conference {\it Days on Diffraction 2014}, St. Petersburg, May 26-30, 2014
for their interest in the talk on the basic results of the present article
which were reported there for the first time,
as well as to Professor Oleg Motygin for his useful remarks.
Special thanks are also to the leader of the Maple-Soft-developers, Edgardo Cheb-Terrap,
for many useful discussions during the last years
on the properties of the Maple-Heun's functions and the computational problems with them.
The author is also thankful to the leadership of BLTP, JINR, Dubna for the support and good working conditions.
This article was also supported by the Sofia University Foundation {\it Theoretical and Computational Physics and Astrophysics}
and by the Bulgarian Agency for Nuclear Regulation, the 2014 grant.
|
\section{INTRODUCTION}
The idea that binary systems of massive stars without compact objects provide suitable environments for efficient particle acceleration and subsequent nonthermal emission has been thoroughly discussed in literature \citep[e.g.,][]{Pollock1987, White1994}. Electrons and protons are thought to be accelerated at the shock fronts tracing the region where the stellar winds collide with supersonic velocities. Analytical models have predicted that the ensuing population of high-energy particles is liable to yield considerable emission of nonthermal radio waves, hard X~rays and $\gamma$~rays via various emission channels \citep[see e.g.;][]{Eichler1993, Benaglia2003, Reimer2006}.
The observational evidence concerning nonthermal photon emission from these particle-accelerating colliding-wind binaries (CBWs) was recently summarized by \cite{Becker2013} who provide a unified census of these systems. Their list of 43 confirmed or suspected binary systems together with their properties helps to pinpoint some important discrepancies between observations and model predictions, such as Fermi-LAT limits on the gamma-ray flux towards WR~140 or WR~147 \citep[see also ][]{Werner2013}, contrasted by the enigmatic properties of the $\eta$~Carinae binary system. The latter shows a comparably strong high-energy $\gamma$-ray signal that has been found to exhibit variability on orbital time scales and a peculiar two-component spectrum \citep[see \textit{Fermi} Large Area Telescope observation in][]{Reitberger2012}.
A better understanding of CWB systems can be obtained by dedicated hydrodynamical (HD) simulations. Studies such as presented in \cite{Pittard2009} have explored the highly dynamical nature of the WCR and its strong dependence on stellar and orbital parameters. The complex density, velocity and temperature structure of the colliding winds have further been used to model the thermal radio and X-ray emission in such systems \citep{Pittard2010,Pittard2010b}. Using magnetohydrodynamic (MHD) simulations \cite{Falceta2012} have recently explored the important role of the magnetic field in the wind collision region (WCR) of these systems and its impact on nonthermal radio emission. Focusing on the properties of a specific binary system, \cite{Madura2013} used three-dimensional (3D) smoothed particle hydrodynamics (SPH) to investigate the turbulent WCR structure in $\eta$~Carinae.
The present study provides simulations of high-energy nonthermal photon emission of CWBs, based on 3D distributions of high-energy particles. The latter has been obtained by combining a 3D HD description of the WCR with the solution of the transport equation of high-energy particles.
In \cite{Reitberger2014} (paper~I) we presented a method to numerically compute the spectral energy distribution (SED) of charged particles within a numerical HD model of the WCR in a binary system of two massive stars. By solving the transport equation including spatial convection, diffusion, diffusive shock acceleration (DSA) and various cooling processes for electrons and protons at every grid point and time step of the HD simulation, we derived the time-dependent 3D spatial distribution of particles at different energies.
In this work, we present the resulting nonthermal photon emission along a given line of sight on the basis of the previously derived electron and proton spectra. The interactions of high-energy particles with the stellar radiation fields and the dense wind material in the WCR give rise to several mechanisms of photon emission, including anisotropic inverse Compton (IC) scattering, relativistic bremsstrahlung and neutral pion decay. We compute the emission throughout the chosen computational domain and present results in terms of two-dimensional (2D) projection maps, spectral energy distributions and total integrated photon flux values.
As not unexpected, the emission is very sensible to the conditions in the WCR, the position of the stars and orbital orientation. We present a case study of three binary systems differing in stellar separation. Non-thermal high-energy photon emission is computed for these systems with respect to different lines of sight.
In Section 2 we provide a detailed description of our calculation of various emission processes, as well as the application of photon photon opacity effects. We investigate the ensuing photon emission in Section~3, first for two stationary stars, then including orbital motion. Sections 4 and 5 provide a summary of our findings as well as an outlook on future developments.
\section{NON-THERMAL HIGH-ENERGY PHOTON EMISSION}
\label{gamma}
In order to estimate the nonthermal high-energy photon emission of a CWB system, we consider the three principal continuum emission processes for energies $E\gtrsim$10 MeV. These are IC scattering of the photons in the dense stellar radiation field to high energies, relativistic bremsstrahlung by the interaction of the energetic electrons with the ions in the wind, and the decay of neutral pions that are produced in hadronic nucleon nucleon collisions.
The computation of all three processes is based on a 3D distribution of high-energy electrons (IC scattering, bremsstrahlung) and protons (neutral pion decay) which we obtain by the method described in paper I. There, all relevant details concerning the hydrodynamics, the transport equation and the numerical implementation are provided.
In principle, the distribution of high-energy electrons can also be used to compute nonthermal low-energy emission such as synchrotron emission. This is highly relevant in modeling specific binary systems, where such observations of a nonthermal signal at radio wavelengths exists \citep[see e.g;][ for the case of WR140]{Williams1994}. A comparison of measured and modelled synchrotron emission yields constraints concerning the magnetic field in conjunction with the electron injection fraction in the WCR and hence the level of IC emission. As it is not the aim of the present study to investigate a specific binary system with fixed astronomical properties, we restrict ourselves to the high-energy part of the nonthermal emission spectrum considering the three emission channels as detailed below.
\begin{enumerate}
\item Anisotropic IC scattering of high-energy electrons on the photons in the radiation field of both stars: This process depends on the local energy density of radiation, the spatial and energetic distribution of electrons and the scattering angle, being the angle between the line of sight to the observer and the direction of the incoming stellar photons. The stars are approximated as monochromatic point sources. The high-energy electrons are considered to have an isotropic distribution function.
\item Relativistic bremsstrahlung of high-energy electrons passing through the wind plasma. This process depends on the number density of the wind plasma and the spatial and energetic distribution of high-energy electrons. We assume a typical interstellar medium (ISM) metallicity of 90\% H, 10\% He.
\item The decay of neutral pions produced in collisions of high-energy protons with nucleons of the wind plasma. This process depends on the number density of the wind plasma and the spatial and energetic distribution of high-energy protons.
\end{enumerate}
We compute the above nonthermal emission components for each numerical grid cell of the computational domain. For the calculation of IC scattering, additional information about the orientation of the system with respect to the observer (in terms of inclination $i$ and argument of periastron of the binary system $\Phi$) must be provided. Bremsstrahlung and neutral pion decay give isotropically distributed radiation.
In a second step we use the resulting 3D distribution of nonthermal photon emission per energy and volume and the information of direction and distance to the observer, in order to compute
\begin{enumerate}
\item 2D projection maps of the nonthermal emission for a given energy interval and line of sight,
\item the SED of the various nonthermal emission components integrated over the whole emission region, and
\item the total integrated flux for various energy bands.
\end{enumerate}
Additionally, we consider effects of photon photon opacity due to the dense radiation fields of the stars in the CWB system. Details are given in Section \ref{opac}. Other possible absorption effects (e.g., via interaction with the ISM) are, presently, not considered since they also depend on the specific system.
In the following, we give a detailed account of how we calculate the individual emission components and derive their 2D projections, SEDs and integrated flux values.
\subsection{Inverse Compton Scattering}
\label{ICs}
The computation of the emission by IC scattering requires information on the high-energy electrons, as well as on the radiation fields involved in the scattering process. For consistency with paper I, we use the approximation of monochromatic radiation from pointlike stars. The differential target photon field of star $j$ can then be expressed as
\begin{equation}
\label{ndelta}
\frac{dn_{\mathrm{ph},j}}{dE_\mathrm{ph}}(\vect{r},E_\mathrm{ph} ,\mu)=n_{0,j}(\vect{r})\delta(E_\mathrm{ph}-E_{T,j})\delta(\mu-\mu_{\mathrm{sc},j})
\end{equation}
where $E_{T,j}=2.7k_BT_{\ast,j}$ is the energy of a single photon from star $j$ with surface temperature $T_{\ast,j}$. The cosine of the scattering angle $\mu_\mathrm{sc}=\cos{\theta_\mathrm{sc}}$ is determined via the inner product of the unit vectors pointing from the position of the cell at $\vect{r}$ toward the observer and toward star $j$. $n_{0,j}(\vect{r})$ is the local photon number density at the position of the cell $\vect{r}$ given by
\begin{equation}
n_{0,j}(\vect{r})=\frac{u_{\mathrm{ph},j}}{E_{T,j}}=\frac{L_{\ast,j}}{4\pi r_j^2cE_{T,j}}
\end{equation} with $r_j$ the distance to the star and $L_{\ast,j}$ its stellar luminosity.
The local differential number density of the scattering electrons is directly obtained from the output of the simulations described in paper~I.
Differential number densities of target photons and high-energy electrons enter into the expression for the IC photon production rate $\frac{d\dot{n}_j(\vect{r},E_\gamma)}{dE_\gamma}$ at energy $E_\gamma$. The complete formula in the context of CWB systems for the case of an isotropic particle distribution is given by \cite{Reimer2006}.
It involves integration over all target photon energies, all target photon incoming directions $\mu$ and all electron energies. By applying the approximations of monochromacity and pointlike stars, only the integral over the energy of the electrons remains. Specific care has to be taken in determining the lower integration limit which marks the lowest energy at which electrons are liable to scatter photons to an energy $E_\gamma$ with regard to the scattering angle $\mu_{\mathrm{sc},j}$. The upper limit is determined by the maximum energy of the electrons. The integration is performed numerically.
The IC flux from a single cell with volume $V_\mathrm{cell}$ reaching an observer along the line of sight at distance $d_l$ (neglecting absorption effects) then is
\begin{equation}
\frac{dF_\mathrm{IC}}{dE_\gamma}(\vect{r},E_\gamma)=\frac{V_\mathrm{cell}}{4\pi d_l^2}
\sum_{j=1}^2\frac{d\dot{n}_j(\vect{r},E_\gamma)}{dE_\gamma}
\end{equation}
where the sum accounts for the contribution of both stars. The integrated IC flux of the total emission region can be determined by a sum over all grid cells and an integration over the photon energy $E_\gamma$.
\subsection{Relativistic Bremsstrahlung}
According to \citet{Blumenthal}, the bremsstrahlung spectrum emitted by a single electron of energy $E_e$ passing through a medium containing various species of ions with number densities $n_i$ is
\begin{equation}
\label{single}
\frac{d\dot{N}_\gamma}{dE_\gamma}(E_e)=c\sum_in_iZ_i^2\frac{d\sigma_i}{dE_\gamma}(E_e)
\end{equation}
where the sum is over all elements relevant for the scattering process.
The differential cross-section $\frac{d\sigma_s}{dE_\gamma}$ is greatly simplified if we assume that all contributing species are fully ionized. This is approximately true in the high temperature plasma inside the WCR. With this assumption (the case of ``weak shielding'') the differential cross-section can be expressed as
\begin{equation}
\frac{d\sigma_i}{dE_\gamma}(E_e)=\alpha r_0^2\frac{\frac{4}{3}E_e^2-\frac{4}{3}E_eE_\gamma+E_\gamma^2}{E_\gamma E_e^2}\phi_\mathrm{weak}
\end{equation}
with
\begin{equation}
\phi_\mathrm{weak}=\left.
\begin{cases}
4\Big[\ln\Big(\frac{2E_e(E_e-E_\gamma)}{E_\gamma m_ec^2}\Big)-\frac{1}{2}\Big],\qquad &\mbox{for }E_\gamma < E_e\\
4\Big[\ln(4E_\gamma)-\frac{1}{2}\Big],\qquad &\mbox{for }E_\gamma \gtrsim E_e\\
\end{cases}
\right.
\end{equation}
We further make the assumption of only dealing with hydrogen and helium (90\% H, 10\% He), allowing to rewrite Equation (\ref{single}) as
\begin{equation}
\frac{d\dot{N}_\gamma}{dE_\gamma}=c(0.9\times1^2+0.1\times2^2)n_\mathrm{H}\frac{d\sigma}{dE_\gamma}
\end{equation}
where the number density $n_\mathrm{H}$ of particles in the wind plasma is obtained via $n_H = \frac{\rho}{m_H}$ from the output of the hydrodynamic simulations described in paper~I.
Computing the emission for isotropically distributed electrons $n_e(E_e)$ requires an integration over their energy spectrum.
\begin{equation}
\frac{d\dot{n}_\gamma}{dE_\gamma}=\int dE_e\frac{dn_e}{dE_e}(E_e)\frac{d\dot{N}_\gamma}{dE_\gamma}=1.3cn_\mathrm{H}\int dE_e\frac{dn_e}{dE_e}(E_e)\frac{d\sigma}{dE_\gamma}
\end{equation}
The integral is solved with the same numerical scheme as discussed in Section \ref{ICs}.
To obtain the resulting flux emitted from a single computational cell at a distance $d_L$, we multiply by the cell volume $V_\mathrm{cell}$ and divide by the respective spherical distance surface.
\begin{eqnarray}
\frac{dF_\mathrm{brems}}{dE_\gamma}=\frac{V_\mathrm{cell}}{4\pi d_l^2}\frac{d\dot{n}_\gamma}{dE_\gamma}
\end{eqnarray}
\subsection{Neutral Pion Decay}
In computing the $\gamma$-ray emission due to the decay of neutral pions, we follow \cite{Kelner2006}
who use a $\delta$-functional approximation for the cross-section $\sigma_{pp}^\pi(E_\pi,E_p)$ with $E_p$ being the total energy of the incident proton (the other one is assumed to be at rest). $E_\pi$ is the energy of the produced pion. For a distribution of protons $\frac{dn_{p}}{dE_{p}}(E_{p})$ the omnidirectional differential neutral pion spectrum can then be expressed as
\begin{align}
\label{pin}
\frac{dn_{\pi^0}}{dE_\pi}(E_\pi)&=
cn_\mathrm{H}\sigma_{pp}^\pi(m_pc^2+\frac{E_\pi}{K_\pi})\frac{dn_{p}}{dE_{p}}(m_pc^2+\frac{E_\pi}{K_\pi})
\end{align}
where $K_\pi$ is the mean fraction of the kinetic energy of the protons $E_\mathrm{kin}$ that is transferred to the secondary neutral pions.
According to \cite{Aharonian2000} the parameterization $K_\pi\approx0.17$ yields results of high accuracy in a broad region from GeV to TeV energies. It can be applied down to the threshold energy of neutral pion production. The total cross-section $ \sigma_{pp}^\pi(E_p)$ is approximated \citep[following][]{Kelner2006} as
\begin{equation}
\sigma_{pp}^\pi(E_p)=(34.3+1.88L+0.25L^2)\Bigg[1-\Big(\frac{E_\mathrm{th}}{E_p}\Big)^4\Bigg]^2\times10^{-31}\;\mathrm{m}^2
\end{equation}
with $L=\ln(E_p/1\mathrm{TeV})$ and $E_\mathrm{th}=m_p+2m_\pi+m_\pi^2/2m_p$.
Equation (\ref{pin}) allows to directly convert the proton spectra (obtained as described in paper I) to the resulting pion spectra. Via interpolation, $\frac{dn_{\pi^0}}{dE_\pi}(E_\pi)$ can be obtained for any given energy $E_\pi$.
The omnidirectional differential photon production rate can then be computed via the integral
\begin{equation}
\frac{d\dot{n}}{dE_\gamma}(E_\gamma)=2\int_{E_\gamma+\frac{m_\pi^2c^4}{4E_\gamma}}^{\infty}dE_\pi\frac{\frac{dn_{\pi^0}}{dE_\pi}(E_\pi)}{\sqrt{E_\pi^2-m_\pi^2c^4}}
\end{equation}
As in case of the bremsstrahlung component, the flux reaching an observer at a distance $d_l$ then is
\begin{equation}
\frac{dF_{\pi^0\mathrm{decay}}}{dE_\gamma}=\frac{V_\mathrm{cell}}{4\pi d_l^2}\frac{d\dot{n}}{dE_\gamma}.
\end{equation}
\subsection{Photon-Photon Opacity in the Stellar Radiation Fields}
\label{opac}
Due to the intense stellar radiation field close to the stars, the emitted nonthermal photon flux above $\sim$~100 GeV may be attenuated significantly via the electron pair production process
\begin{equation}
\gamma+\gamma^\prime\rightarrow e^++e^-.
\end{equation}
This attenuation can be expressed as
\begin{equation}
\frac{dF}{dE_\gamma}(\vect{r},E_\gamma)\longrightarrow\frac{dF}{dE_\gamma}(\vect{r},E_\gamma) \exp\Big(-\tau_{\gamma\gamma}(\vect{r},E_\gamma)\Big)
\end{equation}
with $\tau_{\gamma\gamma}$ being the optical depth due to photon-photon pair production.
For each position within the emission region, the computation of $\tau_{\gamma\gamma}$ requires an integration along the path $\ell$ toward the observer. The optical depth is proportional to the cross-section $\sigma_{\gamma\gamma}$ as well as to the differential photon number density $\frac{dn_{\mathrm{ph}}}{dE_\mathrm{ph}}(E_\mathrm{ph} ,\mu_j)$ of the stellar radiation field. The general expression can be written as
\begin{align}
&\tau_{\gamma\gamma}(\vect{r},E_\gamma)=\nonumber \\
\int d\ell\int dE_\mathrm{ph}\int d\mu \frac{dn_{\mathrm{ph}}}{dE_\mathrm{ph}}&(\vect{x}(\ell),E_\mathrm{ph} ,\mu)\sigma_{\gamma\gamma}(E_\mathrm{ph},E_\gamma,\mu)(1-\mu)\nonumber \\
\;
\end{align}
where the integrals run along the path $\ell$, over the energy of the photons in the stellar radiation fields $E_\mathrm{ph}$, and over the stellar surface described by the cosine of the angle between the two interacting photons $\mu$.
Analogous to the computation of the IC emission in Section \ref{ICs}, we assume the stars to be pointlike and their radiation field to be monochromatic. Thus, the differential photon number density can be expressed as in Equation (\ref{ndelta}) and
\begin{align}
&\tau_{\gamma\gamma}(\vect{r},E_\gamma)=\nonumber\\
\int{d\ell}\sum_j^\mathrm{stars}
n_{0,j}(\vect{x}(\ell))& \sigma_{\gamma\gamma}(E_\mathrm{ph},E_{T,j},\mu_j(\vect{x}(\ell)))(1-\mu_j(\vect{x}(\ell)))
\end{align}
where $\mu_j$ is the angle between the line of sight (LOS) and the direction toward the star j.
The integral along path $\ell$ is solved numerically.
According to \cite{Gould1967} the cross-section of the process $\gamma+\gamma^\prime\rightarrow e^++e^-$ is
\begin{equation}
\sigma_{\gamma\gamma}(E_T,E_\gamma,\mu) = \frac{1}{2}\pi r_0^2(1-\beta^2)\Big[(3-\beta^4)\ln\frac{1+\beta}{1-\beta}-2\beta(2-\beta^2)\Big]
\end{equation}
with
\begin{equation}
\beta=\sqrt{1-s^{-1}}\quad\mathrm{and}\quad s=\frac{E_T E_\gamma}{2m_e^2c^4}(1-\mu)
\end{equation}
The threshold for photon photon absorption is at $E_TE_\gamma\geq(m_ec^2)^2$ (in the most favourable case of $\mu=-1$). For a star of temperature $T = 3\cdot10^4 K$, opacity effects do occur for $E_\gamma\gtrsim$ 40 GeV.
In order to test the validity of the monochromatic approximation, we also computed the absorption assuming a black-body spectrum for both stars. This requires an additional integration over the energies of the stellar photons. We approximate the spectrum by 20 energy bins (equally spaced on a logarithmic scale) in the interval from 0.1 to 100 eV and integrate numerically.
A comparison between the two cases is given in Figure \ref{opacity} where we show the optical depth $\tau$ for a single computational cell as a function of $E_\gamma$ for either the WR star, the B star or both stars (a), as well as the impact of the opacity effects on the SED emitted by a representative fraction of the computational domain (b). The monochromatic approximation is found to overestimate the optical depth near its maximum and underestimates it for lower and higher energies. However, the differences found in the SED are sufficiently small (see Figure \ref{opacity} b)) to warrant the use of the monochromatic approximation, keeping in mind that it has significantly lower computational cost, as the black-body description of the radiation field (with 20 energy bins) increases the overall computation time by a factor of 20.
\section{RESULTS}
\label{results}
In the following, we will investigate three different binary systems with different stellar separations regarding their wind plasma properties, their distribution of charged particles and their ensuing nonthermal high-energy emission. The latter is explored using 2D projection maps, SEDs and integrated flux values for various energy bands. We also present the results for an additional case including effects of orbital motion.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,370)
\put(15,15){
\includegraphics[width=\textwidth]{./f1a}
}
\put(200,0){\footnotesize{log( E in GeV)}}
\put(0,160){\rotatebox{90}{\footnotesize{log($\tau$)}}}
\put(450,60){a)}
\end{picture}
\end{subfigure}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,370)
\put(15,15){
\includegraphics[width=0.96\textwidth]{./f1b}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV cm$^{-3}$ )}}}
\put(440,60){b)}
\end{picture}
\end{subfigure}
\caption{a) Optical depth $\tau$ from photon photon absorption for a single grid cell as a function of the energy of the high-energy photon scattering with the mono-chromatic (solid) or black-body type (dashed) radiation field of the B star (black), the WR star (gray) and both stars (red). \\
b) Photon spectra emitted by a significant fraction of the computational domain for IC scattering (black) and neutral pion decay (red) for a viewing angle where opacity effects are maximal ($i=90^\circ$, $\Phi=0^\circ$). The spectra have been computing using mono-chromatic (solid) and black-body type (dashed) radiation fields. Stellar and orbital parameters are the same as for case C in Section \ref{results}. \label{opacity}}
\end{figure*}
\subsection{HD Model Parameters}
To allow for quick comparison, we consider three binary systems (Wolf--Rayet (WR) + B star) that differ only in their stellar separation $d$ for which we chose 720 (case A), 1440 (case B) and 2880 (case C) R$\odot$. All stellar and stellar wind parameters (listed in Table \ref{params}) are the same as in the parameter studies in \cite{Reimer2006} and in paper I. The important parameter $\eta = \frac{\dot{M}_\mathrm{B}v_{\infty,\mathrm{B}}}{\dot{M}_\mathrm{WR}v_{\infty,\mathrm{WR}}}$ which indicates the point of ram pressure balance between the two stars has a value of 0.1 for the chosen parameters.
In a first approach, we neglect orbital motion and study the three systems in a converged state after all traces of the initial conditions have vanished.
For all three cases we chose the identical computational domain in the shape of a cube with side length 4000 $R_\odot$. The two stars are located on the $x$ axis.
Figure \ref{orient} schematically depicts the computational domain containing the two stars. It also shows the line of centers and various arrows indicating the different viewing angles in respect to which we investigate the nonthermal high-energy emission of the system.
In Section \ref{orbital} we repeat the analysis of case A, now including effects of orbital motion. Again, the two stars are located initially on the $x$ axis and move on a Keplerian orbit in the $x$--$y$ plane.
\subsection{Properties of the Wind Plasma}
\label{hydrod}
As expected, the three different stellar separations show considerable contrasts in terms of properties and structure of the WCR. In Figure \ref{hydros} we show density, absolute velocity and temperature in the $x$-$y$ plane at $z=0$ for cases A, B, and C.
\begin{table}[t]
\begin{center}
\textbf{Stellar and stellar-wind parameters}\\
\end{center}
\centering
\begin{tabular}{r ||c |c |c| l}
& \textbf{B} & \textbf{WR} & \textit{unit}\\ \hline \hline
$M_\ast$ &30 & 30 & M$_{\odot}$\\
$R_\ast$ &20 & 10 & R$_{\odot}$\\
$T_\ast$ &23000 & 40000 & K\\
$L_\ast$ & $10^{5}$ & $2.3\times 10^{5}$ & L$_{\odot}$\\
$\dot{M}$ &$10^{-6}$ & $10^{-5}$ & M$_{\odot}$ yr\textsuperscript{-1}\\
$v_\infty$ & 4000 & 4000 & km s$^{-1}$\\
\end{tabular}
\caption{Stellar and stellar-wind parameters of a typical binary system as considered in this work.\label{params}}
\end{table}
\begin{figure*}
\begin{center}
\includegraphics[trim=0cm 1cm 0cm 0cm, clip=true,width=0.7\textwidth]{./f2}
\end{center}
\caption{Schematic view of the two stars within the computational domain. The line of centers is represented by the horizontal dashed orange line. The various viewing angles (lines-of-sight) in respect to which the emission is computed are indicated as well.
\label{orient}}
\end{figure*}
\paragraph{Shape of WCR}
The thickness of the WCR at the apex widens considerably with increasing stellar separation. Along the line of centers, its width is $\sim$100, $\sim$170 and $\sim$310 $R_\odot$ for cases A, B, and C, respectively. Analogously, the distance between the WCR and the B star (which is generally closer because of $\eta=0.1$) increases. The star is located $\sim$140, $\sim$290 and $\sim$600 $R_\odot$ from the edge, and $\sim$190, $\sim$380 and $\sim$760 $R_\odot$ from the center of the WCR for cases A, B, and C. Positions and curvatures of the WCR in all models agree well with the analytical approximation by \cite{Stevens1992} in which the distance of the apex of the WCR from the B star is determined by $d_B=\frac{\sqrt{\eta}}{1+\sqrt{\eta}}d$ and the curvature of the downstream WCR is approximated by solving the differential equation
\begin{equation}
\frac{dy}{dx}=\frac{\left(\sqrt{\eta}d_\mathrm{WR}^2+d_\mathrm{B}^2\right)y}{\sqrt{\eta}d_\mathrm{WR}^2x+d_\mathrm{B}^2(x-d)}.
\end{equation}
The resulting curves are indicated in the second row of Figure \ref{hydros}.
Contrasting the effect of increasing WCR thickness with increasing stellar separation, the opening angle (or curvature) of the collision region significantly decreases. Among the three systems under investigation, case A has the narrowest WCR at the apex, but the widest WCR close to the edge of the computational domain.
\paragraph{Density}
The maximum number density values within the WCR are 33.1, 7.6 and 1.8 $\times10^{13}$ m$^{-3}$ for cases A, B, and C. The approximate inverse proportionality to $d^2$ is understandable from the fact that, for a strong shock, $\rho_\mathrm{postshock}\sim4\rho_\mathrm{preshock}$ and $\rho_\mathrm{preshock}=\frac{\dot{M}}{4\pi r^2 v_\mathrm{wind}}$. The density contrast of the unshocked winds manifests itself in comparable contrasts within the WCR where the side toward the WR star shows considerable higher density values of the shocked wind plasma.
\paragraph{Temperature}
The maximum temperature reached at the apex of the WCR shows only little variation. It is 2.14, 2.17 and 2.21 $\times$10$^8$ K for cases A, B, and C. In contrast to that the temperature gradient is far more affected by the change of stellar separation. The closer the stars, the larger the drop in temperature with increasing distance from the apex of the WCR. The minimum temperature of the shocked wind plasma reached at the edge of the computational domain varies accordingly, with values of 6.9, 20.9, 61.7 $\times10^6$ K for cases A, B, and C.
\paragraph{Absolute velocity}
In all three systems, the wind is slowed down to near zero velocity at the apex. We observe a considerable difference concerning the size of the region in which the velocity remains comparatively low. The distance downstream from the apex at which the shocked wind plasma is efficiently re-accelerated is significantly shorter for the case of smallest stellar separation.
Another notable difference is the maximum velocity reached within the computational domain in the unshocked wind outside the WCR. For case A, the B wind close to the WCR at the edge of the domain reaches values up to 5100 km s$^{-1}$ in contrast to merely 4700 and 4300 km s$^{-1}$ for case B and C. The reason for these velocity values which lie all above the B wind's terminal velocity of $v_\infty=$4000 km s$^{-1}$ is the radiative wind acceleration mechanism which allows additional acceleration of the unshocked B star's wind due to the radiation of the WR star. The effect increases with increasing proximity of the WR star, thus it is greatest for case A.
\paragraph{Other properties}
The local magnetic field strength and the energy density of radiation are two important properties that critically determine the energy distribution of the particles accelerated at the shocks by influencing energy loss by synchrotron emission and IC cooling. In estimating the magnetic field strength in dependence of the distance of the stars we rely on the approximations following \cite{Usov1992} (details also in paper I). The respective values at the center of the WCR for cases A, B and C are $\sim$0.64 G, $\sim$0.16 G, and $\sim$0.04 G. \\
The energy density of radiation is proportional to the luminosity of the stars and the inverse square of their distance. Its values at the center of the WCR are $\sim$0.77, $\sim$0.19, and $\sim$0.05 J~m$^{-3}$ for cases A, B and C, respectively.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3a}}
\put(120,300){CASE A}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3b}}
\put(120,300){CASE B}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3c}}
\put(120,300){CASE C}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{95\unitlength}
\begin{picture}(95,290)
\put(20,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.3cm, clip=true,height=280\unitlength]{./f3d}}
\put(-5,100){\rotatebox{90}{\footnotesize{log(m$^{-3}$)}}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3e}}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3f}}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3g}}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{85\unitlength}
\begin{picture}(95,290)
\put(20,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.3cm, clip=true,height=280\unitlength]{./f3h}}
\put(-5,100){\rotatebox{90}{\footnotesize{log(K)}}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3i}}
\put(126.7,20){{\color{white}\line(1,0){35.625}}}
\put(126.7,17){{\color{white}\line(0,1){6}}}
\put(162.325,17){{\color{white}\line(0,1){6}}}
\put(126.7,5){\scriptsize{{\color{white}500R$_\odot$}}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3j}}
\put(126.7,20){{\color{white}\line(1,0){35.625}}}
\put(126.7,17){{\color{white}\line(0,1){6}}}
\put(162.325,17){{\color{white}\line(0,1){6}}}
\put(126.7,5){\scriptsize{{\color{white}500R$_\odot$}}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f3k}}
\put(126.7,20){{\color{white}\line(1,0){35.625}}}
\put(126.7,17){{\color{white}\line(0,1){6}}}
\put(162.325,17){{\color{white}\line(0,1){6}}}
\put(126.7,5){\scriptsize{{\color{white}500R$_\odot$}}}
\end{picture}
\end{subfigure}
\begin{subfigure}{85\unitlength}
\begin{picture}(95,290)
\put(20,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.4cm, clip=true,height=280\unitlength]{./f3l}}
\put(-5,100){\rotatebox{90}{\footnotesize{km s$^{-1}$}}}
\end{picture}
\end{subfigure}
\caption{HD quantities in the converged state for three different stellar separations. The plots show the $x$--$y$ plane of a $512\times512\times512$ simulation at z=0.
The WR star is located on the right-hand side of the WCR, and the B star is on its left-hand side. The depicted quantities are particle density (first row) in log(m$^{-3}$), temperature in log(K) (second row), and absolute velocity in km s$^{-1}$ (third row). In the second row, we also plot the analytical approximation for the WRC position following by \cite{Stevens1992} (white solid line).
\label{hydros}}
\end{figure*}
\subsection{High-Energy Particles}
\label{partd}
Along with the hydrodynamic variables of density, velocity and temperature, we consider 200 advected scalar fields containing the number densities of high-energy electrons and protons at different energies accelerated at the shock fronts of the WCR. A transport equation for both electrons and protons in energy space is solved after every hydrodynamical time step. The injection rate of electrons and protons at the shock front is treated proportional to the number density of particles in the wind. An important free parameter is the electron-proton injection ratio for which we chose a typical value of 10$^{-2}$ (for details, see paper I).
Figure \ref{els} depicts the spatial distribution of electrons and protons at two different energies for all three cases A, B, and C.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}[l]{900\unitlength}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4a}}
\put(120,300){CASE A}
\put(180,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4b}}
\put(120,300){CASE B}
\put(180,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4c}}
\put(120,300){CASE C}
\put(180,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4d}}
\put(180,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4e}}
\put(180,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\put(110,-18){electrons}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4f}}
\put(180,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\end{subfigure}
\begin{subfigure}{85\unitlength}
\begin{picture}(85,580)
\put(5,0){\includegraphics[trim=15.9cm 1.4cm 2.3cm 1.5cm, clip=true,height=580\unitlength]{./f4g}}
\put(-20,180){\rotatebox{90}{log(MeV$^{-1}$m$^{-3}$)}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}[l]{900\unitlength}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,310)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4h}}
\put(180,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,310)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4i}}
\put(180,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,310)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4j}}
\put(180,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4k}}
\put(180,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4l}}
\put(180,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\put(110,-18){protons}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f4m}}
\put(180,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\end{subfigure}
\begin{subfigure}{85\unitlength}
\begin{picture}(85,580)
\put(5,0){\includegraphics[trim=15.9cm 1.4cm 2.3cm 1.5cm, clip=true,height=580\unitlength]{./f4n}}
\put(-20,180){\rotatebox{90}{log(MeV$^{-1}$m$^{-3}$)}}
\end{picture}
\end{subfigure}
\caption{Differential number density of electrons (upper two rows) and protons (lower two rows) in MeV$^{-1}$m$^{-3}$ for different values of kinetic particle energy and different stellar separations. The colour maps show the $x$--$y$ plane of a 512$\times$512$\times$\textbf{512} simulation at z=0.
\label{els}}
\end{figure*}
At energies of $\sim$10 MeV, electrons as well as protons reach higher densities for smaller stellar separation. This is due to the proportionality of the electron and proton injection rate to the wind density. The most important difference between the three cases lies in the maximum energy attained by electrons. Whereas electrons of 100 GeV are distributed throughout most of the WCR for case C, they are confined to large stellar distances for case B or even vanish completely for case A. This has two reasons. For smaller stellar separations, the winds collide long before reaching their terminal velocity. This produces lower shock velocities $V_\mathrm{shock}$. As the diffusive shock acceleration term depends on $V_\mathrm{shock}^2$, the maximum attainable energies are significantly reduced for small stellar separations. In addition, close proximity to the stars leads to much higher energy densities of radiation and magnetic field. Together with higher wind plasma densities in the WCR, this greatly increases IC losses, synchrotron losses and bremsstrahlung losses.
As protons are not affected by these loss terms, energies of $\sim$100 GeV are reached for all three cases. However -- mainly due to higher downstream velocities -- they cannot efficiently propagate into the inner downstream regions of the WCR for low stellar separations. This leads to a notable drop in number density of high-energy protons from the shock fronts toward the inner WCR which is most notable for case A. For case C, low downstream velocities allow the protons to fill the WCR homogeneously.
\subsection{The 3D Opacity Structure}
We illustrate the spatial variation of the optical depth due to photon photon absorption caused by the radiation fields of the two stars in Figure \ref{tau} which depicts 3D illustrations of $\log_{10}(\tau_{\gamma\gamma})$ for an incident photon energy of $\sim$200 GeV. All four images represent case B for different orientations of the binary system relative to the observer as indicated in the figure. They cover the entire computational domain.
Each point shows the total value of $\tau$ integrated from the point itself along the line of sight toward the observer. Thus, the region that lies precisely behind a star with respect to the line of sight has a very high opacity. Its emitted flux is reduced to very low levels according to $\frac{dF}{dE}\rightarrow\frac{dF}{dE}e^{-\tau}$.
Photon photon absorption is most effective if the line of centers and the line of sight are aligned such that the opacity contributions of both stars add up. In Figure \ref{tau} b) we see that the high-$\tau$ region notably widens at the $x$-position of the B star.
We find that the photon photon absorption in the radiation field of the stars remains inefficient
below incident photon energies of $\sim$100 GeV. The value of 200 GeV -- chosen for the illustration -- is close to the energy where absorption is at maximum for the given binary system.
\subsection{Photon Emission}
Applying the formalism described in Section \ref{gamma}, we can now compute 2D projection maps, SEDs and integrated flux values of the nonthermal photon emission of the three discussed binary systems. Here, we start with the former of those.
\subsubsection{2D Projection Maps}
Figure \ref{em_0_0} illustrates how the integrated flux above 100 MeV from the individual nonthermal emission channels appears in the face-on configuration ($i=$0$^\circ$,$\Phi=$0$^\circ$) for an ideal observatory at 1 kpc distance with infinite angular resolution and without absorption in the ISM. Several features noted previously in Sections \ref{hydrod} and \ref{partd} concerning the wind structure and the high-energy particle distribution have an impact on the resulting photon emission maps.
The dominant emission channel shifts from IC-scattering to neutral pion decay as the stellar separation decreases. This is because of the lack of high-energy electrons when the WCR is close to the stars (case A).
Figures \ref{em_varic} to \ref{em_varp0} provide further insight by illustrating the 2D projection maps of the individual nonthermal high-energy emission channels for different orientations of the system relative to the observer. The scaling of the colour bars has been kept identical in order to allow quick comparison.
The first and second rows show a case where the line of centers and the line of sight are aligned. Note the different sizes of the occultation region due to the stellar disk for $i=$90$^\circ$,$\Phi=$0$^\circ$ (smaller WR star in front of B star) and $i=$90$^\circ$,$\Phi=$180$^\circ$ (larger B star in front of WR star).
Studying the first and second row clearly reveals the anisotropic character of the IC-component. Both orientations look identical (except for the occultation by the stellar disks) for photons from bremsstrahlung (Figure \ref{em_varbr}) and neutral pion decay (Figure \ref{em_varp0}). This is rather different for photons from IC-scattering where the relative angle of stellar positions and line of sight cause notable contrasts between the two orientations due to a different scattering angle (see first and second row in Figure \ref{em_varic}).
The general shape and extent of the emission region becomes especially clear in the projection for an observer at $i=$45$^\circ$,$\Phi=$45$^\circ$ as it is shown in the third row of Figures \ref{em_varic} to \ref{em_varp0}.
One of the most striking features is the missing photon emission from IC scattering close to the apex of case A, which is due to the lack of high-energy electrons in that region. Again, the dominance of the IC-component for case C in contrast to the dominance of the pion decay component for case A becomes apparent.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{490\unitlength}
\includegraphics[trim=0cm 0cm 0cm 0cm, clip=true,width=490\unitlength]{./f5a}
\end{subfigure}
\begin{subfigure}{490\unitlength}
\includegraphics[trim=0cm 0cm 0cm 0cm, clip=true,width=490\unitlength]{./f5b}
\end{subfigure}
\begin{subfigure}{490\unitlength}
\includegraphics[trim=0cm 0cm 0cm 0cm, clip=true,width=490\unitlength]{./f5c}
\end{subfigure}
\begin{subfigure}{490\unitlength}
\includegraphics[trim=0cm 0cm 0cm 0cm, clip=true,width=490\unitlength]{./f5d}
\end{subfigure}
\begin{subfigure}{990\unitlength}
\centering
\includegraphics[trim=0cm 7cm 0cm 11cm, clip=true,width=990\unitlength]{./f5e}
\end{subfigure}
\caption{Photon-photon opacity log($\tau$) at $\sim$200 GeV integrated for each position along the line of sight through the computational domain for different orientations. a) $i=$0$^\circ$, $\Phi=$0$^\circ$, b) $i=$90$^\circ$, $\Phi=$0$^\circ$, c) $i=$90$^\circ$, $\Phi=$90$^\circ$, d) $i=$45$^\circ$, $\Phi=$45$^\circ$. The box size is $4000\times4000\times4000$ R$_\odot$. A segment is cut out to allow a better view of the structure. The two stars are indicated in red. The line of centers is indicated in orange and yellow, the line of sight in black and gray.
\label{tau} }
\end{figure*}
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6a}}
\put(120,330){CASE A}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength ,]{./f6b}}
\put(120,330){CASE B}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6c}}
\put(120,330){CASE C}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6d}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6e}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6f}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6g}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6h}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f6i}}
\end{picture}
\end{subfigure}\\
\begin{center}
\begin{subfigure}{500\unitlength}
\begin{picture}(500,0)
\put(0,-20){\includegraphics[trim=2cm 2cm 1cm 24cm, clip=true,width=500\unitlength]{./f6j}}
\put(200,25){\footnotesize{ph m$^{-2}$ s$^{-1}$}}
\end{picture}
\end{subfigure}
\end{center}
\caption{Photon flux above 100 MeV at 1 kpc distance with face-on orientation $i=0^\circ$, $\Phi=0^\circ$ for IC-emission (first row), bremsstrahlung (second row) and neutral pion decay (third row).
\label{em_0_0}}
\end{figure*}
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7a}}
\put(120,330){CASE A}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength ,]{./f7b}}
\put(120,330){CASE B}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7c}}
\put(120,330){CASE C}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7d}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7e}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7f}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7g}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7h}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f7i}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{center}
\begin{subfigure}{500\unitlength}
\begin{picture}(500,0)
\put(0,-20){\includegraphics[trim=2cm 2cm 1cm 24cm, clip=true,width=500\unitlength]{./f7j}}
\put(200,25){\footnotesize{ph m$^{-2}$ s$^{-1}$}}
\end{picture}
\end{subfigure}
\end{center}
\caption{Photon flux above 100 MeV at 1 kpc distance for IC-emission with different orientations. first row: $i=$90$^\circ$, $\Phi=$0$^\circ$ (edge-on, along line of centers), second row: $i=$90$^\circ$, $\Phi=$180$^\circ$ (edge-on, along line of centers, opposite of second row),
third row: $i=$45$^\circ$, $\Phi=$45$^\circ$. The spatial dimensions are the same as in Figure \ref{em_0_0}.
\label{em_varic}}
\end{figure*}
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8a}}
\put(120,330){CASE A}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength ,]{./f8b}}
\put(120,330){CASE B}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8c}}
\put(120,330){CASE C}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8d}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8e}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8f}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8g}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8h}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f8i}}
\end{picture}
\end{subfigure}\\
\begin{center}
\begin{subfigure}{500\unitlength}
\begin{picture}(500,0)
\put(0,-20){\includegraphics[trim=2cm 2cm 1cm 24cm, clip=true,width=500\unitlength]{./f8j}}
\put(200,25){\footnotesize{ph m$^{-2}$ s$^{-1}$}}
\end{picture}
\end{subfigure}
\end{center}
\caption{Same as Figure \ref{em_varic} for bremsstrahlung.
\label{em_varbr}}
\end{figure*}
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9a}}
\put(120,330){CASE A}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength ,]{./f9b}}
\put(120,330){CASE B}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9c}}
\put(120,330){CASE C}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9d}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9e}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,320)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9f}}
\put(115,20){\line(1,0){84.4}}
\put(115,17){\line(0,1){6}}
\put(199.4,17){\line(0,1){6}}
\put(115,5){\scriptsize{0.005 arcsec}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9g}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9h}}
\end{picture}
\end{subfigure}
\begin{subfigure}{320\unitlength}
\begin{picture}(320,358)
\put(0,0){\includegraphics[trim=4.2cm 1.5cm 5.1cm 1.5cm, clip=true,width=320\unitlength]{./f9i}}
\end{picture}
\end{subfigure}\\
\begin{center}
\begin{subfigure}{500\unitlength}
\begin{picture}(500,0)
\put(0,-20){\includegraphics[trim=2cm 2cm 1cm 24cm, clip=true,width=500\unitlength]{./f9j}}
\put(200,25){\footnotesize{ph m$^{-2}$ s$^{-1}$}}
\end{picture}
\end{subfigure}
\end{center}
\caption{Same as Figure \ref{em_varic} for for neutral pion decay.
\label{em_varp0}}
\end{figure*}
\subsubsection{Spectral Energy Distribution}
We calculate the SED by summation over all emitting grid cells for a fixed energy throughout the computational domain. This is shown for case A, B, and C and various orientations in Figure \ref{gamspec1} a), b) and c).
Figure \ref{gamspec1} d) explores the variation of the IC-component for a single cell at the apex of the WCR for case B as a function of orientation. It also discriminates between the two components of IC-emission, one being due to the scattering of electrons on the radiation field of the B star (red dotted curves), the other due to scattering on the radiation field of the WR star (dashed black). Whereas the first one is maximal at ($i=$90$^\circ$,$\Phi=$0$^\circ$) and disappears at ($i=$90$^\circ$,$\Phi=$180$^\circ$), the other behaves in the exact opposite manner. This is due to the dependence on the scattering angle $\theta_\mathrm{sc}$. For the case ($i=$90$^\circ$,$\Phi=$0$^\circ$) the scattering angle of photons from the B star is 180$^\circ$ . The corresponding IC emission from the apex of the WCR is therefore at its maximum. For the photon field of the WR star the scattering angle is zero and there is no emission from the apex via IC scattering.
The spectra of photons by neutral pion decay and bremsstrahlung remain the same, as they do not depend on the scattering angle. This last statement excludes the effects of photon photon opacity which is not taken into account in Figure \ref{gamspec1} d).
Studying the spectra in Figure \ref{gamspec1} (a) to (c) (for case A to C) significant differences between small and large stellar separation become apparent.
Case A shows a neutral pion component which dominates the spectrum at E $>$100 MeV for all orientations. The lack of high-energy electrons near the apex of the WCR leads to an early cutoff of the IC-component at $\sim$10 MeV, followed by a steady decline where high-energy electrons in the wings of the WCR contribute a low number of photons at higher energies.
It is interesting to note that amongst the three depicted orientations, the IC flux is highest for ($i=90^\circ$, $\Phi=180^\circ$). The same can be seen in Figure \ref{em_varic} by comparing the first and second row for case A. Because of the lack of high-energy electrons at the apex, it is the wings of the WCR that mainly contribute to the IC emission at energies $>$100 MeV. In these regions, the scattering angle for the radiation field of the WR star is much more favourable for ($i=90^\circ$, $\Phi=180^\circ$) than for ($i=90^\circ$, $\Phi=0^\circ$), whereas the scattering angle for the B-star is unfavourable for both. This is an effect of the curvature of the WCR and the different distance from the WCR to the stars.
Figure \ref{gamspec1} a) also shows effects of photon photon opacity in the spectrum of the neutral pion component which completely dominates the emission at high energies. Owing to the denser radiation field of the WR star, opacity effects are highest for the orientation ($i=90^\circ$, $\Phi=0^\circ$) in which emission from the apex has to pass by close to the star. The higher temperature of the star also causes the onset of photon photon absorption at lower energies than for the orientation ($i=90^\circ$, $\Phi=180^\circ$) for which high-energy photons from the apex come close to the cooler B star.
Due to the gap between the pion-bump and the cutoff of the IC spectrum, the total emitted photon spectrum shows a pronounced dip at $\sim$50 MeV.
Case B is already quite different from case A. Due to higher energies reached by the electrons, the cutoff in the IC component occurs at higher photon energies, causing a broad overlap with the neutral pion component. At the same time, lower plasma densities in the WCR lead to a lower flux for the bremsstrahlung and neutral pion component. The maximum energy of photons by bremsstrahlung increases with the maximum electron energy. As some regions in the wings of the WCR now produce sufficiently high electron energies, both, the IC and the bremsstrahlung component reach up to $\sim$1 TeV. Effects of photon-photon opacity become more pronounced for these components also. Again, they are largest for the case ($i=90^\circ$, $\Phi=0^\circ$) where the emission from the WCR has to pass by the luminous WR star.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,370)
\put(15,15){
\includegraphics[width=\textwidth]{./f10a}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV m$^{-2}$s$^{-1}$ )}}}
\put(450,80){a)}
\end{picture}
\end{subfigure}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,370)
\put(15,15){
\includegraphics[width=\textwidth]{./f10b}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV cm$^{-3}$ )}}}
\put(450,80){b)}
\end{picture}
\end{subfigure}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,400)
\put(15,15){
\includegraphics[trim=0.cm 0.cm 0.cm 0cm,
clip=true,width=\textwidth]{./f10c}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV cm$^{-3}$ )}}}
\put(450,80){c)}
\end{picture}
\end{subfigure}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,400)
\put(15,15){
\includegraphics[trim=0.cm 0.cm 0.cm 0cm,
clip=true,width=\textwidth]{./f10d}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV cm$^{-3}$ )}}}
\put(450,80){d)}
\end{picture}
\end{subfigure}
\caption{a), b), c) Photon spectra emitted by all particles in the whole computational domain of cases A to C for IC scattering (black), bremsstrahlung (gray) and neutral pion decay (red) with $i=$0$^{\circ}$ (solid), $i=90^\circ$, $\Phi=0^\circ$ (dashed) and $i=90^\circ$, $\Phi=180^\circ$ (dotted). The sum of all three components (for $i=90^\circ$, $\Phi=180^\circ$) is indicated in shaded gray.\\
d) Photon spectra emitted by one single cell at the apex of case B. The black dashed lines show the IC flux from scattering in the radiation field of the WR star for an orientation of ($i=90^\circ$, $\Phi=180^\circ$), ($i=45^\circ$, $\Phi=180^\circ$), ($i=0^\circ$, $\Phi=0^\circ$), ($i=45^\circ$, $\Phi=0^\circ$) and ($i=80^\circ$, $\Phi=0^\circ$) in order of decreasing flux. The red dotted lines show the IC flux from scattering in the radiation field of the B star for an orientation of ($i=90^\circ$, $\Phi=0^\circ$), ($i=45^\circ$, $\Phi=0^\circ$), ($i=0^\circ$, $\Phi=0^\circ$), ($i=45^\circ$, $\Phi=180^\circ$) and ($i=80^\circ$, $\Phi=180^\circ$) in order of decreasing flux. The photon flux from neutral pion decay and bremsstrahlung is represented by the dash-dotted and double-dot-dashed line.
\label{gamspec1} }
\end{figure*}
For case C, a large population of high-energy electrons leads to complete dominance of the IC-component. Whereas lower plasma densities lead to still lower flux values for the neutral pion and the bremsstrahlung component, the latter along with the IC component again reaches energies of up to 1 TeV. As the IC flux also decreases with increasing stellar distance (and thus decreasing energy density of radiation), it shows lower flux values with respect to case B below $\sim$1 GeV. For higher energies, flux values unprecedented for lower stellar separations are reached. Effects of photon-photon opacity are visible for all three emission components. For the same reasons as above, they are strongest for ($i=90^\circ$, $\Phi=0^\circ$). Note that the IC component is now maximal for the case ($i=90^\circ$, $\Phi=0^\circ$) which can be understood by the increased density of high-energy electrons at the apex where the radiation field of the B star dominates. The latter has most favourable scattering angles for ($i=90^\circ$, $\Phi=180^\circ$) . Because of the effects of photon photon opacity, the face-on orientation shows the highest resulting IC flux at E$>$100 GeV.
Comparing all three cases, we find that (for an electron-proton injection ratio of $10^{-2}$) growing stellar separation leads to a transition from hadron-dominated emission (neutral pion decay) to lepton-dominated emission (IC) at energies E$>$100 MeV. The latter is inhibited for close stellar separations because of a lack of high-energy electrons at the apex of the WCR. As an additional effect, different viewing angles produce significant variation in the IC emission spectra. These dependencies on stellar separation and viewing angle can be quantified using integrated flux values for various energy bands.
\subsubsection{Integrated Fluxes}
By integrating over energy and space, we obtain the total emitted flux, which we computed for three energy intervals. Table \ref{fluxtable} provides flux values for three different orientations for case A, B, and C. The main characteristics in this table will be further discussed below.
\paragraph{The flux above 10 GeV}
As suggested by the spectral shapes discussed in the previous section, the dominant radiation process at E $>$ 10 GeV changes with stellar separation. From Table \ref{fluxtable} one can assess that, roughly, the total IC and bremsstrahlung fluxes increase by $\sim$ 3 orders of magnitude from case A to C, whereas the neutral pion component decreases on a much smaller scale.
Only for case C, the IC-component dominates over the neutral pion component. Studying effects of changing orientation, we find that the variations in the bremsstrahlung and neutral pion component due to changing magnitude of photon photon opacity are very low.
For the IC component, its anisotropic nature as well as the larger effect of absorption cause flux variations due to the orientation of the system of a factor of 3 for case A and lower factors for cases B and C.
\paragraph{The flux above 100 MeV}
Including also lower energies down to 100 MeV, we find that growing stellar separation has the opposite effect than for the previous case. For E$>$10 GeV, higher maximum energies for electrons for large stellar separation led to an \textit{increase} of the IC and bremsstrahlung photon flux with increasing stellar separation. At lower energies, higher plasma densities in the WCR, along with denser radiation fields for smaller stellar separations, reverse this order and cause IC and bremsstrahlung photon fluxes to \textit{decrease} with increasing stellar separation. For E$>$100 MeV this transition is only partly realised, as we still find an increase in flux from case A to B for the IC component. However, all components decrease in their average flux from case B to C. There is no such reversal of proportionality for the neutral pion component.
We also find that the IC component is now dominant for cases B and C, albeit for some orientations of case B that are least favourable for IC emission, it is just a factor of 1.2 above the neutral pion component. For case A, the latter still dominates all possible orientations.
Variations due to changes in the viewing angle cause differences of a factor of $\sim$8 in the IC flux component for case A with a smaller contrast for cases B and C.
\paragraph{The flux above 1 keV}
Including still lower energies, all components now show the identical trend that increasing stellar separation causes decreasing flux values. Furthermore, the IC component dominates for all three cases. The total flux of the neutral pion component is even below the total flux of the bremsstrahlung component. This can be understood in terms of the lack of photons from pion decay below a few MeV due to the shape of the pion bump.
Effects of varying orientation are still lower (factor $\sim$2) . However the previously observed order of larger fluxes for orientation ($i=$90$^\circ$,$\Phi=$180$^\circ$) than ($i=$90$^\circ$,$\Phi=$0$^\circ$) in case A and the opposite trend in case C remains in place.
\subsection{Effects of Orbital Motion}
\label{orbital}
We finally address the impact of circular orbital motion and the ensuing deformation of the WCR. As the effects increase with orbital velocity, we choose the case of smallest stellar separation (case A) and let it evolve until the WCR adjusts to the new situation and there is no further change in its curvature. We analyse the system after 1.5 orbits when the stars are once more on the $x$ axis. For the given stellar masses and a distance of 720 R$_\odot$, a Keplerian orbit has a period of $\sim$290 days. The orbital velocity of the stars is $\sim$263 km s$^{-1}$. With orbital motion, the two arms of the WCR (henceforth forward arm and trailing arm) develop considerable differences.
\begin{table*}
\begin{center}
\textbf{Integrated flux for various energy bands}\\
\end{center}
\centering
\begin{tabular}{r ccc|ccc|ccc}
\textit{($i$,$\Phi$)}&
\textit{$F_\mathrm{IC}^\mathrm{>10GeV}$} & \textit{$F_\mathrm{br}^\mathrm{>10GeV}$} &
\textit{$F_\mathrm{\pi^0}^\mathrm{>10GeV}$} & \textit{$F_\mathrm{IC}^\mathrm{>100MeV}$} &
\textit{$F_\mathrm{br}^\mathrm{>100MeV}$} & \textit{$F_\mathrm{\pi^0}^\mathrm{>100MeV}$} &
\textit{$F_\mathrm{IC}^\mathrm{>1keV}$} & \textit{$F_\mathrm{br}^\mathrm{>1keV}$} & \textit{$F_\mathrm{\pi^0}^\mathrm{>1keV}$}\\
\textbf{Case A}\\
($0^\circ$, $0^\circ$) &3.2 &1.4 &3.8
&33 &3.1 &3.4
&6.2 &6.5 &5.2 \\
($90^\circ$, $0^\circ$) &1.4 &1.4 &3.7
&6.3 &3.1 &3.4
&4.1 &6.5 &5.2 \\
($90^\circ$, $180^\circ$) &4.6 &1.4 &3.9
&49 &3.1 &3.4
&8.2 &6.5 &5.2 \\
&$\times10^{-5}$ &$\times10^{-7}$ &$\times10^{-2}$
&$\times10^{-2}$ &$\times10^{-2}$ &
&$\times10^{3}$ &$\times10^{1}$ & \\
\textbf{Case B }\\
($0^\circ$, $0^\circ$) &18 &8.4 &1.9
&2.7 &2.9 &1.5
&3.0 &2.5 &2.3 \\
($90^\circ$, $0^\circ$) &9.8 &8.4 &1.8
&1.8 &2.9 &1.5
&2.5 &2.5 &2.3\\
($90^\circ$, $180^\circ$) &24 &8.5 &1.9
&3.6 &2.9 &1.5
&3.6 &2.5 &2.3 \\
&$\times10^{-4}$ &$\times10^{-6}$ &$\times10^{-2}$
& &$\times10^{-2}$ &
&$\times10^{3}$ &$\times10^{1}$ & \\
\textbf{Case C\qquad}\\
($0^\circ$, $0^\circ$) &2.4 &6.3 &5.5
&2.0 &1.1 &4.4
&8.9 &6.8 &6.6 \\
($90^\circ$, $0^\circ$) &2.5 &6.3 &5.4
&2.2 &1.1 &4.4
&9.8 &6.8 &6.6 \\
($90^\circ$, $180^\circ$) &2.2 &6.4 &5.5
&1.9 &1.1 &4.4
&8.7 &6.8 &6.6 \\
&$\times10^{-2}$ &$\times10^{-5}$ &$\times10^{-3}$
& &$\times10^{-2}$ &$\times10^{-1}$
&$\times10^{2}$ & &$\times10^{-1}$ \\
\end{tabular}
\caption{Integrated flux values in units of m$^{-2}$s$^{-1}$ as obtained for the chosen electron-proton injection ratio of $10^{-2}$ for three different stellar separations (case A, B, C), as well as three different viewing angles. \label{fluxtable}}
\end{table*}
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11a}}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{20\unitlength}
\begin{picture}(20,290)
\put(-5,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.3cm, clip=true,height=280\unitlength]{./f11b}}
\put(-19,295){{\footnotesize{log(m$^{-3}$)}}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11c}}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{25\unitlength}
\begin{picture}(20,290)
\put(-5,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.3cm, clip=true,height=280\unitlength]{./f11d}}
\put(-19,295){\footnotesize{log(K)}}
\end{picture}
\end{subfigure}
\begin{subfigure}{290\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11e}}
\put(126.7,20){{\color{white}\line(1,0){35.625}}}
\put(126.7,17){{\color{white}\line(0,1){6}}}
\put(162.325,17){{\color{white}\line(0,1){6}}}
\put(126.7,5){\scriptsize{{\color{white}500R$_\odot$}}}
\end{picture}
\end{subfigure}
\begin{subfigure}{20\unitlength}
\begin{picture}(20,290)
\put(-5,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.4cm, clip=true,height=280\unitlength]{./f11f}}
\put(-19,295){\footnotesize{km s$^{-1}$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{300\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11g}}
\put(20,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{300\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11h}}
\put(20,250){ 1 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{300\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11i}}
\put(20,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{70\unitlength}
\begin{picture}(75,290)
\put(20,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.4cm, clip=true,height=280\unitlength]{./f11j}}
\put(-5,70){\footnotesize{\rotatebox{90}{log(MeV$^{-1}$m$^{-3}$)}}}
\end{picture}
\end{subfigure}\\
\begin{subfigure}{300\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11k}}
\put(20,250){ 10 MeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{300\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11l}}
\put(20,250){ 1 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{300\unitlength}
\begin{picture}(290,290)
\put(0,0){\includegraphics[trim=2.9cm 1.5cm 5.1cm 1.5cm, clip=true,width=290\unitlength]{./f11m}}
\put(20,250){ 100 GeV}
\put(126.7,20){\line(1,0){35.625}}
\put(126.7,17){\line(0,1){6}}
\put(162.325,17){\line(0,1){6}}
\put(126.7,5){\scriptsize{500R$_\odot$}}
\end{picture}
\end{subfigure}
\begin{subfigure}{20\unitlength}
\begin{picture}(75,290)
\put(20,5){\includegraphics[trim=15.9cm 1.4cm 2.4cm 1.4cm, clip=true,height=280\unitlength]{./f11n}}
\put(-5,70){\footnotesize{\rotatebox{90}{log(MeV$^{-1}$m$^{-3}$)}}}
\end{picture}
\end{subfigure}
\caption{First row: HD quantities particle density, temperature and absolute velocity for case A with orbital motion. Second row: Differential number density of electrons for three different values of kinetic particle energy. Third row: the same for protons. The plots show the $x$--$y$ plane of a $256\times256\times256$ simulation at z=0. \label{orbit_hA}}
\end{figure*}
Figure \ref{orbit_hA} shows density, temperature and velocity, as well as particle distributions for different energies on the $x$--$y$ plane (the stars move counter-clockwise). Concerning the hydrodynamic wind variables, the main difference to the stationary case (see Figure \ref{hydros}) lies not in minimal and maximal values but in the now asymmetric shape of the WCR. In paper I, we found that the significant contrasts in the particle distribution between forward and trailing arm are mainly due to the different angle between the stars and the edge of the WCR, resulting in different shock velocities and (as $\dot{E}_\mathrm{DSA}\propto V_\mathrm{Shock}^2$) in different maximum particle energies. For the present case, the ensuing reduction of the DSA rate in the trailing arm of the shock front toward the B star results in a lack of electrons already at 1 GeV (see center plot of Figure \ref{orbit_hA}). In the forward arm, the opposite effect occurs. As wind velocity components are slightly higher close to the edge of the computational domain in the forward arm, an increased DSA rate can accelerate electrons to energies up to 100 GeV. (Note the thin shaded area at $\sim$-6 log(MeV$^{-1}$m$^{-3}$) in the center right plot of Figure \ref{orbit_hA}.) There, electrons reach higher energies than in the case where orbital motion is not taken into account.
Although we also find a certain degree of asymmetry for protons, the effects of orbital motion are less severe.
Figure \ref{orbit_Pha} shows the resulting projected photon emission maps for different orientations with respect to the observer. To simplify comparison of the case including orbital motion with the stationary case, we compensate the reversed order of the stars along the $x$ axis by a redefinition of the angle $\Phi$: $\Phi^\prime= \Phi+180^\circ$ and use the same representation of projection maps as discussed previously for $\Phi^\prime$.
For all three components the face-on orientation (first row) shows considerable asymmetry due to the lack of high-energy particles in the trailing arm of the shock front toward the B star. This produces the apparent lack of emission in the upper left section of the plots. In contrast, a wealth of particles in the trailing arm of the shock front toward the WR star produces significant emission that even dominates the entire emission region for the IC component.
In the case of alignment of the line of centers and the line of sight (second and third row), the differences of trailing and forward arm manifest themselves most strikingly for the IC component. The circular feature bordering the low-emission center region (most prominent for $i=$90$^\circ$, $\Phi^\prime=$180$^\circ$, third row) is caused by the emission of the forward arm toward the B star touching the edge of the computational domain. If seen from the opposite direction (second row), it is highly suppressed because of an unfavourable scattering angle.
The lack of emission next to this feature is due to the low number of high-energy electrons in the forward arm towards the WR star. This becomes also apparent in the corresponding plots for bremsstrahlung and neutral pion decay. For the same reasons as in the stationary case, we find that the configuration ($i=$90$^\circ$,$\Phi^\prime=$180$^\circ$) exhibits considerably higher flux for the IC component than the configuration ($i=$90$^\circ$, $\Phi^\prime=$0$^\circ$).
A comparison of the spectra with and without orbital motion is shown in Figure \ref{orbit_Aspec}. We find a decrease in flux for the neutral pion decay and the bremsstrahlung component. Concerning the IC component we find a new spectral feature emerging at $\sim$100 MeV that persists for all three depicted orientations. We interpret this as the influence of the forward arm toward the B star. There, a population of electrons reaching up to 100 GeV is still sufficiently close to the stars in order to exhibit favourable conditions for IC emission. To a lesser degree, the same region also influences the bremsstrahlung component which shows a similar feature at $>$1 GeV. The above shows that orbital motion has considerable impact on the resulting non-thermal high-energy emission.
\section{DISCUSSION}
We have computed different components of nonthermal high-energy photon emission in colliding wind binaries -- for the first time on the basis of a 3D distribution of high-energy particles that was obtained by hydrodynamic simulations and the simultaneous solution of the transport equation. Our approach includes the radiative acceleration of the stellar winds, radiative cooling in the hot shocked gas, diffusive shock acceleration at the shock fronts bordering the WCR and various energy loss mechanisms affecting the high-energy particles. The resulting nonthermal high-energy photon emission has been explored in terms of 2D projection maps, SEDs as well as integrated flux predictions of nonthermal photon emission components based on 3D distributions of high-energy particles in colliding wind binary systems.
In computing the anisotropic IC-component of the photon emission we take into account the dependence on the scattering angle, as well as the radiation fields of both stars of the binary system. Bremsstrahlung and neutral pion decay components of the photon flux are computed using local wind plasma densities and the present distribution of electrons and protons. The effect of photon photon opacity in the dense radiation fields of the stars has been taken into account for all three nonthermal photon emission mechanisms.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}{990\unitlength}
\includegraphics[width=\textwidth]{./f12a}
\end{subfigure}\\
\begin{center}
\begin{subfigure}{500\unitlength}
\begin{picture}(500,0)
\put(0,-20){\includegraphics[trim=2cm 2cm 1cm 11.7cm, clip=true,width=500\unitlength]{./f12b}}
\put(200,25){\footnotesize{ph m$^{-2}$ s$^{-1}$}}
\end{picture}
\end{subfigure}
\end{center}
\caption{Projected photon flux above 100 MeV at 1 kpc distance for IC-emission (first column), bremsstrahlung (second column) and neutral pion decay (third column) for case A including orbital motion after 1.5 orbits. The rows represent different orientations. First row: $i=0^\circ$, $\Phi^\prime=0^\circ$ (face-on); second row: $i=$90$^\circ$, $\Phi^\prime$=0$^\circ$ (along--line of--center); third row: $i=$90$^\circ$, $\Phi^\prime=$180$^\circ$ (along line of centers, opposite to second row).
\label{orbit_Pha}}
\end{figure*}
Studying three cases of varying stellar separation, we find significant differences in terms of total flux value and also in the identity of the dominant emission component.
We showed that low stellar separations (720 R$_\odot$ for the studied WR-B system, case A) inhibits the acceleration of high-energy electrons at the apex of the WCR due to strong IC and synchrotron losses. This produces an early cutoff in the photon flux ensuing from IC-scattering and bremsstrahlung. Although denser radiation fields and higher plasma densities cause higher photon flux values at lower energies, this early cutoff leads to dominance of the hadronic neutral pion decay component at energies E $\gtrsim$100 GeV for the chosen electron-proton injection ratio of $10^{-2}$. In contrast to that, large stellar separations lead to a dominance of the IC component throughout the studied energy range. As the magnetic and radiation energy density at the WCR are low, the electrons -- and also the ensuing photon fluxes -- reach higher energies. Both, IC and bremsstrahlung components reach up to $\sim$1 TeV if we increase stellar separation by a factor of four (case C). Lower wind plasma densities in the WCR further diminish the significance of the neutral pion decay component. We find that increasing the stellar separation by a factor of four results in an increase of $\sim$ 3 orders of magnitude in the IC and bremsstrahlung components of the photon flux at energies E$>$10~GeV, whereas the neutral pion decay component decreases by about an order of magnitude. For lower energies the importance of high-energy electrons diminishes. Accordingly, the total flux for E$>$1 keV decreases for all three flux components if we increase stellar separation by a factor of four.
The strong dependence on stellar separation suggests that $\gamma$-ray binaries are liable to show a significant variation in their spectra in course of an elliptical orbit with high eccentricity. Spectra can exhibit multiple emission components (as in Figure \ref{gamspec1} a)) for low separations (i.e. during periastron), contrasted by longer periods of one-component spectra (as in Figure \ref{gamspec1} c)) for larger stellar separations. We note that a situation resembling this trend of two apparently distinct emission components has already been observed for the peculiar case of $\eta$ Carinae \citep[see ][]{Takapaper,Walter}. However, this can also be understood in terms of $\gamma$-ray absorption due to the presence a hot X-ray gas \citep[see ][]{Reitberger2012}. The application of our numerical model to other specific candidates for particle-acceleration CWB systems \citep[see catalogue of ][]{Becker2013} will provide predictions for the nonthermal high-energy flux that can then be related to detections and upper limits from observations.
Varying the orientation of the binary system by changing the viewing angles $i$ and $\Phi$, we find a strong dependence of the IC flux component. Photons from bremsstrahlung and neutral pion decay are influenced only indirectly via the changing conditions for photon photon opacity which affect all components at high energies alike. This effect remains small compared to the variation of the IC component due to its dependence on the scattering angle. Since we consider the radiation field of both stars, the IC component does not vanish as the flux from one population of target photons is maximal when the other is minimal. However, since the two radiation fields differ in intensity depending on distance and identity of the star, there remains a significant level of variation. It is largest for the case of small stellar separation where the B star is very close to the WCR. Here, the IC-flux above 100 MeV varies a factor of 8 between the two extreme cases of ($i=90^\circ$, $\Phi=0^\circ$) and ($i=90^\circ$, $\Phi=180^\circ$).
Although this dependence is small for the cases we investigated, it becomes significantly stronger if we have a binary system with greater differences concerning the two stellar radiation fields and the ram pressure of the winds. The IC contribution to the $\gamma$-ray flux will differ by several orders of magnitude in course of an orbital cycle if the thermal photon density of one stellar component is dominant and the orbital inclination is close to the edge-on case of $i~\sim90^\circ$.
\begin{figure*}
\setlength{\unitlength}{0.001\textwidth}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,370)
\put(15,15){
\includegraphics[width=\textwidth]{./f13a}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV m$^{-2}$s$^{-1}$ )}}}
\put(450,80){a)}
\end{picture}
\end{subfigure}
\begin{subfigure}[c]{500\unitlength}
\begin{picture}(500,370)
\put(15,15){
\includegraphics[width=\textwidth]{./f13b}
}
\put(200,0){\footnotesize{log( E in MeV)}}
\put(0,120){\rotatebox{90}{\footnotesize{log( $E^2N$ in MeV cm$^{-3}$ )}}}
\put(450,80){b)}
\end{picture}
\end{subfigure}
\caption{Same as Figure \ref{gamspec1} for case A with a) and without b) orbital motion. The sum of all three components (for $i=90^\circ$, $\Phi=180^\circ$) is indicated in shaded gray.
\label{orbit_Aspec} }
\end{figure*}
The effects of photon photon opacity in the radiation field of the star are restricted to incident photons with $E\gtrsim100$ GeV. Given their low prominence in case of low stellar separation, there is no noticeable impact. However, for larger stellar separation, the absorption leaves a clear signature in the spectral components of IC scattering and bremsstrahlung.
Studying the impact of orbital motion for the case of smaller stellar separation, we find a number of effects that are all significantly less severe than the ones obtained for varying the stellar separation by a factor of two. The distortion of the WCR leads to regions of altered wind velocity normal to the shock and, thus, to notable changes in the DSA acceleration rate. Whereas IC emission is greatly reduced in the trailing arm toward the B star (and also in the forward arm toward the WR star), we find regions of higher maximum energies for electrons (than in the case without orbital motion) in the side of the forward arm toward the B star. In terms of SEDs, this leads to an additional bump in the IC component.
The changes due to orbital motion are less pronounced for the bremsstrahlung and neutral pion decay components.
\section{OUTLOOK}
The presented numerical model can be used to study the high-energy nonthermal flux of specific particle-accelerating CWB systems as a function of time. Possible variations in course of the orbit can be quantitatively assessed and favourable observation periods can be predicted. Compared to observational $\gamma$-ray data (e.g. from the Fermi-LAT instrument), our simulation aims for constraining and refinement of poorly known parameters, such as as the injection fraction of electrons and protons at the shock of the wind collision region. Dependencies on other parameters, such as the magnetic field and the diffusion coefficient, can be studied as well.
The consequent next step will be the application of our simulation procedure to archetypal nonthermal binary systems. We will confront hardly understood discrepancies between existing flux predictions from analytical models with the present absence of detections of particular enigmatic systems in the \textit{Fermi}-LAT data like e.g. WR104 or $\eta$~Carinae.
\acknowledgments
This work is supported through the EU FP7 program by Marie Curie IRG grant 248037; K.R. is supported by the Marietta Blau-Stipendium der OeAD-GmbH, financed by the Austrian Ministry of Science BMWF. AR acknowledges financial support through the Austrian Science Fund (FWF) grant P 24926-N27. Additional support was given by the Austrian Ministry of Science BMWF as part of the UniInfrastrukturprogramm of the Research Platform Scientific Computing at the University of Innsbruck and the Austrian Science Fund (FWF) supported Doctorate School DK+ W1227-N16.
|
\section{INTRODUCTION}
\label{introduction}
Transition disks (TDs) are important for studying the advanced stages
of protoplanetary disk evolution \citep[see, e.g.,][]{espaillat14}.
TDs were originally identified as sources for which the
spectral energy distribution (SED) demonstrated a lack of near-infrared excess
despite the presence of strong mid- to far-infrared excess.
This was attributed to a gap in the inner disk
devoid of small grains \citep[e.g.,][]{strom89}.
Theory suggests gaps in TDs are cleared by close-in companions, with other disk-dispersal mechanisms, e.g.,
grain growth or photoevaporation, happening in parallel \citep[see, e.g.,][]{armitage11,williams11,espaillat14}.
SEDs provide {\em indirect} evidence of gaps in TDs; however, long-baseline
interferometry at (sub)mm wavelengths has revealed their ring-like morphology \citep[e.g.,][]{andrews11}.
The Atacama Large Millimeter/Submillimeter Array (ALMA)
has revealed extreme asymmetries in the dust emission in several systems,
indicative of dust traps triggered by the interaction
between the disk and a close-in companion \citep{casassus12,vandermarel13}
or gravitational instabilities \citep{fukagawa13}.
ALMA observations have also demonstrated that gaps can contain a significant
reservoir of molecular gas \citep{bruderer14}.
We present ALMA Cycle 0 observations
of the transition disk encompassing HD~100546 which reveal the
spatially-resolved gas and dust structure at (sub)mm wavelengths.
\citet{pineda14} have already published these data; however,
we reach different conclusions based on more thorough data processing.
\section{HD~100546}
\label{hd100546}
HD~100546 is a 2.4$\mathrm{M}_\odot$ B9V Herbig Be star located at
103$\pm$6~pc which has a complex circumstellar environment \citep[e.g.,][]{vandenancker98,grady01}.
Coronographic imaging show that the small grains extend to large radii ($\approx$~500~AU)
and reveal evidence of spiral arms and disk brightness asymmetries
\citep[e.g.,][]{pantin00,augereau01,grady01,ardila07,boccaletti13}.
SED models of the dust emission suggest a gap
within $\approx$~10--13~AU
and the presence of an inner tenuous dust disk, $\lesssim$~0.7~AU
\citep{bouwman03,benisty10,tatulli11,panic14}.
Observations of [OI] (6300~\AA) line emission and
OH and CO rovibrational transitions confirm the presence of residual
gas close to the star with the observed dynamical perturbation of
the gas likely induced by a massive close-in companion
\citep{acke06,brittain09,vanderplas09,goto12,liskowsky12,brittain13,bertelsen14}.
Emission at 1.3 and 3.4~mm was detected using the
Swedish-ESO 15~m Submillimeter Telescope (SEST)
and the Australia Telescope Compact Array (ATCA)
yielding a total flux density of 465$\pm$20 and 36$\pm$3~mJy, respectively
\citep{henning94,wilner03}.
A plethora of molecular lines have been observed at far-infrared to
(sub)mm wavelengths including emission from $^{12}$CO, $^{13}$CO, OH, and \ce{CH+}
\citep[see, e.g.,][]{panic10,sturm10,thi11,meeus12,fedele13a}.
These data have allowed contraints on the radial behaviour
of the gas temperature structure, and indicate
thermal decoupling of the gas and dust in the disk atmosphere
\citep{bruderer12,fedele13b,meeus13}.
The detection of significant emission from a point source at a
deprojected radius of 68$\pm$10~AU \citep{quanz13} is of utmost importance
in indicating planet formation around HD~100546.
High-contrast angular differential imaging revealed that
the source emission coincides with a reduction in surface
brightness seen in corresponding polarimetric differential imaging \citep{quanz11}.
\citet{quanz13} conclude that the most likely explanation is a young gas-giant planet
(or protoplanet) caught in the act of formation,
reasoning that a mature massive planet (coeval with the star) would have
had sufficient time to significantly perturb the structure of the disk.
\section{OBSERVATIONS}
\label{observations}
HD~100546 was observed during ALMA Cycle 0 operations on 2012 November 18 using
24 antennas with baseline lengths between 21 and 375~m (program 2011.0.00863.S, P.~I. C.~Walsh).
The source was observed in seven spectral windows in Band~7, each with a bandwidth
of 469~MHz and a channel width of 0.122~MHz (0.24 and 0.21~km~s$^{-1}$
at 300 and 345~GHz, respectively, applying Hanning smoothing).
The central frequencies in each spectral window are
300.506, 301.286, and 303.927~GHz for the first execution,
and 344.311, 345.798, 346.998, and 347.331~GHz for the second execution.
The total on-source observation time was 13 and 14~mins, respectively.
The data were calibrated using the Common Astronomy Software Package (CASA), version 3.4.
The quasar, 3C~297, was used as bandpass calibrator with Titan and a quasar, J1147-6753,
used for amplitude and phase calibration, respectively.
Self-calibration and imaging were performed using CASA version 4.1.
During imaging it was noticed that the telescope pointing
had not taken into account the proper motions of the source
($\alpha_{2000}$~=~11$^\mathrm{h}$~33$^\mathrm{m}$~25\fs44058, $\mu_\alpha$~=~$-38.93$~mas~yr$^{-1}$;
$\delta_{2000}$~=~$-70$\degr~11\arcmin~41\farcs2363, $\mu_\delta$~=~$+0.29$~mas~yr$^{-1}$).
The phase center of the observations was subsequently corrected using the CASA task, {\tt fixvis}.
The continuum bandwidth amounted to 1.48 and 1.83~GHz
averaged at 302 and 346~GHz.
Continuum and line imaging were performed using
the CLEAN algorithm with Briggs weighting (robust~=~0.5) resulting in synthetic beam sizes of
$1\farcs0 \times 0\farcs48$
(23\textdegree) and $0\farcs95 \times 0\farcs42$ (38\textdegree) at 302 and 346~GHz.
The synthesized beam is elongated perpendicular to the
major axis of the disk owing, in part, to the low declination of the source (-70\textdegree).
The continuum was subtracted from line-containing channels using the CASA task, {\tt uvcontsub},
in advance of imaging the CO emission.
The achieved rms for the continuum was 0.4 and
0.5~mJy~beam$^{-1}$ at 302 and 346~GHz, respectively, with
an rms of 19 mJy~beam$^{-1}$~channel$^{-1}$ attained for the CO-containing channels.
\section{RESULTS}
\label{results}
Figure~\ref{figure1} presents the CO $J$=3--2 first moment map overlaid with
contours of the integrated intensity and the continuum emission at 870~$\mu$m.
The integrated intensity was determined between $-12$ and $+12$~km~s$^{-1}$
relative to the source velocity (constrained by these data to 5.7~km~s$^{-1}$),
corresponding to channels containing significant emission ($\gtrsim$~3$\sigma$).
The CO emission is detected with a peak signal-to-noise of 163 in the channel maps.
The continuum emission is detected with a peak signal-to-noise of 1525 and 1320
and a total flux density of 0.980 and 1.240~Jy (summing over all flux $\gtrsim$~3~$\sigma$)
at 302 and 346~GHz, respectively.
The estimated absolute flux calibration uncertainties are $\approx$~10\%.
These flux densities are consistent with previous mm observations \citep{henning94,wilner03} and yield a
dust spectral index ($F_\nu$~$\propto$~$\nu^{\beta+2}$), $\beta$~$\approx$~0.7--0.8 between
3.4 and 1.0~mm, that falls to $\approx$~$-0.4$ between 1.0~mm and 870~$\mu$m, indicating that
the continuum emission is entering the optically thick regime at submillimeter wavelengths.
The total dust mass, M$_\mathrm{dust}$~$\approx$~$D^2 F_\nu / \kappa_\nu B_\nu(T_\mathrm{dust})$,
is $\approx$~0.035~M$_\mathrm{Jup}$, assuming
$\kappa_\nu = 10$~cm$^2$~g$^{-1}$ at 300~GHz, and $T_\mathrm{dust} = 60$~K
\citep[see, e.g.,][]{andrews11,bruderer12}.
Figure~\ref{figure2} shows the continuum flux density at 346~GHz
and CO integrated intensity along the major axis of the disk.
The data confirm the radius of the molecular disk, 390$\pm$20~AU (the error
corresponds to half the width of the synthesised beam).
The CO emission is similar in extent to the scattered light images from
\citet{ardila07} suggesting that the molecular gas and micron-sized grains are well mixed in the disk atmosphere.
The CO brightness distribution follows a $r^{-2}$ behaviour similar
to that seen for the micron-sized grains.
The size of the molecular disk is approaching the largest resolvable angular scale;
hence, the drop beyond $3\arcsec$ may be caused by spatial filtering.
However, the total integrated CO flux in these data is 151~Jy~km~s$^{-1}$ which is around
92\% of the flux measured with APEX \citep{panic10}.
Hence, it is unlikely that the disk extends significantly beyond the radius derived here.
The mm continuum emission extends to only 230$\pm$20~AU and
has two components: strong emission from the inner disk ($\lesssim 1\arcsec$) and a weaker
outer component (1\arcsec--2\farcs2) with a peak flux density $\approx$~4--5\% of the central flux.
The self-calibration procedure (using a mask containing only the strong continuum component)
significantly increased the dynamic range of the observations improving the peak signal-to-noise at
346 GHz from 150 to 1320, allowing the weak extended emission to be revealed.
All subsequent analysis is conducted in the visibility domain.
This allows a search for evidence of gaps or cavities which are not visible in the images.
As a first step, the CASA task {\tt uvmodelfit} was used to fit the continuum
visibilities assuming the emission arises from a elliptical disk.
This resulted in an inclination of 44$\pm$3\textdegree and a position angle
(measured East from North) of 146$\pm$4\textdegree, respectively, in excellent
agreement with previous observations
\citep[see, e.g.,][]{pantin00,augereau01,grady01,ardila07,panic10}.
Without any further knowledge on the structure we assume a circular-symmetric
surface-brightness distribution.
Visibilities of such distributions depend
only on the deprojected baseline length,
$r_{uv} = \sqrt{u_\phi^2\cos^2 i + v_\phi^2}$, where
$u_\phi = u\cos\phi + v\sin\phi$ and
$v_\phi = -u\sin\phi + v\cos\phi$ assuming the $u$-axis is aligned with right ascension \citep[see, e.g.,][]{berger07}.
Here, $(u,v)$ are the observed visibility coordinates,
$i$ is the source inclination, and $\phi$ is the disk position angle.
Figure~\ref{figure3} presents the binned visibilities
(in 10~k$\lambda$ bins) as a function of $r_{uv}$.
The error bars correspond to the standard error in each bin.
The imaginary components show very small scatter around zero $\lesssim$~250--300~k$\lambda$,
confirming the assumption of a symmetric brightness distribution {\em a posteriori}
(a point-symmetric brightness distribution has zero imaginary components).
For $r_{uv}\gtrsim 250\mathrm{-}300$~k$\lambda$,
the scatter in the imaginary components increases, which may indicate
an asymmetry in the continuum emission; however, this may also be caused by coarser the $uv$
coverage at long baselines.
Higher spatial resolution observations are needed to confirm
any asymmetry at small spatial scales.
The real components of the visibilities decrease as a function
of the deprojected baseline indicating the continuum emission is resolved
and there is a zero crossing (null) at 290~k$\lambda$.
The Fourier transform of an infinitesimally narrow ring is
a Bessel function of the first kind, $J_0$: a null suggests the emission
originates from a ring with a finite width \citep[e.g., ][]{berger07,hughes07}.
\citet{pineda14} determine a null position at 250~k$\lambda$.
This is likely due to an incorrect deprojection related to the convention of the
direction of the $u$-axis relative to right ascension \citep{berger07,hughes07}.
For a ring, the real component of the visibilities is given by
\begin{equation}
V_{\mathrm{Re}}(r_{uv}) = 2\pi \int_{0}^{\infty} I(\theta)J_0(2 \pi r_{uv}\theta )\theta \, \mathrm{d}\theta
\end{equation}
\citep{berger07}.
The intensity, $I(\theta)$, is modelled as a power-law,
\begin{equation}
I(\theta) =
\begin{cases}
C \cdot \theta^{-\gamma} \quad \mathrm{for} \quad \theta \ge \theta_\mathrm{in} \quad \mathrm{and} \quad \theta \le \theta_\mathrm{out} \\
0 \quad \quad \mathrm{otherwise}.
\end{cases}
\end{equation}
The flux scaling factor, $C$, is determined using the total observed flux,
$V_{\mathrm{Re}}(0)$, i.e, $C = V_\mathrm{Re}(0)/\int_{0}^{\infty} I(\theta)J_0(0)\theta \mathrm{d}\theta$.
$\theta_\mathrm{in}$ and $\theta_\mathrm{out}$
were varied between 0 and 50~AU and 20 and 400~AU, respectively,
for $\gamma$~=~0, 1, and 2, using a small step size (1~AU) to adequately
sample the parameter space.
The best-fit model has an inner and outer radius of 16 and 51~AU,
and a power-law index of 2 (see Figure~\ref{figure3}).
This model corresponds to a deep global minimum in the $\chi^2$ value indicating that the
estimated uncertainties are smaller than the step size of the grid (1~AU).
The model residuals were imaged using an identical $uv$ coverage as
the observations (see Figure~\ref{figure4}).
The residuals in both the visibility and image domains
are large ($\gg$~3$\sigma$) indicating a poor fit.
The images reveal significant extended, weak
continuum emission (peak residuals~=~6~--~8$\sigma$) .
\citet{pineda14} do not see this extended emission because
no self calibration of the data was performed.
To include this more extended component, the model was adapted to include i)
a compact ring with a gaussian brightness distribution,
\begin{equation}
I(\theta) = C \exp{\left( \frac{-(\theta - \theta_\mathrm{peak})^2}{2\theta_\mathrm{width}^2}\right)},
\end{equation}
\citep{perez14}
and ii) an extended disk/ring with a flat brightness distribution ($\gamma$~=~0).
A low-resolution grid was run (5~AU) to determine
the location of the global minimum in the $\chi^2$ value.
A subsequently denser grid was run in which
$\theta_\mathrm{peak}$ and $\theta_\mathrm{width}$ were varied
between 20 and 50~AU and 5 and 20~AU, respectively, and
$\theta_\mathrm{in}$ and $\theta_\mathrm{out}$ between 10 and 200~AU
and 100 and 400~AU.
This grid included models composed of both overlapping and distinct rings.
A small step size of 0.5~AU was chosen to allow quantification of the
errors via Bayesian inference.
For simplicity, the total flux contribution from the compact and extended
components were fixed at 0.962 and 0.024~Jy at 302~GHz and
1.190 and 0.048~Jy at 346~GHz, respectively.
This was set by the flux in the residual images.
The visibilities are best reproduced by a
compact ring with a peak brightness at 26~AU and a FWHM of 21~AU and
an outer ring with a width of 75$\pm$3~AU centered at 190$\pm$3~AU
(see Figure~\ref{figure3}).
The data exclude overlapping rings in favour of two distinct rings of emission.
The estimated dust masses for the inner and outer rings
are $\approx$~2.5~$\times$~10$^{-2}$ and
$\approx$~1.4~$\times$~10$^{-3}$~M$_\mathrm{Jup}$,
assuming disk temperatures of 80 and 40~K at
$\approx$~30 and $\approx$~190~AU \citep[see, e.g.,][]{bruderer12}.
Figure~\ref{figure4} shows the residual images for the ``double-ring" model.
The peak residuals at 302 and 346~GHz are 1.5~mJy (3.8$\sigma$)
and 4.1~mJy (8.2$\sigma$).
However, these are restricted to small regions and
are likely owing to deviations from circular symmetry also suggested by non-zero
imaginary components on long baselines (see~Figure~\ref{figure3}).
\section{DISCUSSION}
\label{discussion}
Previous observations show that the mm-sized grains are not
necessarily cospatial with the molecular gas in protoplanetary disks
\citep[][]{isella07,andrews12,degregoriomonsalvo13}.
This can be explained by {\em radial drift}:
dust grains feel a drag force as they move through
the sub-Keplerian gas causing a loss of angular momentum and
migration inwards towards the star \citep[e.g.,][]{birnstiel10}.
When a massive companion opens a gap in the disk \citep[e.g.,][]{kley12}, this halts
the migration of grains because of the presence of a positive pressure
gradient at the outer edge of the gap.
Grains can accumulate and grow in this ``pressure trap" with the peak
and structure of the pressure profile dependent
on the disk viscosity and the location and mass of the companion \citep{pinilla12}.
Observations of HD~100546 support the presence of a
close-in companion \citep[e.g.,][]{acke06,liskowsky12,brittain13}.
\citet{mulders13} derived a lower limit of 20~M$_\mathrm{Jup}$ for the
companion mass and constrained the disk viscosity to
$\alpha_\mathrm{turb} \gtrsim 2\times 10^{-3}$.
A potential protoplanet has also been observed at 68$\pm$10~AU \citep{quanz13}.
The ALMA observations suggest the mm-sized grains are located in two rings:
one between the proposed companions and the other beyond the outer protoplanet.
To investigate the influence of companions on the dust evolution in HD~100546,
we model the dust growth and migration for two scenarios \citep{birnstiel10,pinilla12}:
(a) a 20~M$_\mathrm{Jup}$ companion at 10~AU only, and
(b) both a 20~M$_\mathrm{Jup}$ companion and a protoplanet (15~M$_\mathrm{Jup}$) at 68~AU.
We assume an initial particle size, 1$\mu$m,
a disk viscosity, $\alpha_\mathrm{turb} = 2\times 10^{-3}$,
a stellar mass, 2.4~$\mathrm{M}_\odot$, and a dust mass, $5.0 \times 10^{-4}$~$\mathrm{M}_\odot$ \citep{mulders13}.
The model from \citet{mulders13} is extrapolated to larger radii (400~AU)
using a power-law and assuming a gas-to-dust mass ratio of 100.
The younger protoplanet is injected into the simulations at 1~Myr.
Figure~\ref{figure5} presents the surface density of mm-sized and $\mu$m-sized
grains at different simulation times.
For the single-companion scenario, long evolution times are required, $\approx$~10~Myr,
for the grains to grow to mm sizes in the outer regions (100--400~AU)
and migrate inwards to accumulate in a radial pressure trap with a peak at
$\approx$~30~AU and a width of $\approx$~20~AU.
For the two-companion scenario, the surface density
decreases sharply in the region between the two companions upon introduction
of the protoplanet after 1~Myr of dust evolution.
The resulting steep pressure gradient causes
grains to migrate inwards on shorter timescales, $\approx$~0.2~Myr.
The injection of the protoplanet at 68~AU generates a second ring
at $\gtrsim$~100~AU with a surface density $\approx$~100--1000 times
lower than that for the inner ring.
Around 1.0~Myr of additional evolution is required
for this ring to narrow to a width $\lesssim$~100~AU.
These results are qualitatively in agreement with the ALMA observations
which also show a contrast of $\sim$~100 between emission from the inner and outer ring.
The $\mu$m-sized grains extend from $\approx$~12--13~AU to $\approx$~400~AU
which is consistent with scattered light observations and SED models of the source.
These observations and simulations support the presence of a massive companion
orbiting within the inner gap and a protoplanet embedded within the outer disk.
Numerical models of dust evolution including two companions recreate the inner
ring of mm emission and the extended weaker emission seen in the ALMA data.
Particle trapping by the inner companion alone cannot explain the nature
of the outer ring.
\acknowledgments
This paper makes use of the following ALMA data: ADS/JAO.ALMA\#2011.0.00863.S.
ALMA is a partnership of ESO (representing its member states),
NSF (USA) and NINS (Japan), together with NRC (Canada) and NSC and
ASIAA (Taiwan), in cooperation with the Republic of Chile.
The Joint ALMA Observatory is operated by ESO, AUI/NRAO and NAOJ.
The authors thank E.~F.~van~Dishoeck, C.~P.~Dullemond, N. van~der~Marel,
and M.~Schmalzl for useful discussions, and G.~D.~Mulders for sharing the results of his
hydrodynamical simulations.
C.~W. acknowledges support from the
Netherlands Organisation for Scientific Research (NWO, program number 639.041.335).
This work was also supported by EU A-ERC grant 291141 CHEMPLAN and a KNAW prize.
T.~B. acknowledges support from NASA Origins of Solar Systems grant NNX12AJ04G.
Astrophysics at QUB is supported by a grant from the STFC.
M.~R.~H., A.~J., and G.~S.~M. acknowledge support from the Netherlands Organization for
Scientific Research (NWO) to Allegro, the European ALMA Regional Center node in the Netherlands.
|
\section{Introduction}
We study the local H\"older regularity for viscosity solutions of possibly degenerate and singular non-local equations of the form
\begin{equation*
\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))K(x,y)\, dy = f(x),
\end{equation*}
where $f$ is bounded and $K(x,y)$ essentially behaves like $|y|^{-n-sp}$. Here $\operatorname{PV}$ stands for the \emph{principal value}.
This type of equations is one possible non-local counterpart of equations of $p$-Laplace type and arises for instance as the Euler-Lagrange equation of functionals in fractional Sobolev spaces. Solutions can also be constructed directly via Perron's method (cf. \cite{IN10}). In the case $K(y)=|y|^{-n-sp}$, when properly rescaled, solutions converge to solutions of the $p$-Laplace equation
$$
\Delta_p u=\div (|\nabla u|^{p-2}\nabla u)=0
$$ as the parameter $s$ tends to 1, see \cite{IN10}.
Our first and main result is that bounded viscosity solutions (see Section \ref{sec:visc}) of the homogeneous equation are locally H\"older continuous, see the theorem below. Throughout the paper we denote by $B_r$, the ball of radius $r$ centered at the origin.
\begin{thm}\label{thm:main} Assume $K$ satisfies $K(x,y)=K(x,-y)$ and there exist $\Lambda\geq \lambda>0$, $M>0$ and $\gamma>0$ such that
\begin{align*}
\frac{\lambda}{|y|^{n+sp}}\leq &K(x,y)\leq \frac{\Lambda}{|y|^{n+sp}}, \text{ for } y\in B_2,x\in B_2, \\%\Lambda\geq \lambda>0,\\
0\leq &K(x,y) \leq \frac{M}{|y|^{n+\gamma}}, \text{ for } y\in \mathbb{R}^n\setminus B_\frac14, x\in B_2,
\end{align*}
where $s\in (0,1)$ and $p\in (1,\infty)$. In the case $p<2$ we require additionally $p>1/(1-s)$. Let $u\in L^\infty(\mathbb{R}^n)$ be a viscosity solution of
$$
Lu:=\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))K(x,y)\, dy =0\text{ in }B_2.
$$
Then $u$ is H\"older continuous in $B_1$ and in particular there exist $\alpha$ and $C$ depending on $\lambda,\Lambda,M,p,s$ and $\gamma$ such that
$$
\|u\|_{C^\alpha(B_1)}\leq C\|u\|_{L^\infty(\mathbb{R}^n)}.
$$
\end{thm}
In particular, Theorem \ref{thm:main} applies for the fractional $p$-Laplace equation
$$
\operatorname{PV} \int_{\mathbb{R}^n}\frac{|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))}{|y|^{n+sp}}\, d y=0.
$$
We are also able to prove H\"older estimates for inhomogeneous equations with variable exponents, see below:
\begin{thm}\label{thm:main2} Assume $K$ satisfies $K(x,y)=K(x,-y)$ and there exist $\Lambda\geq \lambda>0$, $M>0$ and $\gamma>0$ such that
\begin{align*}
\frac{\lambda}{|y|^{n+s(x)p(x)}}\leq &K(x,y)\leq \frac{\Lambda}{|y|^{n+s(x)p(x)}}, \text{ for } y\in B_2,x\in B_2,\\% \Lambda\geq \lambda>0,\\
0\leq &K(x,y) \leq \frac{M}{|y|^{n+\gamma}}, \text{ for } y\in \mathbb{R}^n\setminus B_\frac14, x\in B_2,
\end{align*}
where $0<s_0<s(x)<s_1<1$ and $1<p_0<p(x)<p_1<\infty$. In the case $p(x)<2$ we require additionally that there is $\tau>0$ such that
$$
p(x)(1-s(x))-1>\tau.$$
Let $f\in C(B_2)\cap L^\infty(B_2)$ and let $u\in L^\infty(\mathbb{R}^n)$ be a viscosity solution of
$$
Lu:=\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(x+y)|^{p(x)-2}(u(x)-u(x+y))K(x,y)\, dy = f(x)\text{ in } B_2.$$
Then $u$ is H\"older continuous in $B_1$ and in particular there exist $\alpha$ and $C$ depending on $\lambda,\Lambda,M,p_0,p_1,s_0,s_1,\gamma $ and $\tau$ such that
$$
\|u\|_{C^\alpha(B_1)}\leq C\left(\|u\|_{L^\infty(\mathbb{R}^n)}+\max\left(\|f\|_{L^\infty(B_2)}^\frac{1}{p_0-1},\|f\|_{L^\infty(B_2)}^\frac{1}{p_1-1}\right)\right).
$$
\end{thm}
\begin{rem} It might seem odd that the two conditions on $K$ in our main theorems are supposed to be satisfied in overlapping regions, $B_2$ and $B_\frac14$. This is only for notational convenience. It would be sufficient to have the first condition satisfied in $B_\rho$ for some $\rho>0$ and the second one satisfied outside $B_R$ for some large $R$ as long as we ask $K$ to be bounded in $B_R\setminus B_\rho$.
\end{rem}
\begin{comment}We can also formulate our result for the so called maximal and minimal operators
$$
M^+ u :=\int_{\mathbb{R}^n}\left(\Lambda \delta^+(u,x,y)-\lambda \delta^-(u,x,y)\right)\,\frac{d y}{|y|^{n+sp}}
$$
and
$$
M^- u :=\int_{\mathbb{R}^n}\left(\lambda \delta^+(u,x,y)-\Lambda \delta^-(u,x,y)\right)\,\frac{d y}{|y|^{n+sp}},
$$
where
$$
\delta(u,x,y)=\frac{1}{2} \left(|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))+|u(x)-u(x-y)|^{p-2}(u(x)-u(x-y))\right),
$$
$$
\delta^\pm(u,x,y) = \max(\pm \delta(u,x,y),0).
$$
We have the following result:
\begin{thm}\label{thm:main2} Assume $p\in (1,\infty)$ and $p(1-s)>1$ if $p<2$. If $u\in L^\infty(\mathbb{R}^n)$ and $u$ satisfies
$$
M^+ u\geq -A,\quad M^-u\leq A \text{ in $B_2$},
$$
then there are constants $\alpha$ and $C$ depending only on $s$, $p$ such that
$$
\|u\|_{C^\alpha(B_1)}\leq C(\|u\|_{L^\infty(\mathbb{R}^n)}+A).
$$
\end{thm}
\end{comment}
\subsection{Known results}
Equations similar to the ones in Theorem \ref{thm:main} were, to the author's knowledge, introduced in \cite{IN10}, where existence and uniqueness is established. It is also shown that the solutions converge to solutions of the $p$-Laplace equation, as $s\to 1$. Similar equations were also studied in \cite{CLM12}, where the focus lies in the asymptotic behaviour as $p\to \infty$. Related equations have also been suggested to be used in image processing and machine learning, see \cite{EDL12} and \cite{EDLL14}.
Recently, in \cite{CKP13} and \cite{CKP14}, H\"older estimates and a Harnack inequality were obtained for weak solutions of a very general class of equations of this type. The difference between these results and the ones in the present paper can be seen as the difference between equations in divergence form and those in non-divergence form in the non-local setting. In other words, their results are more in the flavour of Di Giorgi-Nash-Moser (cf. \cite{DG57}, \cite{Nas58} and \cite{Mos61}) while the results in this paper are more in the flavour of Krylov-Safonov (cf. \cite{KS79}).
In the case $p=2$, corresponding to equations of the form
\begin{equation}\label{eq:fraclap}
\operatorname{PV} \int_{\mathbb{R}^n} \frac{u(x)-u(x+y)}{|y|^{n+2s}}\, dy =f(x),
\end{equation}
a similar development has already taken place. In \cite{Sil06}, a surprisingly simple proof of H\"older estimates for viscosity solutions were given for a very general class of equations corresponding to equations of non-divergence form. An adaptation of the method used therein is used in the present paper. In \cite{Kas09}, H\"older estimates were obtained for weak solutions for a class of equations corresponding to equations of divergence form, including equations of the form \eqref{eq:fraclap}.
Related is also \cite{BCF12a} and \cite{BCF12b}, where another type of degenerate (or singular) non-local equation is studied. H\"older estimates and some higher regularity theory are established. It is also proved that these equations approach the $p$-Laplace equation in the local limit.
\subsection{Comments on the equation}
Let us very briefly point out the difference between the class of equations considered in \cite{CKP13} and \cite{CKP14}, and the class of equations considered here (see also \cite{Sil06} for a similar discussion). There, weak solutions are considered, in the sense that
\begin{equation}\label{eq:weak}
\int_{\mathbb{R}^n}\int_{\mathbb{R}^n}|u(x)-u(y)|^{p-2}(u(x)-u(y))(\phi(x)-\phi(y))G(x,y)\,dx dy=0
\end{equation}
for any $\phi\in C_0^\infty(B_2)$, where $G(x,y)$ behaves like $|x-y|^{-n-sp}$. These solutions arise for instance as minimizers of functionals of the form
$$
\int_{\mathbb{R}^n}\int_{\mathbb{R}^n}|u(x)-u(y)|^{p}G(x,y)\,dx dy.
$$
In the most favorable of situations, we are allowed to change the order of integration and write \eqref{eq:weak} as
$$
\int_{\mathbb{R}^n}\int_{\mathbb{R}^n}|u(x)-u(y)|^{p-2}(u(x)-u(y))(G(x,y)+G(y,x))\phi(x)\,dx dy=0,
$$
and conclude
$$
\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(y)|^{p-2}(u(x)-u(y))(G(x,y)+G(y,x))\,dy=0.
$$
The change of variables $y=z+x$ yields
$$
\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(z+x)|^{p-2}(u(x)-u(z+x))(G(x,z+x)+G(z+x,x))\,dz=0,
$$
or
$$
\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(z+x)|^{p-2}(u(x)-u(z+x))K(x,z)\,dz=0,
$$
where $K(x,z)=G(x,z+x)+G(z+x,x)$. Then necessarily $K(x,z-x)=K(z,x-z)$. Moreover, we are not always allowed to perform the transformations above. Hence, the two types of equations overlap but neither is contained in the other. In other words, the results in \cite{CKP13} and \cite{CKP14} do not always apply to the equations considered in this paper, and vice versa, the results in this paper do not always apply to the equations studied therein.
Another important remark is that the estimates obtained in this paper are not uniform as $s\to 1$, i.e., in the limit in which the equation becomes local. This is also the case in \cite{Sil06}. For fully nonlinear equations of fractional Laplace type, uniform estimates as $s\to 1$ have been obtained (see for instance \cite{CS09}), but they are more involved, and they follow the same strategy as the estimates for fully nonlinear (local) equations.
In our case, if $\phi \in C^2_0$ and $p>2$, then
$$
(1-s)\operatorname{PV} \int_{\mathbb{R}^n}\frac{|\phi(x)-\phi(x+y)|^{p-2}(\phi(x)-\phi(x+y))}{|y|^{n+sp}}\,d y \to -C_{p,n}\Delta_p \phi,
$$
as $s\to 1$. If we instead have a kernel of the form
$$G\left(\frac{y}{|y|}\right)\frac{1}{|y|^{n+sp}},$$
then
\begin{align*}
&(1-s)\operatorname{PV} \int_{\mathbb{R}^n}\frac{|\phi(x)-\phi(x+y)|^{p-2}(\phi(x)-\phi(x+y))G(\frac{y}{|y|})}{|y|^{n+sp}}\,d y \\
&\to -C_{p,n}|\nabla \phi|^{p-2}a_{ij}(\nabla \phi)D_{ij}^2 \phi,
\end{align*}
as $s\to 1$, where the matrix $(a_{ij})(\nabla \phi)$ is positive definite and can be given explicitly as integrals over the sphere in terms of $G$. This type of degenerate (or singular) equations of non-divergence form, remained fairly unstudied until quite recently. Starting with \cite{BID04}, these equations have attracted an increasing amount of attention. See also \cite{Imb11} and \cite{IS13} where $C^\alpha$ and $C^{1,\alpha}$-estimates are established, respectively.
\section{Viscosity solutions}\label{sec:visc}
In this section, we introduce the notion of viscosity solutions (as in \cite{CS09}) and prove that viscosity solutions can be treated almost as classical solutions.
\begin{de} Let $D$ be an open set and let $L$ be as defined in Theorem \ref{thm:main} or Theorem \ref{thm:main2}. A function $u\in L^\infty(\mathbb{R}^n)$ which is upper semicontinuous in ${D}$ is a subsolution of
$$
L u\,\leq C \text{ in $D$}
$$
if the following holds: whenever $x_0\in D$ and $\phi\in C^2({B_r(x_0)})$ for some $r>0$ are such that
$$
\phi(x_0)=u(x_0), \quad \phi(x)\geq u(x) \text{ for $x\in B_r(x_0)\subset D$}
$$
then we have
$$
L\phi_r\, (x_0)\leq 0,
$$
where
$$
\phi_r =\left\{\begin{array}{lr}\phi \text{ in }B_r(x_0),\\
u\text{ in }\mathbb{R}^n\setminus B_r(x_0).
\end{array}\right.
$$
A supersolution is defined similarly and a solution is a function which is both a sub- and a supersolution.
\end{de}
The following result verifies that whenever we can touch a subsolution from above with a $C^2$ function, we can treat the subsolution as classical subsolution. The proof is almost identical to the one of Theorem 2.2 in \cite{CS09}.
\begin{prop}\label{prop:pw} Assume the hypotheses of Theorem \ref{thm:main} or Theorem \ref{thm:main2}. Suppose $Lu\leq C$ in $B_1$ in the viscosity sense and that $x_0\in B_1$ and $\phi\in C^2({B}_r(x_0))$ is such that
$$
\phi(x_0)=u(x_0),\quad \phi(x)\geq u(x)\text{ in }B_r(x_0)\subset B_1,
$$
for some $r>0$. Then $Lu$ is defined pointwise at $x_0$ and $Lu\, (x_0)\leq C$.
\end{prop}
\begin{proof} Since the result is only concerned with the behavior at one fixed point $x_0$, we see that there is no difference between assuming the hypotheses of Theorem \ref{thm:main} or Theorem \ref{thm:main2}. Hence, we give the proof under the hypotheses of Theorem \ref{thm:main}. For $0<s\leq r$, let
$$
\phi_s=\left\{\begin{array}{lr} \phi \text{ in }B_s(x_0),\\
u\text{ in }\mathbb{R}^n\setminus B_s(x_0).
\end{array}\right.
$$
Since $u$ is a viscosity subsolution, $L \phi_s\,(x_0)\leq C$. Now introduce the notation
\begin{align*}
\delta(\phi_s, x,y)=&\frac{1}{2}|\phi_s(x)-\phi_s(x+y)|^{p-2}(\phi_s(x)-\phi_s(x+y))\\&+\frac{1}{2}|\phi_s(x)-\phi_s(x-y)|^{p-2}(\phi_s(x)-\phi_s(x-y)),
\end{align*}
$$
\delta^\pm(\phi_s,x,y) = \max(\pm \delta(\phi_s,x,y),0).
$$
By simply interchanging $y\to-y$ we have
\begin{equation}\label{eq:deltasubsol}
\int_{\mathbb{R}^n}\delta(\phi_s,x_0,y)K(x_0,y)\, dy\leq C,
\end{equation}
since one can easily see that the integral is well defined since $\phi_s$ is $C^2$ near $x_0$. Moreover,
$$\delta(\phi_{s_2},x_0,y)\leq \delta(\phi_{s_1},x_0,y)\leq \delta(u,x_0,y)\text{ for }s_1< s_2< r,
$$
so that
$$
\delta^-(u,x_0,y)\leq |\delta(\phi_r,x_0,y)|.
$$
Since $|\delta(\phi_r,x_0,y)K(x,y)|$ is integrable, so is $\delta^-(u,x_0,y)K(x,y)$. In addition, by \eqref{eq:deltasubsol}
$$
\int_{\mathbb{R}^n}\delta^+(\phi_s,x_0,y)K(x_0,y)\, dy\leq \int_{\mathbb{R}^n}\delta^-(\phi_s,x_0,y)K(x_0,y)\, dy+C.
$$
Thus, for $s_1<s_2$
\begin{align}\label{eq:srineq}
\int_{\mathbb{R}^n}\delta^+(\phi_{s_1},x_0,y)K(x_0,y)\, dy&\leq\int_{\mathbb{R}^n}\delta^-(\phi_{s_1},x_0,y)K(x_0,y)\, dy+C\\
&\leq\int_{\mathbb{R}^n}\delta^-(\phi_{s_2},x_0,y)K(x_0,y)\, dy+C<\infty.\nonumber
\end{align}
Since $\delta^+(\phi_s,x_0,y)\nearrow \delta^+(u,x_0,y)$, the monotone convergence theorem implies
$$
\int_{\mathbb{R}^n}\delta^+(\phi_s,x_0,y)K(x_0,y)\, dy \to \int_{\mathbb{R}^n}\delta^+(u,x_0,y)K(x_0,y)\, dy,
$$
and by \eqref{eq:srineq}
\begin{equation}\label{eq:deltaplus}
\int_{\mathbb{R}^n}\delta^+(u,x_0,y)K(x_0,y)\, dy\leq \int_{\mathbb{R}^n}\delta^-(\phi_s,x_0,y)K(x_0,y)\, dy+C<\infty,
\end{equation}
for any $0<s<r$. We conclude that $\delta^+(u,x_0,y)K(x_0,y)$ is integrable. By \eqref{eq:srineq} and the bounded convergence theorem, we can pass to the limit in the right hand side of \eqref{eq:deltaplus} and obtain
$$
\int_{\mathbb{R}^n}\delta (u,x_0,y)K(x_0,y)\, dy=\lim_{s\to 0}\int_{\mathbb{R}^n}\delta (\phi_s,x_0,y)K(x_0,y)\, dy\leq C.
$$
This implies that $Lu\,(x_0)$ exists in the pointwise sense and $Lu\, (x_0)\leq C$.
\end{proof}
\section{H\"older regularity for constant exponents}
In this section we give the proof of our main theorem for the case of constant $s$ and $p$. This is based on Lemma \ref{lem:key}, sometimes referred to as the oscillation lemma. Throughout this section, $L$ denotes an operator of the form in Theorem \ref{thm:main}, i.e.,
$$
Lu\,(x):=\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))K(x,y)\, dy.
$$
Let us also, by abuse of notation, introduce the function
$$
\beta(x)=\beta(|x|)=\left((1-|x|^2)^+\right)^2.
$$
The exact form of $\beta$ is not important, we could have chosen any radial function which is $C^2$ and zero outside $B_1$ and non-increasing along rays from the origin.
We start with a couple of auxiliary inequalities. Here $a,b\in \mathbb{R}$.
\begin{lem}\label{lem:pineq1} Let $p\geq 2$. Then
$$
\big||a+b|^{p-2}(a+b)-|a|^{p-2}a\big |\leq (p-1)|b|(|a|+|b|)^{p-2}.
$$
\end{lem}
\begin{proof}
We have
\begin{align*}
\big||a+b|^{p-2}(a+b)-|a|^{p-2}a\big|&\leq \int_0^{|b|}\Big| \frac{d}{ds} (|a+s|^{p-2}(a+s)\Big|\,ds\\
&= \int_0^{|b|}(p-1)|a+s|^{p-2}\, ds\\
&\leq (p-1)|b|(|a|+|b|)^{p-2}.
\end{align*}
\end{proof}
\begin{lem}\label{lem:pineq2}
Let $p\in (1,2)$. Then
$$
\big||a+b|^{p-2}(a+b)-|a|^{p-2}a\big|\leq (3^{p-1}+2^{p-1})|b|^{p-1}.
$$
\end{lem}
\begin{proof}
We split the proof into two cases.
\noindent {\bf Case 1: $|a|\leq 2|b|$.} Then
$$
\big| |a+b|^{p-2}(a+b)-|a|^{p-2}a\big|\leq |a+b|^{p-1}+|a|^{p-1}\leq (3^{p-1}+2^{p-1})|b|^{p-1}.
$$
\noindent {\bf Case 2: $|a|> 2|b|$.} Then for $|s|\leq |b|$
$$
|a+s|\geq |a|-|s|>2|b|-|b|=|b|,
$$
so that
$$
\big| |a+b|^{p-2}(a+b)-|a|^{p-2}a\big|\leq \int_0^{|b|}(p-1)|a+s|^{p-2} \, ds \leq (p-1)|b|^{p-1}.
$$
Since $p-1\leq 3^{p-1}+2^{p-1}$, this concludes the proof.
\end{proof}
\begin{lem}\label{lem:pest}
Let $p\geq 2$ and assume $a+b\geq 0$. Then
$$
|a+b|^{p-2}(a+b)\leq 2^{p-2}(|a|^{p-2}a+|b|^{p-2}b).
$$
\end{lem}
\begin{proof} The inequality is trivial for $p=2$ so we assume $p>2$. Since $a+b\geq 0$, $|a|^{p-2}a+|b|^{p-2}b\geq 0$. Without loss of generality we can assume $a>0$ and define $t=b/a$. The statement of the lemma is then equivalent to
$$
|1+t|^{p-2}(1+t)\leq 2^{p-2}(1+|t|^{p-2}t), \text{ for }t\geq -1.
$$
This is trivially true for $t=-1$. Hence we are lead to study the function
$$
f(t):=\frac{|1+t|^{p-2}(1+t)}{1+|t|^{p-2}t}, \text{ for }t> -1.
$$
We find that $f$ has critical points at $t=1$ and $t=0$. In addition,
$$
f(1)=2^{p-2}, \lim_{t\searrow -1} f(t)=0, f(0)=1, \lim_{|t|\to \infty} f(t)=1.
$$
We conclude that $f(t)\leq 2^{p-2}$ for all $t\geq -1$, and the result follows.
\end{proof}
Below we prove that a kernel $K$ behaving like $y^{-n-sp}$ satisfies certain inequalities that might look strange at a first glance, but they are exactly the ones that will appear in the proof of our key lemma later.
\begin{prop}\label{prop:fixp}
Assume $K$ satisfies $K(x,y)=K(x,-y)$ and there exist $\Lambda\geq \lambda>0$, $M>0$ and $\gamma>0$ such that
\begin{align*}
\frac{\lambda}{|y|^{n+sp}}\leq & K(x,y)\leq \frac{\Lambda}{|y|^{n+sp}}, \text{ for } y\in B_2,x\in B_2,\\
0\leq & K(x,y) \leq \frac{M}{|y|^{n+\gamma}}, \text{ for } y\in \mathbb{R}^n\setminus B_\frac14,x\in B_2,
\end{align*}
where $s\in (0,1)$ and $p\in (1,\infty)$. In the case $p<2$ we require additionally $p>1/(1-s)$. Then for any $\delta>0$ there are $1/2\geq k>0$ and $\eta>0$ such that for $p\in (2,\infty)$
\begin{align}\label{eq:kassp2}
&2^{p-2}k^{p-1}\operatorname{PV} \int_{x+y\in B_1}|\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(y+x))K(x,y)\, dy \nonumber \\
&+2^{p-2}\int_{y\in \mathbb{R}^n\setminus B_\frac14}|k\beta(x)+2(|8y|^ \eta-1)|^{p-1} K(x,y)\, dy\\ \nonumber
&+2^{p-1}\int_{y\in \mathbb{R}^n\setminus B_\frac14}(|8y|^ \eta-1)^{p-1} K(x,y)\, dy<2^{1-p}\inf_{A\subset B_2,|A|>\delta}\int_A K(x,y)\,d y
\end{align}
and for $p\in (1/(1-s),2)$
\begin{align}\label{eq:kassp1}
(3^{p-1}+2^{p-1})k^{p-1}\int_{\mathbb{R}^n}|\beta(x)-\beta(x+y)|^{p-1}K(x,y)\, dy\\ \nonumber
+2^{p-1}\int_{\mathbb{R}^n\setminus B_\frac14}(|8y|^\eta-1)^{p-1}K(x,y)\,dy<2^{1-p}\inf_{A\subset B_2,|A|>\delta}\int_A K(x,y)\,d y,
\end{align}
for any $x\in B_{3/4}$.
Here $k$ and $\eta$ depend on $\lambda,\Lambda,M,p,s,\gamma$ and $\delta$.
\end{prop}
\begin{proof}
The proof is split into two different cases.\\
\noindent {\bf Case 1: $p>2$}\\
The first term in the left hand side of \eqref{eq:kassp2} reads
\begin{align*}
&2^{p-2}k^{p-1}\operatorname{PV} \int_{x+y\in B_1}|\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(y+x))K(x,y)\, dy\\
=&2^{p-2}k^{p-1}\operatorname{PV} \int_{x+y\in B_1,y\not\in B_\frac14}|\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(y+x))K(x,y)\, dy\\
&+2^{p-2}k^{p-1}\operatorname{PV} \int_{y\in B_\frac14}|\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(y+x))K(x,y)\, dy\\
&=I_1+I_2.
\end{align*}
Since $\beta$ is uniformly bounded by a constant $C$, we can, using the upper bound on $K$ outside $B_{1/4}$, obtain
\begin{equation}\label{eq:case1est1a}
|I_1|\leq |2kC|^{p-1}\int_{\mathbb{R}^n\setminus B_\frac14} K(x,y)\,dy\leq |2kC|^{p-1} M\int_{\mathbb{R}^n\setminus B_\frac14} \frac{dy}{|y|^{n+\gamma}},
\end{equation}
which is finite and converges to zero as $k\to 0$.
For $I_2$ we proceed as follows
\begin{align*}
I_2&=2^{p-2}k^{p-1}\operatorname{PV} \int_{y\in B_\frac14}|\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(y+x))K(x,y)\, dy\\
&=2^{p-3}k^{p-1}\operatorname{PV} \int_{y\in B_\frac14}|\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(y+x))K(x,y)\,dy \\
&+2^{p-3}k^{p-1}\operatorname{PV} \int_{y\in B_\frac14}|\beta(x)-\beta(-y+x)|^{p-2}(\beta(x)-\beta(-y+x))K(x,y)\,dy.
\end{align*}
Introducing the notation
$$
F=-(\beta(x)-\beta(x-y)),\quad G=(\beta(x)-\beta(x-y))+(\beta(x)-\beta(x+y)),
$$
$I_2$ can be written as
\begin{align*}
&2^{p-3}k^{p-1}\int_{y\in B_\frac14}\left(|F+G|^{p-2}(F+G)-|F|^{p-2}F\right) K(x,y) \,dy\\
&\leq 2^{p-3}k^{p-1}(p-1)\int_{y\in B_\frac14}|G|(|F|+|G|)^{p-2}K(x,y)\,dy,
\end{align*}
by Lemma \ref{lem:pineq1}. Since $\beta$ is $C^2$, $|F|\leq C|y|$ and $|G|\leq C|y|^2$. Invoking the upper bound on $K$ in $B_2$ yields the estimate
\begin{equation}\label{eq:case1est1}
I_2\leq C^{p-1}2^{p-3}k^{p-1}(p-1)\Lambda\int_{y\in B_\frac14}|y|^{p-n-sp}\, dy\leq \frac{C^{p-1}2^{p-3}k^{p-1}(p-1)\Lambda \left(\frac14\right)^{p(1-s)}}{p(1-s)},
\end{equation}
where $C$ only depends on the $C^2$-norm of $\beta$, which is fixed. Clearly the left hand side of \eqref{eq:case1est1} goes to zero as $k\to 0$.
For the rest of the terms in the left hand side we observe first that if $\eta<\gamma/(p-1)$ then from the upper bound on $K$ outside $B_{1/4}$
\begin{equation}\label{eq:case1est2}
\int_{\mathbb{R}^n\setminus B_\frac14}\left(|8y|^\eta-1\right)^{p-1} K(x,y)\, dy\leq M\int_{\mathbb{R}^n\setminus B_\frac14}\left(|8y|^\eta-1\right)^{p-1} \frac{dy }{|y|^{n+\gamma}},
\end{equation}
which is uniformly bounded and tends to zero as $\eta \to 0$, by the dominated convergence theorem.
In addition, since $\beta$ is uniformly bounded by some constant $C>0$ we have
\begin{equation}\label{eq:case1est3}
\int_{\mathbb{R}^n\setminus B_\frac14}|k\beta(x)|^{p-1} K(x,y)\,dy\leq k^{p-1}C^{p-1}M\int_{\mathbb{R}^n\setminus B_\frac14} \frac{dy}{|y|^{n+\gamma}},
\end{equation}
which is finite and converges to zero as $k\to 0$, where we again have used the upper bound on $K$ outside $B_{1/4}$.
Thus, if we choose $\eta$ and $k$ small enough (depending on $\Lambda$, $M$, $p$, $s$ and $\gamma$) we can make all the terms in the left hand side as small as desired.
Now we turn our attention to the right hand side. We have, due to the lower bound on $K$ in $B_2$
$$
2^{1-p}\inf_{A\subset B_2,|A|>\delta}\int_{A}K(x,y)\,dy \geq \frac{2^{1-p}\lambda \delta}{2^{n+sp}}.
$$
Then it is clear that we can choose $\eta$ and $k$, depending only on $\lambda$, $\Lambda$, $M$, $p$, $s$, $\gamma$ and $\delta$, so that the left hand side is larger than the right hand side.
\noindent {\bf Case 2: $1/(1-s)<p<2$}\\
The only difference from the case $p>2$ is the first term in the left hand side. We need to show that for $k$ small enough, the term
$$(3^{p-1}+2^{p-1})k^{p-1}\int_{\mathbb{R}^n}|\beta(x)-\beta(x+y)|^{p-1}K(x,y)\, dy,$$
is small. We split the integral into two parts, one in $B_1$ and one in $\mathbb{R}^n\setminus B_1$. We have $|\beta(x)-\beta(x+y)|\leq C|y|$ for $y\in B_1$ and $|\beta(z)|\leq C$ for all $z\in \mathbb{R}^n$. Hence,
\begin{align}
&(3^{p-1}+2^{p-1})k^{p-1}\int_{B_1}|\beta(x)-\beta(x+y)|^{p-1}K(x,y)\, dy\nonumber \\\label{eq:case2est1}
&\leq \Lambda C^{p-1}(3^{p-1}+2^{p-1})k^{p-1}\int_{B_1}|y|^{p-1-n-sp}\,dy\\\nonumber
&\leq \Lambda C^{p-1}(3^{p-1}+2^{p-1})k^{p-1}\frac{1}{p(1-s)-1},
\end{align}
where we have used the upper bound on $K$ in $B_2$. For the part outside $B_1$ we have
\begin{align}
&(3^{p-1}+2^{p-1})k^{p-1}\int_{\mathbb{R}^n\setminus B_1}|\beta(x)-\beta(x+y)|^{p-1}K(x,y)\, dy\nonumber \\\label{eq:case2est2}
&\leq C^{p-1}M(3^{p-1}+2^{p-1})k^{p-1}\int_{\mathbb{R}^n\setminus B_1}\frac{dy}{|y|^{n+\gamma}}\\\nonumber
&\leq C^{p-1}M(3^{p-1}+2^{p-1})k^{p-1}\gamma^{-1},
\end{align}
from the upper bound on $K$ outside $B_{1/4}$.
By choosing $k$ small (depending on $\Lambda$, $M$, $p$, $s$, $\gamma$) we can make both of these terms as small as desired. Hence, the result follows as in the case $p>2$.
\end{proof}
\begin{rem}
We remark that in the proof above, nothing would change if the exponents would depend on $x$, since $x$ is a fixed point. This is important later when we redo the proof for the case of variable exponents.
\end{rem}
The lemma below is the core of this paper. The proof is an adaptation of the proof of Lemma 4.1 in \cite{Sil06}.
\begin{lem}\label{lem:key} Assume the hypotheses of Proposition \ref{prop:fixp}. Suppose
\begin{align*}
Lu\leq 0\text{ in }B_1,\\
u\leq 1\text{ in }B_1,\\
u(x)\leq 2|2x|^\eta-1\text{ in }\mathbb{R}^n\setminus B_1,\\
|B_1\cap \{u\leq 0\}|>\delta,
\end{align*}
where $\eta$ is as in Proposition \ref{prop:fixp}.
Then $u\leq 1-\theta$ in $B_{1/2}$, where $\theta =\theta(\lambda,\Lambda,M,p,s,\gamma,\delta)>0$.
\end{lem}
\begin{proof} We argue by contradiction. Let
$$
\theta=k\left(\beta(1/2)-\beta(3/4)\right),
$$
where $k$ is as in Proposition \ref{prop:fixp}. If there is $x_0\in B_{1/2}$ such that $u(x_0)>1-\theta$, then
$$
u(x_0)+k\beta(1/2)>1+k\beta(3/4).
$$
Moreover, for any $y\in B_1\setminus B_{3/4}$ there holds
$$
u(x_0)+k\beta(x_0)>u(x_0)+k\beta(1/2)>1+k\beta(3/4)\geq u(y)+k\beta(y).
$$
Hence, the maximum of $u+k\beta$ in $B_1$ is attained inside $B_{3/4}$ and it is strictly larger than 1. Suppose that the maximum is attained at the point $x$.
The rest of the proof is devoted to estimating $L(u+k\beta)\,(x)$ from above and from below in order to obtain a contradiction with Proposition \ref{prop:fixp}. At this point, we remark that $-k\beta+(u+k\beta)(x)$ touches $u$ from above at $x$. Hence, by Proposition \ref{prop:pw}, $Lu\,(x)\leq 0$ in the pointwise sense.
We first estimate $L(u+k\beta)\, (x)$ from below. We split the integrals into two parts and write
\begin{align*}
L(u+k\beta)\, (x)&=\operatorname{PV} \int_{x+y\in B_1}+\int_{x+y\not\in B_1}\\
&=\lim_{r\to 0}\int_{x+y\in B_1,y\not\in B_r}+\int_{x+y\not\in B_1}=\lim_{r\to 0} I_r+I_2,
\end{align*}
where there is no need for the principal value in the second integral, since $x\in B_{3/4}$. Using that $u(x)+k\beta(x)>1$ is the maximum of $u+k\beta$ in $B_1$ we see that the integrand in $I_r$ is non-negative and we have the estimate
\begin{align*}
I_r\geq \int_{A_0}(1-k\beta(x+y))^{p-1}K(x,y)\, dy,
\end{align*}
where
$$A_0=\{x+y\in B_1,\quad u(x+y)\leq 0\}.$$
Since $\beta \leq 1$ and $k\leq 1/2$ we conclude
$$
I_r\geq \frac{1}{2^{p-1}}\inf_{A_0\subset B_2,|A_0|>\delta}\int_{A_0}K(x,y)\,d y.
$$
Now we estimate $I_2$ from below. Using that $u(x)+k\beta(x)>1$ and $u(z)\leq 2|2z|^\eta-1$ for $z\in \mathbb{R}^n\setminus B_1$ and $\beta=0$ in $\mathbb{R}^n\setminus B_1$, we have
\begin{align*}
I_2&\geq \int_{x+y\not\in B_1} 2^{p-1}\Big|1-|2(x+y)|^\eta\Big|^{p-2}(1-|2(x+y)|^\eta) K(x,y)\, dy\\
&\geq 2^{p-1}\int_{y\not\in B_\frac14}\Big|1-\Big|2\left(|y|+\frac34\right)\Big|^\eta\Big|^{p-2}\left(1-\Big|2\left(|y|+\frac34\right)\Big|^\eta\right)K(x,y)\, dy\\
&\geq -2^{p-1}\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p-1} K(x,y)\, dy.
\end{align*}
Adding the two estimates together we can summarize
\begin{align}\label{eq:Lfrombelow}
&L(u+k\beta)\,(x)\geq \\
\nonumber & \frac{1}{2^{p-1}}\inf_{A_0\subset B_2,|A_0|>\delta}\int_{A_0}K(x,y)\,d y-2^{p-1}\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p-1} K(x,y)\, dy.
\end{align}
The next step is to estimate $L(u+k\beta)\, (x)$ from above. This part of the proof is split into two cases: $p\geq 2$ and $p<2$.
\noindent {\bf Case 1: $p\geq 2$}\\
Again we split the integral defining $L(u+k\beta)\, (x)$ into two parts
$$
L(u+k\beta)\, (x)=\operatorname{PV} \int_{x+y\in B_1}+\int_{x+y\not\in B_1}:=I_1+I_2,
$$
where again, there is no need for the principal value in the second integral. We first treat $I_1$ by noting that when $x+y\in B_1$, we know
$$
u(x)+k\beta(x)-u(x+y)-k\beta(x+y)\geq 0,
$$
recalling that $u+k\beta$ attains its maximum (in $B_1$) at $x$.
From Lemma \ref{lem:pest}
\begin{align*}
|u(x)&-u(x+y)+ k\beta(x)-k\beta(x+y)|^{p-2}(u(x)-u(x+y)+k\beta(x)-k\beta(x+y))\leq \\
&2^{p-2}|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))\\
+&2^{p-2}|k\beta(x)-k\beta(x+y)|^{p-2}(k\beta(x)-k\beta(x+y)).
\end{align*}
Hence,
\begin{align*}
I_1&\leq 2^{p-2}\operatorname{PV}\int_{x+y\in B_1} |u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))K(x,y)\,dy \\
&+ 2^{p-2}k^{p-1}\operatorname{PV}\int_{x+y\in B_1} |\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(x+y)) K(x,y)\,dy.
\end{align*}
Now we turn our attention to $I_2$. We note that when $x+y\not\in B_1$, we cannot apply Lemma \ref{lem:pest} directly, but we still have from the hypothesis
$$
u(x)+k\beta(x)>1,\quad u(x+y)+k\beta(x+y)\leq 2|2(x+y)|^\eta-1.
$$
In other words,
$$
u(x)-u(x+y)+k\beta(x)-k\beta(x+y)>2(1-|2(x+y)|^\eta).
$$
By adding the term $2(|2(x+y)|^\eta-1)>0$ to the the expression, we increase the integrand, and we also make the integrand non-negative so that we can, once more, apply Lemma \ref{lem:pest}. It follows that
\begin{align*}
I_2&\leq \int_{x+y\not\in B_1}|u(x)-u(x+y)+k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1)|^{p-2}\times \\
&(u(x)-u(x+y)+k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1))K(x,y)\, dy\\
&\leq 2^{p-2}\int_{x+y\not\in B_1} |u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))K(x,y)\,dy\\
&+2^{p-2}\int_{x+y\not\in B_1}|k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1)|^{p-2}\times \\
&(k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1))K(x,y)\, dy.
\end{align*}
Adding the estimates for $I_1$ and $I_2$ together we arrive at
\begin{align}\nonumber
&L(u+k\beta)\,(x)\leq 2^{p-2}Lu\, (x)\\\nonumber
&+2^{p-2}k^{p-1}\operatorname{PV}\int_{x+y\in B_1} |\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(x+y)) K(x,y)\,dy\\\nonumber
&+2^{p-2}\int_{x+y\not\in B_1}|k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1)|^{p-2}\times\\ \label{eq:Lfromabove1}ß
&(k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1))K(x,y)\, dy\\\nonumber
&\leq 2^{p-2}k^{p-1}\operatorname{PV}\int_{x+y\in B_1} |\beta(x)-\beta(x+y)|^{p-2}(\beta(x)-\beta(x+y)) K(x,y)\,dy\\\nonumber
&+2^{p-2}\int_{x+y\not\in B_1}|k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1)|^{p-1}K(x,y)\, dy,
\end{align}
since $Lu\,(x)\leq 0$.
\noindent {\bf Case 2: $\frac{1}{1-s}<p<2$}\\
From Lemma \ref{lem:pineq2}
\begin{align*}
|u(x)-u(x+y)+k\beta(x)-k\beta(x+y)|^{p-2}(u(x)-u(x+y)+k\beta(x)-k\beta(x+y))\leq \\
|u(x)-u(x+y)|^{p-2}(u(x)-u(x+y))+(3^{p-1}+2^{p-1})k^{p-1}|\beta(x)-\beta(x+y)|^{p-1}
\end{align*}
from which it follows that
\begin{align}
L(u+k\beta)\,(x)&\leq Lu\,(x)+k^{p-1}(3^{p-1}+2^{p-1})\int_{\mathbb{R}^n}|\beta(x)-\beta(x+y)|^{p-1} K(x,y)\,d y\label{eq:Lfromabove2}\\\nonumber
&\leq k^{p-1}(3^{p-1}+2^{p-1})\int_{\mathbb{R}^n}|\beta(x)-\beta(x+y)|^{p-1} K(x,y)\,d y.
\end{align}
Finally, we arrive at a contradiction by observing that \eqref{eq:Lfrombelow} combined with either \eqref{eq:Lfromabove1} or \eqref{eq:Lfromabove2} results in a contradiction with \eqref{eq:kassp2} or \eqref{eq:kassp1} in Proposition \ref{prop:fixp}.
\end{proof}
Once the lemma above is established, the proof of the H\"older regularity is standard. We follow the lines of the proof of Theorem 5.1 in \cite{Sil06}.
\begin{proof}[~Proof of Theorem \ref{thm:main}]We first rescale $u$ by the factor
$$
\frac{1}{2\|u\|_{L^\infty(\mathbb{R}^n)}}.
$$
Then the new $u$ satisfies
$$
L u =0\text{ in }B_1,\quad \operatorname{osc}_{\mathbb{R}^n}u \leq 1.
$$
We will now show that for $j=0,1,\ldots$
$$
\operatorname{osc}_{B_{2^{-j}(x_0)}} u\leq 2^{-j\alpha}, \text{ for any }x_0\in B_1,
$$
where $\alpha$ is chosen so that
$$
\frac{2-\theta}{2}\leq 2^{-\alpha}\text{ and } \alpha\leq \eta,
$$
where $\theta$ is from Lemma \ref{lem:key} and $\eta$ is from Proposition \ref{prop:fixp}, with $\delta=|B_1|/2$. This will imply the desired result with $C=2^{\alpha}$.
In what follows we will find constants $a_j$ and $b_j$ so that
\begin{equation}\label{eq:akbk}
b_j\leq u\leq a_j\text{ in }B_{2^{-j}(x_0)},\quad |a_j-b_j|\leq 2^{-j\alpha}.
\end{equation}
We construct these by induction. For $j\leq 0$, \eqref{eq:akbk} holds true with $b_j=\inf_{\mathbb{R}^n} u $ and $a_j=b_j+1$.
Assume \eqref{eq:akbk} holds for all $j\leq k$. We need to construct $a_{k+1}$ and $b_{k+1}$. Put $m=(a_k+b_k)/2$. Then
$$
|u-m|\leq 2^{-k\alpha-1}\text{ in $B_{2^{-k}}(x_0)$.}
$$
Let
$$
v(x)=2^{\alpha k+1}(u(2^{-k}x+x_0)-m).
$$
Then
$$
\operatorname{PV} \int_{\mathbb{R}^n}|v(x)-v(x+y)|^{p-2}(v(x)-v(x+y))K_{x_0,2^{-k}}(x,y)\, dy= 0 \text{ in }B_1
$$
and
$$
|v|\leq 1 \text{ in }B_1,
$$
where
$$
K_{x_0,2^{-k}}(x,y)=2^{-k(n+sp)}K(2^{-k}x+x_0,2^{-k}y),$$
which satisfies the same assumptions as $K$ itself. We also remark for $|y|>1$ such that $2^\ell\leq |y|\leq 2^{\ell+1}$ we have
\begin{align*}
v(y)= 2^{\alpha k+1}(u(2^{-k}y+x_0)-m)&\leq 2^{\alpha k+1}(a_{k-\ell-1}-m)\\
&\leq 2^{\alpha k+1}(a_{k-\ell-1}-b_{k-\ell-1}+b_{k}-m)\\
&\leq 2^{\alpha k+1}(2^{-\alpha(k-\ell-1)}-\frac12 2^{-k\alpha})\\
&\leq 2^{1+\alpha(\ell+1)}-1\leq 2|2y|^\alpha-1\\
&\leq 2|2y|^\eta-1,
\end{align*}
where we have used that \eqref{eq:akbk} holds for $j\leq k$. Suppose now that \mbox{$|\{v\leq 0\}\cap B_1|\geq |B_1|/2$} (if not we would apply the same procedure to $-v$). Then $v$ satisfies all the assumptions of Lemma \ref{lem:key}, with $\delta =|B_1|/2$ and we obtain
$$
v(x)\leq 1-\theta\text{ in }B_\frac12,
$$
where $\theta=\theta(\lambda,\Lambda,M,p,s,\gamma)$, since $\delta$ is fixed. Scaling back to $u$ this yields
\begin{align*}
u(x)&\leq 2^{-1-\alpha k}(1-\theta)+m\leq 2^{-1-k\alpha}(1-\theta)+\frac{a_k+b_k}{2}\\
&\leq b_k+2^{-1-\alpha k}(1-\theta)+2^{-1-\alpha k}\\
&\leq b_k+2^{-\alpha (k+1)}
\end{align*}
by our choice of $\alpha$. Hence, if we let $b_{k+1}=b_k$ and $a_{k+1}=b_k+2^{-\alpha(k+1)}$ we obtain \eqref{eq:akbk} for the step $j=k+1$ and the induction is complete.
\end{proof}
\section{Variable exponents}
In this section we show that our results also apply to the case when both $p$ and $s$ vary with $x$. In particular we prove Theorem \ref{thm:main2}. Throughout this section $L$ denotes the operator
$$
Lu\,(x):=\operatorname{PV} \int_{\mathbb{R}^n}|u(x)-u(x+y)|^{p(x)-2}(u(x)-u(x+y))K(x,y)\, dy.
$$
We follow the same strategy as in the case of constant exponents and prove slightly modified versions of Proposition \ref{prop:fixp} and Lemma \ref{lem:key}. The proof of H\"older continuity is then similar.
\begin{prop}\label{prop:varp}
Assume $K$ satisfies $K(x,y)=K(x,-y)$ and there exist $\Lambda\geq \lambda>0$, $M>0$ and $\gamma>0$ such that
\begin{align*}
\frac{\lambda}{|y|^{n+s(x)p(x)}}\leq & K(x,y)\leq \frac{\Lambda}{|y|^{n+s(x)p(x)}}, \text{ for } y\in B_2,x\in B_2,\\
0\leq & K(x,y) \leq \frac{M}{|y|^{n+\gamma}}, \text{ for } y\in \mathbb{R}^n\setminus B_\frac14,x\in B_2,
\end{align*}
where $0<s_0<s(x)<s_1<1$ and $1<p_0<p(x)<p_1<\infty$. In the case $p(x)<2$ we require additionally that there is $\tau>0$ such that
$$
p(x)(1-s(x))-1>\tau .$$
Then for any $\delta>0$ there are $1/2 \geq k>0$ and $\eta>0$ such that for $p\in (2,\infty)$
\begin{align}\label{eq:varkassp2}
&2^{p(x)-2}k^{p(x)-1}\operatorname{PV} \int_{x+y\in B_1}|\beta(x)-\beta(x+y)|^{p(x)-2}(\beta(x)-\beta(y+x))K(x,y)\, dy \nonumber \\
&+2^{p(x)-2}\int_{y\in \mathbb{R}^n\setminus B_\frac14}|k\beta(x)+2((|8y|^ \eta-1)|^{p(x)-1} K(x,y)\, dy\\ \nonumber
&+2^{p(x)}\int_{y\in \mathbb{R}^n\setminus B_\frac14}((|8y|^ \eta-1)|^{p(x)-1} K(x,y)\, dy<2^{1-p(x)}\inf_{A\subset B_2,|A|>\delta}\int_A K(x,y)\,d y
\end{align}
and for $p(x)\in (1/(1-s),2)$
\begin{align}\label{eq:varkassp1}
(3^{p(x)-1}+2^{p(x)-1})k^{p(x)-1}\int_{\mathbb{R}^n}|\beta(x)-\beta(x+y)|^{p(x)-1}K(x,y)\, dy\\ \nonumber
+2^{p(x)}\int_{\mathbb{R}^n\setminus B_\frac14}(|8y|^\eta-1)^{p(x)-1}K(x,y)\,dy<2^{1-p(x)}\inf_{A\subset B_2,|A|>\delta}\int_A K(x,y)\,d y,
\end{align}
for any $x\in B_{3/4}$.
Here $k$ and $\eta$ depend on $\lambda,\Lambda,M,p_0,p_1,s_0,s_1,\gamma,\tau$ and $\delta$.
\end{prop}
\begin{proof}
We point out the differences to the proof of Proposition \ref{prop:varp} and briefly explain how they can be dealt with.
\noindent{\bf Case 1: $p(x)\geq 2$}\\
By the exact same computation as in \eqref{eq:case1est1a}, \eqref{eq:case1est1}, \eqref{eq:case1est2} and \eqref{eq:case1est3} in the proof of Proposition \ref{prop:fixp} (since the computation is made for a fixed $x$), we can conclude that the left hand side is bounded by
\begin{align*}
|2kC|^{p-1} M\int_{\mathbb{R}^n\setminus B_\frac14} \frac{dy}{|y|^{n+\gamma}}+\frac{C^{p(x)-1}2^{p(x)-3}k^{p(x)-1}(p(x)-1)\Lambda \left(\frac14\right)^{p(x)(1-s(x))}}{p(x)(1-s(x))}
\end{align*}
plus terms involving the quantities
\begin{align*}
M\int_{\mathbb{R}^n\setminus B_\frac14}\left(|8y|^\eta-1\right)^{p(x)-1} \frac{dy }{|y|^{n+\gamma}}
\end{align*}
and
\begin{align*}
k^{p(x)-1}C^{p(x)-1}2^{p(x)-1}M\int_{\mathbb{R}^n\setminus B_\frac14} \frac{dy}{|y|^{n+\gamma}}.
\end{align*}
Due to the assumptions on $p$ and $s$, the terms are all uniformly bounded. Thus, if we choose $\eta$ and $k$ small enough (depending on $\Lambda$, $M$, $p_0$, $p_1$, $s_0$, $s_1$ and $\gamma$) we can make all the terms in the left hand side as small as desired.
For the right hand side, we again have
$$
2^{1-p(x)}\inf_{A\subset B_2,|A|>\delta}\int_{A}K(x,y)\,dy \geq \frac{2^{1-p(x)}\lambda \delta}{2^{n+s(x)p(x)}}\geq \frac{2^{1-p_1}\lambda \delta}{2^{n+s_1p_1}}.
$$
Then it is clear that we can choose $\eta$ and $k$, depending only on $\lambda$, $\Lambda$, $M$, $p_0$, $p_1$, $s_0$, $s_1$, $\gamma$ and $\delta$, so that the left hand side is larger than the right hand side.
\noindent{\bf Case 2: $1/(1-s(x))<p(x)<2$}\\
We can again estimate the left hand side by
\begin{align*}
&\Lambda C^{p(x)-1}(3^{p(x)-1}+2^{p(x)-1})k^{p(x)-1}\frac{1}{p(x)(1-s(x))-1}\\&+C^{p(x)-1}M(3^{p(x)-1}+2^{p(x)-1})k^{p(x)-1}\gamma^{-1},
\end{align*}
as in \eqref{eq:case2est1} and \eqref{eq:case2est2}. By choosing $k$ small (depending on $\Lambda$, $M$, $p_0$, $p_1$, $\tau$, $\gamma$) we can make both these terms as small as desired. The result follows also in this case.
\end{proof}
\begin{lem}\label{lem:key2} Assume the hypotheses of Proposition \ref{prop:varp}. Suppose
\begin{align*}
Lu\leq \varepsilon\text{ in }B_1,\\
u\leq 1\text{ in }B_1,\\
u(x)\leq 2|2x|^\eta-1\text{ in }\mathbb{R}^n\setminus B_1,\\
|B_1\cap \{u\leq 0\}|>\delta,
\end{align*}
where $\eta$ is as in Proposition \ref{prop:varp} and
\begin{align*}
\varepsilon=\min(2,2^{p(x)-1})\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p(x)-1} K(x,y)\, dy.
\end{align*}
Then $u\leq 1-\theta$ in $B_{1/2}$, where $\theta = \theta(\lambda,\Lambda,M,p_0,p_1,s_0,s_1,\gamma,\tau,\delta)>0$.
\end{lem}
\begin{proof}
The first part of the proof is exactly the same as the one of Proposition \ref{prop:fixp}. Then it comes to estimating $L(u+k\beta)\, (x)$ from below. Since $x$ is a fixed point throughout all the calculations, we obtain as in \eqref{eq:Lfrombelow}
\begin{align}\label{eq:Lfrombelowvar}
&L(u+k\beta)\,(x)\geq \\
\nonumber & \frac{1}{2^{p(x)-1}}\inf_{A_0\subset B_2,|A_0|>\delta}\int_{A_0}K(x,y)\,d y-2^{p(x)-1}\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p(x)-1} K(x,y)\, dy.
\end{align}
The next step is then to estimate $L(u+k\beta)\, (x)$ from above. We obtain almost the same estimate as in \eqref{eq:Lfromabove1} and \eqref{eq:Lfromabove2}. The difference is that we instead of $Lu\,(x)\leq 0$ use $Lu\,(x)\leq \varepsilon$ and obtain an extra term
$$
\min(2,2^{p(x)-1})\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p(x)-1} K(x,y)\, dy.
$$
Hence the estimate reads in the two different cases:\\
\noindent{\bf Case 1: $p(x)\geq 2$}
\begin{align}\nonumber
&L(u+k\beta)\,(x) \\&\label{eq:Lfromabove1var} \leq 2^{p(x)-2}k^{p(x)-1}\operatorname{PV}\int_{x+y\in B_1} |\beta(x)-\beta(x+y)|^{p(x)-2}(\beta(x)-\beta(x+y)) K(x,y)\,dy\\
&+2^{p(x)-2}\int_{x+y\not\in B_1}|k\beta(x)-k\beta(x+y)+2(|2(x+y)|^\eta-1)|^{p(x)-1}K(x,y)\, dy \nonumber
\\
&+2^{p(x)-1}\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p(x)-1} K(x,y)\, dy .\nonumber
\end{align}
\noindent{\bf Case 2: $1/(1-s(x))<p(x)<2$}
\begin{align}
L(u+k\beta)\,(x)&\leq k^{p(x)-1}(3^{p(x)-1}+2^{p-1})\int_{\mathbb{R}^n}|\beta(x)-\beta(x+y)|^{p(x)-1} K(x,y)\,d y \label{eq:Lfromabove2var}\\
&\nonumber +2^{p(x)-1}\int_{y\not\in B_\frac14}(|8y|^\eta-1)^{p(x)-1} K(x,y)\, dy.
\end{align}
The combination of \eqref{eq:Lfrombelowvar} with either \eqref{eq:Lfromabove1var} or \eqref{eq:Lfromabove2var} is a contradiction to \eqref{eq:varkassp2} or \eqref{eq:varkassp1}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:main2}]
The proof is very similar to the proof of Theorem \ref{thm:main}. We first rescale $u$ by the factor
$$
\left(2\|u\|_{L^\infty(\mathbb{R}^n)}+2^{\frac{p_1-1}{p_0-1}}\max\Big\{\left(\frac{\|f\|_{L^\infty(B_2)}}{\varepsilon}\right)^\frac{1}{p_0-1},\left(\frac{\|f\|_{L^\infty(B_2)}}{\varepsilon}\right)^\frac{1}{p_1-1}\Big\}\right)^{-1},
$$
where $\varepsilon$ is chosen as in Lemma \ref{lem:key2} with $\delta=|B_1|/2$. Then one readily verifies that
$$
L u =\tilde f\text{ in }B_2,\quad \|\tilde f\|_{L^\infty(B_2)}\leq \frac{\varepsilon}{2^{p_1-1}},\quad \operatorname{osc}_{\mathbb{R}^n}u \leq 1.
$$
Next we proceed as before: we find $a_j$ and $b_j$ such that
\begin{equation}\label{eq:akbkvar}
b_j\leq u\leq a_j\text{ in }B_{2^{-j}(x_0)},\quad |a_j-b_j|\leq 2^{-j\alpha},
\end{equation}
where we require from $\alpha$ that
$$
\frac{2-\theta}{2}\leq 2^{-\alpha}, \alpha\leq \eta \text{ and } \alpha\leq \frac{s_0p_0}{p_1-1},
$$
where $\beta$ is from Lemma \ref{lem:key2} and $\eta$ from Proposition \ref{prop:varp}, with $\delta =|B_1|/2$. As before, \eqref{eq:akbkvar} is satisfied for $j\leq 0$ with the choice $b_j=\inf_{\mathbb{R}^n} u$ and $a_j=b_j+1$. Now, given that \eqref{eq:akbkvar} holds for $j\leq k$ we construct $a_{k+1}$ and $b_{k+1}$. Define
$$
v(x)=2^{\alpha k+1}(u(2^{-k}x+x_0)-m),\quad \text{ with } m=\frac{a_k+b_k}{2}.
$$
Then
\begin{align*}
&\operatorname{PV} \int_{\mathbb{R}^n}|v(x)-v(x+y)|^{p(x)-2}(v(x)-v(x+y))K_{x_0,2^{-k}}(x,y)\, dy \\
&=2^{(\alpha k+1)(p(2^{-k}x+x_0)-1)-k(s(2^{-k}x+x_0)p(2^{-k}x+x_0))}\tilde f\text{ in }B_1,
\end{align*}
and
$$
|v|\leq 1 \text{ in }B_1.
$$
As before,
$$
K_{x_0,2^{-k}}(x,y)=2^{-k(n+s(2^{-k}x+x_0)p(2^{-k}x+x_0))}K(2^{-k}x+x_0,2^{-k}y)
$$
satisfies the same assumptions as $K$. From our choice of $\alpha$ it also follows that
$$
\Big|2^{(\alpha k+1)(p(2^{-k}x+x_0)-1)-k(s(2^{-k}x+x_0)p(2^{-k}x+x_0))}\tilde f\Big|\leq \varepsilon \text{ in $B_1$}.
$$
Supposing that $|\{v\leq 0\}\cap B_1|\geq |B_1|/2$ and observing that as before
$$
v(y)\leq 2|2y|^\eta-1, \quad \text{ for $|y|>1$},
$$
we see that $v$ satisfies all the assumptions of Lemma \ref{lem:key2}. The choice $\delta = |B_1|/2$ yields
$$
v(x)\leq 1-\theta \text{ in }B_\frac12,$$
which again implies
$$
u(x)\leq b_k+2^{-\alpha(k+1)}.
$$
Thus the choice $b_{k+1}=b_k$ and $a_{k+1}=b_k+2^{-\alpha(k+1)}$ settles \eqref{eq:akbkvar} for the step $j=k+1$. Hence, we arrive at the estimate
$$
\operatorname{osc}_{B_r(x_0)} u\leq 2^\alpha r^\alpha .
$$
Recalling our rescaling factor in the beginning and rescaling back to our original $u$ yields
\begin{align*}
&\operatorname{osc}_{B_r(x_0)} u\\
&\leq 2^\alpha\left(2\|u\|_{L^\infty(\mathbb{R}^n)}+2^{\frac{p_1-1}{p_0-1}}\max\Big\{\left(\frac{\|f\|_{L^\infty(B_2)}}{\varepsilon}\right)^\frac{1}{p_0-1},\left(\frac{\|f\|_{L^\infty(B_2)}}{\varepsilon}\right)^\frac{1}{p_1-1}\Big\}\right)^{-1} r^\alpha \\
&\leq C\left(\|u\|_{L^\infty(\mathbb{R}^n)}+\max\left(\|f\|_{L^\infty(B_2)}^\frac{1}{p_0-1},\|f\|_{L^\infty(B_2)}^\frac{1}{p_1-1}\right)\right) r^\alpha,
\end{align*}
which is the desired result.
\begin{comment}
Put $m=(a_k+b_k)/2$. Then
$$
|u-m|\leq 2^{-k\alpha-1}\text{ in $B_{2^{-k}}(x_0)$.}
$$
Let
$$
v(x)=2^{\alpha k+1}(u(2^{-k}x+x_0)-m).
$$
Then
$$
L_{2^{-k},x_0}v = 0 \text{ in }B_1, \quad |v|\leq 1 \text{ in }B_1.
$$
Suppose for the moment that $|\{v\leq 0\}\cap B_1|\geq |B|/2$ (if not we would apply the same procedure to $-v$), then for $|y|>1$ such that $2^j\leq |y|\leq 2^{j+1}$ we have
\begin{align*}
v(y)= 2^{\alpha k+1}(u(2^{-k}y+x_0)-m)&\leq 2^{\alpha k+1}(a_{k-j-1}-m)\\
&\leq 2^{\alpha k+1}(a_{k-j-1}-b_{k-j-1}+b_{k}-m)\\
&\leq 2^{\alpha k+1}(2^{-\alpha(k-j-1)}-\frac12 2^{-k\alpha})\\
&\leq 2^{1+\alpha(j+1)}-1\leq 2|2y|^\alpha-1.
\end{align*}
Hence, $v$ satisfies all the assumptions of Lemma \ref{lem:key}. With $\delta =|B_1|/2$ we then obtain
$$
v(x)\leq 1-\gamma,\text{ in }B_\frac12.
$$
Scaling back to $u$ this yields
\begin{align*}
u(x)&\leq 2^{-1-\alpha k}(1-\gamma)+m\leq 2^{-1-k\alpha}(1-\gamma)+\frac{a_k+b_k}{2}\\
&\leq b_k+2^{-1-\alpha k}(1-\gamma)+2^{-1-\alpha k}\\
&\leq b_k+2^{-\alpha (k+1)}
\end{align*}
by our choice of $\alpha$. Hence, if we let $b_{k+1}=b_k$ and $a_k+1=b_k+2^{-\alpha(k+1)}$ we obtain \eqref{eq:akbk} for the step $k+1$ and the induction is complete.
\end{comment}
\end{proof}
|
\section{Introduction}\label{sect:introduction}
Birman and Menasco introduced arc-presentation of links in \cite{BM},
and Cromwell formulated it in \cite{C}.
Dynnikov pointed out in \cite{D1} and \cite{D2}
that Cromwell's argument in \cite{C} almost shows
that any arc-presentation of a split link
can be deformed into one which is $\lq\lq$visibly split"
by a finite sequence of elementary moves
which do not changes number of arcs of arc-presentations.
He also showed
that any arc-presentation of the trivial knot
can be deformed into the trivial one with only two arcs
by a finite sequence of merge elementary moves
without increasing number of arcs.
Since there are only finitely many arc-presentations
with a fixed number of edges,
these results give finite algorithms for the decision problems.
As is shown in page 41 in \cite{C},
an arc-presentation is almost equivalent to a rectangular diagram.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=45mm]{s-TrivialKnot8arcs.eps}
\end{center}
\caption{A rectangular diagram of the trivial knot with $8$ vertical edges}
\label{fig:TrivialKnot8arcs}
\end{figure}
A {\it rectangular diagram} of a link
is a link diagram in the plane ${\mathbb R}^2$
which is composed of vertical line segments and horizontal line segments
such that no pair of vertical line segments are colinear,
no pair of horizontal line segments are colinear,
and the vertical line segment passes over the horizontal line segment
at each crossing.
See Figure \ref{fig:TrivialKnot8arcs}.
These vertical line segments and horizontal line segments
are called {\it edges} of the rectangular diagram.
Every rectangular diagram
has the same number of vertical edges and horizontal edges.
It is known that every link has a rectangular diagram
(Proposition in page 42 in \cite{C}).
In \cite{HK}, A. Henrich and L. Kauffman announced
an upper bound of the number of Reidemeister moves
needed for unknotting (Theorem 8)
by applying Dynnikov's theorem to rectangular diagrams.
Let $D$ be an oriented link diagram in Morse form,
and $c(D)$, $b(D)$ the numbers of crossings and maxima.
Lemma 2 in \cite{HK} states
that we can obtain a rectangular diagram
with at most $2b(D)+c(D)$ vertical edges
from $D$ by an ambient isotopy of the plane ${\mathbb R}^2$.
In this paper,
we consider the number of Seifert circles $s(D)$
instead of that of maxima.
Note that $s(D)$ does not change under isotopy of ${\mathbb R}^2$.
We obtain an estimation depending only on $c(D)$, too.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-smoothing.eps}
\end{center}
\caption{smoothing operation}
\label{fig:smoothing}
\end{figure}
Let $D$ be an oriented link diagram in the plane ${\mathbb R}^2$.
If we perform smoothing operations at all the crossings
as shown in Figure \ref{fig:smoothing},
then we obtain a disjoint union of oriented circles in ${\mathbb R}^2$
as in Figure \ref{fig:Seifert},
which we call {\it Seifert circles}.
This operation is introduced by Seifert in \cite{S}
to construct an orientable surface spanning a knot.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-Seifert.eps}
\end{center}
\caption{Seifert circles}
\label{fig:Seifert}
\end{figure}
A crossing $x$ of a link diagram $D$ in the plane ${\mathbb R}^2$
is called {\it nugatory}
if there is a circle $C$ in ${\mathbb R}^2$
which intersects $D$ only in a single point at $x$.
See Figure \ref{fig:nugatory}.
Link diagrams with a nugatory crossing are often left out from consideration
since we can get rid of a nugatory crossing
by rotating the part of the link inside $C$ through $180^{\circ}$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=70mm]{s-nugatory.eps}
\end{center}
\caption{a nugatory crossing}
\label{fig:nugatory}
\end{figure}
\begin{theorem}\label{theorem:c+2s}
Let $D$ be an oriented link diagram in the plane ${\mathbb R}^2$,
and $c(D)$, $s(D)$ the numbers of crossings and Seifert circles of $D$
respectively.
Then an adequate ambient isotopy of ${\mathbb R}^2$
deforms $D$ into a rectangular diagram
with at most $c(D) + 2 s(D)$ vertical edges.
If $c(D) \ge 1$ and $D$ has no nugatory crossings,
then $c(D) + 2 s(D) -2$ vertical edges are enough.
\end{theorem}
A link diagram $D$ in the plane ${\mathbb R}^2$
is said to be {\it connected}
if it is connected
when the underpasses are restored at all the crossings.
An {\it undirected smoothing operation} at a crossing
is a smoothing operations neglecting orientation of a ink.
It may or may not respect the orientation of the link
when we orient the link.
By an adequate undirected smoothing operations,
as shown in Figure \ref{fig:RMCAS},
we obtain a single circle from any connected link diagram (Lemma \ref{lemma:monadic}).
This leads to the next theorem.
\begin{theorem}\label{theorem:2c}
Let $D$ be a connected link diagram in ${\mathbb R}^2$,
and $c(D)$ the number of crossings of $D$.
Then $D$ can be deformed
into a rectangular diagram with at most $2c(D) - w(D)+2$ vertical edges
by an adequate ambient isotopy of ${\mathbb R}^2$,
where $w(d)$ is a non-negative integer defined as the width of $D$ as below.
\end{theorem}
In order to show Theorems \ref{theorem:c+2s} and \ref{theorem:2c},
we observe a system of circles obtained
by performing smoothing operations
at all the crossings of a link diagram
and arcs corresponding to crossings.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=110mm]{s-smoothing2.eps}
\end{center}
\caption{Seifert circles and arcs system}
\label{fig:smoothing2}
\end{figure}
After we perform smoothing operations at all the crossings,
we place a line segment connecting Seifert circles
as a substitute for each crossing as shown in Figure \ref{fig:smoothing2} (1).
Then we obtain a union of circles and arcs,
which we call {\it Seifert circles and arcs system}.
In Figure \ref{fig:smoothing2} (2),
the one which is obtained
from the knot diagram in Figure \ref{fig:Seifert} is depicted.
Note that Seifert circles and arcs system does not have an arc
with its both endpoints in the same Seifert circle.
(Otherwise, we would have a contradiction on orientation of the link.)
Moreover,
the orientations of the two circles containing the endpoints of an arc
are both clockwise or both anti-clockwise
if and only if one circle is contained in the disk bounded by the other.
When we apply undirected smoothing operations,
we obtain a system of circles and arcs,
where circles do not have orientations,
and there may be an arc
which has its both endpoints in the same circle.
In general,
let $C$ be a disjoint union of circles in the plane ${\mathbb R}^2$,
and $A$ a disjoint union of arcs in ${\mathbb R}^2$
such that $A \cap C = \partial A$,
where $\partial A$ denotes the set of endpoints of arcs of $A$.
Then we call the union $C \cup A$
{\it circles and arcs system} in ${\mathbb R}^2$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-depth.eps}
\end{center}
\caption{depth}
\label{fig:depth}
\end{figure}
Let $S = C \cup A$ be a circles and arcs system in ${\mathbb R}^2$.
We divide ${\mathbb R}^2$ into regions by the circles of $C$.
A circle $Z$ of $C$ is said to be of {\it depth} $1$
if $Z$ is contained in the boundary of the infinitely large region,
and of {\it depth} $i$
if $Z$ is contained in the boundary of a region $R$
such that its boundary $\partial R$ contains a single circle, say $Z'$, of depth $i-1$
and $R$ is inside $Z'$.
See Figure \ref{fig:depth}.
Every circles and arcs system can be moved to be $\lq\lq$beautiful"
as shown in the next theorem.
Let $\pi_x: {\mathbb R}^2 \ni (x,y) \mapsto (x,0) \in {\mathbb R}^2$
and $\pi_y: {\mathbb R}^2 \ni (x,y) \mapsto (0,y) \in {\mathbb R}^2$
be projections.
We say subsets $A$ and $B$ of ${\mathbb R}^2$
{\it overlap each other under $\pi_x$ (resp. $\pi_y$)}
if $\pi_x(A) \cap \pi_x(B) \ne \emptyset$ (resp. $\pi_y(A) \cap \pi_y(B) \ne \emptyset$).
\begin{theorem}\label{theorem:RCAS}
Let $C \cup A$ be a circles and arcs system in ${\mathbb R}^2$,
where $C$ is the union of circles and $A$ the union of arcs.
Then it can be deformed by an ambient isotopy of ${\mathbb R}^2$
so that
(1) circles are rectangles composed
of two vertical line segments and two horizontal line segments, and
(2) each arc either (a) is a vertical line segment,
or
(b) has its both endpoints in the same circle
and is composed of three line segments:
one horizontal line segment $s$
and two vertical line segments in the same side of $s$.
Moreover,
the isotopy can be taken
so that no pair of rectangular circles of the same depth
overlap each other under $\pi_y$.
\end{theorem}
We call a circles and arcs system {\it rectangular}
if it satisfies the conditions (1) and (2) in the above theorem.
For an example of rectangular circles and arcs system,
see Figure \ref{fig:RCAS}.
We say that an arc of $A$ as in (2)(a) in the theorem is of {\it type I}
and an arc as in (2)(b) is of {\it type $\sqcup$}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-RCAS.eps}
\end{center}
\caption{rectangular circles and arcs system}
\label{fig:RCAS}
\end{figure}
\begin{corollary}\label{corollary:RSCAS}
Let $D$ be an oriented link diagram in the plane ${\mathbb R}^2$.
The Seifert circles and arcs system for $D$
can be deformed by an ambient isotopy of ${\mathbb R}^2$
so that Seifert circles are
rectangles composed
of two vertical line segments and two horizontal line segments
and arcs are vertical line segments.
Moreover,
the isotopy can be taken
so that no pair of rectangular circles of the same depth
overlap each other under $\pi_y$.
\end{corollary}
For an example of deformation as in the above corollary,
see Figure \ref{fig:RSCAS}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=110mm]{s-RSCAS.eps}
\end{center}
\caption{deforming a Seifert circles and arcs system to be rectangular}
\label{fig:RSCAS}
\end{figure}
A cirlces and arcs system is called {\it monadic}
if it has only one circle.
Every connected link diagram admits a system of smoothing operations
which yields a monadic circles and arcs system $S$.
This is shown in Lemma \ref{lemma:monadic}.
Let $S = C \cup A$ be a monadic circles and arcs system.
An arc $\beta$ with $\beta \cap C = \partial \beta$ is called a {\it ruler}
if it
is free from the endpoints $\partial A$,
is contained in the disk bounded by $C$,
and intersects every arc of $A$ transversely in at most one point.
The number of intersection points of $\beta$ and $A$ is called the {\it length} of $\beta$.
Then the {\it width} of $S$ is the maximal number among lengths of all rulers for $S$,
and let $w(S)$ denote it.
The width of a connected link diagram $D$, denoted by $w(D)$,
is the maximal width $w(S)$
over all systems of undirected smooth operations on $D$
yielding a monadic circles and arcs system $S$.
The next is a corollary of Theorem \ref{theorem:RCAS} and Lemma \ref{lemma:monadic}.
\begin{corollary}\label{corollary:RMCAS}
Let $D$ be a link diagram in the plane ${\mathbb R}^2$.
Then an adequate system of undirected smoothing operations
and an adequate ambient isotopy of ${\mathbb R}^2$
deform $D$ into a monadic circles and arcs system $C \cup A$
such that (1) the circle $C$ is a rectangle composed
of two vertical line segments and two horizontal line segments,
and (2) each arc of $A$ is of type I or $\sqcup$.
\end{corollary}
For an example of deformation as in the above corollary,
see Figure \ref{fig:RMCAS}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-RMCAS.eps}
\end{center}
\caption{deforming a link diagram to a rectangular monadic circles and arcs system}
\label{fig:RMCAS}
\end{figure}
\section{Proof of Theorem \ref{theorem:2c}}\label{section:2c}
\begin{lemma}\label{lemma:monadic}
From any connected link diagram $D$,
we can obtain a monadic circles and arcs system
by adequately applying undirected smoothing operations
at all the crossings of $D$.
\end{lemma}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=70mm]{s-MakeMonadic.eps}
\end{center}
\caption{changing the way of undirected smoothing at a crossing}
\label{fig:MakeMonadic}
\end{figure}
\begin{proof}
We apply arbitrary undirected smoothing operations at all the crossings of $D$,
to obtain a circles and arcs system $S = C \cup A$.
Note that $S = C \cup A$ is connected because $D$ is connected.
If the set of circles $C$ consists of a single circle, we are done.
If it contains plural circles,
then there is an arc of $A$ which connects two distinct circles of $C$.
We change the way of undirected smoothing at the crossing corresponding the arc,
and obtain a circles and arcs system with one less circles.
See Figure \ref{fig:MakeMonadic}.
Repeating this, we obtain a monadic circles and arcs system.
\end{proof}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=125mm]{s-s-merge.eps}
\end{center}
\caption{straight merge}
\label{fig:s-merge}
\end{figure}
\begin{lemma}\label{lemma:s-merge}
Let $R$ be a rectangular diagram with $n$ vertical edges.
Suppose that $R$ has a vertical edge $e$
such that $e$ is free from the crossings of $R$,
and that the horizontal edges $f, g$ sharing an endpoint with $e$
are in the opposite sides of $e$ to each other.
Then an adequate ambient isotopy of ${\mathbb R}^2$
deforms $f \cup e \cup g$ into a single horizontal edge,
and $R$ into a rectangular diagram with $n-1$ vertical edges.
\end{lemma}
We call the ambient isotopy as in the proof below a {\it straight merge} operation at $e$.
A similar thing holds also when $e$ is horizontal.
\begin{proof}
Let $y_f$, $y_g$ be the ordinates of $f$, $g$ respectively,
and $x_e$ the abscissa of $e$.
We assume, without loss of generality, that $y_f > y_g$,
and that $f$ is in the left side of $e$ and $g$ is in the right side of $e$.
We take a small positive real number $\epsilon$
so that there is no horizontal edge of $R$
at any ordinate in $(y_g -\epsilon, y_g) \cup (y_g, y_g+\epsilon) \cup (y_f, y_f+\epsilon)$
and that there is no vertical edge of $R$
at any abscissa in $(x_e, x_e+\epsilon)$.
Then there is an ambient isotopy of ${\mathbb R}^2$ as in the conclusion of this lemma
which shrinks
the area $[y_g+\epsilon, y_f+\epsilon] \times [x_e+\epsilon, \infty)$
in the vertical direction
so that it is deformed
into the area $[y_f+(\epsilon/2), y_f+\epsilon] \times [x_e+\epsilon, \infty)$
and expands
the area $[y_g-\epsilon, y_g+\epsilon] \times [x_e+\epsilon, \infty)$
in the vertical direction
so that it is deformed
into the area $[y_g-\epsilon, y_f+(\epsilon/2)] \times [x_e+\epsilon, \infty)$.
See Figure \ref{fig:s-merge}.
\end{proof}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=70mm]{s-restore.eps}
\end{center}
\caption{restoring the link diagram}
\label{fig:restore}
\end{figure}
\begin{proof}
We prove Theorem \ref{theorem:2c}.
Let $D$ be a connected link diagram,
and $c(D)$ the number of crossings of $D$.
We perform undirected smoothing operations at all the crossings
so that we obtain a monadic circles and arcs system $S=C\cup A$
with $w(D) = w(S)$.
Let $\beta$ be a ruler for $S$
such that its length gives the width $w(D)$.
We can isotope $S \cup \beta$
so that (1) $\beta$ is a horizontal line segment,
that (2) the circle $C$ forms the boundary circle of a tubular neighbourhood of $\beta$
and consists of two horizontal line segments and two vertical line segments,
and that (3) arcs intersecting $\beta$ in a single point are vertical line segments
and the other arcs of $A$
consist of two vertical line segments and a single horizontal line segment.
Then we restore the diagram $D$ from $S$
by replacing arcs of $A$ with crossings as shown in Figure \ref{fig:restore}.
For each of $w(D)$ vertical arcs of $A$,
we use a vertical line segment as an overpass,
and a union of two vertical line segments and a single horizontal line segment as an underpass
to form a crossing.
Note that the two vertical lines are free from the crossing,
and horizontal lines with which they share endpoints are in the opposite sides of them.
For each of the other arcs of $A$,
we use a union of two vertical line segments and a single horizontal line segment
as both of an overpass and an underpass.
Note that the three vertical lines are free from the crossing.
For two of them,
horizontal lines with which they share endpoints are in the opposite sides of them.
Thus we obtain a link diagram
composed of
$2+3w(D)+4(c(D)-w(D))=2+4c(D)-w(D)$ vertical line segments.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=70mm]{s-SmergeAtVerticalLine.eps}
\end{center}
\caption{straight merge operations at vertical line segments
in an underpass}
\label{fig:SmergeAtVerticalLine}
\end{figure}
By slightly perturbing ordinates of horizontal line segments,
and abscissa of vertical line segments,
we obtain a rectangular diagram with
$2+4c(D)-w(D)$ vertical edges.
Then we perform straight merge operations at vertical edges as above,
twice per each crossing
as shown in Figure \ref{fig:SmergeAtVerticalLine},
to obtain a rectangular diagram
with $(2+4c(D)-w(D))-2c(D)=2+2c(D)-w(D)$ vertical edges.
\end{proof}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-notation.eps}
\end{center}
\caption{Crossings are denoted by arcs with a slash.}
\label{fig:notation}
\end{figure}
In the rest of this section,
crossings are denoted by arcs with a slash
as shown in Figure \ref{fig:notation}
when we need to specify
over or under information of them.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=100mm]{s-MergeAtHorizontalEdge.eps}
\end{center}
\caption{horizontal line segments at which we perform straight merge operations}
\label{fig:MergeAtHorizontalEdge}
\end{figure}
After applying straight merge operations at vertical line segments
twice per each crossing as in the above proof,
we can perform straight merge operations
at certain kinds of horizontal line segments
if there are.
Let $C_h$ be the union of the top and bottom horizontal line segments of $C$.
The endpoints of $A$ divide $C_h$ into shorter horizontal line segments.
Among them,
those which are depicted in bold lines in Figure \ref{fig:MergeAtHorizontalEdge}
are available for straight merge operations.
For type (5), for example,
the operation is performed as shown in Figure \ref{fig:MergeAtHorizontalEdge2}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-MergeAtHorizontalEdge2.eps}
\end{center}
\caption{a straight merge at a horizontal line segment}
\label{fig:MergeAtHorizontalEdge2}
\end{figure}
In case of the circles and arcs system in Figure \ref{fig:cannot},
we cannot apply straight merge operation
at any horizontal line segment
after the straight merge operations at vertical line segments.
(However, the resulting rectangular diagram is not minimal
with respect to the number of vertical edges.)
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-cannot.eps}
\end{center}
\caption{We cannot apply a straight merge at any horizontal line segment to this system.}
\label{fig:cannot}
\end{figure}
If there is an arc of type $\sqcup$ inside $C$,
and the link diagram does not have a monogon face,
then we can apply a straight merge at a horizontal line segment
as in the lemma below.
Let $\alpha$ be an arc of type $\sqcup$ inside $C$ (resp. outside $C$),
and $Q$ the disk bounded by the circle
formed by $\alpha$ and a subarc of $C_h$.
We call $\alpha$ is {\it innermost}
among the arcs of type $\sqcup$ inside $C$ (resp. outside $C$)
if $Q$ does not contain such an arc other than $\alpha$.
\begin{lemma}\label{lemma:InnermostSqcup}
{\rm
Let $D, S, C, A$ as in the proof of Theorem \ref{theorem:2c},
$C_h$ as above,
and $m$ the number of innermost arcs of type $\sqcup$ inside $C$.
Suppose that $D$ has no monogon region.
After performing straight merge operations
at vertical line segments twice per each arc of $A$,
we can apply at least $m$ straight merge operations
at hirozontal line segments as in Figure \ref{fig:MergeAtHorizontalEdge} (6).
}
\end{lemma}
\begin{proof}
Let $\alpha$ be an arc of type $\sqcup$ which is innermost inside $C$,
and $\beta$ the subarc of $C_h$ between the endpoints $\partial \alpha$.
If int\,$\beta$ is disjoint from endpoints of $A$,
then $D$ has a monogon region, a contradiction.
Hence int\,$\beta$ contains an endpoint of $A$.
Let $\gamma$ be an innermost arc of $A$ outside $C$
with at least one of its endpoints in $\beta$.
If $\gamma$ has both its endpoints in $\beta$,
then $\gamma$ gives a monogon region of $D$,
which is a contradiction.
Hence $\gamma$ has precisely one of its endpoints in $\beta$.
Then $\alpha$ and $\gamma$ together form the pattern (6) in Figure \ref{fig:MergeAtHorizontalEdge},
and we can apply a straight merge operation there.
\end{proof}
\section{Proof of Theorem \ref{theorem:c+2s}}\label{section:c+2s}
In this section,
we prove Theorem \ref{theorem:c+2s}
using Corollary \ref{corollary:RSCAS}.
(Corollary \ref{corollary:RSCAS} is a corollary of Theorem \ref{theorem:RCAS}
which is proven in the next section.)
\begin{proof}
We prove Theorem \ref{theorem:c+2s}.
Let $D$ be an oriented link diagram,
and $c(D)$ and $s(D)$ the numbers of crossings and Seifert circles of $D$
respectively.
We perform smoothing operations at all the crossings,
so that we obtain a Seifert circles and arcs system $S=C \cup A$.
Using Corollary \ref{corollary:RSCAS},
we can deform $S$
by an ambient isotopy of ${\mathbb R}^2$
into a rectangular Seifert circles and arcs system $R$.
We restore the diagram $D$ from $R$
by replacing arcs of $A$ with crossings
as shown in the upper half of Figure \ref{fig:restore}.
The resulting link diagram has $2s(D)+3c(D)$ vertical line segments.
Then we perform straight merge operations
as described in the proof of Lemma \ref{lemma:s-merge}
twice per each crossing
as shown in Figure \ref{fig:SmergeAtVerticalLine}.
Note that each straight merge operation
decreases the number of vertical line segments by one.
After an adequate small ambient isotopy of ${\mathbb R}^2$,
no pair of line segments are colinear.
Thus we obtain a rectangular diagram with $2s(D)+3c(D)-2c(D)=2s(D)+c(D)$ vertical edges.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-MergesAtLeftmostArc.eps}
\end{center}
\caption{straight merge or generalized merge}
\label{fig:MergesAtLeftmostArc}
\end{figure}
We consider the case
where $c(D) \ge 1$ and $D$ has no nugatory crossings.
If $c(D)=1$, then $D$ would have a nugatory crossing.
Hence $c(D) \ge 2$.
We observe the leftmost arc $\alpha_l$ of $A$.
In the argument in the previous paragraph,
we have performed straight merge operations twice
at two vertical line segments in underpass of the crossing
corresponding to $\alpha_l$.
The second arrows in Figure \ref{fig:MergesAtLeftmostArc}
show the deformation as above.
Now, we perform deformations
described by the third and the forth arrows in Figure \ref{fig:MergesAtLeftmostArc}.
Note that $\alpha_l$ has its endpoints in distinct circles of $C$
since $R$ is a Seifert circles and arcs system.
In Cases (1) and (2) in Figure \ref{fig:MergesAtLeftmostArc},
$\alpha_l$ has its endpoints in circles of the same depth,
while it does not in Cases (3) and (4).
In Cases (1), (2) and (4) in Figure \ref{fig:MergesAtLeftmostArc},
we apply a straight merge operation at a horizontal line segment drawn in a bold line.
In Case (3) in Figure \ref{fig:MergesAtLeftmostArc},
we shrink circles of $C$
which overlap $\alpha_l$ under $\pi_x$ and do not contain an endpoint of $\alpha_l$
to the right direction
so that they are in the right side of the vertical line segment
forming the overpass of the crossing corresponding $\alpha_l$,
and then,
we perform a $\lq\lq$generalized merge operation"
as shown by the forth arrow.
We do a similar operation for the rightmost arc of $A$,
to obtain a rectangular diagram with $2s(L)+c(L)-2$ vertical edges.
\end{proof}
\section{Rectangular circles and arcs system}\label{section:RCAS}
We prove Theorem \ref{theorem:RCAS} in this section.
\begin{definition}\label{definition:BlackBox}
Let $S= C \cup A$ be a rectangular circles and arcs system
with $C$ being the union of circles and $A$ the union of arcs.
Let $R$ be a rectangular disk
whose boundary circle $\partial R$ is composed of two vertical line segments
and two horizontal line segments.
We call $S \cap R$ a {\it black box} in $S$
if $\partial R \cap C = \emptyset$,
the vertical lines of $\partial R$ are disjoint from $S$,
the horizontal lines of $\partial R$ intersect $A$ transversely
in one or more points.
\end{definition}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=60mm]{s-BlackBoxLemma.eps}
\end{center}
\caption{Deformation in Lemma \ref{lemma:BlackBox}}
\label{fig:BlackBoxLemma}
\end{figure}
\begin{lemma}\label{lemma:BlackBox}
Let $S, C, A, R$ be as in Definition \ref{definition:BlackBox} above.
Let $\tilde{R}$ be a rectangular disk
such that it forms a regular neighbourhood of $R$,
that its boundary circle $\partial{\tilde{R}}$
is composed of two vertical line segments and two horizontal line segments,
and that $S \cap (\tilde{R}-{\rm int}\,R)$ consists of vertical line segments,
say, $\gamma_1, \gamma_2, \cdots, \gamma_k$ from the right
above the top horizontal line segment of $\partial R$,
and $\alpha_1, \alpha_2, \cdots, \alpha_j, \beta_1, \beta_2, \cdots, \beta_i$
from the right
below the bottom horizontal line segment of $\partial R$
for some non-negative integers $i,k$ and a positive integer $j$.
Suppose that there is no vertical line segment in $A$
connecting
one of $\gamma_1, \cdots, \gamma_k$
and one of $\alpha_1, \cdots, \alpha_j, \beta_1, \cdots, \beta_i$,
and that no pair among $\alpha_1, \cdots, \alpha_j, \beta_1, \cdots, \beta_i$ is contained
in the same arc of type $\sqcup$.
See Figure \ref{fig:BlackBoxLemma}.
Then there is an ambient isotopy $H: {\mathbb R}^2 \times [0,1] \rightarrow {\mathbb R}^2$
with a homeomorphism $H_t : {\mathbb R}^2 \ni x \mapsto H_t(x)=H((x,t)) \in {\mathbb R}^2$
fot all $t \in [0,1]$
such that $H_t (R) = R$ and $H_t(\tilde{R})=\tilde{R}$ for all $t \in [0,1]$,
that $H_t (p) = p$
for all points $p$ in the left vertical line segments of $\partial R \cup \partial \tilde{R}$
and for all $t \in [0,1]$,
that $H_1(\alpha_m)$ is a vertical line segment
above the top horizontal line segment of $\partial R$
for all $m \in \{ 1, 2, \cdots, j \}$,
and that $H_1(S \cap R)$ is a black box
in some rectangular circles and arcs system.
\end{lemma}
\begin{remark}
This lemma does not mention deformation of $S$ outside $\tilde{R}$
which should occur accompanied by the deformation within $\tilde{R}$.
The isotopy in this lemma may not keep $S$ rectangular.
\end{remark}
\begin{proof}
We say that
an ambient isotopy $H: {\mathbb R}^2 \times [0,1] \rightarrow {\mathbb R}^2$ is {\it good}
if $H_t (R) = R$ and $H_t(\tilde{R})=\tilde{R}$ for all $t \in [0,1]$,
and $H_t (p) = p$
for all points $p$ in the vertical line segments of $\partial R \cup \partial \tilde{R}$
and for all $t \in [0,1]$.
We can move the circles and arcs system $S$
by a good ambient isotopy of ${\mathbb R}^2$
so that the rectangular circles of $C$ are thinned in the vertical direction,
and no pair of rectangular circles of the same depth overlap under $\pi_y$,
since the arcs of $A$ intersects $C$
in endpoints of vertical line segments in $A$.
It is enough to show this lemma in the case $j=1$.
Applying the result for the case $j=1$ repeatedly,
we obtain the desired conclusion also for the case $j>1$.
Set $X = S \cap R$, the black box.
The vertical line segment of $A$ containing $\alpha_1$ has an endpoint $p$
in the bottom horizontal line segment $b$
of some circle $Z$ of $C$ in $X$.
The point $p$ divides $b$ into two segments,
one of which, say $b_r$, lies in the right side of $p$.
Let $t$ be the top horizontal line segment of $Z$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-nCX1.eps}
\end{center}
\caption{The case $c(X)=1$}
\label{fig:nCX1}
\end{figure}
Let $n_C(X)$ denote the number of circles of $C \cap X$.
The proof proceeds by induction on $n_C(X)$.
We first consider the case $n_C(X)=1$.
Then $Z$ is the only circle of $C$ contained in $X$.
We move arcs of $A$ by a good ambient isotopy of ${\mathbb R}^2$.
If the interior of $b_r$ contains endpoints of arcs of $A$,
we move the arcs near the endpoints
along a subarc of $b_r$ and the right vertical line segment of $Z$
so that the endpoints are contained in the top horizontal line segment of $Z$.
See Figure \ref{fig:nCX1}.
Because $j=1>0$,
no arc of $A$ outside $Z$ has
an endpoint in $t$
and the other in $b-b_r$ or below the bottom horizontal line segment of $\partial R$.
Any arc with both endpoints in $t$
cobounds a disk with a subarc of $t$ inside or outside $Z$.
Then we move arcs of $A$
so that $A \cap R$ is a union of arcs of types I and $\sqcup$.
There is no obstruction because $n_C(X)=1$.
Thus we can assume
that int\,$b_r$ does not contain such an endpoint of $A$.
Then we can move $\alpha_1$ so that it is above the top horizontal line segment of $Z$
as shown in Figure \ref{fig:nCX1a1},
which shows the lemma in the case of $n_C(X)=1$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-nCX1a1.eps}
\end{center}
\caption{Bring $\alpha_1$ above $t$}
\label{fig:nCX1a1}
\end{figure}
We consider the case $n_C(X) > 1$.
Let $y_t, y_b$ (resp. $\eta_t, \eta_b$) be
the ordinates of the top and the bottom horizontal line segments of $Z$ (resp. $\partial R$),
and $x_l, x_r$ (resp. $\xi_l, \xi_r$)
the abscissae of the left and the right vertical line segments of $Z$ (resp. $\partial R$).
Let $x_{\alpha}$ be the abscissa of $\alpha_1$.
We consider rectangles
\newline
$R_a=[\xi_l +\epsilon, \xi_r -\epsilon] \times [y_t +\epsilon, \eta_t -\epsilon]$
and
$R_b=[x_{\alpha} +\epsilon, \xi_r -\epsilon] \times [\eta_b +\epsilon, y_b -\epsilon]$,
and black boxes $X_a = S \cap R_a$ and $X_b = S \cap R_b$,
where $\epsilon$ is a small positive real number
such that there is no horizontal line segment in $S$ with ordinate
in the union of open intervals
$(\eta_b, \eta_b +\epsilon) \cup (y_b -\epsilon,y_b)
\cup (y_t, y_t +\epsilon) \cup (\eta_t -\epsilon, \eta_t)$
and there is no vertical line segment in $S$ with abscissa in
$(\xi_l, \xi_l +\epsilon) \cup (x_{\alpha}, x_{\alpha}+\epsilon)
\cup (\xi_r -\epsilon, \xi_r)$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=130mm]{s-ne0.eps}
\end{center}
\caption{The case where int\,$b_r$ is free from endpoints}
\label{fig:ne0}
\end{figure}
We consider first the case
where int\,$b_r$ does not contain an endpoint of an arc of $A$ outside $Z$.
In this case,
the proof proceeds
by induction on the number, say $n_e$, of endpoints of arcs of $A$ contained in int\,$b_r$.
Such arcs are inside $Z$.
When $n_e=0$,
we deform $S$ as in Figure \ref{fig:ne0}.
We shrink $X_b$ in the horizontal direction
by a good ambient isotopy of ${\mathbb R}^2$
so that it is contained in $(x_r, \xi_r -\epsilon] \times [\eta_b + \epsilon, y_b - \epsilon]$.
Then, shrinking the arcs connecting $X_b$ and $X_a$ and thinning $X_b$ in the vertical direction,
we move $X_b$ upward so that the ordinates of points in $X_b$ are within the interval $(y_t, y_t +\epsilon)$.
Now, there is nothing in the right side of $Z$ within $R$.
We lengthen $Z$ to the right direction
so that the abscissa of the right vertical line segment of $Z$
is a little larger than $\xi_r - \epsilon$.
Then we can move the arc $\alpha_1$
so that it forms a vertical line segment above the top horizontal line segment of $Z$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-DeformTypeI.eps}
\end{center}
\caption{The case where $\delta$ is of type I and with $\partial \delta$ in $Z$}
\label{fig:DeformTypeI}
\end{figure}
We consider the case $n_e > 0$.
Let $q$ be the rightmost endpoint among those of arcs of $A$ in int\,$b_r$,
and $\delta$ the arc with $q \in \partial \delta$.
Note that $\delta$ is inside $Z$.
If $\delta$ is an arc of type I with its both endpoints in $Z$,
then we move $\delta$ near the point $q$
along a subarc of $b_r$ and the right vertical line segment of $Z$,
so that $\delta$ is deformed into an arc of type $\sqcup$.
See Figure \ref{fig:DeformTypeI}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-DeformTypeSqcup.eps}
\end{center}
\caption{The case where $\delta$ is of type $\sqcup$}
\label{fig:DeformTypeSqcup}
\end{figure}
If $\delta$ is an arc of type $\sqcup$,
then
we can move $\delta$ as shown in Figure \ref{fig:DeformTypeSqcup}
so that it forms an arc of type I.
We deform $\delta$ near the point $q$
along a subarc of $b_r$ and the right vertical line segment of $Z$,
and then perform deformations similar to straight merges
as described in the proof of Lemma \ref{lemma:s-merge}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=100mm]{s-DeformTypeI2.eps}
\end{center}
\caption{The case where $\delta$ connects $Z$ and other circle of $C$}
\label{fig:DeformTypeI2}
\end{figure}
In the remaining case,
$\delta$ has the other endpoint in a circle of $C$ other than $Z$.
Let $R_Z$ be the rectangular disk bounded by $Z$ in ${\mathbb R}^2$.
The arcs of type I with their both endpoints in $Z$ divides $R_Z$ into subdisks.
Let $R'_r$ be the rightmost one
with $R'_r = [x_I, x_r] \times [y_b, y_t]$ for some real number $x_I$.
If there are no such arc of type I,
then we set $x_I = x_l$ and $R'_r = R_Z$.
Then, set
$R_r = [x_I +\epsilon', x_r -\epsilon'] \times [y_b +\epsilon', y_t -\epsilon']$,
where the positive real number $\epsilon'$ is taken to be small
so that there is no horizontal line segment in $S$ with ordinate in
$(y_b, y_b +\epsilon') \cup (y_t -\epsilon', y_t)$
and there is no vertical line segment in $S$ with abscissa in
$(x_I, x_I +\epsilon') \cup (x_r -\epsilon', x_r)$.
Let $X_r = S \cap R_r$, the black box.
If $X_r$ contains an arc $\lambda$ of type $\sqcup$ with its both endpoints in $b$,
then we deform the rectangle cobounded by $\lambda$ and a subarc of $b$
to be very thin in the vertical direction
so that $\lambda$ gets out from $X_r$.
See Figure \ref{fig:DeformTypeI2}.
Note that $n_C(X_r) < n_C(X)$.
By the hypothesis of induction,
we can apply Lemma \ref{lemma:BlackBox} to $X_r$.
Then the black box $X_r$ is deformed to some black box, say $X'_r$,
and the arc $\delta$ is deformed to an arc
which connects $b_r$ and the top horizontal line segment of $\partial R_r$.
Then $\delta$ can be deformed to a vertical line segment
connecting the top horizontal line segment of $Z$
and that of $\partial R_r$.
In each of the three cases above we can decrease $n_e$,
and the lemma follows by induction.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=100mm]{s-outsideZ1.eps}
\end{center}
\caption{The case where $\delta$ is outside $Z$}
\label{fig:outsideZ1}
\end{figure}
We consider the case
where the interior of $b_r$ contains an endpoint of an arc of $A$ outside $Z$.
If $X_b$ contains an arc $\mu$ of type $\sqcup$ with its both endpoints in int\,$b_r$,
then we deform the rectangle cobounded by $\mu$ and a subarc of $b_r$
to be very thin in the vertical direction
so that $\mu$ gets out from $X_b$.
Note that $n_C(X_b) < n_C(X)$.
The hypothesis of induction allows us
to apply Lemma \ref{lemma:BlackBox} to $X_b$
so that the vertical line segments connecting $b$ and the top line segment of $\partial R_b$
are deformed to be arcs connecting $b$ and the bottom line segment of $\partial R_b$.
Let $X'_b$ be the black box obtained from $X_b$ by this deformation.
See Figure \ref{fig:outsideZ1}.
We move $X'_b$ as in this figure.
We shrink $X'_b$
so that the abscissa of one of the leftmost points of $X'_b$ is
larger than that of the right vertical line segment of $Z$,
and lift $X'_b$ up
so that the ordinate of one of the bottom points of $X'_b$ is
larger than that of the top horizontal line segment of $Z$,
and then lengthen it
so that the abscissa of one of the leftmost points of $X'_b$ is
smaller than that of the right vertical line segment of $Z$.
At this stage, the circles and arcs system is not rectangular
since the arcs
connecting $b$
and the bottom line segment of the boundary of the rectangle bounding $X'_b$
are neither of type I nor of type $\sqcup$.
We will deform them into arcs of type I one by one from now.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=120mm]{s-outsideZ2.eps}
\end{center}
\caption{The case where $\delta$ is outside $Z$ and of type I}
\label{fig:outsideZ2}
\end{figure}
Let $q$ be the rightmost endpoint among those of arcs of $A$ in $b_r$.
If $q$ is an endpoint of an arc inside $Z$,
then we perform one of deformations
as in Figures \ref{fig:DeformTypeI} through \ref{fig:DeformTypeI2} above.
We consider the case where $q$ is an endpoint of an arc $\delta$ outside $Z$.
If $\delta$ connects $Z$ and another circle of $C$,
then we can deform $\delta$ to be a vertical line segment connecting $Z$ and $X'_b$
as in Figure \ref{fig:outsideZ2}.
The last deformation in this figure is similar
to the straight merge described in the proof of Lemma \ref{lemma:s-merge}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=90mm]{s-outsideZ3.eps}
\end{center}
\caption{The case where $\delta$ is outside $Z$ and of type $\sqcup$}
\label{fig:outsideZ3}
\end{figure}
When $\delta$ has its both endpoints in $Z$ and is of type $\sqcup$,
a subarc of $b_r$ and $\delta$ cobound a disk, say $R_{\sqcup}$.
We shrink $R_{\sqcup}$ to be very small,
and move it along a subarc of $b_r$ and the right vertical line segment of $Z$
as in Figure \ref{fig:outsideZ3}.
If the arc $R_{\sqcup} \cap b_r$ contains endpoints of arcs of $A$ inside $Z$,
then we also deform inside $Z$
as in Figures \ref{fig:DeformTypeI} through \ref{fig:DeformTypeI2}.
Repeating such deformations,
int\,$b_r$ becomes free from endpoints of $A$
and $S$ becomes rectangular again.
Then, similarly to Figure \ref{fig:ne0},
we lengthen $Z$ to the right direction,
and move $\alpha_1$
so that it forms a vertical line segment above the top horizontal line segment of $Z$.
This completes the proof of the lemma.
\end{proof}
\begin{proof}
We prove Theorem \ref{theorem:RCAS}.
The proof proceeds by induction on the number of arcs of $A$.
If $A$ is empty, then the theorem is very clear.
We assume that the theorem holds when $A$ consists of $n-1$ arcs,
and consider the case where $A$ consists of $n$ arcs.
Let $A'$ be the union of arbitrary $n-1$ arcs of $A$.
We can deform the circles and arcs system $S=C\cup A$
by an ambient isotopy of ${\mathbb R}^2$
so that $C \cup A'$ is rectangular,
and that the arc $\alpha = A-A'$ is composed
of $m$ vertical line segments and $m-1$ horizontal line segments
for some positive integer $m$.
Note that $A \cap C\ (= \partial A)$
are endpoints of vertical line segments in $A$.
Hence we can deform the rectangular circles of $C$ very thin in the vertical direction
so that no pair of rectangular circles of the same depth overlap under $\pi_y$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-n_th_arc.eps}
\end{center}
\caption{$\alpha$}
\label{fig:n_th_arc}
\end{figure}
If $\alpha$ has a vertical line segment (resp. horizontal line segment) $e$,
which has two horizontal line segments (resp. vertical line segments)
sharing an endpoint with $e$
in both sides of $e$,
then we can perform a deformation
similar to the straight merge operation in the proof of Lemma \ref{lemma:s-merge},
to decrease the number of line segments forming $\alpha$.
Hence, without loss of generality, we assume
that $\alpha$ does not contain such a line segment
and is of one of the forms shown in Figure \ref{fig:n_th_arc}
or images of them
by a reflection in a vertical line or a horizontal line,
a rotation through $180^{\circ}$ or their composition.
Thus, if the arc $\alpha$ has its both endpoints in the same circle of $C$
and is inside the circle,
then $\alpha$ is already of type I or $\sqcup$,
and the theorem follows.
Hence we can assume
that either (1) $\alpha$ has its both endpoints in the same circle, say $Z$,
and is outside $Z$,
or (2) $\alpha$ connects distinct circles of $C$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=90mm]{s-GetCircleOut.eps}
\end{center}
\caption{get the circles intersecting $\partial R$ out of $R$}
\label{fig:GetCircleOut}
\end{figure}
We consider first Case (1).
Then $m=2, 3$ or $4$.
If $m=2$, then $\alpha$ is of type $\sqcup$, and we are done.
When $m=3$,
the arc $\alpha$ has two horizontal line segments,
say $s_t$ and $s_b$,
with their ordinates $t_{\alpha}, b_{\alpha}$ satisfying $t_{\alpha} > b_{\alpha}$.
See Figure \ref{fig:GetCircleOut}.
We can assume, without loss of generality,
that the vertical line segment, say $s_r$, in $\alpha$ between $s_t$ and $s_b$
connects right endpoints of $s_t$ and $s_b$.
Let $\ell_Z$ be the abscissa of the left vertical line segment of $Z$,
$r_{\alpha}$ the abscissa of the vertical line segment $s_r$,
and $R$ the rectangular disk
$[\ell_Z - \epsilon, r_{\alpha}-\epsilon] \times [b_{\alpha}+\epsilon, t_{\alpha}-\epsilon]$
for a very small positive real number $\epsilon$.
We take $\epsilon$
so that there is no vertical line segment with its abscissa
in $(\ell_Z - \epsilon, \ell_Z) \cup (r_{\alpha}-\epsilon, r_{\alpha})$,
and there is no horizontal line segment with its ordinate
in $(b_{\alpha}, b_{\alpha}+\epsilon) \cup (t_{\alpha} -\epsilon, t_{\alpha})$.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=80mm]{s-GetRidOfSqcup.eps}
\end{center}
\caption{get the arcs of type $\sqcup$ away from $\partial R$}
\label{fig:GetRidOfSqcup}
\end{figure}
If $\partial R$ intersects circles of $C$ other than $Z$
or the left vertical line segment of $\partial R$ intersects horizontal line segments in $A$,
then we perform deformations
at left parallel copies of the two vertical line segments
$\alpha-(s_b \cup s_r \cup s_t)-{\rm int}\,N(\partial \alpha)$
similar to the straight merge in the proof of Lemma \ref{lemma:s-merge}
to get such circles and horizontal line segments out of $R$,
where $N(\partial \alpha)$ is a very small regular neighborhood of $\partial \alpha$
in $\alpha$.
See Figure \ref{fig:GetCircleOut}.
If the bottom horizontal line segment of $\partial R$ intersects arcs of type $\sqcup$,
then we shrink them and what are surrounded by them and subarcs of $C$
in the vertical direction (upward or downward)
to cancel the intersection points.
See Figure \ref{fig:GetRidOfSqcup}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=125mm]{s-Case1m3.eps}
\end{center}
\caption{Case (1), $m=3$}
\label{fig:Case1m3}
\end{figure}
Let $q$ be the intersection point of $\alpha$
and the bottom horizontal line segment of $\partial R$.
We can apply Lemma \ref{lemma:BlackBox} to the black box $X=S \cap R$,
to bring $q$
to the top horizontal line segment of $\partial R$.
See Figure \ref{fig:Case1m3}.
Then we can deform $\alpha$ to an arc of type $\sqcup$,
and the theorem follows in this case.
When $m=4$,
a similar argument as above decreases $m$ to $3$,
and then the theorem follows by the above argument.
See Figure \ref{fig:Case1m4}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=125mm]{s-Case1m4.eps}
\end{center}
\caption{Case (1), $m=4$}
\label{fig:Case1m4}
\end{figure}
We consider Case (2).
We proceed by induction
on the number $m$ of vertical line segments in $\alpha$.
If $m=1$, then $S$ is rectangular, and we are done.
Hence we assume that $m \ge 2$,
and that the theorem holds
if $\alpha$ has at most $m-1$ vertical line segments
as the hypothesis of induction.
If the two circles of $C$ connected by $\alpha$ are of distinct depths,
then let $Z$ be the circle of larger depth.
If $\alpha$ connects two circles of $C$ of the same depth,
then we take $Z$ as below.
When $m \ge 3$,
let $Z$ be the circle of $C$
which contains one of the endpoints of $\alpha$
whose abscissa is between abscissae of some two vertical line segments in $\alpha$.
We can assume, without loss of generality,
that $\alpha$ has an endpoint
in the bottom horizontal line segment of $Z$
rather than in the top horizontal line segment.
When $m=2$,
we can assume, without loss of generality,
both endpoints of $\alpha$ are contained in the bottom horizontal line segment
of circles of $C$.
In this case,
let $Z$ be one of the circles connected by $\alpha$
such that the ordinate of the top horizontal line segment of $Z$
is smaller than
that of the bottom horizontal line segment of the other circle.
(Recall that circles of the same depth do not overlap under $\pi_y$.)
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=90mm]{s-Case2.eps}
\end{center}
\caption{Case (2)}
\label{fig:Case2}
\end{figure}
Let $s_Z$ be the vertical line segment in $\alpha$
with one of its endpoints in $Z$,
$s_h$ the horizontal line segment in $\alpha$
which share an endpoint with $s_Z$,
and $s_v$ the vertical line segment in $\alpha$ other than $s_Z$
such that $s_v$ and $s_h$ share an endpoint.
We assume, without loss of generality,
that $s_Z \cap s_h$ is the left endpoint of $s_h$.
Let $x_1$ be the abscissa of the left vertical line segment of $Z$,
$x_2$ the abscissa of the vertical line segment $s_v$,
$y_1$ the ordinate of the horizontal line segment $s_h$,
$y_2$ the ordinate of the top horizontal line segment of $Z$,
and $R$ the rectangular disk
$[x_1-\epsilon, x_2-\epsilon] \times [y_1 + \epsilon, y_2 + \epsilon]$ in ${\mathbb R}^2$
for a very small real number $\epsilon$.
We take $\epsilon$
so that $S$ has no vertical line segment with its abscissa
in $(x_1 - \epsilon, x_1) \cup (x_2-\epsilon, x_2)$,
and $S$ has no horizontal line segment with its ordinate
in $(y_1, y_1+\epsilon) \cup (y_2, y_2 +\epsilon)$.
The top horizontal line segment and the right vertical line segment of $\partial R$
do not intersect a circle of $C$.
See Figure \ref{fig:Case2}.
If $\partial R$ intersects circles of $C$
or the left vertical line segment of $\partial R$ intersects horizontal line segments in $A$,
then we perform a deformation
at left parallel copy of $s_Z - {\rm int}\,N(\partial \alpha)$
similar to the straight merge as in the proof of Lemma \ref{lemma:s-merge},
to get such circles of $C$ and horizontal line segments out of $R$.
If the bottom horizontal line segment of $\partial R$
intersects arcs of type $\sqcup$,
then we shrink them and what are surrounded by them and subarcs of $C$
in the vertical direction (upward or downward)
to cancel the intersection points.
Then we can deform $S$ as shown in Figure \ref{fig:Case2deformation}.
Precisely,
let $q$ be the intersection point $\alpha \cap \partial R$.
We can apply Lemma \ref{lemma:BlackBox} to the black box $X=S \cap R$,
to bring $q$
to the top horizontal line segment of $\partial R$.
Note that $Z$ is lengthened to the right by the isotopy in the proof of Lemma \ref{lemma:BlackBox}.
Then we can deform $\alpha$ to an arc with less number of vertical line segments.
The theorem follows by the hypothesis of induction.
\end{proof}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=125mm]{s-Case2deformation.eps}
\end{center}
\caption{deformation in Case (2)}
\label{fig:Case2deformation}
\end{figure}
\section*{Acknowledgments}
The authors would like to thank Nobuya Satoh for helpful comments.
He is the advisor of the first author's master's thesis
at Graduate School of Science, Rikkyo University.
The second author is partially supported
by JSPS KAKENHI Grant Number 25400100.
|
\section{Conclusion}
\label{sec:conclusion}
This paper discusses the needs for adapting the way of using MDE techniques created for design time activities, to support the use of models at runtime.
To this end, several requirements are described, and the {\em de facto} standard in the MDE community, {\em i.e.} the Eclipse Modeling Framework (EMF), is evaluated against these requirements.
After this evaluation, this paper presents the Kevoree Modeling Framework (KMF) as an alternative solution to create modeling tools more suitable for runtime purposes.
This paper has presented the evolutions of the KMF generated code, according to what was presented in \cite{Fouquet:2012fk}.
The new version of KMF offers better performances with a reduced footprint, allowing to embed KMF more easily in resource-constrained JVM such as Android or Embedded Java.
This new version of KMF also introduces several new features.
KMF now provides a way to handle concurrency at the level of the model, with the ability to define fragments as read-only.
This latter option is particularly useful when combined with the new clone operator, making it possible to factorize read-only fragments among model clones (hence saving memory) while the mutable part is specific to each clone.
Finally, KMF also provides a query language to directly access a model element by providing a path, inspired by well-established work in the database domain.
This new version of KMF has been compared to EMF and the initial version of KMF~\cite{Fouquet:2012fk}.
These evaluations clearly show the gain offered by the KMF generated API in several domains, with an average gain of 28\% in time with respect to the EMF implementation.
Then we highlighted the need to rethink some modeling operators such as cloning and lookup to make them more efficient at runtime.
The empirical evaluation clearly shows that the use of unique identifiers and KMFQL-PS saves lots of computation time.
This work finds several application cases in domains like Internet of Things, Cloud management or dynamic adaptation of software systems.
The opportunities of application are also augmented by the good performances of the KMF generated code and the reduction of the memory required compared to EMF generated code.
These good properties should enable a wider use of models at runtime in industrial scale projects.
In future work, we plan to port KMF to JavaScript, to make the generated API available directly in web-browsers, making it possible to develop web-based, massively distributed, collaborative modeling environment.
Finally, also plan use and extend KMF in the context of Big Data models.
Because of their particularly big size, these models cannot be entirely loaded in memory and require seamless memory swapping between memory and disk.
Based on existing works~\cite{pagan2011morsa,Barmpis:2012:CAD:2467307.2467314}, several approaches are currently being tested to offer MDE WITH seamless manipulation solutions for these big models, mostly inspired by noSQL database approaches.
\subsection{Reduction of the memory footprint}
\label{sec:contrib:memoryFootprint}
The static footprint reduces the memory available for the dynamic creation of objects necessary for business processing. It is thus important to watch both static and dynamic memory requirements of the generated code.
\paragraph{Static memory footprint}~\\
To limit the dependencies and thus the static memory footprint, we decided to restrict the inheritance relationships of our generated code to sibling classes and elements from standard libraries only.
In a first attempt~\cite{Fouquet:2012fk}, the source code generated by KMF already allowed to save some 8~MB of dependencies, and up to 13~MB with the help of a code shrinker (ProGuard).
Now, KMF generates Kotlin code, with an even lighter static footprint.
On the same FSM metamodel, KMF generates 76~KB of code (including the core API, as well as model loader/serializer, and a query API) to be compared to the 55~KB of EMF pure code.
Including the dependencies ({\em i.e.} the Kotlin standard library), the standalone generated code grows up to 488~KB, which can again be reduced to 335~KB by removing unused classes from the standard Kotlin library.
Table~\ref{fig:memory_summary_static} summarizes the gain of the successive versions of KMF {\em w.r.t.} EMF.
\begin{table}
\vspace{-1em}
\center
{\renewcommand{\arraystretch}{1.2}
\begin{tabular}{|c| >{\centering\arraybackslash}p{1.7cm}| >{\centering\arraybackslash}p{2.2cm}| >{\centering\arraybackslash}p{2.5cm}|}
\hline Tool (Language) & Effective Code (KB) & Standalone Package (MB) & Reduced Standalone (MB)\\
\hline EMF (Java) & ~55 & 15 & No Data\\
\hline KMF (Kotlin) & 76 & 0,488 & 0,335 \\
\hline
\end{tabular}
}
\caption{EMF and KMF static memory footprint comparison}\label{fig:memory_summary_static}
\vspace{-1em}
\end{table}
Thanks to this drastically reduced footprint, KMF has successfully been used to generate APIs able to run on a large offer of JVM, including mobile and embedded ones: Dalvik~\footnote{http://www.dalvikvm.com/}, Avian~\footnote{http://oss.readytalk.com/avian/}, JamVM\footnote{http://jamvm.sourceforge.net/} or JavaSE for embedded Oracle Virtual Machine~\footnote{http://www.oracle.com/technetwork/java/embedded/downloads/javase/index.html}.
\paragraph{Dynamic memory footprint}
~\\
To further improve the dynamic memory usage, we paid more attention to the creation of temporary objects. One of the main consequence of that is the dropping of {\em scala.Option}~\cite{scala} (compared to our previous contribution).
The new version of KMF in Kotlin gets rid of these {\em scala.Option} because (1) these objects were among the most frequently created, generating a huge overhead; and (2) Kotlin does not support such a mechanism but proposes {\em Nullable} variables instead. This nullable mechanism relies on a static check at compile time introducing then no overhead at runtime.
Table~\ref{fig:memory_summary_dynamic} compares the sizes of heap memory used to create the experimental FSM model (presented in section~\ref{sec:sota} in EMF, KMF with Scala~\cite{Fouquet:2012fk} and finally, KMF with Kotlin.
\begin{table}[h!]
\center
{\renewcommand{\arraystretch}{1.2}
\begin{tabular}{|c| >{\centering\arraybackslash}p{1.8cm}| >{\centering\arraybackslash}p{2cm}| >{\centering\arraybackslash}p{2.2cm}|}
\hline Tool (Language) & EMF (Java) & KMF (Scala) & KMF (Kotlin)\\
\hline Heap Memory Used & 140~MB & 61~MB & 52,2~MB\\
\hline
\end{tabular}
}
\caption{EMF and KMF dynamic memory footprint comparison}\label{fig:memory_summary_dynamic}
\vspace{-1.5em}
\end{table}
This reduction of the memory used also reduces the need for garbage collection, because less objects are created and disposed.
Since the garbage collection in Java is a rather costly operation, this also participates to the efficiency of the overall approach.
\subsection{Save and Load}
\label{sec:contrib:saveLoad}
Load and save operations are two of the most used operations when dealing with models at runtime.
Indeed, any change in a model may have to be distributed to other devices taking part in the software system, or stored for history to be potentially restored later on.
The performance of these operations can thus generate a considerable overhead for little manipulations.
As described in section~\ref{sec:sota}, the reflexivity of loaders and the creation of temporary structures (objects) are two important bottlenecks.\\
~\\
{\bf Flat reflexivity}~\\
Thanks to the close world assumption, we decided in KMF to generate domain specific loaders and serializers. This flat generation removes all reflexive calls from the loading and saving processes making them more efficient.\\
{\bf Avoid temporary objects creation}~\\
Usually, (un)marshalling operations rely on an intermediate structure like a DOM to simplify the mapping from a structure in memory into a persistent structure ({\em e.g.} XML file). However, this implies the creation of complex temporary structures in memory.\\
To avoid this costly temporary structure, the loader works directly from the stream, and creates the final model elements directly. This is possible because each element of the stream is predictable (close world).
Similarly, KMF generates serializers that directly print in a stream buffer, to avoid unnecessary creation and concatenations of Strings.\\
{\bf Minimize class loadings}~\\
Previously~\cite{Fouquet:2012fk}, for each concept of the metamodel (domain), a specific class providing specific uniquely named loading/saving methods were generated and merged at compile time.
However, if this kind of generation makes a better organization of the code and eases the reading by a human being, it creates a lot of classes and requires, at runtime, many costly class-loading operations.
Thus, the new version of KMF presented in this paper groups all the loading (respectively saving) methods in a single file to drastically reduce the class loading time part of these operations.\\
{\bf Several exchange formats}~\\
The generative approach enables the generation of loaders and serializers for several standard formats.
At this time, KMF can generate I/O tools for the standard XMI and JSON, from files or streams (easier to load on some resource constrained platforms). They also provide a rich API to enable marshalling and unmarshalling in compressed formats.\\
~\\
{\bf Efficiency measurement}~\\
Table~\ref{fig:contrib::saveload:summary} summarizes the time required for loading and saving a model from/to a serialized formed with classical EMF/Java tools, former KMF version and the new KMF.
Again, the measures presented in the table are the results from the loading and serialization of the experimental FSM model presented in section~\ref{sec:sota}.
\begin{table}[h!]
\center
{
\renewcommand{\arraystretch}{1.2}
\begin{tabular}{|c| >{\centering\arraybackslash}p{2cm}| >{\centering\arraybackslash}p{2cm}|}
\hline Tool (Language) & Loading (ms) & Saving (ms)\\
\hline EMF (Java) & 1214 & 7021\\
\hline KMF (Scala) & 1193 & 799\\
\hline KMF (Kotlin) & 999 & 802\\
\hline
\end{tabular}
}
\caption{EMF and KMF load and save time}\label{fig:contrib::saveload:summary}
\vspace{-1.5em}
\end{table}
\subsection{Efficient Model navigation : KMF Path}
\label{sec:contrib:path}
As illustrated by the example in section~\ref{sec:generic:nav}, activities to be carried on models@runtime need an efficient way to look up and navigate across model elements.
In particular, during the model comparison steps, the merge and check operations require an efficient tool to reach a specific element in the model.
To enable this efficient research, KMF leverages the notion of Unique Identifier from relational databases, which should be declared in the metamodel using the `id` field, already present in Ecore.
This section introduces the Path Selector (KMFQL-PS) of the Kevoree Modeling Framework Query Language (KMFQL), which uses the id attribute specified in the metamodel, as the unique key to find a model element by following model relationships.
\paragraph{Approach}~\\
Directly inspired by the select operator of relational databases and by XPath\cite{Consortium:2007fk}, KMFQL-PS defines a query syntax aligned with the MOF concepts.
The navigation through a relationship can be achieved with the following expression: \textbf{relationName[IDAtrributeValue]}.
This expression defines the \textbf{PATH} of an element in a MOF relationship.
Several expressions can be chained to recursively navigate the nested models.
Each expression is delimited by a \textbf{/}.
It is thus possible to build a unique path to a model element, by chaining all sub-path expressions needed to navigate to it from the root model element, via the containment relationships.
In our illustrating example, the ``name'' attribute is defined as the ID attribute of \textit{NamedElement}, and we know precisely where the component instance is hosted ({\em i.e.} the path to the model element in the model).
The numerous nested loops presented in section~\ref{sec:sota} are now reduced to the piece of code presented in listing~\ref{lst:kmfql_ps}.
\begin{scriptsize}
\begin{lstlisting}[frame=L, language=Java, caption={Collect of the FakeConsole component instance with KMFQL-PS.}, label={lst:kmfql_ps}]
ComponentInstance fConsole2 = model.findByPath("nodes[node6]"
+ "/hosts[node7]/hosts[node8]/hosts[node4]"
+ "/components[FakeConso380]",
ComponentInstance.class);
\end{lstlisting}
\end{scriptsize}
\paragraph{Implementation details}~\\
KMF generates hash-tables in place of simple collections, for each relationship.
Then, each time a model element is added in/removed from a relationship, it is stored in this hash-table.
Hash keys are computed from the ID attributes, or are automatically generated if this ID is not defined.
The keys in the hash-tables are equivalent to indexes in noSQL databases. Also, even if the use of hash-tables introduces a slight memory overhead, it considerably speeds up the resolution of the paths ({\em i.e.} the retrieval of a model element).
Moreover, the hash-tables allow to get rid of the high number of temporary objects created when looping, which favours the scalability on limited execution environments such as in the IoT domain.\\
{\bf Experimental validation\footnote{The code and more documentation about this experiment can be found within the KMF Github repository: https://github.com/dukeboard/kevoree-modeling-framework/blob/master/doc/kmf\_path.md}}~\\
We performed an evaluation of this feature on the model presented in Section~\ref{sec:generic:nav} (Figure~\ref{fig:kevoree_pathexample}).
The APIs have been generated with EMF (2.7) and KMF, and the XMI model has been loaded prior.
Then, as explained in section~\ref{sec:generic:nav}, we collected the ``FakeConsole380'' model element with both plain Java approach (listing~\ref{lst:plainJava}) and KMFQL-PS (listing~\ref{lst:kmfql_ps}).
For this resolution, the plain Java approach with \textbf{EMF} takes~\textbf{954$\mu$s} to execute, while the same resolution in \textbf{KMFQL-PS} takes~\textbf{37$\mu$s} on the same hardware (i7 processor 2.6Ghz and Oracle JVM 7). The resolution of already known model elements with KMFQL-PS is thus 25.7 times faster than the approach using nested for-loops applied on EMF models.
Also the algorithmic complexity has been significantly reduces compared to the nested for-loops approach (see section~\ref{sec:generic:nav}), to O($\log(n) \times \log(o) \times \log(p) \times \log(q) \times \log(r)$), which confers a better scalability to the approach.
\section{The Kevoree Modeling Framework}
\label{sec:contrib}
This section presents the Kevoree Modeling Framework (KMF), an alternative to the Eclipse Modeling Framework to address the requirements on the generated code listed in section~\ref{sec:requirements}, imposed by the use of models at runtime .
This section is divided into four subsections according to the four key requirements.
As an introduction to the contribution, the section~\ref{sec:contrib:overview} gives an overview of the main principles that drove our work.
Section~\ref{sec:contrib:memoryFootprint} then describes how we reduced the memory footprint of the generated code.
Section~\ref{sec:contrib:saveLoad} presents the improvements made on I/O operations (load and save) for models.
The mechanisms to manage concurrency are explained in Section~\ref{sec:contrib:clone}.
Finally, Section~\ref{sec:contrib:path} introduces the KMF Query Language (KMFQL) to efficiently reach a specific known element in a model.
\subsection{General approach}
\label{sec:contrib:overview}
We followed several principles and guidelines during the development of KMF in order to meet the performance requirements. Here is the list.
\begin{enumerate}[A)]
\item {\em Avoid the creation of temporary objects as much as possible}.
While the creation of objects in the Java virtual machine is not very costly, their destruction by the garbage collector\footnote{responsible for freeing the memory of not referenced objects} is a more complex and costly operation\footnote{See http://www.oracle.com/technetwork/java/gc-tuning-5-138395.html}.
We optimize the memory by reusing objects already loaded instead of creating temporary objects.
\item {\em Sharing immutable objects}. In Java, several collections\footnote{See http://docs.oracle.com/javase/tutorial/essential/concurrency/immutable.html} and primitive objects can define an immutable state.
Leveraging immutability, we automatically apply a flyweight pattern~\cite{gof94} to share these immutable objects among several models to reduce the overall memory footprint.
\item {\em Flat reflexivity}. By relying on a sound generative approach, any modification of the metamodel implies a regeneration of all the tools, but it also allows using the closed world assumption~\cite{reiter1978closed}. Thus, model manipulations at runtime rely on a finite and well-defined set of types. This allows generating a flat reflexive layer, composed of a finite set of matching functions, avoiding costly reflexive calls at runtime.
\item {\em No fancy dependency}. To be embeddable in the wider range of hosting nodes, the modeling framework must limit its dependencies. The code generated with KMF depends only on no third-party library.
\item {\em Direct Memory Access for models I/O}. KMF uses as much as possible Direct Memory Access to load and save models from the network or the disk to the local memory.
This principle reduces the need for buffering during model I/O operations.\\
\end{enumerate}
\input{Contribution/contrib_memoryFootprint}
\input{Contribution/contrib_saveLoad}
\input{Contribution/contrib_concurrency}
\input{Contribution/contrib_path}
\subsection{Concurrent read/write access}
\label{sec:contrib:clone}
\vspace{-0.5em}
The protection against concurrent accesses can be fine-grained, on each individual model element, or coarse-grained, on the whole model.
This section first details fine-grained protection mechanisms for concurrent reads, then presents coarse-grained protection with efficient runtime clone operation for concurrent writes.\\
{\bf Concurrent shared model}~\\
To allow concurrent read access to model elements and relationships, the code generated by KMF (1) only uses standard Java thread-safe collections instead of ad-hoc EMF EList (or similar), which have issues on concurrent usages, and (2) exposes cloned immutable lists only via its public methods.\\
{\bf Model clone for independent modifications}~\\
To allow modifications of models, a simple strategy consists in providing each thread with a copy (clone) of the model, so that it can manipulate it independently from the other threads, with no need for a fine-grained synchronization.
Cloning is a very costly operation, basically because it duplicates entire structures in memory.
In the Java world, this means that every object from the model A yields a new equivalent object in the model A'.
To optimize the cloning of models, KMF relies on a flat reflexivity model to implement a very efficient, single pass, model-cloning operator. Moreover, all clones share the immutable Java objects like String resources to avoid duplication of invariant objects.\\
{\bf Towards immutable models}~\\
The KMF framework offers a method to turn individual model elements into read-only state.
The read-only state cannot be reverted to a read/write structure to ensure that no process will break this assumption while the model is being processed.
The only way to obtain a mutable structure is to duplicate the model into a new semantically identical model by cloning.\\
In addition, if an element is set as read-only, we also set all its contained elements and their references to read-only.
This ensures that no part of the sub-graph remains mutable when the root is immutable.
To do so, the generated API relies on the flat reflexivity and close world assumption to automatically call the $setReadOnly$ method on each sub element.\\
{\bf Partial clone operator}~\\
Cloning is a very costly operation, because it duplicates the graph structure in memory, similarly to the fork operation on a Linux process.
Despite the semantic of clone design implies this duplication, runtime optimizations can minimize the memory usage while keeping the right semantics.
To this end, KMF proposes a partial clone operation.\\
The basic idea is to duplicate only the mutable parts of model, while sharing immutable parts between the original model and its clones.
Such a mechanism greatly improves the memory usage while leaving a lot of flexibility to the user. Indeed, mutable and immutable parts can be defined precisely using the \textit{readOnly} operation.
\begin{figure}[h!]
\center
\includegraphics[width=0.6\textwidth]{figures/partialClone.pdf}
\caption{Partial clone operator illustration}\label{partialClone}
\vspace{-1em}
\end{figure}
Figure~\ref{partialClone} illustrates the partial clone operator where a model $A$ is cloned twice, and the immutable zone is shared between the clones $A'$ and $A''$.
The cloned model has a smaller size than the original model, due to the shared memory zone.
Moreover, since the clones reference the immutable zone, they remain consistent if the garbage collector frees the model $A$ ({\em i.e.:} the immutable zone will not be collected while at least one clone has a reference on it).
Finally the clone operation uses the $readOnly$ flag to prevent from navigating or cloning model elements that do not need to be duplicated, thus improving the cloning time.\\
{\bf Experimental validation}~\\
In order to evaluate this smart clone, the experiment takes a model $A$ from the Cloud domain, containing 400 nodes, and tries two strategies:
\begin{enumerate}
\item One node is mutable and we perform a very specific modification, for example searching for component placement inside this node.
\item all the 400 nodes are mutable, but we fix all static information (mainly related to provisioning).
\end{enumerate}
The hardware used in this experiment is a Macbook Pro (i7 processor 2.6Ghz and Oracle JVM 7) and results are presented in table~\ref{fig:partialCloneTab}.
\begin{table}[h!]
\center
{\renewcommand{\arraystretch}{1.2}
\begin{tabular}{|c| >{\centering\arraybackslash}p{2cm}| >{\centering\arraybackslash}p{2cm}|}
\hline experimental configuration & time to clone (ms) & memory per model\\
\hline Full clone (1 mutable node) & 18 & 705kb\\
\hline Partial clone (1 mutable node) & 0.86 & $<$1kb\\
\hline Full clone (400 mutable nodes) & 19 & 705kb\\
\hline Partial clone (400 mutable nodes) & 5.36 & 304kb\\
\hline
\end{tabular}
}
\caption{Partial and Full clone comparison}\label{fig:partialCloneTab}
\vspace{-1em}
\end{table}
The clone time goes down from 18 ms to 0.86 ms, while the memory used for each clone drops from 705kb to less than 1kb per clone, which highlights the significant gain of memory.
Also, even if we take a larger zone of mutable elements, the gain is still significant with a time reduced from 19 ms to 5.36 ms and each clone takes twice as less memory as the original.
Improving clone performance enables the use of coarse grain strategies (model unit) instead of relying on costly synchronized model mutators.
\section{Evaluations}
\label{sec:evaluation}
\subsection{Overview}
Each requirement has previously been evaluated in the dedicated sub-sections of Section~\ref{sec:contrib}.
The aim of this section is to provide more insights on the performances of KMF over different models, from different domains, with different sizes and features.
\subsection{Experimental protocol}
The experimental protocol is composed by the following steps, which compose a standard process when using models at runtime. For each model / domain:
\begin{enumerate}
\item The model, previously serialized in the XMI standard format, is loaded from a file (repeated 10 time),
\item The model is then cloned (100 repetitions) to prepare a search/comparison,
\item One clone and the original model are compared (1 time), element by element,
\item Finally, the clone is serialized (10 repetitions) in a file, in XMI format.
\end{enumerate}
This process is repeated on models of various sizes, conforming to three representative metamodels:
\begin{enumerate}
\item {\bf Kevoree}: the platform leveraging Model@Runtime previously detailed in this paper~\cite{fouquet:hal-00688707,Fouquet:2012:DCM:2304736.2304759}
\item {\bf Kermeta}: a model-oriented language developed by the Triskell Team to define the operational semantics of models~\cite{kermeta}
\item {\bf ThingML}: an operational DSL dedicated to the Internet of Thing (embedded systems connected to the Internet)~\cite{thingml}
\end{enumerate}
All the experiments are run on the same hardware as the one used to evaluate EMF in the requirements section ({\em i.e.} MacBook pro, Intel Core i7 2.6Go, 16Go RAM and SSD drive).
The JVM used is the Oracle standard JVM 7 and all files and code used during the experiment are publicly available on the project Github\footnote{https://github.com/dukeboard/kevoree-modeling-framework/tree/master/sosym-evaluations}.
\subsection{Results}
{\bf Generalization of performance improvement on different metamodels}~\\
The benchmark is executed on a first set of models conforming to Kermeta and ThingML metamodels.
Their sizes span from 100 to 6000 elements, and have different structures implied by their metamodels.\\
The results of KMF and EMF when dealing with Kermeta and ThingML models are graphed in Figure~\ref{eval:KermetaComparison:fig} and Figure~\ref{eval:ThingMLComparison:fig}.\\
It is important to note that the metamodels (as well as the models) of Kermeta and ThingML are used as-is and have not been modified to leverage the specific capabilities of KMF (in particular the definition of IDs).
\begin{figure}
\center
\includegraphics[width=1\textwidth]{figures/evalA3_1.pdf}
\caption{Evaluation of KMF and EMF on Kermeta models}\label{eval:KermetaComparison:fig}
\vspace{-0.5em}
\end{figure}
\begin{figure}
\center
\includegraphics[width=1\textwidth]{figures/evalA3_2.pdf}
\caption{Evaluation KMF and EMF on ThingML models}\label{eval:ThingMLComparison:fig}
\vspace{-1em}
\end{figure}
Overall, the code generated by KMF tends to be more efficient than the one with EMF (in addition to other benefits such as thread safety, discussed in Section~\ref{sec:contrib}), with an average gain of 28\% on our benchmarks.
This confirms that KMF is generalizable, even if the metamodels have not been optimized for KMF generation. \\
Also, we observe in these experiments that the optimization of the generated API can not go beyond the limit of 30\% of gain, because of the graph complexity of models.
This advocates for the need of new features to reduce more significantly the algorithmic complexity, such as partial clone and query language which are evaluated in the next two sections.\\
{\bf Performance evaluation of the KMFQL-PS}~\\
The metamodels of Kermeta and ThingML do not define IDs and thus, do not leverage KMF optimized paths.
A second set of models, describing the topology of cloud stacks (software, platform and infrastructure), is created.
These models have different size, ranging from 876 to 33644 elements.
This second experiment varies the use of ID definitions and the size of the models in order to evaluate the impact of the new KFM concepts.
Three cloud models have been selected depending of their percentage of ID definition (18, 28 and 99\%).
As a reminder the KMF lookup method reuse ID information to speed-up the resolution.
The results are presented in Figure~\ref{eval:SecondPack}.
\begin{figure}[h!]
\vspace{-1em}
\center
\includegraphics[width=1\textwidth]{figures/evalB3.pdf}
\caption{KMF results on Kevoree models (models@runtime)}\label{eval:SecondPack}
\vspace{-1.5em}
\end{figure}
This evaluation clearly highlights the benefit of defining ID attributes on model elements.
The code generated with KMF already allows gaining time compared to the EMF approach, with a very low percentage of ID definition (18\%) as illustrated by the first grey line in figure~\ref{eval:SecondPack}.
This gain is made even more important as the percentage of model elements that have an ID grows (grey lines 2 and 3 of the figure, with respectively 28\% and 99\% of elements with IDs).
The most significant gain is clearly on the lookup resolution, extensively used for comparison, merging, composition, {\em etc}.
From a wider point of view, the global process of loading, cloning, searching and saving is reduced from 35s, with EMF, to less than 2.5s with the optimized KMF.\\
{\bf Performance evaluation of the Partial Cloning mechanism}~\\
The final experimentation aims at assessing the improvements brought by the use of partial clones.
To highlight the gain, a cloud model is used.
In this model, several nodes are containing other nodes and components.
The \textit{readOnly} option offered by KMF is progressively applied on 20, 40, 60, 80 and 99\% of the model.
For each percentage, the clone in KMF is compared with the clone in EMF.\\
The results are graphed in figures~\ref{fig:readOnlyComparison:fig} and \ref{fig:readOnlyComparison:fig2}.
We observe that the reduction of time to clone the model is close to be inversely proportional to the percentage of readOnly elements({\em i.e.} the more readOnly elements, the less time required).\\
The Figure~\ref{fig:readOnlyComparison:fig2} presents the result of the same experiment with EMF without any readOnly optimization.
Such results are taken only for reference, because EMF can only perform a full clone, time is then not impacted by the percentage of readOnly elements, and remains to a value around 204ms.
\begin{figure}[hb]
\vspace{-1em}
\centering
\begin{minipage}{.45\textwidth}
\centering
\includegraphics[width=0.90\linewidth]{figures/partialClone2.pdf}
\captionof{figure}{Time to clone w.r.t. percentage of read-only elements in the model}
\label{fig:readOnlyComparison:fig}
\end{minipage}%
\hspace{1em}
\begin{minipage}{.45\textwidth}
\centering
\includegraphics[width=0.99\linewidth]{figures/partialClonerExperiment.pdf}
\captionof{figure}{Time to clone w.r.t. percentage of read-only elements in the model}
\label{fig:readOnlyComparison:fig2}
\end{minipage}
\end{figure}
The results clearly show that the individual features of KMF presented and preliminary assessed in Section{sec:contrib}, also significantly improve the performances of a complete models@runtime process. The performance gains compared to the first experiment clearly emphasize the role of the KMF-specific features (partial cloning and path) in this performance improvement.
\section{Introduction}
\label{sec:intro}
Model-Driven Software Engineering investigates how models, as abstract representations, can support and enhance the authoring of software systems.
In particular, Domain-Specific Modeling Languages (DSMLs) aim at providing software developers with dedicated languages and tools specifically created for their business domain.
DSMLs usually leverage an Object-Oriented software API, generated from a metamodel, facilitating the authoring and exchange of domain-specific concepts between several tools, applications and stakeholders.
This API can for example support graphical or textual editors to edit models, or specific load and save algorithms to serialize and exchange models.
The primary focus of DSMLs was to ease the comprehension and manipulation of concepts during the design phase of an application.
In this area, the Eclipse Modeling Framework (EMF) has rapidly become the \textit{defacto} standard in the MDSE for building DSMLs and associated tools using generative techniques.
As the boundaries between design-time and runtime become more and more blurry, DSMLs are increasingly embedded at runtime to monitor, manage, maintain and evolve running applications, leading to the emerging Models@Runtime paradigm~\cite{BlairBF09}.
However, the integration and use of EMF APIs and tools in a Models@Runtime context reached several limitations.
For example, when applied to the domain of the Internet of Things~\cite{thingml}, the execution environments are usually constrained in terms of memory and computational power, making it difficult to embed the DSML tools generated by EMF.
When applied to the Cloud Computing domain, the models created with a specific DSML for cloud are extensively exchanged, cloned and compared in order to manage the maintenance and adaptation of large-scale distributed applications. These operations call for very efficient means to clone and access model elements, which EMF fails at providing.
Also, the dynamic adaptation of a software system ({\em e.g.} to manage the scalability of Cloud-based system), requires efficient search and validation techniques to maximize the continuity of service, and minimize the downtime when reconfiguring the system.
This paper originates from the application track of MODELS'12~\cite{Fouquet:2012fk}. Our main contribution is to integrate best development practices, largely accepted in the Software Engineering community and largely disregarded in the MDSE community, in order to get improved performances on the generated code, compared to standard modeling tools and frameworks (and EMF in particular). This contribution is integrated into the Kevoree Modeling Framework (KMF), an alternative to EMF still compatible.
This paper extends our previous work in three main aspects:
\begin{enumerate}[i)]
\item It elicits the requirements and the state of practice in using DSLs and models during the execution of a software, in three domains.
\item It highlights the last improvements made in KMF to improve efficiency. In particular it focuses on the new features introduced to decrease the memory footprint, to decrease the time required to load or save models, to support a partial clone operator and to query a find model elements in a model using ``KMF Query Language - Path Selector''.
\item It evaluates KMF and EMF against the properties identified as required to use models during runtime in several domains.
\end{enumerate}
The outline of this paper is the following.
Section~\ref{sec:requirements} presents use cases in three different domains, in order to illustrate the requirements raised by the use of models as first-class entities at run-time.
Section~\ref{sec:sota} gives a state of practice of using design-time tools like EMF for runtime manipulations of models.
The contribution of this paper, the Kevoree Modeling Framework(KMF), is described in Section~\ref{sec:contrib}.
Section~\ref{sec:evaluation} evaluates our alternative implementation in comparison to EMF.
Section~\ref{sec:conclusion} concludes this work and presents future work.
\section{Requirements for using models at runtime}
\label{sec:requirements}
This section introduces briefly three domains in which models are used during the system runtime. For each domain, a specific DSL is used and highlights different requirements for the object-oriented API and the associated tools generated. Then, several properties are identified and listed as mandatory to enable a consistent use of models during runtime.
\vspace{-1em}
\subsection{Use cases}
\label{sec:usecases}
\vspace{-1em}
We present in this section three cases in which the use of models during runtime, for different purposes, requires some properties to be guaranteed by the tools generated for the DSL.
\subsubsection{Internet of Things}~\\
The domain of the Internet of Things relates to the interconnection of communicating object.
The variability and heterogeneity of devices to be considered in this domain challenges the possibilities of interoperate the devices and the development of software systems to offer new services.
In this domain, models are used during the software execution to reflect the configuration of devices, their links to siblings or services.
Also, the sporadic presence of devices requires dynamic adaptations of the running system to install and remove pieces of software to control each specific device present in the range of control.
The models are very useful in this context to describe what are the services available on each device, where are the binaries to connect it, etc.
In addition, model structures are useful to express composition of services operations, inherently necessary to compose data collected asynchronously.
The main concern here is about the memory restrictions and time of adaptations. Indeed, the kind of execution environment used in this domain are usually constrained in terms of memory and computational power.
Thus, attention has to be paid to the size of the generated code and the memory necessary for its execution, as long as to the time necessary to read (load) models.
\subsubsection{Cloud Computing}~\\
Cloud computing is characterized by a dynamic provisioning of services on computation resources.
Also called elasticity, this intelligent process must dynamically place several software to adapt the computation power to the real consumption of customers' software.
However, costs limits imposed by cloud customers force the elasticity mechanism of cloud to perform some choices, sometime contradictory.
For instance, cost and power can not be optimize at same time, because of their intrinsic dependency.
This leads to multi-objective optimisations, which consider lot of information to compute a placement of software offering the best tradeoff.
In this context, models are perfect candidates to represent the state of the cloud infrastructure and ease the work of these optimisation algorithms.
Starting from the model reflecting the current state of the cloud infrastructure, these algorithms explore and evolve lot of configurations to find the most suitable tradeoffs between all cloud metrics.
In this process, the high number of models (configurations as used by Ferry~\textit{and al}~\cite{ferry2013towards}) created and dropped calls for very efficient mechanisms for allocation, deletion, clone and mutation, because they constitute the basic operations of evolutionary algorithms~\cite{frey2013search}.
\subsubsection{Software systems self-adaptation using Models@Runtime}~\\
Models@runtime is an emerging paradigm aiming at making the evolution process of applications more agile and dynamic, still leveraging the key benefits of modeling: simplicity, efficiency and safety.
It is inspired by solid research in reflective software engineering~\cite{DBLP:conf/reflection/2001} ({\em e.g.} Meta-Object Protocol~\cite{Kiczales1991}) and makes available, at runtime, an abstraction of the system for the system itself or for external human/programmatic actors. This abstraction helps in supporting efficient decisions related to the adaptation or evolution of the system by hiding (simplifying) irrelevant details.
Kevoree~\cite{DBLP:conf/models/MorinFBJSDB08,Fouquet:2012:DCM:2304736.2304759} is an open-source project leveraging the Models@Runtime (M@RT) paradigm to support the continuous design of complex, distributed, heterogeneous and adaptive systems. The overhead implied by this advanced reflection layer (Models@Runtime) has to be minimized to reduce as much as possible the resources (computation and memory) mobilized, and should enable the efficient dissemination of models among nodes~\cite{fouquet:hal-00688707}, to manage the widest possible types of node (including Android, Java Embedded).
If the M@RT overhead is too heavy, the process tends to be centralized, impacting the overall resilience of the software system (single point of failure, etc).\\
Current MDE tools were however designed and engineered without any consideration for runtime constraints, since they were originally thought for design-time use only, where time and resource constraints are much softer. The following section gives a more detailed description of the properties to comply with when targeting a use at runtime.
\subsection{Generic requirements for using models at runtime}
\label{sec:generic:requirements}
In the three domains presented in~\ref{sec:usecases}, applications need to perform common operations on models, where performance and safety are critical issues.
The memory footprint is the first performance concern.
Indeed, the generated Object-Oriented API and all its dependencies must be compatible with the hosting device in terms of memory footprint, to enable the inclusion on low-resource IoT nodes.
Secondly, models should come with efficient (un)-marshaling techniques to enable the rapid exchange of models between distributed cloud nodes.
Also, model cloning and concurrent, yet safe, manipulations of models should be supported by the framework to enable distributed reasoning, {\em e.g.} based on exploratory techniques such as genetic algorithms for cloud management.
Finally, models should be traversable efficiently to enable rapid computation on already known elements, for comparison or merge of models for instance, which are basic operations.
Each of these key requirements for M@RT is further described in the following sections.
\subsubsection{Reduced memory footprint}
~\\
The memory overhead introduced by models at runtime has to be considered from both static and dynamic point of view.\\
The \textbf{static memory footprint} basically depends on the complexity (size) of the metamodel that defines the concepts and relationships among these concepts to be manipulated in the models.
Basically, the more complex a metamodel is, the bigger the generated Object-Oriented(OO) API is.
Also, metamodels themselves conform to a meta-metamodel (M3), such as Executable Meta-Object Facility (EMOF)\cite{OMG:2013} or ECORE, which can also participate in the static memory footprint. However, the footprint of the M3 remains stable independently of the size of the metamodels.\\
The number and size of the dependencies required by the generated OO API have to be minimized, because each device must provision these dependencies to be able to run the API.
Heavy dependencies would indeed increase the initialization or update time of a device and reduce the memory space available for the models.
The \textbf{dynamic memory footprint} required to manipulate a model using the generated OO API is linked to the size of the manipulated model.
Thus, the number and size of the objects required to instantiate a model element ({\em i.e.} the concept and its relationship) have to be watched, because of their impact on the number of model elements that can be kept in memory.
If this footprint is too high for a device, the model will be swapped ({\em i.e.} unloaded from memory to disk), consuming a lot of resources to perform this operation.
All this implies that the memory used to create a model element has to be controlled to allow the manipulation of big models on devices with limited resources.
\subsubsection{Efficient models (un)marshalling}
~\\
The efficiency of marshaling and unmarshaling operations is a key criteria for models@runtime, because they are the basic input/ouput operations on a model at runtime, and should not ``freeze'' the system while they are serialized or loaded.
Models are used in memory for validation, verification and system manipulation.
They are also extensively exchanged across local processes and remote nodes to disseminate decisions, and stored to disk for later use ({\em e.g.} system restore, {\em post mortem} analysis as with a flight data recorder (black box)).
To describe this requirement, the Figure~\ref{fig:MARSaveLoad} depicts two critical use cases of models I/Os ({\em i.e.} (un)marshalling operations) in Kevoree. The first is the exchange of models across nodes (1) to build a shared reflexive layer; the second concerns backup and restore operations of models (2).
\vspace{-2em}
\begin{figure}[h!]
\centering
\includegraphics[scale=0.40]{figures/MRTsaveload.pdf}
\caption{Model@Runtime (un)marshalling operation case study illustration}
\label{fig:MARSaveLoad}
\vspace{-2.5em}
\end{figure}
\subsubsection{Concurrent read/write usage of models}
~\\
When used at runtime, models are generally at the centre of a highly concurrent environment.
In the IoT domain for instance, different probes integrated in several devices can update a shared context model. Obviously, this shared model must offer safe and consistent read and write operations.
\vspace{-1em}
\begin{figure}[h!]
\centering
\includegraphics[scale=0.37]{figures/marSharedOps.pdf}
\caption{Shared Model@Runtime protection strategies}
\label{fig:MARSharedOps}
\vspace{-1em}
\end{figure}
Like shared memory~\cite{chase1994sharing}, the generated OO API must implement a protection strategy in order to offer safe and efficient concurrent read or concurrent read-and-write operations for the various producers writing in the model and consumers of these data.
In case of a concurrent read strategy, the generated API must ensure that the multiple threads can reliably, independently and consistently access and iterate over model elements.
Similarly in case of a concurrent write strategy, the framework must ensure the atomicity of each modification.
This is depicted on the left side (Atomic Read/Write) of Figure~\ref{fig:MARSharedOps} where {\it Reasoners} components are both producers and consumers.
The cloning of models is at the centre of this concurrency problem.
It acts as a concurrent protection for each producer/consumer, which works on its own clone of the current model, as presented in Figure~\ref{fig:MARSharedOps} that depicts a Cloud management use case where search-based algorithms use small mutations as basic operations to explore a space of potential configurations.
The memory and computational efficiency of the cloning strategies are thus key requirements for the generated API.
\subsubsection{Efficient navigation across model elements}
\label{sec:generic:nav}
The purpose of a DSL is to offer a \textbf{simple} way to represent and manipulate domain specific models.
Models can create very complex graph structures in which the navigation is complicated.
So complicated, that iterating over relations and model elements in order to find a specific element create serious issues on performance and code complexity.
To exemplify this problem, we consider a small excerpt of the Kevoree meta-model (Fig.~\ref{fig:kevoree_MM_path}) which is used to organize the placement of components and hosted nodes, on execution nodes.
\begin{figure}[hb!]
\centering
\begin{subfigure}[b]{.6\textwidth}
\centering
\includegraphics[width=\textwidth]{figures/minikev2.png}
\caption{Kevoree Metamodel exerpt}\label{fig:kevoree_MM_path}
\end{subfigure}%
\begin{subfigure}[b]{.4\textwidth}
\centering
\includegraphics[width=.8\textwidth]{figures/deepModel.png}
\caption{Kevoree Model example}\label{fig:kevoree_Model_path}
\end{subfigure}
\caption{Illustration of the case with Kevoree}\label{fig:kevoree_pathexample}
\end{figure}
With this meta-model one can easily build the Kevoree model presented in figure~\ref{fig:kevoree_Model_path}, where a component instance is hosted on a node called 'node4' itself hosted on 'node8', hosted on 'node7', hosted on 'node6'.\\
Let now imagine the software system is trying to place this component instance on a node according to several constraints.
To perform this task, several threads are running different algorithms on clones of the model.
The need for an efficient navigation becomes obvious when the system has to compare, select and merge the solutions from all threads.
\section{State of the practice: Using EMF to generated DSLs for runtime usage}
\label{sec:sota}
A natural way to create a DSL is to rely on tools and techniques well established in the MDE community, and in particular, the {\em de facto} EMF standard.
This section provides a brief overview of EMF and then discusses the suitability of this modeling framework to generate a Object-Oriented API usable at runtime, with respect to the requirements identified in the previous section.
\vspace{-1em}
\subsection{Overview}
\par EMF is a Java-based EMOF implementation and code generation facility for building languages, tools and other applications based on a structured data model.
From a metamodel conforming to the Ecore meta-metamodel (which is largely inspired by/aligned with the EMOF meta-metamodel), EMF provides tools and support to create Domain Specific Languages (DSLs) on top of the Eclipse IDE. Additional languages and tools like EMFText or GMF make it possible to rapidly define textual or graphical syntaxes for these DSLs.\\
\begin{figure}[h!]
\vspace{-2em}
\center
\includegraphics[width=0.8\textwidth]{figures/fsmmetamodel2.pdf}
\caption{Finite State Machine Metamodel used for Experiments}\label{fsmmeta}
\vspace{-2em}
\end{figure}
In order to evaluate the code generated by EMF, we consider one of the simplest well known meta-model within the MDE community: the Finite State Machine (FSM).
This metamodel (Fig.~\ref{fsmmeta}) is composed of four meta-classes: FSM, State, Transition and Action.
A FSM contains the States of the machine and references to initial, current and final states.
Each State contains its outgoing Transitions, and Transitions can contain Actions.
To measure the dynamic memory usage and time consumption on the different requirements, we use the code generated by EMF to programmatically create a flat FSM (each state has exactly one outgoing transition and one incoming transition, except the initial and final states) model composed of 100,000~State instances, 99,999 transitions and an action for each transition. The experimentations have been performed on a Dell Precision E6400 with a 2.5~GHz iCore~I7 and 16~GB of memory.
The following sections discuss the position of EMF {\em w.r.t.} the identified requirements.
\subsection{Memory aspects}
For this evaluation, we first generate the modeling API (source code) of the FSM metamodel with the {\it EMF genmodel} tool, in a project called \textit{testemfdependencies}.
\begin{figure}[h!]
\vspace{-2em}
\includegraphics[width=1.\textwidth]{figures/emfdependencies1.png}
\caption{Dependencies for each new metamodel generated code}\label{dependencies}
\vspace{-1em}
\end{figure}
Figure~\ref{dependencies} shows the dependencies needed for the generated modeling API to compile and run.
The analyse of these dependencies shows that the generated code is tightly coupled to the Eclipse environment and to the Equinox runtime (Equinox is the version of OSGi by the Eclipse foundation).
Although this is not problematic when the generated API is used in the Eclipse environment,
these dependencies are more difficult to resolve and provision in a standalone context outside Eclipse.
In the case of the FSM, the business code generated from the FSM metamodel is only {\bf 55~KB}, but requires {\bf 15~MB} of dependencies when packed in a standalone executable JAR. This represents an overhead of {\bf 99,6\%}. Moreover, the reflexive calls extensively used in the generated API makes it difficult to reduce the size with tools like ProGuard\footnote{ProGuard is a code shrinker (among other features not relevant for this paper) for Java bytecode: http://proguard.sourceforge.net/}. Beyond the technical for memory optimization at runtime, we can extract the following generality: by introducing dynamic reflexive call (e.g dynamic because came from dynamically created string) we forbid to use dead code analyser and shrinker, then we introduce a runtime overhead.
As for the dynamic memory, the creation of the experimental FSM model lasts for 376~ms and uses 104~MB of the heap memory.
This overhead is one of the main limitations of EMF when the generated API must be embedded at runtime.
\subsection{Load and Save operations}
EMF uses the XMI format\cite{OMG:2011fk} as default serialization strategy to allow any tool supporting this interchange format ({\em e.g.} Xtext~\cite{Merkle:2010:TMT:1869542.1869564,Eysholdt:2010:XIY:1869542.1869625}, EMFText~\cite{emftext}, GMF~\cite{gmf} or ObeoDesigner~\footnote{http://www.obeodesigner.com/}) to load/save the models.\\
EMF provides a generic loader that uses reflexive mechanisms.
It navigates the metamodel and makes reflexive calls to classes and methods of the generated API to load a model.
This loader is not generated, which limits the size of the generated code, but its genericness and reflexive approach have several drawbacks.
First, both the model and the metamodel have to be parsed and loaded in memory, which is consuming a lot of time and memory. The reflexive calls are also very costly.\footnote{See the Java tutorial on reflection, in particular the discussion on the performance overhead: http://docs.oracle.com/javase/tutorial/reflect/}
To hide and avoid some of this algorithmic complexity, the EMF loader implements a {\em lazy loading} mechanism which actually loads the attributes and references of model elements only if the model element is actually accessed.
As for the marshaling (save) operation, the EMF standard serializer works in two steps. It first transforms the model to be serialized into a Document Object Model (DOM). The size of this temporary DOM structure is directly linked to the size of the model. Then it prints this DOM in a file.
The serialization of our experimental FSM model to a file lasts 7021~ms and the loading of the model from this file lasts 5868~ms, which is a lot considering the power of the computer used for the experimentation.
\subsection{Concurrency in the generated code}
Many runtime platforms used to support dynamic software architectures implement their own class loaders to keep control of resources. It is the case in OSGi, Frascati or Kevoree for instance.
Then, software running on these platforms must use these class loaders in order for the platform to offer an efficient and reliable dynamic class loading.
However, the use of static registries in EMF leads to incompatibilities with runtime platforms using multi-class loaders.
Moreover, EMF provides no guarantee of thread safety of the generated code~\footnote{http://wiki.eclipse.org/EMF/FAQ\#Is\_EMF\_thread-safe.3F}.
Even worse, when a collection on a model element is accessed by two threads in parallel, the iterator on the collection may be shared.
Nevertheless, this need for a proper concurrency management is essential, because several IoT sensors may access the model at the same time or several cloud-management threads may put their results in the model concurrently.
EMF is thus not adapted for this kind of use.
As for the clone of models, we used the \texttt{EcoreUtil} to clone the FSM experimental model, and the process took 3588~ms, which is not optimal.
\vspace{-1em}
\subsection{Navigation in models}
The code generated by EMF provides an embedded visitor pattern and an observer pattern~\cite{gof94}.
Nevertheless, the navigation through model elements and the lookup of a specific known element can be very costly.
To illustrate that, we consider the previously defined Kevoree component lookup example detailed in section~\ref{sec:generic:nav} (Figure~\ref{fig:kevoree_Model_path}), and use the plain Java API generated by EMF.
The code required to access the component instance element in the model is presented in Listing~\ref{lst:plainJava}.
Let consider $n$ the number of nodes at the root of the model, $o$ the number of nodes hosted by 'node6', $p$ the number of nodes in 'node7', $q$ the number of 'node8', $r$ the number of component instance in 'node4'; the algorithmic complexity of that piece of code is O($n \times o \times p \times q \times r$) which creates a combinatorial explosion with the growth of nodes at each level, and the depth of nested nodes.
\begin{scriptsize}
\begin{lstlisting}[frame=L, language=java, caption={Collect of the FakeConsole component instance in plain Java.}, label={lst:plainJava}]
for (ContainerNode node : model.getNodes()) {
if (node.getName().equals("node6")) {
for (ContainerNode node2 : node.getHosts()) {
if (node2.getName().equals("node7")) {
for (ContainerNode node3 : node2.getHosts()) {
if (node3.getName().equals("node8")) {
for (ContainerNode node4 : node3.getHosts()) {
if (node4.getName().equals("node4")) {
for (ComponentInstance i : node4.getComponents()) {
if (i.getName().equals("FakeConso380")) {
fConsole = i;break;
}
[...]
}
\end{lstlisting}
\end{scriptsize}
Functional languages ({\em e.g.} Scala), or languages dedicated to model query/transformations ({\em e.g.} OCL, QVT) can of course be used instead of a plain Java approach to facilitate the writing of this code, but come with additional integration costs and worse performances, typically due to the use of an interpreter.
A common workaround for this performance issue is the creation of cache or similar buffer approaches.
Such techniques introduce new problems like eventual consistency, which are often in opposition with multi-thread access.
As a consequence, common modeling projects use helpers with high computational complexity to index and search model elements, which has serious performance penalty at runtime.
\vspace{-1em}
\subsection{Synthesis}
This section shows that MDE tools, initially developed for design-time activities can be used to create proof-of-concept Models@Runtime platforms.
However, several limitations (in terms of memory and time efficiency) and drawbacks (thread safety not guaranteed) make their use sub-optimal or even impossible at runtime.
In order to improve MDE tools for a runtime usage, and make them generally more efficient, we focus the work described in this paper on four questions:
\begin{enumerate}
\item How to reduce the number of dependencies and intermediate objects created, to reduce the memory requirements ?
\item How to take advantage of the information available in the metamodel to generate more efficient loaders and serializers?
\item How to offer an efficient support for concurrent access to models at runtime?
\item How to efficiently access a specific model element?
\end{enumerate}
In the next section, we describe the Kevoree Modeling Framework which implements our contributions to address these questions.
|
\section{Introduction}
For explicit calculations of instationary solutions to hyperbolic
conservation laws, the timestep is dictated by the CFL condition due
to Courant, Friedrichs and Lewy \cite{CourantFriedrichsLewy1928},
which requires that the numerical speed of propagation should be at
least as large as the physical one. For implicit schemes, the CFL
condition does not provide a restriction, since the numerical speed of
propagation is infinite. Depending on the equations and the scheme,
restrictions may come in via the stiffness of the resulting
nonlinear problem. These restrictions are usually not as strict as in
the explicit case, where the CFL number should be below unity. For
implicit calculations, CFL numbers of 10, 100 or even 1000 may well be
possible. Therefore, it is a serious question how large the timestep,
i.e. the CFL number, should be chosen.
We are particularly interested in timestep control which is based upon
computable, a-posteriori error estimates. In \cite{KroenerOhlberger2000,
Ohlberger2001} Kr\"oner and Ohlberger based their space-time adaptivity upon
$L^1$, Kuznetsov type estimates for scalar conservation laws. In
\cite{Johnson2,Johnson3,Johnson4,Johnson5,Johnson6}, Eriksson and Johnson developed space-time
adaptive methods for parabolic pde's. These a-posteriori error estimates require
the solution of an adjoint problem. A space-time projection of the adjoint
solution makes it possible to consider spatial and temporal error separately.
They closed the error estimates by an a-priori bound on the dual solution. In
\cite{Sueli, SueliHouston2003}, S\"uli and Houston developed an analogous
approach for hyperbolic transport equations.
The work of Eriksson and Johnson has been extended by many authors, see, for
example, the review articles of Becker and Rannacher \cite{BeckerRannacher1996,
BeckerRannacher2000}
and of Hoffman and Johnson \cite{HoffmanJohnson}. We would like to mention that we learned
a lot about these developments from the unpublished thesis of Ralf Hartmann
\cite{Hartmann1998}. Instead of relying upon an (usually pessimistic) a-priori
error estimate for the adjoint solution, Hartmann and others
\cite{HartmannHoustonSISC2002, SueliHouston2003} {\em computed} the adjoint solution and hence
obtained an (in principle exact) error representation.
More recently these methods have also been developed
for hyperbolic problem by Barth, Hartmann, Houston, Giles, S\"uli,
Schwab and others. An excellent collection of review papers may
be found in \cite{BarthDeconinck2003}.
Let us briefly summarize the space-time splitting of the adjoint error
representation (see \cite{Johnson2,Johnson3,Johnson4,Johnson5,Johnson6,
BeckerRannacher1996,Sueli,Hartmann1998} for details). The error representation
expresses the error in a target functional as a scalar product of the finite
element residual with the dual solution. This error representation is decomposed
into separate spatial and temporal components. The spatial part will decrease
under refinement of the spatial grid, and the temporal part under refinement of
the timestep. Technically, this decomposition is achieved by inserting an
additional projection. Usually, in the error representation, one subtracts from
the dual solution its projection onto space-time polynomials. Now, we also
insert the projection of the dual solution onto polynomials in time having
values which are $H^1$ functions with respect to space.
This splitting can be used to develop a strategy for a local choice of
timestep. Here we add to the
results in \cite{Sueli,Hartmann1998} by studying a weakly instationary
solution to Burgers' equation, for which the timestep will be very large
(and we will quantify this) in regions of stationary flow, and become
small when a perturbation enters the flow field. We believe that this type
of flow is a prime example where the space-time splitting can become useful.
Besides applying adjoint techniques which are already well-established to a new
test problem, we also add a new ingredient which simplifies and accelerates the
computation of the dual problem. Due to Galerkin orthogonality, the dual
solution $\varphi$ does not enter the error representation as such. Instead, the
relevant term is the difference of the dual solution and its projection to the
finite element space, $\varphi-\varphi_h$. We can show that it is therefore sufficient to
compute the spatial gradient of the dual solution, $w=\nabla \varphi$. This gradient
satisfies a conservation law instead of a transport equation, and it can
therefore be computed with the same algorithm as the forward problem, and in the
same finite element space.
Our goal here is time step adaptation. Ultimately, this will become a
building block of an aerodynamic and aeroelastic solver which is
currently being developed by the SFB 401 research group at RWTH Aachen
\cite{BramkampLambyMueller2004}. In that solver, multiscale analysis
is used to compress data, coarsen and refine the spatial grid. Time stepping for
instationary problems is done by a methods of lines approach, using
explicit or implicit Runge-Kutta schemes. The latter is, of course, a
standard set-up used for aerodynamic, or conservation law, solvers.
In the aerodynamical applications which we have in mind, we may have
to resolve many different features of the flow, more than can be
controlled by a small number of functionals like drag and
lift. Therefore, an adaptive monitoring of the complete flow field, as
done by the multiscale analysis, is very desirable.
Here we develop our strategy for a test case. Since we focus on timestep
adaptation we will use uniformly refined meshes in space.
Starting with a very coarse spatial mesh and CFL below unity, we
gradually establish sequences of timesteps which are well adapted to
the physical problem at hand. The scheme detects stationary regions,
where it switches to very high CFL numbers, but reduces the time
steps appropriately as soon as a perturbation enters the flow field.
Depending on the CFL number and the cost of the nonlinear solver, the adaptive
scheme chooses either explicit or implicit timesteps. For reasons of efficiency,
very small timesteps $C\!F\!L \ll 1$ may be merged into a single step. This strategy
is detailed in Section \ref{sec.applications.strategy}.
Once we arrive at the fine spatial mesh, on which we really want to
compute and where most of the work is being done, we already work with
a very efficient time step. Moreover, we have a rational criterion
what the finest grid should be.
The paper is organized as follows: in Section \ref{sec.theory} we review the
theoretical background for our adaptive timestep control: DG and FV methods,
control of target functionals, error representation, space-time splitting, error
estimates. The new conservative approach for solving the dual problem is
presented in Section~\ref{sec.newapproach}. In Section \ref{sec.applications} we
define our adaptive strategy and apply it to compute perturbations of a
stationary shock. Some conclusions are drawn in Section \ref{sec.conclusions}.
\section{Derivation of space-time-split error estimates}
\label{sec.theory}
In this section, we recall some of the theoretical background of
adjoint error control, and we represent the extensions needed in our
time adaptive strategy.
In Section~\ref{sec.dg_method} we introduce
the DG method used in the paper. In Section
\ref{sec.error_representations} we state the adjoint based error
representation for target functionals. In
Section~\ref{sec.space_time_splitting} we introduce a variant of the
projections in
space and time which lead to a splitting of the error representation.
One part decreases when the spatial grid is refined, and
the other part decays with the timestep.
The corresponding decay rates are a crucial ingredient of the time-adaptation strategy. This
strategy and its application will be presented in
Section~\ref{sec.applications} below.
\subsection{Discontinuous Galerkin methods for conservation laws}
\label{sec.dg_method}
Let $D$ be an open connected subset of $\mathbb R^d$, $d\geq 1$, let
$I := (0,T)$ be the time interval and let $\Omega := D\times I$ be the
space-time domain, with boundary $\Gamma$ and outside unit normal
$\nu$. We consider the system
\begin{align}
\partial_t u +\nabla f(u) &=0 \quad \mbox{in } \Omega, \\
\quad f_\nu(u) &= \gamma_\nu \quad \mbox{on }\Gamma_{in}, \label{eq.cl}
\end{align}
where $u=(u_1,\ldots,u_m)^T :\Omega \rightarrow \mathbb R^d$ is the vector of conservative variables
and $f(u) = (f_1(u), \dots, f_d(u))$ the flux matrix, with $f_i \in
C^1(\mathbb R^m, \mathbb R^m)$. The vector
\begin{align*}
f_\nu(u) := (f(u),u)\cdot\nu
\end{align*}
is the space-time normal flux across the boundary, and for scalar
equations, the inflow boundary is given by
\begin{align*}
\Gamma_{in} := \{(x,t) \in \Gamma \mid \frac{d}{du}((f(u),u)\cdot\nu < 0 \}.
\end{align*}
Note that \eqref{eq.cl} includes initial data, since $D \times \{0\} \subset \Gamma_{in}$, and
$f_\nu(u(x,0)) = u(x,0)$. For systems of conservation laws, the definition of in- and outflow
boundaries may be generalized via characteristic decompositions
\cite{HartmannHoustonJCP2002,SueliHouston2003}.
Let us define a partition of our time interval $I$ into subintervalls
$I_n = (t_{n-1}, t_n)$, where
\begin{align*}
0=t_0 < t_1 < \ldots < t_n < \ldots < t_N=T.
\end{align*}
Later on this partition will be defined automatically by the
adaptive algorithm. Furthermore we define a regular polygonal
spatial grid $\mathcal {T}_D = \bigcup_j \{D_j\}$ such
that $\overline D = \bigcup_j \overline{D_j}$. We denote the
corresponding space time prisms by
$$
\Omega_j^n := D_j \times I_n.
$$ For future reference, we denote the outward unit normal vector to
$\Omega_j^n$ by $\nu_j^n$ or simply $\nu$. Thus we have constructed a
subdivision
\begin{align*}
\mathcal{T}_{\Omega} = \bigcup_{j,n} \{\Omega_j^n\}
\end{align*}
of the computational domain $\Omega$. The spatial discretisation
$\mathcal T_D$ can change adaptively from timestep to timestep,
and for each fixed time interval $I_n$, the timestep is global
(i.e. it is the same for all spatial cells $D_j$).
\begin{remark}
We do not admit local timesteps, since we want to couple our time-adaptive strategy
to standard Runge-Kutta Finite Volume methods and Runge-Kutta
Discontinuous Galerkin methods.
\end{remark}
On this grid we define the following function spaces: First, let $S_h(\Omega)$
be the mesh dependent broken space of discontinuous piecewise $H^1$
functions defined on $\mathcal{T}_{\Omega}$,
\begin{align}
S_h(\Omega) := \left\lbrace u \mid u_{\mid_{\Omega_j^n}}\in H^1(\Omega_j^n), \forall
\Omega_j^n \in \mathcal{T}_{\Omega}\right\rbrace .
\end{align}
Furthermore we denote by $S_h^{s,r}(\Omega)$ the (locally) finite dimensional
space consisting of discontinuous piecewise polynomial functions of
degree $s$ in space and $r$ in time defined on $\mathcal{T}_h$
\begin{align}
S_h^{s,r}(\Omega) := \left\lbrace u_h \mid u_{h}(\cdot,t)\in
P_{s}(D_j), \forall t\in {I_n},
u_{h}(x,\cdot)\in
P_{r}(I_n), \forall x\in {D_j},
\forall D_j \times I_n\in \mathcal{T}_{\Omega}\right\rbrace ,
\end{align}
where $P_{r}(I_n)$ denotes the space of polynomials of degree $r$ on
$I_n$ and $P_{s}(D_j)$ the space of polynomials of degree $s$ on
$D_j$. Given a cell $\Omega_j^n$ and a point $(x,t)\in\Gamma_j^n$, we define the
inner ($u^+$) and outer ($u^-$) values of a function $u \in S_h(\Omega)$ with
via
\begin{align}\label{eq.fnpm}
u^\pm(x,t) := \lim _{{\delta \searrow 0 +} \atop {\delta \nearrow 0 -}}
u((x,t)-\delta \nu_j^n).
\end{align}
Defining the DG method for nonlinear conservation laws, whose
solutions in general contain shock waves, requires a careful
application of the theory of weak solutions, which states that
for a weak solution $u$ and a continuously differentiable test
functions $v$,
\begin{align*}
-(u,\partial_t v)_{\Omega_j^n} -(f(u),\nabla v)_{\Omega_j^n} +
(f_\nu(u),v)_{\Gamma_j^n} = 0 \quad \forall j,n.
\end{align*}
Thus we have to define the normal flux $f_\nu(u)$ at the cell
boundaries, where the approximate solution $u_h$ is
discontinuous. This can be done with the help of numerical flux
functions, which we denote by $f_\nu^\ast$. So suppose that
$(x,t)\in\Gamma_j^n\setminus\Gamma$ is contained in an interior edge. If
$(x,t) \in \partial D_j \times I_n$, so that the normal points into
the spatial direction, then the canonical choice for $f_\nu^\ast$ is an
approximate Riemann solver
\begin{align}\label{eq.fn1}
f_\nu^\ast := f(u_h^+,u_h^-,n_j),
\end{align}
where $n_j$ is the outer normal to $D_j$ (i.e. $\nu_j^n=(n_j,0)$).
We require that the flux $f_\nu^\ast$ is consistent and conservative
in the sense of Lax.
If, on the other hand, $(x,t)\in D_j \times \partial I_n$, so that the
normal points into the time direction and $f_\nu(u)=u$, then we simply
require that $f_\nu^\ast$ be a convex combination of $u_h(x,t^\pm)$. More
specifically, suppose that $t=t_n$. Then we set
\begin{align}\label{eq.fn2}
f_\nu^\ast := u_h^\ast(x,t_n) := (1-\theta) u_h(x,t_n^+) + \theta u_h(x,t_n^-)
\end{align}
for some value $\theta \in [0,1]$.
Different values of $\theta$ will yield different time discretisations, e.g.
explicit Euler for $\theta =0$, implicit Euler for $\theta =1$, if we work
with piecewise constant ansatz functions.
On the boundary of the domain, i.e. for $(x,t)\in\Gamma$, we set
\begin{align}\label{eq.fn3}
f_\nu^\ast := \left\{ \begin{array}{ll}
\gamma_\nu & \mathrm{if} \; (x,t)\in\Gamma_{in} \\
f_\nu(u^+_h(x,t)) & \mathrm{if} \; (x,t)\in\Gamma_{out}
\end{array} \right.
\end{align}
In the following definition we simply state the
resulting DG(s,r) method, which is a discontinuous method both in
space and time. This definition is very similar to, see e.g.
\cite{BarthLarson2002,CockburnHouShu1990,HartmannHoustonSISC2002,
SueliHouston2003} and the references therein.
\begin{definition}
(i) The abstract semilinear form
$\mathcal N : S_h(\Omega) \times S_h(\Omega) \to \mathbb R$ is given by
\begin{align}\label{eq.dg_form}
\mathcal N \left(u_h, v_h\right)
:= \sum_{j,n}
\left\lbrace
(\partial_t u_h + \nabla f(u_h),v_h)_{\Omega_j^n}
+ (f_\nu^\ast - f_\nu(u_h^+),v_h^+)_{\partial\Omega_j^n}
\right\rbrace.
\end{align}
(ii) Now the DG(s,r) finite element method for the system of
hyperbolic conservation laws \eqref{eq.cl} is defined as follows: Find
$u_h \in S_h^{s,r}(\Omega)$, such that
\begin{align}
\mathcal N \left(u_h, v_h\right) =0 \quad \forall v_h \in S_h^{s,r}(\Omega).\label{eq.dg}
\end{align}
\end{definition}
As usual, the variational formulation \eqref{eq.dg_form},
\eqref{eq.dg} can be exploited as follows: Given $u_h \in S_h(\Omega)$,
$\mathcal N (u_h,\cdot)$ is a linear functional on $S_h(\Omega)$. Thus it can
be represented by an element of $S_h(\Omega)$, which we call $R(u_h)$, the
{\em residual}. On the interior of a cell $\Omega_j^n$ we introduce the
{\em cell residual}
\begin{equation}\label{eq.r1}
R_h := \partial_t u_h + \nabla f(u_h)
\end{equation}
and on the boundaries $\Gamma_j^n$ the {\em edge residual}
\begin{equation}\label{eq.r2}
r_h := f_\nu^\ast - f_\nu(u_h^+).
\end{equation}
Then \eqref{eq.dg_form} can be rewritten as
\begin{equation}\label{eq.r3}
(R(u_h),v_h)
= \sum_{j,n}
\left\lbrace
(R_h,v_h)_{\Omega_j^n} + (r_h,v_h^+)_{\partial\Omega_j^n}
\right\rbrace.
\end{equation}
The $DG(s,r)$ solution $u_h\in S_h^{s,r}(\Omega)$ of \eqref{eq.dg}
is now given by
\begin{equation}\label{eq.galerkin_orthogonality}
(R(u_h),v_h) = 0 \quad \forall v_h \in S_h^{s,r}(\Omega),
\end{equation}
which is the classical Galerkin orthogonality: the residual $R(u_h)$ is
orthogonal to the test space $S_h^{s,r}(\Omega)$.
In the following, we mostly work with the $DG(0,0)$ and $DG(1,1)$ methods, both
in their explicit ($\theta=0$) and implicit ($\theta=1$) form. The DG(0,0)
method is equivalent to a first order accurate finite volume scheme, using
explicit order implicit Euler scheme for the time integration.
\iffalse
For higher oder
accurate finite volume schemes we refer to Barth and Larson
\cite{BarthLarson2002}.
\fi
In \cite{BarthLarson2002}, Barth and Larson derive a weak formulation of the
form \eqref{eq.galerkin_orthogonality} for higher oder accurate finite volume schemes.
Therefore, the techniques presented in this paper can be
applied to finite volume and Discontinuous Galerkin methods.
\subsection{Adjoint error representation for target functional}
\label{sec.error_representations}
In this section we define the class of target functionals treated in
this paper, state the corresponding adjoint problem and recall the
classical error representation which we will later use for
adaptive time step control.
Our objective is to estimate the error in a user specified functional
$J(u)$, which can be expressed as a sum of weighted integrals over the
domain $\Omega$ and the outflow boundary $\Gamma_{out}$.
Typical examples of such functionals are the lift or the drag of a
body immersed into a fluid.
To simplify matters we consinder functionals of the following form:
\begin{align*}
J(u) = (u,\psi)_{\Omega} - (f_\nu(u),\psi_\Gamma)_{\Gamma_{out}}
\end{align*}
Our purpose is to control the error
\begin{align*}
J(u)-J(u_h) .
\end{align*}
In order to derive the classical error representation one linearizes
the evolution equation satisfied by the error $u-u_h$ and works with
the adjoint equation of the linearized error equation. Thus we
introduce an approximate Jacobian $\overline{a}(u;u_h)$ of $f$ by
\begin{align}
\overline{a}(u;u_h) &:= \int\limits_0^1 \frac{d}{d\tau}f(u_h+\tau(u-u_h))d\tau. \label{eq.ab}
\end{align}
Note that
\begin{align*}
f(u)-f(u_h) &= \overline{a}(u;u_h) (u-u_h).
\end{align*}
In practice we linearize around the approximate solution.
A direct calculation yields the following theorem:
\begin{theorem}\label{theorem.error_representation}
Suppose $\varphi \in H^1(\Omega)$ solves the adjoint problem
\begin{align}\label{eq.adjoint_1}
\varphi &= \psi_\Gamma \quad \mathrm{on} \; \Gamma_{out} \\
\label{eq.adjoint_2}
\partial_t \varphi + \overline{a}(u;u_h) \nabla \varphi &= \psi \quad \; \; \mathrm{in} \; \Omega.
\end{align}
Then for all $\varphi_h \in S_h^{s,r}(\Omega)$, the error in the target functional satisfies
\begin{align}\label{eq.error_representation}
J(u)-J(u_h) = ( R(u_h),\varphi-\varphi_h).
\end{align}
\end{theorem}
Equivalently one can also define the adjoint solution via a
variational formulation (see e.g.
\cite{BarthLarson2002,HartmannHoustonSISC2002,SueliHouston2003}).
In \cite{Tadmor1991} Tadmor proves the well-posedness of the
adjoint problem \eqref{eq.adjoint_1} -- \eqref{eq.adjoint_2}
for scalar, convex, one-dimensional conservation laws.
The key observation is that, if the forward solution $u$ has jump
discontinuities, then due to the entropy condition the jump of the
transport coefficient $\overline{a}(u;u_h)$ has a distinct sign. This makes it
possible to follow the characteristics of the adjoint problem
backwards in time.
Identity \eqref{eq.error_representation} is the error representation
which we discussed in the introduction and onto which we are going to
base our adaptive strategy. By definition \eqref{eq.r1} -
\eqref{eq.r3} of the residual $ R(u_h)$, the error
representation may be decomposed as a sum over the cells and edges of
inner products of the local residuals with the solution of our dual
problem. Due to Galerkin orthogonality
\eqref{eq.galerkin_orthogonality}, we can subtract an arbitrary
test function $\varphi_h$, which is very convenient when we derive local
error estimates later on.
\subsection{Space-time splitting and the error estimate}
\label{sec.space_time_splitting}
The error representation \eqref{eq.error_representation} is not yet
suitable for time adaptivity, since it combines space and time
components of the residual and of the difference $\varphi-\varphi_h$ of the dual
solution and the test function. The main result of this section is an error estimate whose components
depend either on the spatial grid size $h$ or the time step $k$, but
never on both. The key ingredient is a space-time splitting of
\eqref{eq.error_representation} based on $L^2$ projections.
Similar space-time projections were introduced previously
in \cite{Hartmann1998, Sueli}.
Here we adapt them to the finite element spaces used in
the error representation \eqref{eq.error_representation}.
Let $P_{s,r}(\Omega_j^n) = P_s(D_j)\times P_r(I_n)$ be the space of polynomials of
degree $s$ on ${D_j}$ and $r$ on $I_n$.
Furthermore let
$\hat P_{I_n}^r(\Omega_j^n)=\{w\in L^2(\Omega_j^n)|w(x,\cdot)\in P_r(I_n),\forall x\in D_j\}$, and
$\hat P_{D_j}^s=\{w\in L^2(\Omega_j^n)|w(\cdot,t)\in P_s(D_j),\forall t\in I_n\}$.
For $r \geq 0$ define the $L^2$ projection
$\Pi_{I_n}^r: L^2(\Omega_j^n) \to \hat P_{I_n}^r(\Omega_j^n)$ via
\begin{align}
(u-\Pi_{I_n}^r u, \varphi)_{I_n} = 0 \quad \forall \varphi\in \hat P_{I_n}^r(\Omega_j^n), \forall x\in D_j,
\end{align}
and for $s \geq 0$ define the $L^2$ projection
$\Pi_{D_j}^s: L^2(\Omega_j^n) \to \hat P_{D_j}^s(\Omega_j^n)$ via
\begin{align}
(u-\Pi_{D_j}^s u, \varphi)_{D_j} = 0 \quad \forall \varphi\in \hat P_{D_j}^s(\Omega_j^n), \forall t\in I_n.
\end{align}
Similarly let the $L^2$ projection
$\Pi_{\Omega_j^n}^{s,r}: = L^2(\Omega_j^n) \to P_{s,r}(\Omega_j^n)$
be defined via
\begin{align}
(u-\Pi_{\Omega_j^n}^{s,r}u,\varphi)_{\Omega_j^n}=0\quad \forall \varphi\in
P_{s,r}(\Omega_j^n).
\end{align}
Note that $\Pi_{\Omega_j^n}^{s,r} = \Pi_{{D_j}}^s \Pi_{I_n}^r$.
First we choose $\varphi_h$ in the error representation
(\ref{eq.error_representation}) to be $\varphi_h =\Pi_{h,k}^{s,r} \varphi$,
i.e. $\varphi_h\mid_{\Omega_j^n} =\Pi_{{D_j}}^s \Pi_{I_n}^r \varphi= \Pi_{I_n}^r \Pi_{{D_j}}^s
\varphi$, with $\Pi_{I_n}^r$ and $\Pi_{{D_j}}^s$ as defined above.
Using the identity
\begin{align*}
\varphi-\Pi_{h,k}^{s,r}\varphi =\varphi-\Pi_{I_n}^r \varphi + \Pi_{I_n}^r
\varphi -\Pi_{h,k}^{s,r}\varphi = (id-\Pi_{I_n}^r
)\varphi+(id-\Pi_{D_j}^s )\Pi_{I_n}^r \varphi
\end{align*}
we obtain the following splitting of the error representation:
\begin{align}\label{eq.error_representation2}
J(u)-J(u_h) &= ( R(u_h),(id-\Pi_{I_n}^r
)\varphi+(id-\Pi_{D_j}^s )\Pi_{I_n}^r \varphi)\\
&= \sum_{j,n}
\{ \underbrace{(R_h,(id-\Pi_{I_n}^r)\varphi)_{\Omega_j^n} +
(r_h,(id-\Pi_{I_n}^r)\varphi^+)_{\partial\Omega_j^n}}_{\eta_{k}^{jn}}\\
&\phantom{ = \sum_{j,n}}+\underbrace{(R_h,(id-\Pi_{D_j}^s )\Pi_{I_n}^r \varphi)_{\Omega_j^n} +
(r_h,(id-\Pi_{D_j}^s )\Pi_{I_n}^r \varphi^+)_{\partial\Omega_j^n}
}_{\eta_{h}^{jn}}\}\\ \label{eq.error_representation3}
&=: \eta_k +\eta_h,
\end{align}
where $\eta_k$ is the time-component and $\eta_h$ the space-component of the
error representation $\eta$.
In this paper we consider grids, which are locally tensor products of a
spatial grid $\mathcal {T}_{D_j}$ and a timestep $I_n$. For an implicit Runge-Kutta Finite Volume Method
$\eta_{k}^{jn}$ and $\eta_{h}^{jn}$ then take the form
\begin{align*}
\eta_{k}^{jn}=&
(R_h, (id-\Pi_{I_n}^r)\varphi)_{D_j \times I_n}
+(f_\nu^\ast - f_\nu(u_h^+),(id-\Pi_{I_n}^r)\varphi^+)_{\partial D_j \times I_n} \\
&+ (\left[ u_h \right]_{n-1} ,(id-\Pi_{I_n}^r)\varphi_{n-1}^+)_{D_j}\\
\eta_{h}^{jn} =& (R_h, (id-\Pi_{D_j}^s )\Pi_{I_n}^r \varphi)_{D_j \times I_n}
+(f_\nu^\ast - f_\nu(u_h^+),((id-\Pi_{D_j}^s )\Pi_{I_n}^r \varphi)^+)_{\partial D_j \times
I_n} \\
&+ (\left[ u_h \right]_{n-1} ,((id-\Pi_{D_j}^s )\Pi_{I_n}^r
\varphi)_{n-1}^+)_{D_j}
\end{align*}
where the flux difference on the spatial boundaries $\partial D_j \times I_n$
and the jump of $u_h$ on the time boundary $D_j\times\{t_{n-1}\}$ are
realizations of the residual term $r_h$ in \eqref{eq.r2}.
For future reference, we also introduce the quantities
\begin{align}
\eta_k &:= \sum_{j,n}\eta_{k}^{jn}&
\eta_h &:= \sum_{j,n}\eta_{h}^{jn}\\
\label{eq.bar_eta_kn}
\bar\eta_k^n&:= \frac{1}{k_n}\sum_{{D_j}\in \mathcal T}|\bar\eta_{k}^{jn}|&
\bar\eta_h^n&:= \frac{1}{k_n}\sum_{{D_j}\in \mathcal T}|\bar\eta_{h}^{jn}|\\
\bar\eta_k &:= \sum_{n}k_n|\bar\eta_{k}^{n}|&
\bar\eta_h &:= \sum_{n}k_n|\bar\eta_{h}^{n}| \\
\bar\eta &:=\bar\eta_k +\bar\eta_h.&
\end{align}
In Section \ref{sec.SEP} we will show numerically, that the error terms
$\bar\eta_{k}$ and $\bar\eta_{h}$ depend on $k$ and $h$.
\section{A new approach to solving the adjoint problem}
\label{sec.newapproach}
The error representation \eqref{eq.error_representation} assumes that the exact
solution $\varphi$ of the dual problem \eqref{eq.adjoint_2} is available. This
is, of course, not the case. All we can do is to compute an approximation
$\varphi_\sharp$ of $\varphi$. An important question is in which space we should choose
the approximation $\varphi_\sharp$ (let us call this space $S_\sharp$ ). If we choose $S_\sharp
\subseteq S_h^{s,r}$, then - due to Galerkin orthogonality of the residual - the
error representation \eqref{eq.error_representation} would return zero.
Therefore, $S_\sharp$ should not be contained in $S_h^{s,r}$.
There are essentially three approaches in the literature to compute an
approximate solution to the dual problem. The first approach is to keep the
polynomial degrees $r$ and $s$ fixed, but compute the solution to the dual
problem on a finer grid $\mathcal {T}_{D_j} \subset \mathcal {T}_{D _{j+1}} $.
The second approach is to compute the dual solution using higher order finite
elements and using projections to get $\varphi_h$:
\begin{align*}
\mathrm{Compute\!:} \;\; \varphi_\sharp \in S_h^{s+1,r+1}(\Omega)
\qquad\rightsquigarrow\qquad
\varphi_h:= \Pi \varphi_\sharp,
\end{align*}
where $\Pi$ is the projection from the higher order finite element space onto the test space $S_h^{s,r}(\Omega)$.
The third way is to compute a solution in the test space of the forward problem, which means to use
the same order finite elements, and then do a higher order reconstruction $R$.
\begin{align*}
\mathrm{Compute\!:}\;\;\varphi_h \in S_h^{s,r}(\Omega)
\qquad\rightsquigarrow\qquad
\varphi_\sharp := R\varphi_h
\end{align*}
In the following, we describe a fourth approach, which avoids to approximate
$\varphi$ alltogether. Instead, we approximate the spatial gradient
$\nabla_x\varphi$. The remarkable fact is that this gradient satisfies a
conservation law instead of a nonlinear transport equation, and its numerical
approximation is therefore very robust in the presence of shocks.
In the present paper we limit our presentation to first order schemes in one
space dimension. Our approach can be applied to the dual problem, if the forward
problem is approximated by a first order DG method, or a Finite Volume method.
The backward problem can then be computed by the same method as the forward
problems. The generalization of our ansatz to higher order schemes is relatively
straightforward in one space dimension.
Let us look at the details: Due to Galerkin orthogonality, the dual solution
$\varphi$ does not enter the error representation as such. Instead, the relevant
term is the difference of the dual solution and its projection to the finite
element space, $\varphi-\varphi_h$. Using one of the three methods described
above, one needs additional degrees of freedom to compute an approximation
$\varphi$ to the dual problem, and some computed information will never be used,
since only the difference $\varphi -\varphi_h$ enters the error representation.
Therefore we suggest to compute the spatial gradient of the dual solution.
To illustrate our approach (still in one spatial dimension), we assume that
$\varphi_h$ is the piecewise constant function satisfying
\begin{eqnarray*}
\varphi_h(x,t) &\equiv& \varphi(x_0, t_0) \quad\mathrm{for}\quad (x,t)\in
D_j\times I_n.
\end{eqnarray*}
for some given point $(x_0, t_0)\in D_j\times I_n$ (e.g. the midpoint).
Expanding $\varphi$ around $(x_0,t_0)$,
\begin{eqnarray*}
\varphi(x,t) &=&\varphi(x_0,t_0)+(x-x_0) \partial_x \varphi(x_0, t_0) +(t-t_0)
\partial_t
\varphi(x_0, t_0)+O(h^{2}+k^2),
\end{eqnarray*}
and using the adjoint equation \eqref{eq.adjoint_2}, we obtain that
\begin{align*}
\varphi-\varphi_h(x,t) &=(x-x_0) \partial_x \varphi(x_0, t_0) +(t-t_0)
\partial_t \varphi(x_0, t_0)+O(h^{2}+k^2)\\
&=(x-x_0) \partial_x \varphi(x_0, t_0)+(t-t_0)
(\psi-\overline{a}(u;u_h) \partial_x \varphi(x_0,t_0 )) + O(h^{2}+k^2)\\
&=[(x-x_0)+(t-t_0)(\psi-\overline{a}(u;u_h))] \partial_x \varphi(x_0, t_0)
+ O(h^{2}+k^2).
\end{align*}
Since $\psi$ and $\overline{a}(u;u_h)$ are assumed to be known, the only unknown function is $\partial_x
\varphi(x_0, t_0)$.
In order to derive the differential equation which is satisfied by
$\partial_x\varphi$, we differentiate the adjoint equation \eqref{eq.adjoint_2},
\begin{align*}
\varphi &= \psi_\Gamma \quad \mathrm{on} \; \Gamma_{out} \\
\partial_t \varphi + \overline{a}(u;u_h) \partial_x \varphi &= \psi \quad \; \; \mathrm{in} \; \Omega
\end{align*}
with respect to $x$ and obtain
\begin{align}
w &= \partial_x \psi_{\mathit{\Gamma}} \quad \mathrm{on} \; \Gamma_{out} \label{eq.dp2.1}\\
\partial_t w + \partial_x (\overline{a}(u;u_h) w) &= \partial_x \psi \quad \; \; \mathrm{in} \;
\Omega,
\label{eq.dp2.2}
\end{align}
where $w:= \partial_x \varphi$.
Therefore it is not necessary to compute the approximations
$\varphi_\sharp$ and $\varphi_h$ of $\varphi$, but it is sufficient to compute an
approximation $w_\sharp \in S_h^{s,r}(\Omega)$ of $\partial_x \varphi$.
\begin{remark}
It is striking to note that the gradient $w= \partial_x \varphi$
actually satisfies a conservation law, \eqref{eq.dp2.1}-\eqref{eq.dp2.2},
instead of a linear transport equation,
\eqref{eq.adjoint_1}-\eqref{eq.adjoint_2}. Therefore, $w_\sharp$ can be
computed with the same algorithm as the forward problem, and in the same finite
element space. This leads to an efficient and robust solver: for discontinuous
$\overline{a}(u;u_h)$, finite difference schemes for \eqref{eq.adjoint_1}-\eqref{eq.adjoint_2}
may suffer from serious stability problems. Due
to their upwind nature, finite volume schemes for the conservation law
\eqref{eq.dp2.1}-\eqref{eq.dp2.2} handle discontinuous coefficients easily.
In work in progress, we are analysing the efficiency of the new approach in more
detail, generalize it to higher order and several space dimensions, and study
related issues like boundary conditions for compressible fluid flows.
\end{remark}
We will use this new approach in the numerical examples in Section
\ref{sec.applications}.
\section{Time adaptive strategy and application to perturbed shocks}
\label{sec.applications}
In this section we describe the strategy for adaptive time step
control, define a suitable numerical experiment and present first
numerical results which demonstrate the potential of this approach.
\subsection{The adaptive strategy}
\label{sec.applications.strategy}
In many applications, there are canonical target functionals which are of great interest to the user, like the lift and drag in aerodynamics. In some cases, an error margin may be prescribed for a given application. In other cases, it is less clear which accuracy should be and can be provided by a numerical computation, and with reasonable resources. In the following, we suggest prototype strategies to deal with both situations, where the tolerance may be, or may not be, prescribed. Many equally valid variants of these could be proposed, as well. As pointed out before, we focus on the time adaptation. For clarity of exposition, we therefore use uniformly refined spatial grids.
In the present paper, we only treat Burgers' equation. In a paper in preparation, we extend this to the
Euler equations of gas dynamics. We begin by computing the forward and the dual solution as well as the
error estimator on a relatively coarse spatial grid ($L=0$). Usually this spatial grid is much coarser than
the grid we actually want to compute on. Since we want to compute a solution with accuracy comparable to an explicit solution, we prescribe a uniform CFL number below unity in this first computation (e.g. CFL=0.8).
After evaluating the error representation, we have to take two decisions:
\begin{enumerate}
\item
the refinement level $L$ of the next spatial grid. In some cases we will gradually increase the level by
one. This careful approach may be important if it is not clear whether the dynamics of the solution is already captured on the present grid. In other cases (including the example treated below), the time dynamics is already resolved very well on level $L=0$, and we can immediately proceed to the finest grid level.
\item
the tolerance $T\!ol_k(L)$ for the temporal component of the error,
$\bar\eta_k$. The choice of $T\!ol_k(L)$ will be based on assumptions of the
asymptotic decay of the error. If, as in Figure \ref{fig.burgers_1_etabar},
the error decays to first order, then we may choose $T\!ol_k(L+1)=0.5\,T\!ol_k(L)$.
\end{enumerate}
Now we adapt the timestep locally in order to equidistribute the error
densities $\bar\eta_k^n$. Recall from \eqref{eq.bar_eta_kn} that
$$
\bar\eta_k^n = \frac{1}{k_n}\sum_{{D_j}\in \mathcal T}|\bar\eta_{k}^{jn}|,
$$
and
$$
\sum\limits_n k_n \bar\eta_k^n = \bar\eta_k.
$$
If the $\bar\eta_k^n$ were already equidistributed with respect to $n$, then
they would satisfy
$$
\bar\eta_k^n = \bar\eta_k/T \quad\mathrm{for \; all} \; n.
$$
Now, instead of aiming at local error densities of $\bar\eta_k/T$, we
target at an equidistribution of
$$
\bar\eta_k^n \approx T\!ol_k(L+1)/T \quad\mathrm{for \; all} \;
n,
$$
where $T\!ol_k(L+1)$ is a given tolerance on grid $(L+1)$.
Assuming once more that the time component of the error varies
linearly with the time step, we compute the new timestep $k_m$ (on level
$(L+1)$) as
\begin{align}
\label{eq.k_m}
k_m :=k_n \; \frac{
T\!ol_k(L+1)/T}{\bar\eta_k^n} \; .
\end{align}
Using this new timestep distribution we perform a new computation
on the finer spatial grid. Note that due to the linear decay of the error with the timestep, often the new
distribution has a similar number of timesteps as the previous one.
If a total tolerance for the error, $|J(u)-J(u_h)|<T\!ol_{tot}$ is prescribed,
then the above loop is stopped once
$$
\bar\eta_k+\bar\eta_h< T\!ol_{tot}.
$$
Our experience so far is the
following: already on very coarse grids, the method detects the areas of
stationary and instationary flow quite well, and chooses the time steps
accordingly.
We would like to call the approach which combines \eqref{eq.k_m} with an
implicit solver the {\em adaptive, fully implicit strategy}. A possible drawback
of this strategy is that it may lead to extremely small timesteps (CFL $\ll 1$)
when strong instationary waves pass the computational domain. Therefore, in the
second and third example, we restrict the time step size from below. When the
equidistribution of the error suggests $C\!F\!L<5$, we switch to an explicit solver
with $C\!F\!L =0.8$. This saves a considerable number of timesteps. We call this
approach the {\em adaptive, implicit/explicit strategy}.
\subsection{Test problem and asymptotic decay rates}
\label{sec.applications.test_problem}
Now we set up an instationary test case, which is almost stationary, such that
an implicit (or implicit/explicit) scheme might be superior to a fully
explicit one. Our choice is a perturbed stationary shock for Burgers' equation
\begin{align*}
u_t +(\frac{1}{2} u^2)_x =0
\quad \mathrm{for} \; x\in[0,1] \; \mathrm{and} \; t\in[0,48].
\end{align*}
The initial data and corresponding unperturbed solution are given by
\begin{align*}
u(x,t) =
\left\{ \begin{array}{rrl}
1 & \mathrm{for} & x<0.5 \\
-1 & \mathrm{for} & x>0.5
\end{array}\right. .
\end{align*}
Then we place a disturbance at the left boundary of the
domain, which makes the stationary shock move. The new
shock position as a function of time is given by
\begin{align*}
s(t) =
\left\{ \begin{array}{rrl}
0.5 & \mathrm{for} & t<12 \\
0.5+\theta_1(t) sin(\frac{2\pi}{3}(t-12)) & \mathrm{for} & 12<t<18 \\
0.5 & \mathrm{for} & 18<t<30 \\
0.5+\theta_2(t) sin(\frac{2\pi}{3}(t-30)) & \mathrm{for} & 30<t<36 \\
0.5 & \mathrm{for} & 36<t
\end{array}\right..
\end{align*}
where
\begin{align*}
\theta_1(t) = 7.5 \cdot 10^{-3} (t-12)^4(t-18)^4 / {6561} \\
\theta_2(t) = 0.5 \cdot 10^{-3} (t-30)^4(t-36)^4 / {6561}
\end{align*}
Using characteristic theory, we can derive the perturbed left boundary
condition, which is displayed in Figure~\ref{fig.burgers_bsp3}. Note that
the magnitude of the first perturbation is about 1.5 percent of the
shock strength and that of the second perturbation about 0.1 percent.
\begin{figure}[ht]
\begin{center}
\epsfig{file=pic3/fbound.eps, width=0.49\linewidth,height=0.15\textheight}
\end{center}
\caption{Burgers' equation: Left boundary data for perturbed stationary shock.}
\label{fig.burgers_bsp3}
\end{figure}
The functional $J(u)$ is a weighted mean value in space
and time of the solution,
\begin{align*}
J(u) := \int_0^T\int_{0.25}^{0.65}u(x,t)\exp\left(-\frac{1}{1-y(x)^2}\right)dxdt,
\end{align*}
where $y(x):= (x-0.45)/0.2$. Note that the integration area completely
covers the domain containing the shock.
\subsubsection{Asymptotic decay rates
\label{sec.SEP}
Since the adaptive strategy outlined in Section \ref{sec.applications.strategy}
above depends on assumptions on the assymptotic behavior of the error, we first
try to estimate these decay rates. There is no analytical result which shows how the error terms
$\bar\eta_k$ and $\bar\eta_h$ depend on $k$ and $h$. Therefore, we estimate this
dependence numerically. We compute the perturbed shock described in
Section~\ref{sec.applications.test_problem} with a first order
finite volume method with Engquist-Osher flux, which is equal to a DG(0,0) method.
We compare the two approaches:
\begin{itemize}
\item refinement only time
\item and refinement only space.
\end{itemize}
\begin{figure}[hbtp]
\begin{center} {
\subfigure[uniform refinement in time]
{\epsfig{file= pic3/ref_time.eps, width=0.49\linewidth}}
\subfigure[uniform refinement in space]{\epsfig{file= pic3/ref_space.eps, width=0.49\linewidth}}
}
\end{center}
\caption{Error representation for Burgers equation,
first order method, $\bar{\eta_k}$ and $\bar{\eta_h}$ versus level
of refinement. (a) uniform refinement in time. (b) uniform refinement in space.}
\label{fig.burgers_1_etabar}
\end{figure}
\iffalse
\begin{figure}[hbtp]
\begin{center} {
\subfigure[uniform refinement in time, level of refinemet]{\epsfig{file= pic3/ref_time2.eps, width=0.49\linewidth}}
\subfigure[uniform refinement in space, level of refinemet]{\epsfig{file= pic3/ref_space2.eps, width=0.49\linewidth}}
}
\end{center}
\caption{Error representation for Burgers equation,
first order method, ${\eta_k}$, ${\eta_h}$.} \label{fig.burgers_1_eta}
\end{figure}
\fi
\begin{table}
\begin{tabular}{c|c|c|c|c|c|c|c|c|c}
$L$ &$dx$ & $dt$ & $\bar{\eta_k}$ &$\bar{\eta_h}$&$\eta_k$ &$\eta_h$ &$
J(u_h)$ &$\eta_h+\eta_k$& $\theta$ \\ \hline
1& 0.050000& 0.038795& 1.96e-03& 2.02e-01& 1.29e-04& 2.00e-01& 1.72e+00& 2.01e-01& 5.57e+00 \\ \hline
2& 0.025000& 0.019398& 9.81e-04& 4.83e-02& 1.83e-05& 4.75e-02& 1.74e+00& 4.75e-02& 5.49e+00 \\ \hline
3& 0.012500& 0.009699& 4.81e-04& 1.21e-02& 3.57e-06& 1.17e-02& 1.75e+00& 1.17e-02& 5.71e+00 \\ \hline
4& 0.062550& 0.004849& 2.37e-04& 3.10e-03& 1.30e-06& 2.89e-03& 1.75e+00& 2.89e-03& 7.06e+00
\end{tabular}\\
\caption{Efficiency $\theta = \frac{\eta_h+\eta_k}{J(u)-J(u_h)}$of the error representation}
\end{table}
Each of the plots in Figure~\ref{fig.burgers_1_etabar} show the error
estimators $\bar\eta_k$ (error in time) and $\bar\eta_h$ (error in space). In the
Figure~\ref{fig.burgers_1_etabar}(a) we refined only in time. Here the spatial
error remains constant, while the time error still decreases with first order.
The second Figure \ref{fig.burgers_1_etabar}(b) shows the refinement only in space.
The time error $\bar\eta_k$ is almost constant, while the spatial error is decreasing with second order.
Numerically the terms $\bar\eta_t$ and $\bar\eta_h$ behave as expected.
They depend either on $k$ or on $h$, but never on both. The behaviour of
$\eta_h$ and $\eta_k$ is very similar, and not displayed here.
\begin{remark}
The numerically validated results can be used for adaptive
grid refinement. The error estimator $\bar\eta_h$ can be used as an indicator for spatial
adaption and the estimator $\bar\eta_k$ for time step control.
\end{remark}
\subsection{Computational results}
\label{sec.applications.results}
{\bf Example 1:} The first computation ($L=0$) is done on a grid with 20
spatial cells and a uniform CFL number of 0.8 using explicit timesteps. It needs
$N=1238$ timesteps, reaching a total error of $\bar\eta =0.204$ and a relative
error of $|\bar\eta/J(u_h)| = 11.9 \%$, but a temporal error of
$|\bar\eta_k/J(u_h)| = 0.13 \%$. Our adaptive strategy now aims at a time step
distribution on the next grid with tolerance $T\!ol_k(L+1) = \bar\eta_k(L)$.
Based on the assumption that the time component of the error varies linearly
with the time step (which is motivated by Fig.~\ref{fig.burgers_1_etabar}), the
scheme chooses new timesteps on the next grid according to the equidistribution
rule \eqref{eq.k_m}.
The second row of Table~\ref{table.perturbed_shock.1}, for level $L=1$, gives
also $N=1238$ time steps, now using adaptive implicit timesteps. Now the
relative temporal error is $|\bar\eta_k/J(u_h)| = 0.073 \%$, and it is dominated
by the spatial error $|\bar\eta_h/J(u_h)| = 2.2 \%$.
\begin{table}
\begin{tabular}{c|c|c |c|c}
L & N &$\bar\eta_k/J(u_h)$ &$\bar\eta_h/J(u_h)$ &$\bar\eta/J(u_h)$ \\ \hline
0 &1238 &1.34e-03 &1.17e-01 &1.19e-01 \\
1 &1238 &7.37e-04 &2.20e-02 &2.27e-02 \\
\end{tabular}\\
\caption{Example 1: Perturbed shock for Burgers' equation.
From left to right: level $L$, number of time steps $N$, time component
of error estimator $\bar\eta_k/J(u_h)$, spatial component
of error estimator $\bar\eta_h/J(u_h)$, total error estimator
$\bar\eta/J(u_h)$.}
\label{table.perturbed_shock.1}
\end{table}
Important additional information can be gained by looking at the plots in
Figure~\ref{fig.perturbed_shock.tol_0.01}, showing the CFL distribution on each
time interval $I_n$ and the normalized time components of the error estimator
$\bar\eta_k^n$, both in logarithmic scale. The stationary and instationary regions are
separated by the estimator. In particular, note that
\begin{itemize}
\item
the time component of the error varies over more than 14 orders
of magnitude.
\item
in the three stationary regions, $\bar\eta_k^n$ is very close to zero.
\item
the two instationary waves are distinguished very clearly. The second wave is about one order of magnitude smaller than the first wave. This corresponds closely to the different magnitudes of the inflow perturbations.
\item
furthermore, one can clearly identify an initial layer, where
$\bar\eta_k^n=\mathcal O(1)$ at the inflow boundary $t=0$, and $\bar\eta_k^n$
decays exponentially for time $t>0$ until it reaches machine accuracy.
\end{itemize}
We advance to level $L=1$, Figures~\ref{fig.perturbed_shock.tol_0.01}(c) and
(d). We observe that
\begin{figure}[hbtp]
\begin{center} {
\subfigure[$C\!F\!L(t_n)$, L=0, uniform explicit timestep]
{\epsfig{file= pic3/1.eps, width=0.49\linewidth,height=0.16\textheight}}
\subfigure[$\bar\eta_k^n(t_n)$, L=0, uniform explicit timestep]
{\epsfig{file= pic3/2.eps, width=0.49\linewidth,height=0.16\textheight}}
}
\end{center}
\vspace{0.5cm}
\begin{center}{
\subfigure[$C\!F\!L(t_n)$, L=1, adaptive implicit timestep]
{\epsfig{file= pic3/3.eps, width=0.49\linewidth,height=0.16\textheight}}
\subfigure[$\bar\eta_k^n(t_n)$, L=1, adaptive implicit timestep]
{\epsfig{file= pic3/4.eps, width=0.49\linewidth,height=0.16\textheight}}
}
\end{center}
\caption{Example 1: Perturbed shock for Burgers' equation with equidistributed time error.
Left column: CFL(t); right column: $\bar\eta_k^n(t_n)$. Upper row: level $L=0$,
fully explicit scheme, uniform timestep. Lower row: level $L=1$,
fully implicit scheme, adaptive timestep.
(from top to bottom).} \label{fig.perturbed_shock.tol_0.01}
\end{figure}
\begin{itemize}
\item
the error on level $L=1$ varies by less than 2 orders of magnitude, 12
orders of magnitude less than on level $L=0$. The magnitude of the
maximal error has decreased by almost two orders of magnitude.
Therefore the solution is much better resolved in the instationary regions,
and the computational recources are clearly distributed more efficiently.
\item
in the initial layer, the CFL number starts with $\mathcal{O}(10^{-2})$.
Then it grows at least exponentially until it reaches a maximal value of about
500. At the same time, the error $\bar\eta_k^n$ decays roughly by two orders of
magnitude. Thus, these initial steps can be seen as a preprocessing of the initial
data, to translate a prescribed steady shock on the pde level into a
steady discrete shock layer. The initial layer is also clearly visible
in the plot of the error distribution (and this will never disappear).
Indeed, the initial data, a sharp jump from 1 to -1, are a steady
shock only on the level of the exact solution. Numerically, the scheme has
to converge towards a discrete shock layer, and this will always need
a few time steps. In fact this is an instance of a scheme converging
towards a numerical steady state solution, using adaptive time steps.
\item
in the stationary region between the initial layer and the first
perturbation, the error is more than three orders of magnitudes
smaller than in the following flow field. Observe that the whole stationary region is
computed by a single time step. Thus the scheme is only held back from
choosing a larger time step by the appearance of the instationary
perturbation. If we had introduced this perturbation at a later time, the
time step and thus the local CFL number would haves been
correspondingly larger.
\item
The next region of stationary flow is again bridged by a single time step,
and correspondingly the local error is somewhat below the equidistributed
one.
\item
For the two perturbations, the normalized error is already close to being uniformly distributed.
\item
Using large timesteps does not mean that each timestep has higher computational costs. Since the adaptation
chooses large timesteps, where the solution is(nearly) stationary, these timesteps have low computational
costs.
\end{itemize}
We would also like to point out one drawback of the equidistribution strategy
for the timestep. In the first (and larger) instationary wave, the proposed CFL
number is often much smaller than unity, e.g. $\min\limits_n(C\!F\!L(t^n)) =
0.009$ in Figure~\ref{fig.perturbed_shock.tol_0.01}. It is well-known that
lowering the CFL number much below unity smears the solution. Therefore, while
such small timesteps may improve the temporal accuracy somewhat, they will deteriorate
the spatial accuracy considerably. Moreover, they increase the number of
timesteps, and hence the computational cost. In the following example, we
discuss a more efficient strategy.
\vspace*{2ex}\noindent
{\bf Example 2:}
This example is a modification of the first example which used a fully implicit
strategy for the timestep. Here we introduce a mixed {\em implicit/explicit strategy}.
We still want the equidistribute the error, but we will give up this goal
partially when the local $C\!F\!L$ number drops below a certain threshhold.
As discussed above, choosing timestep sizes with $C\!F\!L$ much less than unity
seems to be inefficient both for explicit and for implicit schemes. For
implicit methods, even timesteps with $C\!F\!L <5$ are not efficient, since we have
to solve a nonlinear system of equations at each timestep. Thus, the new
implicit/explicit strategy switches to the cheaper (and less dissipative)
explicit method, if $C\!F\!L < 5$, computing perhaps a few more timesteps if
$0.8<C\!F\!L<5$, and saving timesteps if $C\!F\!L<0.8$. (Of course, we could choose
other thresholds than $C\!F\!L=0.8$ and 5.)
As we can see in Table~\ref{table.perturbed_shock_cfl}, the new strategy
requires only 449 timesteps, instead of 1238 with the direct equidistribution
in Example 1. Out of these, only 166 are implicit and hence expensive. This
leads to considerable speed-up.
\begin{table}
\begin{tabular}{c|c|c|c|c}
L & N(expl) &$\bar\eta_k/J(u_h)$ &$\bar\eta_h/J(u_h)$ &$\bar\eta/J(u_h)$ \\ \hline
1 & 449\,(283) &8.50e-04 &2.14e-02 &2.23e-02
\end{tabular}\\ [1ex]
\caption{Example 2: Same as Table~\ref{table.perturbed_shock.1}, but CFL restriction from below
(implicit/explicit strategy).}
\label{table.perturbed_shock_cfl}
\end{table}
\begin{figure}[hbtp]
\begin{center}{
\subfigure[$C\!F\!L(t_n)$, L=1, adapt.~impl./expl.~timestep]
{\epsfig{file= pic3/5.eps, width=0.49\linewidth,height=0.16\textheight}}
\subfigure[Error $\bar\eta_k^n$, L =1, adapt. impl./expl. timestep]
{\epsfig{file= pic3/6.eps, width=0.49\linewidth,height=0.16\textheight}}
}
\end{center}
\caption{Example 2: Same as Figure~\ref{fig.perturbed_shock.tol_0.01} but
adaptive implicit/explicit strategy (CFL restriction from below).}
\label{fig.perturbed_shock_cfl}
\end{figure}
\vspace*{2ex}\noindent
{\bf Example 3:}
Table~\ref{table.perturbed_shock_cfl_2} and Figure
~\ref{fig.perturbed_shock_cfl_3} show three extensions of Example 2. We used the
same implicit/explicit strategy as in Example 2, but after the explicit
reference computation on the coarse grid (L=0, error $\bar\eta_k^{ref}$), we
proceed directly to a finer grid with 320 cells (L=4). We compare an explicit
and two implicit/explicit computations on the fine grid.
The first row shows results of the fully explicit scheme with uniform refinement
in time and space for L=4. As expected, the errors are about $2^4$ times smaller
than those on the original coarse grid. Now suppose we wanted to reach
comparable errors on level $L=4$ using adaptive timestepping. Then we should set
the tolerance to be $\;T\!ol(4)=2^{-4}\bar\eta_k^{ref}\,$. The results of this
computation are shown in the second row of
Table~\ref{table.perturbed_shock_cfl_2}. The three components of the error are
comparable with those of the fully explicit computation, but the number of
timesteps is only 3975 instead of 19200. Out of these 3975 steps, only 1235 are
implicit.
Another strategy for equidistributing the error might be to fix any constant
tolerance, for example $T\!ol(4)=\bar\eta_k^{ref}$ itself. The results of this
computation are displayed in the last row of the table. The error in time is
now a factor 5-8 higher than for the other two computations, while the spatial
error is comparable. Remarkably, this computation needs only 1780 timesteps, and
only 507 of these are implicit.
Both of these calculations show that considerable savings are possible with the
implicit/exp\-licit, time-adaptive strategy.
\begin{table}
\begin{tabular}{c|c|c|c|c|c}
strategy & $T\!ol_k$ & N(expl) &$\bar\eta_k/J(u_h)$ &$\bar\eta_h/J(u_h)$ &$\bar\eta/J(u_h)$ \\ \hline
fully expl. & -- & 19200\,(19200) &7.11e-05 &4.57e-04 &5.28e-04 \\
impl./expl. & $2^{-4}\bar\eta_k^{ref}$ & \phantom{1}3975\,\phantom{1}(2740) &1.06e-04 &3.71e-04 &4.77e-04 \\
impl./expl. & $\bar\eta_k^{ref}$ & \phantom{1}1780\,\phantom{1}(1243) &5.46e-04 &3.66e-04 &9.12e-04
\end{tabular}\\[1ex]
\caption{Example 3: Same as Table~\ref{table.perturbed_shock_cfl}, but on level
L =4 and with different tolerances.}
\label{table.perturbed_shock_cfl_2}
\end{table}
\begin{figure}[hbtp]
\begin{center} {
\subfigure[$C\!F\!L(t_n)$, uniform explicit]
{\epsfig{file= pic3/CFL3.eps, width=0.49\linewidth,height=0.16\textheight}}
\subfigure[$\bar\eta_k^n(t_n)$, uniform explicit]
{\epsfig{file= pic3/Fehler3.eps, width=0.49\linewidth,height=0.16\textheight}}
}
\end{center}
\vspace{0.5cm}
\begin{center}{
\subfigure[$C\!F\!L(t_n)$, adapt.~impl./expl., $T\!ol_k =2^{-4}\bar\eta_k^{ref}$]
{\epsfig{file= pic3/9.eps, width=0.49\linewidth,height=0.16\textheight}}
\subfigure[$\bar\eta_k^n(t_n)$, adapt.~impl./expl., $T\!ol_k =2^{-4}\bar\eta_k^{ref}$]
{\epsfig{file= pic3/10.eps, width=0.49\linewidth,height=0.16\textheight}}
}
\end{center}
\vspace{0.5cm}
\begin{center}{
\subfigure[$C\!F\!L(t_n)$, adapt.~impl./expl., $T\!ol_k =\bar\eta_k^{ref}$]
{\epsfig{file= pic3/7.eps, width=0.49\linewidth,height=0.16\textheight}}
\subfigure[$\bar\eta_k^n(t_n)$, adapt.~impl./expl., $T\!ol_k =\bar\eta_k^{ref}$]
{\epsfig{file= pic3/8.eps, width=0.49\linewidth,height=0.16\textheight}}
}
\end{center}
\caption{Example 3: Same as Figure~\ref{fig.perturbed_shock_cfl} but on Level L =4 and with different tolerances. \label{fig.perturbed_shock_cfl_3}
}\end{figure}
\section{Conclusions}
\label{sec.conclusions}
In this paper, we combine space- and time-projections of S\"uli, Houston and Hartmann
to split the classical adjoint based error representation formula for
target functionals into space and time components. Based on a numerical study
of these components we design an adaptive strategy which attempts to minimize the number of
time steps by equidistributing the time components of the error. We apply the adaptive
scheme to a weak perturbation of a stationary shock.
Already on a very coarse mesh of 20 points the error representation formula
precisely gives the location and strength of the instationary perturbations.
This can be translated into efficient timestep distributions, which respect a
desired accuracy. We show that these timestep distributions can be applied
successfully to much finer spatial grids.
We never compute implicit timesteps below CFL=5. Instead, when the error
analysis suggests a timestep below CFL=5, we switch to an explicit scheme with
CFL=0.8. This implicit/explicit strategy gives considerable savings.
For nonlinear perturbations of a stationary shock, we have demonstrated that our
strategy does reach its goals: it separates initial layers, stationary regions
and perturbations cleanly and chooses just the right timestep for each of them.
Besides building upon well-established adjoint techniques, we have also added
a new ingredient which simplifies the computation of the dual problem. We show
that it is sufficient to compute the spatial gradient of the dual solution,
$w=\nabla \varphi$, instead of the dual solution $\varphi$ itself. This gradient
satisfies a conservation law instead of a transport equation, and it can
therefore be computed with the same algorithm as the forward problem, and in the
same finite element space. For discontinuous transport coefficients, the new
conservative algorithm for $w$ is more robust than our previous transport
schemes for $\varphi$.
In ongoing work, we are adapting this strategy to aerodynamic
problems. First test calculations show a promising speed-up.
\medskip
{\bf Acknowledgement:} The authors would like to thank Ralf Hartmann, Paul
Houston, Mario Ohlberger and Endre S\"uli for stimulating discussions.
The work of both authors was supported by DFG grant
SFB 401 at RWTH Aachen. Part of the work was completed while the first author
was in residence at the Center of Mathemtics for Applications (CMA) at Oslo University. Both authors would like to thank
CMA its members for their generous hospitality.
|
\section{\texorpdfstring{B\"uhlmann, Meier and van de Geer.}{B\"uhlmann, Meier and van de Geer}}\label{secbuhl}
We thank Professors B\"uhlmann, Meier and van de Geer for their
extensive and detailed discussion---they raise many interesting
points. Before addressing these, there are a few issues worth
clarifying.
\begin{itemize}
\item These authors rewrite the covariance test in what they claim
is an alternate form. Sticking to the notation in our original
paper, the quantity that they consider is
\begin{eqnarray*}
T(A,\lambda_{k+1}) &=& \bigl(\bigl\|y-X_A \tilde{
\beta}_A(\lambda_{k+1})\bigr\|_2^2 +
\lambda_{k+1}\bigl\|\tilde{\beta}_A(\lambda_{k+1})
\bigr\|_1 \bigr)/\sigma^2
\\
&&{} -\bigl(\bigl\|y-X \hbeta(\lambda_{k+1})\bigr\|_2^2 +
\lambda_{k+1}\bigl\|\hbeta(\lambda_{k+1})\bigr\|_1 \bigr)/
\sigma^2.
\end{eqnarray*}
(In\vspace*{1pt} B\"uhlmann et al. the quantities $A$ and $\lambda_{k+1}$ above are
written as $\hat{A}_{k-1}$ and $\hat{\lambda}_{k+1}$.) This is not actually
equivalent to the covariance statistic. Expanding the
above expression yields
\begin{eqnarray*}
&& 2 \bigl( \bigl\langle y, X\hbeta(\lambda_{k+1}) \bigr\rangle- \bigl
\langle y,X_A \tbeta_A(\lambda_{k+1}) \bigr
\rangle\bigr) / \sigma^2
\\
&&\quad {}+ \bigl\|X_A\tilde{\beta}_A(\lambda_{k+1})
\bigr\|_2^2 - \bigl\|X\hbeta(\lambda_{k+1})
\bigr\|_2^2 + \lambda_{k+1} \bigl(\bigl\|\tilde{
\beta}_A(\lambda_{k+1})\bigr\|_1 - \bigl\|\hbeta(
\lambda_{k+1})\bigr\|_1 \bigr),
\end{eqnarray*}
which is (two times) the covariance test statistic plus several
additional terms (the difference in squared $\ell_2$ norms of fitted
values, plus $\lambda_{k+1}$ times the difference in $\ell_1$ norms of
coefficients). These additional terms can be small,
especially if $\lambda_k$ and $\lambda_{k+1}$ are close together,
since in this case the solutions $\hbeta(\lambda_{k+1})$ and
$\tilde{\beta}_A(\lambda_{k+1})$ can themselves be close---however,
they are certainly not zero. We wonder whether this discrepancy has
affected the simulation results of B\"uhlmann et al., in Section~5 of
their discussion.
\item The authors note that the asymptotic null distributions that we
derive for the covariance test statistic require that $\mathbb{P}(B)
\rightarrow1$ as $n,p \rightarrow\infty$, for a particular event
$B$.
This event is defined slightly differently in Section~3.2, which
handles the orthogonal $X$ case, than it is in Section~4.2, which
handles the general $X$ case. Regardless, the event $B$ can be
roughly interpreted as follows: ``the lasso active model at the
given step $k$ converges to a fixed model containing the truth
(i.e., its active set contains the truly active variables, and
its active signs match those of the truly active coefficients).''
B\"uhlmann et al. comment that, to ensure that
$\mathbb{P}(B)\rightarrow1$, we assume a ``beta-min'' condition and an
``irrepresentable-type'' condition. However, this is not quite
correct. The main result of our paper, Theorem 3 in Section~4.2,
\textit{assumes that} $\mathbb{P}(B) \rightarrow1$, and uses an
irrepresentable-type condition to ensure that the conditions of the
critical Lemma 8 are met---namely, that each quantity
$M^+(j_k,s_k)$
diverges to $\infty$ quickly enough. There is no beta-min
condition employed here. If we were to have additionally assumed
a beta-min
type condition, then from this we could have shown that
$\mathbb{P}(B)\rightarrow1$. Instead, we left $\mathbb
{P}(B)\rightarrow1$ as a
direct assumption, for good reason: as described in the remarks
following Theorem 3, we believe there are weaker sufficient
conditions for $\mathbb{P}(B) \rightarrow1$
that do not require the true coefficients to be well separated from
zero---remember, for the event $B$ to hold we only need the
computed active set to contain the set of the true variables, not
equal it.
This distinction---between exact variable recovery and correct
variable screening---is an important one.
Figure~1 in the discussion by B\"uhlmann et al. shows empirical
probabilities of exact variable recovery by the lasso. It
demonstrates that, as the size $k_0$ of the true active set
increases, the
minimum absolute value of true nonzero coefficients must be quite
high in order for the lasso to recover the exact model with high
probability. But the story is quite different when we look at
variable screening; see our Figure~\ref{figscreening} below, which
replicates the simulation setup of B\"uhlmann et al., but now
records the empirical probabilities that the computed lasso model
contains the true model, after some number of steps $k \geq k_0$.
We can see that the story here is much more hopeful. For example, while
the underlying model with $k_0=10$ truly nonzero coefficients
cannot be consistently recovered after $k=10$ lasso steps, even
when beta-min is large (middle panel), this model is indeed
consistently contained in the computed lasso model after $k=20$
steps, even for very modest values of beta-min. What this means
for the covariance test, in such a setup: the asymptotic $\Exp(1)$
null distribution of the covariance statistic kicks in at some step
$k \geq k_0$, and we start to see large $p$-values. Then, by
failing to reject the null hypothesis, we correctly
screen out a sizeable proportion of truly inactive
variables.\footnote{To be fair, we are certain that B\"uhlmann et
al. are familiar with the screening properties of the lasso,
given some of these authors' own pioneering work on the subject.
Our intention here is to clarify the assumptions made in the
covariance test theory, and in particular, clarify what it means
to consider $\mathbb{P}(B) \rightarrow1$. B\"uhlmann et al. do discuss
variable screening, and remark that achieving such a property
in practice seems unrealistic, referring to their Figure~1
as supporting evidence. However, as explained above,
their Figure~1 examines the probability of exact model recovery,
and not screening.}
\begin{figure
\includegraphics{1175rejf01.eps}
\caption{Replication of the simulation setup considered in
the discussion by B\"uhlmann et al., but now with attention being
paid to correct variable screening, rather than exact variable
recovery. Here, $k_0$ denotes the true number of nonzero
coefficients, and $k$ the number of chosen lasso predictors
(steps along the lasso path). We see that, with high probability,
the true model is contained in the first 5, 15 or 20 chosen
predictors.}\label{figscreening}
\end{figure}
In any case, it is important to point out that the newer sequential
testing procedure in \citet{spacing} and the fixed-$\lambda$
testing procedure in \citet{howlong} do not assume a beta-min or
irrepresentable condition whatsoever, and do not require any
conditions like $\mathbb{P}(B) \rightarrow1$.
\end{itemize}
Now we respond to some of the other points raised. One of the remarks
that we made after the main result in Theorem 3 of our original paper
claims that this result can be extended to cover just the ``strong''
true variables (ones with large coefficients), and not necessarily
the ``weak'' ones (with small coefficients). B\"uhlmann et al. comment
that such an extension would likely require a ``zonal''
assumption, that bounds the number of small true coefficients, as
in \citet{hdscreen}.
As a matter of fact, we know a number of examples, with many small
nonzero coefficients, for which the conclusions of Theorem 3 continue
to hold. In any case, we emphasize that the newer sequential test of
\citet{spacing} does not need to make any assumption of this
sort, and
neither does the fixed-$\lambda$ test of \citet{howlong}.
Properly interpreting the covariance test $p$-values, as B\"uhlmann et
al. point out, can be tricky. But we believe this comes with the
territory of a conditional test for adaptive regression, since the
null hypothesis is random (and, as B\"uhlmann et al. note,
is an unobserved event). Consider the wine dataset from Section~7.1
of our original paper as an example. Looking at the $p$-values in right
panel of Table~5, one might be tempted to conclude from the $p$-value
of 0.173 in the fourth line that the constructed lasso model with
{\tt alcohol}, {\tt volatile acidity} and {\tt sulphates} contains
all of the truly active variables. There are potentially several
problems with such an interpretation (one of which being that we have
no reason to believe that the true model here is actually linear), but
we will focus on the most flagrant offense: because the constructed
lasso model is random, the $p$-value of 0.173 reflects the test of a
random null hypothesis, and so we cannot generally use it to draw
conclusions about the specific variables {\tt alcohol},
{\tt volatile acidity},
{\tt sulphates} that happened to have been selected in the current
realization. The $p$-value of 0.173 does, however, speak to the
significance of the 3-step lasso model, that is, the lasso model after 3
steps along the lasso path---said differently, we can think of this
$p$-value as reflecting the significance of the 3 ``most important''
variables as deemed by the lasso. This properly accounts for the
random nature of the hypothesis (as in any realization, the
identity of these first 3 active variables may change), and is an
example of valid post-selection inference.
Of course, one may ask: is this really what
should be tested? That is, instead of inquiring about the significance of
the 3-step lasso procedure, would a practitioner not actually want to
know about the significance of the variables {\tt alcohol},
{\tt volatile acidity} and {\tt sulphates} in particular? In a
sense, this is really a question of philosophy, and the answer
is not clear in our minds. Here, though, is a possibly helpful
observation: when we consider a single wine data set, testing the
significance of the (fixed) variables {\tt alcohol}, {\tt volatile
acidity} and {\tt sulphates} (after these 3 have been selected as
active by the lasso) seems more natural; but when we consider a
sequence of testing problems, in which we observe a new wine data set
and rerun the lasso for 3 steps on each of a sequence of days
$1,2,3,\ldots,$ testing the significance of the (random) 3-step lasso
procedure seems more appropriate.
[As an important side note, in the finite sample spacing test given in
\citet{spacing}, one can argue that both interpretations are valid,
since our inference in this work is based on conditioning on the
value of the selected variables.]
In Tables~1 and 2 of their discussion, B\"uhlmann et al. compare their
approach to
the covariance test in terms of false positive and false negative
rates. In our original paper, we had not specified a sequential
stopping rule for the covariance test, and it is not clear to us that
the two they used were reasonable. (Additionally, we are not sure
what form they assumed for the covariance test, as the representation
they present, based on the difference in lasso criterion values, is
not equivalent to the covariance statistic; see the first
clarification bullet point above.) B\"uhlmann et al. kindly sent us
their R code for their procedure, and we applied it to a subset of
their examples,
corresponding to the setup in the second row of each of their Tables~1
and 2. The results of 1000 simulations are shown in Table~\ref{tabbuhl}.
There are two setups: $n=100$, $p=80$, and $n=100$, $p=200$. In
line 1 of each, we applied their de-sparsification technique, using
the same estimate of $\sigma$ as in their
discussion. We found that this commonly overestimates
$\sigma$ by $>$100\%, so in line 2 we use the true value, $\sigma=1$.
Line 3 uses the covariance test with the ForwardStop rule of
\citet{fdrlasso}, and the true $\sigma=1$, designed to control
the FDR
at~5\%. We see that the de-spars rule does well with the inflated
estimate of $\sigma$, but produces far too many false positives when
the true value of $\sigma$ is used. Reliance on a inflated variance
estimate does not seem like a robust strategy, but perhaps there is a
way to resolve this issue. (In all fairness, B\"uhlmann et al. told us
that they are aware of this.)
The covariance test with ForwardStop does a reasonable job of
controlling the FDR, while capturing just under half of the true
signals.
\begin{table}
\tabcolsep=0pt
\caption{Results of a simulation study, repeating
the setup in the second row of each of Tables~1 and 2 from the
B\"uhlmann et al. discussion. Shown are the average number of
predictors called significant (out of $p=80$ or $200$), the average
number of false and true positives, the familywise error rate and
the false discovery rate}\label{tabbuhl}
\begin{tabular*}{\tablewidth}{@{\extracolsep{\fill}}@{}lccccc@{}}
\hline
& \textbf{Ave number called signif.} & \textbf{Ave FP} & \textbf{Ave TP} & \textbf{FWER} & \textbf{FDR} \\
\hline
\multicolumn{6}{@{}c@{}}{$n=100$, $p=80$} \\
(1) de-spars (estimated $\sigma$) & \phantom{0}6.89& \phantom{0}0.05& 6.84& 0.04& 0.01\\
(2) de-spars (true $\sigma$) & 17.29& \phantom{0}7.36& 9.93& 0.98& 0.43\\
(3) covTest/forwStop & \phantom{0}4.81& \phantom{0}0.25& 4.55& 0.28& 0.05
\\[3pt]
\multicolumn{6}{@{}c@{}}{$n=100$, $p=200$} \\
(1) de-spars (estimated $\sigma$) & \phantom{0}3.35 & \phantom{0}0.04& 3.30& 0.04 &0.01\\
(2) de-spars (true $\sigma$) & 44.52& 34.81& 9.71& 1.00& 0.78\\
(3) covTest/forwStop & \phantom{0}4.29 & \phantom{0}0.31& 3.97& 0.26& 0.07\\
\hline
\end{tabular*}
\end{table}
\section{\texorpdfstring{Interlude: Conditional or fixed hypothesis testing?}{Interlude: Conditional or fixed hypothesis testing}}\label{secfinal}
We would like to highlight some of the differences
between conditional and fixed hypothesis testing. This section is
motivated by the comments of B\"uhlmann et al., as well as the referees
and Editors of our original article, and personal conversations with
Larry Wasserman.
Though it has been said before, it is worth repeating:
the covariance test does not give $p$-values for classic tests of fixed
hypotheses,
such as $\beta^*_S=0$ for a fixed subset $S \subseteq\{1,\ldots, p\}$;
however, it was not designed for this purpose. As we see it:
conditional hypothesis tests like the covariance test, and fixed
hypothesis tests like that of \citet{vdgsignif} and many others (see
the references in Section~2.5 of our original paper) are two
principally different approaches for assessing significance in
high-dimensional modeling. The motivation behind the covariance test
and others is that often a practitioner becomes interested
in assessing the significance of a variable only \textit{because} it has
been entered into the active set by a fitting procedure like the lasso. If
this matches the actual workflow of the practitioner, then the
covariance test or other conditional tests seem to be best-suited to
his or her needs. A resulting complexity is that interpretation here
must be drawn out carefully (refer back to Section~\ref{secbuhl}).
On the other hand, the idea behind fixed tests like that of
\citet{zhangconf}, \citet{vdgsignif} and \citet{montahypo2},
(or at least, a typical use case in our
view) is to compute $p$-values for all fixed hypothesis $\beta_j^*=0$,
$j=1,\ldots, p$, and then perform a multiple testing correction at the
end to determine global variable significance. Even though the lasso
may have been used to construct such $p$-values, the practitioner is to
pay no attention to its output---in particular, to its active set.
And of course, the final model output by this testing procedure (which
contains the variables deemed significant) may or may not match the
lasso active set. The appeal of this approach lies in the simplicity
and transparency of its conclusions: each computed $p$-value is
associated with a familiar, classical hypothesis test, $\beta_j^*=0$
for a fixed $j$. In fact,\vspace*{2pt} we too like this approach, as it is very
direct. One drawback is that it is unclear how this might be used
for post-selection inference, if that is what is desired by the
practitioner.
We note the conditional perspective is not really a foreign one, as it is
indeed completely analogous to the (proper) interpretation of
cross-validation errors for the lasso or forward stepwise regression.
In this setting, to estimate the expected test error of a $k$-step
model computed by, say,
the lasso, we rerun the lasso for $k$ steps on a fraction of the
data set, record the observed validation error on held-out data, and
repeat this a number of times. This yields a final estimate of the
expected test error for the $k$-step lasso model; but importantly, in
each iteration of cross-validation, the selected variables will likely have
changed (since the lasso is being run on different data sets),
and so it is really only appropriate to regard cross-validation as
producing as error estimate for the $k$-step lasso procedure, not for
the particular realized model of size $k$ that was fit on the entire
data set.
Lastly, we draw attention to a connection between
our work on post-selection inference, and the de-biasing techniques
pursued by \citet{zhangconf}, \citet{vdgsignif} and
\citet{montahypo2}. In Section~7.1 of \citet{howlong}, we
show how
the framework developed in this paper can be used to form intervals or
tests for the components of a de-biased version of the true
coefficient vector, that is, something like a \textit{population
analog} of
the de-biased estimator studied by these authors. Under the
appropriate sufficient conditions [e.g., the same as those in
\citet{montahypo2}], these population de-biased coefficients converge
to the true ones, so these tests and intervals are also valid for the
underlying coefficients as well.
\section{\texorpdfstring{Buja and Brown.}{Buja and Brown}}
We thank Professors Buja and Brown for their scholarly summary
of inference in adaptive regression. We learned a great deal from it
and we enthusiastically recommend it to readers. They discuss in
detail the forward stepwise approach, and outline many
different ways to carry out inference in this setting.
To explore the $t_{\max}$ proposal
in their discussion, we carried out a simulation study.
It turns out that this is helpful in illustrating the special
properties of the covariance test with null distribution $\Exp(1/k)$,
as well as the spacing $p$-values [\citet{spacing}].
\begin{figure}
\includegraphics{1175rejf02.eps}
\caption{Simulation of $p$-values for the first four
steps of using the test in \protect\eqref{eqtstat} with forward stepwise
regression (top row), the covariance test (middle row), and the
spacing test (bottom row). Details are given in the text.}\label{figbujaplot}
\end{figure}
With $n=50$, $p=10$, we generated standardized Gaussian predictors,
the population correlation between predictors $j$ and $j'$
being $0.5^{|j-j'|}$. The true coefficients were $\beta^*=0$, and
the marginal error variance was $\sigma^2=1$. The middle
and bottom panels of Figure~\ref{figbujaplot} show quantile--quantile
plots of the covariance $p$-values and spacing $p$-values for the first
four steps of the least angle regression path [see \eqref{eqspacings}
for the spacing test in the first step, and \citet{spacing} for
subsequent steps]. In the top panel, we have applied forward
stepwise regression, using the test statistic
\begin{equation}
\label{eqtstat} t_{\max}(y)=\max_{j=k,\ldots, p}
\bigl|t^{(j)}(y)\bigr| \qquad\mbox{where } t^{(j)}(y)=\frac{\langle X_{j\cdot
A},y\rangle}{\| X_{j\cdot A}\|_2},
\end{equation}
per the proposal of Buja and Brown. Here, $A$ is the set of active
variables currently in the model and $X_{j\cdot A}$ denotes the $j$th
predictor orthogonalized with respect to these variables.
Note that we have used the true value $\sigma^2=1$ in \eqref{eqtstat}
(and in the covariance and spacing tests as well). As suggested by
Buja and Brown, we simulated $\varepsilon\sim N(0,I)$ in order to
estimate the $p$-value $\mathbb{P}(t_{\max}(\varepsilon) >
t_{\max}(y))$.
All three tests look good at the first step, but the forward stepwise
test based on \eqref{eqtstat}
becomes more and more conservative for later steps.
The reason is that the covariance test and the spacing test (even
moreso) properly account for the selection events up to and including
step $k$. To give a concrete example, the forward stepwise test
ignores the fact that at the second step, the observed
$t_{\max}$ is the \textit{second} largest value of the statistic
in the data, and erroneously compares it to a null distribution of
\textit{largest} $t_{\max}$ values. This creates a conservative
bias in the $p$-value. If predictor $j$ were chosen at the first step
of the forward stepwise procedure, then a correct numerical
simulation for
$t_{\max}$ at the second step would generate
$y^*=X_j \hbeta_j + \varepsilon$ (with $\hbeta_j$ being
the least squares coefficient on variable $j$), and only keep those
$y^*$ vectors for which predictor $j$ is chosen at the first
step, using these to compute $t_{\max}(y^*)$ [which
equals $t_{\max}(\varepsilon)$]. Such a simulation setup might be
practical for a few steps, but would not be practical beyond
that, though there do exist efficient algorithms for
sampling from such distributions. Remarkably, the covariance and
spacing tests are able carry out
this conditioning analytically.
On a separate point, we agree with Buja and Brown that inferences
should not typically focus on the true regression coefficients when
predictors are highly correlated,
and even the definition of FDR seems debatable in that setting.
In \citet{uvr}, we propose an alternative definition of FDR,
called the
``Uninformative Variable Rate'' (UVR), which tries to finesse this
issue by projecting the true mean $X\beta^*$ onto the set of
predictors in the current model. A selection is deemed a false
positive if it has a zero coefficient in this projection.
For example, in a model with $\beta_1^*=5$, $\beta^*_2=0$ and
$\operatorname{Cor}(X_1,X_2)=0.95$, the selection of $X_2$ by itself would
be considered a false positive in computing the FDR. But this does not
seem reasonable, and the UVR would instead consider it a true
positive.
As Buja and Brown mentioned, we have proposed a method for
combining sequential $p$-values to achieve FDR control in
\citet{fdrlasso}. But we believe there is more to do, especially in
light of the last point just raised.
Finally, as they remark, our tests will not be valid if the
practitioner uses them in combination with other selection techniques,
or as they put it, the data analyst is ``arbitrarily informal in their
meta-selection of variable selection methods.''
As they point out, the POSI methods they propose in \citet{posi}
are valid even in that situation. This is a very nice property, but of
course the pressing question is: are the inferences too conservative
as a result of protecting the type I error in such a broad sense?
\section{\texorpdfstring{Cai and Yuan.}{Cai and Yuan}}
We are grateful to Professors Cai and Yuan for their suggestion of an
alternative test based on the Gumbel distribution. In the most basic
setting, testing at the first step (i.e., global null hypothesis) in
the orthogonal $X$ setting, both our proposal and theirs stem from the
same basic arguments. To see this,
suppose that $V_1 \geq\cdots\geq V_p>0$ are the ordered absolute
values of a sample from a standard normal distribution. Then, as
$p \to\infty$,
\begin{equation}
\label{eqablim} b_p(V_1 - a_p) \stackrel{d} {
\rightarrow}\operatorname{Gumbel}(0,1),
\end{equation}
where
\[
a_p = \Phi^{-1} \bigl(1-1/(2p) \bigr) = \sqrt{2 \log p} -
\frac{\log\log p +\log\pi}{2\sqrt{2\log p}} + o(1/\sqrt{\log p})
\]
and
\[
b_p = \sqrt{2\log p} \bigl(1+o(1) \bigr).
\]
We used this and the fact that
$b_p(V_1-V_2) \stackrel{d}{\rightarrow}\Exp(1)$
to handle the orthogonal $X$ case.
Dividing \eqref{eqablim} by $b_p^2$, we see that
\[
\frac{V_1+a_p}{b_p} =\frac{V_1-a_p}{b_p} + 2 + o(1) \rightarrow2
\]
and multiplying by \eqref{eqablim}, we get
\[
V_1^2-a_p^2 \stackrel{d} {
\rightarrow}\operatorname{Gumbel}(0,2),
\]
which may be rearranged to give Cai and Yuan's observation [since
$a_p^2 = 2\log p -\log\log p - \log\pi+o(1)$]. Hence for the
orthogonal case, under the global null, we are basically using the
same extreme value theory.
But for a general predictor matrix $X$, even if we stick to
testing at the first
step, we believe the Gumbel test does not share the same kind of
parameter-free asymptotic behavior of the covariance test.
Specifically, take $X^TX$ to be the $p \times p$ matrix with 1 on the
diagonal and every off diagonal element equal to some fixed
$\rho\in(0,1)$. In this case, we can show that under the global null,
\[
V_1 - \sqrt{2\rho\log p}\stackrel{d} {\rightarrow}\bigl|N(0,1-\rho)\bigr|,
\]
so the asymptotic distribution depends on $\rho$,
and the procedure suggested by Cai and Yang must fail.
Figure~\ref{figcai} shows an example with $n=100$, $p=50$, and
$\rho=0.7$. The Gumbel approximation is poor, while the $\Exp(1)$
distribution for the covariance test statistic still works well.
\begin{figure}
\includegraphics{1175rejf03.eps}
\caption{Quantile--quantile of Gumbel test (left panel) and
covariance test (right panel) with features having pairwise
correlation $0.7$.}\label{figcai}
\end{figure}
Another important point is that, for a general $X$, the test proposed
by Cai and Yuan does not apply to the sequence of variables entered
along the lasso path. Cai and Yuan assume that,
given a current active set $A$, the variable $j$ to be entered is that
which maximizes the drop in residual sum of squares. (In their
notation, the representation $R_j = \max_{m \notin A} R_m$ is what
allows them to derive the asymptotic Gumbel null distribution for
their test.)
While this is true at each lasso step in the orthogonal $X$ case, it
is certainly not true in the general $X$ case. Meanwhile, for an
arbitrary $X$, it is true in forward stepwise regression (by
definition).
\section{\texorpdfstring{Fan and Ke.}{Fan and Ke}}
Professors Fan and Ke extend the covariance test and its null
distribution to the SCAD and MCP penalties, in the orthogonal $X$
case. This is very exciting. We wonder whether this can be extended to
arbitrary $X$, and whether the spacing test [\citet{spacing}] can be
similarly generalized.
Fan and Ke (and also B\"uhlmann, Meier and van de Geer) also study the
important issue of the power of the covariance test,
relative to the ``RSSdrop'' and ``MaxCov'' statistics.
The discussants here have honed in on
the worst case scenario for the covariance test, in which two
predictors have large and equal coefficients. In this situation, the
LARS algorithm
takes only a short step after the first predictor has been
entered, before entering the second predictor, and the hence the
$p$-value for the first step is not very small. For this reason, better
power can be achieved by constructing functions of more than one
covariance test $p$-value, as illustrated by Figure~4 in the discussion
of Fan and Ke. We note, however, that neither RSSdrop nor MaxCov have
tractable null distributions in the general $X$ case, and it is not
even clear how to approximate these null distributions by simulation
except in the global null setup. Power concerns were also part of the
motivation for our development of the sequential tests in
\citet{fdrlasso}. Also, it is worth mentioning that the framework of
\citet{spacing} actually allows for combinations of the knots
$\lambda_j$, $j=1,\ldots, k$ from the first $k$ steps, so that we could
form an exact test based on, for example, $\sum_{j=1}^k
\lambda_j$ is
this was seen to have better power. Overall, the issue of the ``most
powerful sequential test'' remains an open and important one.
Continuing on the topic of power, Fan and Ke (and again, B\"uhlmann et
al.) raise asymptotic concerns.\vspace*{1pt} They suggest that power against
coefficients on the
order $O(n^{-1/2})$ is desirable. A first clarification: if elements of
$y$ and the rows of $X$ are generated by i.i.d. sampling, then the
matrix $X^TX$ grows like $n$; our standardization, in which $X^TX$
has 1 in each diagonal entry, corresponds to multiplying $\beta^*$ by
$\sqrt{n}$ in this i.i.d. sampling context. The rate $O(n^{-1/2})$
mentioned then becomes
$O(1)$. Power results will generally depend on $X$, and a complete
discussion would be outside of the scope of this discussion, but some
insight into what is possible or what is reasonable to expect may be
gained by considering the orthogonal
case. Consider now the problem of testing the global null against the
alternative $\beta^*_{j_0} \neq0$ and $\beta^*_j=0$ for all
$j\neq j_0$, with $j_0$ known. For $|\beta^*_{j_0}|=\nu$,
fixed, we get nontrivial limiting power by rejecting if
$|U_{j_0}| = |X_{j_0}^Ty|>z_{\alpha/2}$, as usual.\vspace*{2pt} But
realistically, ${j_0}$~will
not be known and it will be sensible to ask about the average power
over all ${j_0}\in\{1,\ldots,p\}$. The problem of testing $\beta^*=0$
against the hypothesis that there is a unique ${j_0}\in\{1,\ldots,p\}
$ for
which $\beta^*_{j_0} \neq0$ is invariant under permutations
of the entries in $U$. Let ${\mathcal T}_p$ denote the class of all
permutation invariant tests $T(U)$; our test $T_1$ and any other tests
which are functions of the order statistics of $U_j$, $j=1,\ldots, p$ are
permutation invariant. Let $B_p(\nu)$ be the set of $\beta^*$ with
exactly one nonzero entry satisfying $|\beta^*_j| \le\nu$. We can
prove that if $\nu_p$ is any sequence of constants with
\[
\sqrt{2\log p }-\nu_p \to\infty,
\]
then
\[
\sup_{T \in{\mathcal T}_p, \beta^*\in B_p(\nu_p)} \bigl|\operatorname
{Power}\bigl(T,\beta^*\bigr) -
\operatorname{Level}(T)\bigr| \to0.
\]
For tests which are not permutation invariant, we can prove
\[
\sup_{T, \beta^*\in B_p(\nu_p)} \bigl|\operatorname{AveragePower}\bigl
(T,\beta^*\bigr)
- \operatorname{Level}(T)\bigr| \to0,
\]
where now AveragePower denotes, for a given $\beta^*\in B_p(\nu_p)$,
the average over the $p$ vectors obtained by permuting the
entries of $\beta^*$. In other words, unless $\beta^*$ has an entry on
the order of $\sqrt{2\log p}$, there is no permutation invariant way to
distinguish the null from the alternative. On the other hand, if
$a_p=\sqrt{2\log p} - \log(\log p)/(2\sqrt{2\log p})$ and
\[
a_p(a_p-\nu_p) \to-\infty,
\]
then our test has limiting power 1 in this context.
This $\sqrt{2\log p}$ rate, then, cannot be substantially improved in
general. The same conclusion holds if $B_p(\nu)$ is replaced by the
intersection of the $O(1)$ ball $\{\beta\dvtx\|\beta\|_2 \le\Delta\}$
with $\{\beta\dvtx|\bar\beta| \le\varepsilon_p/\sqrt{p}\}$.
Here $\Delta$ is any fixed constant,
$\bar\beta= \sum_{j=1}^p \beta_j/p$,
and $\varepsilon_p$ is any sequence shrinking to 0. Notice that if
$\beta^*$ in this set is known then using a likelihood ratio test
against that alternative achieves nontrivial asymptotic power
(provided $\|\beta^*\|_2$ stays away from 0). If the permutation group
is expanded to the signed permutation group, then the condition on
$\bar\beta$ may be deleted; natural procedures will have this added
sign invariance in the orthogonal case.
\section{\texorpdfstring{Lv and Zheng.}{Lv and Zheng}}
Professors Lv and Zheng explore extensions of these ideas to nonconvex
objective functions, for example, a combination of Lasso and the SICA
penalty. This is interesting but seems difficult, as even the
computation of the global solution is infeasible in general. However,
the existing asymptotic results for these methods suggest that
inference tools might also prove to be tractable. Regarding the
significance of each active predictor conditional on the set of all
remaining active predictors: the spacing theory in \citet{spacing}
provides a method for doing this.
Lv and Zheng also suggest extra shrinkage, replacing
$\lambda_{k+1}$ in our, and their, test statistics by
$c\lambda_{k+1}$, in the hopes that a better choice of $c<1$ will lead
to an improved $\Exp(1)$ approximation. In knot form, this would look
like
\[
C(A,s_A,j,s)\lambda_k(\lambda_k - c
\lambda_{k+1}) = T_k +(1-c)C(A,s_A,j,s)
\lambda_k\lambda_{k+1}.
\]
Typically, $\lambda_k$, $\lambda_{k+1}$ are drifting to $\infty$ with
$p$, so the shrinkage factor $c$ will have to be chosen carefully in
order to control the second term above;
it seems that $c \rightarrow1$ is needed whenever the limit of
$T_k$ is $\Exp(1)$.
\section{\texorpdfstring{Wasserman.}{Wasserman}}
Professor Wasserman appropriately points out the stringency of
assumptions made in our paper, assumptions that
are common to much of
the theoretical work on high-dimensional regression. We would like to
reiterate that three of the offending assumptions in his
list---that is, the assumptions that the true model is linear and is
furthermore
sparse, and that the predictors in $X$ are weakly correlated---are not
needed in the newer works of \citet{spacing} and \citet{howlong}.
In general, though, we do agree that the rest of assumptions in his list
(implying independent, normal, homoskedastic errors) are used for as a
default starting point for theoretical analysis, but are certainly
suspect in
practice.
Wasserman outlines a model-free approach to inference in adaptive
regression based on sample-splitting and the increase in predictive
risk due to setting a coefficient to zero. The proposal is simple and
natural, and we can appreciate model-free approaches that use sample
splitting like this one. However, we worry about
the loss in power due to splitting the data in half, especially when
$n$ is small relative to $p$. As he says, this may be the
price to
pay for added robustness to model misspecification. How steep is
this price? It would be interesting to investigate, both
theoretically or empirically, the precise power lost due to sample
splitting.
Also, we note that
the random choice of splits will also influence the results, perhaps
considerably. Therefore, one would need to take multiple random
splits, and somehow combine the results at the end;\vadjust{\goodbreak} but then the
interpretation of the final ``conditional'' test seems challenging.
We are eager to read a completed manuscript on this interesting idea.
His discussion of conformal prediction is fascinating; this is an area
completely new to us. And finally, we thank him for his clearly
expressed reminder of the difficulties of determining causality from a
standard statistical model.
\section{\texorpdfstring{Thanks.}{Thanks}}
We thank all the discussants again for their contributions. They
have given us much to think about. We hope that our original paper,
the subsequent discussions and this response will be a valuable
resource for researchers interested in inference for adaptive
regression.
|
\section{Introduction}\label{sec:intro}
Early large-scale, cosmological simulations of galaxy formation in the $\Lambda$CDM paradigm suggested that the spherically-averaged density profiles of dark matter halos follow a universal profile across a large dynamic range in mass \citep{navarro96}. Since then, higher resolution simulations --- both with and without baryons --- have produced dark matter halos that (1) are permeated with substructure on many scales, (2) are triaxial in shape, and (3) have shapes and orientations that vary with radius \citep{dubinski91, jing02, kuhlen07, veraciro11}. Dark-matter-only simulations produce triaxial halos \citep{jing02} with large density fluctuations \citep{zemp09}. Inclusion of baryons tends to soften the triaxiality and graininess in the inner galaxy through a combination of dissipative infall \citep{dubinski94} or cooling \citep{bryan13}. These processes combined with the gravity from a baryonic disk or ellipsoid can act to make the inner halo more oblate or spherical, however they do not seem to erase the clumpy, triaxial nature of the outer halo \citep[e.g.,][]{pontzen12}. This can lead to radially-dependent axis ratios, orientation, and smoothness, and suggests that the true mass distributions around Milky Way-like galaxies are not easily represented by simple, time-independent potentials. Methods that seek to measure the gravitational potentials around such galaxies must be flexible enough to handle generic potential forms where finding simple analytic approximations or computing actions may not be possible.
The bulk of the baryonic matter in galaxies spans roughly 5-10\% of the spatial extent of the host dark matter halo. Hence, the brightest and most easily observable components of a galaxy are sensitive to the inner portion of the host halos mass distribution. For example, the rotation curves of disk galaxies trace the inner mass with exquisite sensitivity since matter in disks can be assumed to move on nearly circular orbits. Measurements of the dark matter distribution at large radii is complicated by the low density of visible tracers, observational difficulties of measuring kinematics of stars at large distances, and unknown orbits. Around external galaxies, the extended mass distribution has been studied using a variety of approaches \citep[see][for a a complete and detailed review]{courteau13}. For example, the kinematics of tracer populations such as globular clusters or planetary nebulae can be used to derive mass estimates under the assumptions that these satellite systems are on random orbits and are well-mixed in orbital phase \citep[early investigations include][]{mendez01,cote03}. Simple, parameterized fits to both the mass and orbit distribution have been simultaneously constrained using such data \citep[e.g.][]{napolitano11,deason12c}. Alternatively, the statistical properties of gravitationally lensed background sources around a galaxy can be used to constrain the \emph{projected} shape, orientation, and radial profile of mass \citep[see, for example, the Lens Structure and Dynamics Survey described in][]{koopmans02}. Of course, lensing reconstructions can only be performed for galaxies which closely intersect our line of sight to background sources, but the advent of large photometric catalogues has allowed automatic searches for such chance alignments and significant increases in the number of objects studied in this way \citep[e.g. the Sloan Lens ACS Survey, see][]{bolton06}.
Within the Milky Way our unique vantage point allows us a three-dimensional view of stars within our own dark matter halo. Our proximity allows us to use individual stars as kinematic tracers and hence build much larger samples that probe deeper into the halo than the globular cluster and planetary nebula studies of external galaxies. For example, \cite{deason12a} used halo BHB stars selected from the Sloan Digital Sky Survey \cite[SDSS;][]{york00} as a random tracer population to measure the mass and slope of a power-law fit to the potential. Such studies assume that the tracer orbits are randomly sampled from a smooth distribution function and are fully phase mixed. However, large photometric surveys such as the SDSS and 2MASS \citep{skrutskie06} have discovered copious amounts of substructure --- in streams and kinematic associations of stars --- in the Milky Way halo \citep[e.g.,][]{belokurov06, rochapinto04}, thus demonstrating that the stellar distribution is neither on random orbits nor fully phase-mixed. Substructure in the form of stellar streams and clouds is known to bias mass and velocity inferences from random tracer methods by several tens of percent \citep{yencho06}.
Another approach to using halo stars as potential measures is to exploit the non-random nature of halo. Tidal streams are dynamically cold systems --- debris typically have small distributions of energy and angular momentum --- and thus require orders of magnitude fewer tracers than a random sample to get constraints of comparable accuracy to Jeans analysis. For example, in the simplest case we might assume that debris stars are actually still on the same orbit as their progenitor system (a \emph{wrong} assumption, see below). This information about the orbits combined with measurements of the full-space velocities ${\bf v}$ at different points ${\bf x}$ in the structure (e.g., along a stream) would give us a direct measure of differences in a potential, $\Phi$.
\citet[][LM10]{law10} used N-body simulations of the disruption of the Sgr dwarf galaxy to simultaneously fit a model to the available data on the debris stars and a triaxial, analytic Milky Way potential. By varying parameters of the Milky Way potential, they found ``best-fit'' parameters by comparing the properties of observed Sgr stars to their simulated debris. The computational costs of running N-body simulations limited their search to a grid of potential parameters and forced the authors to fix many other parameters (e.g., properties of the disk and bulge). Nevertheless, they were able to constrain the 3D shape that their assumed potential model must take in order to best represent the Milky Way out to $\sim$70 kpc and found that the best-fitting halo has a nearly oblate (only mildly triaxial) shape flattened in a direction close to the Sun-Galactic center line. Though an unlikely orientation for the halo --- \cite{debattista13} find that the disk of the Milky Way would not remain stable in such a configuration --- LM10 showed that the data is at a state where such inference is possible.
The computational costs associated with N-body simulations has motivated the development of many methods that approximately model tidal streams. The simplest alternative is to fit a single orbit to observed debris \citep[e.g.,][]{koposov10, deg13}. Though this is known to be incorrect and leads to biases in inferred properties of the underlying potential \citep[e.g.,][]{eyre11, lux13, sanders13a}, \cite{deg14} and \cite{lux13} have used orbit fitting to demonstrate the power of combining multiple streams in dynamical inference. To account for the offset between the orbit of the progenitor and the orbits of the debris stars, methods have been proposed that add some dispersion or offset around a single orbit either in phase space \citep[e.g.,][]{eyre09a, varghese11, kuepper12} or action-angle coordinates \citep{eyre11, sanders13b, bovy14, sanders14}. Other statistical methods have been proposed \citep[][]{johnston99a, penarrubia12, sanderson14} that may prove powerful when applied to, for example, data from the \project{Gaia}\, mission, where full 6D coordinates will be known for large samples of stars in the halo --- and therefore many debris structures --- but stream membership is not known for all stars.
Motivated by the strengths of previous work, we identify a minimum set of issues that any approximate stream model or potential recovery method should address (see Section~\ref{sec:discussion} for a more detailed discussion of these points in the context of this work):
\begin{enumerate}
\item \textbf{Observational uncertainties:} The known debris structures are $\gtrsim$10 kpc from the Sun where distance and proper motion measurement errors are significant. Thus it is critical for any method that uses tidal debris to incorporate observational uncertainties and missing dimensions in a consistent and justified way.
\item \textbf{Form of the potential:} There is large uncertainty in the radial profile, shape, orientation, and graininess of the outer halo and the constancy of these parameters over distance. Properly describing the potential may require non-parametric techniques or complex analytic functions so the method should not rely on the existence of or ability to compute conserved orbital properties such as actions.
\item \textbf{Multiple debris structures:} Near-future photometric surveys such as \project{Gaia}\, and the \project{LSST} will likely discover many new streams and kinematic associations of stars. Potential recovery methods should be able to simultaneously use multiple streams and incorporate other dynamical constraints.
\item \textbf{Comparing models to data:} Matching generated streams to observed stellar densities is difficult; models that rely on this must account for observational biases, internal properties of the progenitor, and background halo stars.
\item \textbf{Computational expense:} Full N-body simulations are expensive to run; incorporating an N-body simulation into a likelihood function evaluation and then performing a parameter search would be computationally intensive and, presently, intractable.
\end{enumerate}
In \citet{apw13}, we introduced a simple method based on the work of \citet{johnston99a} for using individual stars associated with tidal debris combined with knowledge about the mass and orbit of the progenitor to constrain properties of the host galaxy potential. The method exploits the relationship between the phase-space distribution of debris and (measurable) properties of the progenitor system (e.g., the tidal radius and escape velocity). Specifically, a better potential is one in which orbits of test particle stars (integrated backwards from their present position) came close (within the tidal radius in position, and escape velocity in velocity) to the orbit of the progenitor at some time in the past.
In this paper, we present a fully probabilistic model (\texttt{Rewinder}) for tidal streams that builds on these simple scalings, which depend only on the mass and orbit of the progenitor and parent potential. By ``probabilistic model'' we mean a justified, parametrized likelihood function with priors on the parameters. \texttt{Rewinder}\ relies on numerical orbit integration in ordinary phase-space and can thus incorporate arbitrarily complex forms for the parent potential (such as time dependence and significant substructure). Individual stars are constraints on the potential, thus \texttt{Rewinder}\ can handle very small samples of well-measured stars (e.g., 4-16). Incorporating multiple streams, debris structures, and other kinematic information is trivial but will be explored in future work.
In Section~\ref{sec:sims} we describe a suite of N-body simulations that span a range of progenitor masses on a characteristic, mildly eccentric orbit and then use them in Section~\ref{sec:method} to motivate a new, flexible model for tidal streams that works entirely in phase-space. In Section~\ref{sec:experiments} we demonstrate how \texttt{Rewinder}\ can be used to measure properties of a non-trivial Galactic potential by performing several experiments with simulated observations of data from N-body simulations. In Section~\ref{sec:discussion}, we discuss the results of these experiments and the extent to which we address the above points 1-5. We conclude in Section~\ref{sec:conclusion}.
\section{Simulations}\label{sec:sims}
We performed a set of N-body simulations with the Self-Consistent Field (SCF) basis function expansion code \citep{hernquist92} to build realistic models of streams for studying the phase-space distribution of debris as it is stripped from its progenitor. We later use one of these simulations to test \texttt{Rewinder}. In each simulation, a $10^5$ particle NFW-profile satellite was inserted at the apogalacticon of its orbit in a static, multi-component galaxy with a triaxial host halo, described below. The satellite was evolved first in isolation, then the host potential was turned on slowly over 10 satellite internal dynamical times to reduce artificial gravitational shocking. The initial position and velocity was obtained by integrating the orbit of the Sagittarius dwarf galaxy from present-day coordinates ${\bf r} =(19.0, 2.7, -6.9)\ \mathrm{kpc}, \ {\bf v} = (230., -35., 195.) \ \mathrm{km\ s^{-1}}$ \citep{law10} backwards for $\sim$6 Gyr. The N-body satellite was then reintegrated from that initial position for the same interaction time and number of time-steps. This orbit has its apogalacticon at approximately 59 kpc and perigalacticon near 12 kpc with an average orbital period of 930 Myr, although all of these quantities vary over the course of the simulation due to the halo's triaxiality. Total energy is conserved to $\sim$2\% of the satellite internal potential energy.
To capture the character of streams across a range of merger mass ratios, four satellites with masses $\mathrm{m} = 2.5 \times 10^6,\ 2.5 \times 10^7,\ 2.5 \times 10^8,\ 2.5 \times 10^9\ \mathrm{M}_\odot$, where m is the mass enclosed within 35 NFW scale radii, were evolved in the setup described above. Their scale radii, $r_0$, were adjusted to maintain a constant density across all simulations which results in identical fractional mass loss rates: 74\% of the initial mass is lost by the end of each simulation. The base value is $\mathrm{r_0}=0.01565$ kpc at $\mathrm{m} = 2.5 \times 10^6\ \mathrm{M}_\odot$.
We take care to record the time at which each particle is unbound from the satellite in the simulations: we locate the position of the remnant iteratively by first calculating the satellite potential with all particles, removing particles with kinetic energies sufficient to escape, then recalculate the potential without those particles until the system converges. There is no spatial restriction on where a particle may become unbound and all particles (both bound and unbound) contribute their gravity during normal time-steps, presuming that they are resolved by the basis functions.
In all simulations, the host potential is taken to be a three-component sum of a Miyamoto-Nagai disk \citep{miyamoto75}, Hernquist spheroid, and a triaxial, logarithmic halo \citep[e.g.,][]{law10}:
\begin{align}
&\Phi_{\rm disk}(R,z) = -\frac{GM_{\rm disk}}{\sqrt{R^2 + (a + \sqrt{z^2 + b^2})^2}}\\
&\Phi_{\rm spher}(r) = -\frac{GM_{\rm spher}}{r + c}\\
&\Phi_{\rm halo}(x,y,z) = v_{\rm h}^2 \ln(C_1 x^2 + C_2 y^2 + C_3 xy + (z/q_z)^2 + r_h^2)
\end{align}
where $C_1$, $C_2$, and $C_3$ are combinations of the $x$ and $y$ axis
ratios ($q_1$, $q_2$) and orientation of the halo with respect to the
baryonic disk ($\phi$):
\begin{align}
C_1 &= \frac{\cos^2\phi}{q_1^2} + \frac{\sin^2\phi}{q_2^2}\\
C_2 &= \frac{\sin^2\phi}{q_1^2} + \frac{\cos^2\phi}{q_2^2}\\
C_3 &= 2\sin\phi\cos\phi \left(q_1^{-2} - q_2^{-2}\right).
\end{align}
The total potential then is just
\begin{equation}
\Phi_{\rm tot} = \Phi_{\rm disk} + \Phi_{\rm spher} + \Phi_{\rm halo}\label{eq:lm10}.
\end{equation}
\begin{figure*}[!ht]
\begin{center}
\includegraphics[width=\textwidth]{potentials.pdf}
\caption{ Equipotential contours for the LM10 potential (Eq.~\ref{eq:lm10}) in Galactocentric, cartesian coordinates for various halo parameter choices. For all panels, $v_{\rm h}=121.858~\mathrm{km}/\mathrm{s}$, $r_h=12~\mathrm{kpc}$, and $q_2=1$. Left to right, each column represents a new choice of parameters. If not specified, other parameters are fixed to $q_1=q_z=1$ and $\phi=0^\circ$ (far left panels). Panels on far right show the best-fit parameter values from LM10. }\label{fig:potential}
\end{center}
\end{figure*}
We use this potential not because we think it is a realistic representation of the Galactic potential, but because successful inference with this potential demonstrates that it is possible to recover information about non-trivial potential forms. Figure~\ref{fig:potential} shows equipotential slices in the Galactic X-Z plane (Y=0) and Y-Z plane (X=0) for a few choices of $q_1$, $q_z$, and $\phi$ while holding all other parameters fixed, as described in the figure caption. The potential parameters used for the simulations are shown in the ``Truth'' column of Table~\ref{tbl:params}.
\begin{figure*}[!ht]
\begin{center}
\includegraphics[width=\textwidth]{simulated_streams.pdf}
\caption{ Particle positions (grey dots) in Galactocentric cartesian coordinates from the final time-step of four N-body simulations (Section~\ref{sec:sims}) with the same progenitor orbit initial conditions over a range of progenitor masses (columns). Black crosses indicate four particles chosen from each mass simulation and used in the experiments described in Section~\ref{sec:experiments}.}\label{fig:sims}
\end{center}
\end{figure*}
\section{Method (\texttt{Rewinder})}\label{sec:method}
\texttt{Rewinder}\ integrates stars and the progenitor back from their present-day, observed positions to the time at which they become unbound from the satellite, where we evaluate the likelihood for each star. Below we (1) motivate a simple model for the debris just as it disrupts and (2) present this idea within a probabilistic framework for incorporating observational uncertainties and missing data. For a satellite galaxy orbiting within a static host potential, mass loss is driven by a combination of tidal stripping by the steady tidal field of the parent system and tidal shocking by the rapidly changing tidal field as the progenitor system moves through pericenter \citep[e.g.,][]{choi09}. On a mildly eccentric orbit, the tidal shocking is subdominant to the steady disruption. The interplay between these two processes as a function of orbital properties is outside the scope of this paper and will be explored in future work; we focus here on the slow removal of stars by steady tidal forcing.
\subsection{Motivation: tidal debris}\label{sec:debris}
Consider a point-mass satellite with mass $m$ on a circular orbit with frequency $\Omega$ around a more massive ``host'' mass $M\gg m$ \citep[the ``restricted three-body problem''; e.g., \S 8.3][]{binneytremaine}. In a frame rotating with the orbital frequency of the satellite, the satellite remains fixed and the static, effective potential around the satellite has (amongst others) two unstable optima located around galactocentric radii $\sim R \pm r_J$, where $R$ is the orbital radius of the satellite and $r_J$ is the Jacobi or tidal radius
\begin{equation}
r_J \sim R\left(\frac{m}{3M}\right)^{1/3}.\label{eq:ptmass}
\end{equation}
Particles that would be bound to the satellite in isolation may have enough (Jacobi) energy to overcome the effective potential barrier at these Lagrange points (Fig.~8.6 of BT) and thus will be preferentially stripped from the satellite at these points. For a spherical, extended parent mass distribution the tidal radius instead scales with the enclosed mass at the instantaneous orbital radius, $R(t)$, with additional terms that account for the local slope of the density profile. In general the satellite mass also depends on time. The tidal radius of the debris then follows
\begin{equation}
r_{\rm tide}(t) = R(t)\left(\frac{m(t)}{3M_{\rm enc}(R)}\right)^{1/3}\label{eq:tidalradius}
\end{equation}
where $M_{\rm enc}$ is the instantaneous enclosed mass of the parent system within orbital radius $R$.\footnote{Note that this is the instantaneous orbital radius, \emph{not} the orbital radius at pericenter.}
In general, the Lagrange points may not be symmetric about the center of the satellite and may deviate from the tidal radius by factors of order unity.
The bulk velocity of the satellite will be of order $V\sim \sqrt{GM_{\rm enc}/R}$. If the velocity dispersion of the satellite is $\sigma_v \sim \sqrt{Gm/r_J}$, then it follows
\begin{equation}
\sigma_v(t) \sim V\left(\frac{m(t)}{M_{\rm enc}(t)}\right)^{1/3}\label{eq:velscale}.
\end{equation}
\citep[as pointed out in][]{binneytremaine} and hence we expect the debris-star velocities also to scale with $(m/M_{\rm enc})^{1/3}$. These scalings assume that the satellite is spherical with isotropic velocities; non-random, internal satellite orbits (e.g., a disk) will break these assumptions.
In a triaxial potential, the orbital plane of the satellite is not fixed but we still expect there to be \emph{effective} Lagrange points for a spherical satellite system along the line of centers connecting the origin of the parent potential to the satellite --- that is, we expect the stars to be stripped at some characteristic distance from the satellite (near the tidal radius) along the instantaneous position vector of the satellite with some dispersion about these points. We proceed by defining a coordinate system relative to the position and velocity of the progenitor, ($\boldsymbol{r}_p$,$\boldsymbol{v}_p$), rotated into the instantaneous orbital plane defined by $\boldsymbol{L} = \boldsymbol{r}_p \times \boldsymbol{v}_p$. The (time-dependent) basis vectors are given by
\begin{align}
\hat{\boldsymbol{x}}_1 &= \frac{\boldsymbol{r}_p}{\|\boldsymbol{r}_p\|}\label{eq:x1}\\
\hat{\boldsymbol{x}}_2 &= \frac{\boldsymbol{L} \times \hat{\boldsymbol{x}}_1}{\|\boldsymbol{L} \times \hat{\boldsymbol{x}}_1\|}\\
\hat{\boldsymbol{x}}_3 &= \hat{\boldsymbol{L}} = \frac{\boldsymbol{L}}{\|\boldsymbol{L}\|} = \frac{\boldsymbol{r}_p \times \boldsymbol{v}_p}{\|\boldsymbol{r}_p \times \boldsymbol{v}_p\|}\label{eq:x3}.
\end{align}
\begin{figure}[p]
\begin{center}
\includegraphics[width=\textwidth]{Lpts_r.pdf}
\caption{ Orbits of 2000 randomly-drawn, disrupted particles projected into the instantaneous orbital plane coordinates (Eqs.~\ref{eq:x1}-\ref{eq:x3}), normalized by the tidal radius (Eq.~\ref{eq:tidalradius}), and only shown within half a satellite crossing time around $t=t_{\rm ub}$ for each of the four progenitor masses. The orbits were integrated backwards from their present-day positions (final time step of the N-body simulations) as test particles without the potential of the progenitor. Horizontal black line shows $x_2=0$ (top panels) and $x_3=0$ (bottom panels), and the unit circle (black circle) illustrates the classical disruption radius in these coordinates. }\label{fig:lpts_r}
\end{center}
\end{figure}
Figure~\ref{fig:lpts_r} shows sections of particle orbits (for particles that are stripped) from the N-body simulations described above, projected into this coordinate system. The positions are normalized by the instantaneous tidal radius (Eq.~\ref{eq:tidalradius}) and velocities are normalized by the instantaneous velocity scale (Eq.~\ref{eq:velscale}). Sections of the orbits symmetric around their unbinding times, $t_{\rm ub}$, are shown for each of the four progenitor masses. In these coordinates, the classical Lagrange points would be located at $x_1\approx\pmr_{\rm tide},x_2=0,x_3=0$ (illustrated by the intersection of the black unit circles and horizontal lines in Figure~\ref{fig:lpts_r}). Though not exactly centered on the point-mass Lagrange points derived for a circular orbit, the location of and dispersion about the effective Lagrange points in these scaled coordinates remain remarkably consistent across the range of progenitor masses explored. Figure~\ref{fig:lpts_v} shows the velocity orbits of each star also projected into these coordinates and normalized. The dispersion in velocity is well-normalized by the velocity scale of Eq.~\ref{eq:velscale}, though it is clear that the velocity dispersion along $\hat{x}_1$, the radial vector to the satellite position, is larger than that in other dimensions, possibly because of mild tidal shocking.
\begin{figure}[ph]
\begin{center}
\includegraphics[width=\textwidth]{Lpts_v.pdf}
\caption{ Same as Figure~\ref{fig:lpts_r} but for particle velocities normalized by the velocity scale (Eq.~\ref{eq:velscale}). }\label{fig:lpts_v}
\end{center}
\end{figure}
We conclude that even on a non-circular orbit in a complex potential, the dispersion --- in position and velocity --- of tidally stripped debris as it comes unbound from the satellite scales with the mass ratio $(m / M_{\rm enc})^{1/3}$. This motivates a model for tidal debris in which each star was ``released'' at the instantaneous, effective Lagrange point at its unbinding time, $t_{\rm ub}$, with a dispersion in position and velocity, all of which depend only on the mass and orbit of the progenitor and the parent potential. We present this model in detail below.
\subsection{Probabilistic model}
Suppose we observe the 6D position of a star, ${\bf D} = (l,b,d,\mu_l,\mu_b,v_r)$, in heliocentric coordinates --- e.g., the measured position on the sky, ($l$, $b$); distance, $d$; proper motions, ($\mu_l$, $\mu_b$); and line-of-sight velocity, $v_r$ --- and have determined through some other means that this star was once part of a progenitor system (e.g., satellite galaxy) with mass $m(t)$ that is disrupting and forming a cold debris structure in the potential, $\Phi$, of the parent galaxy (e.g., the Milky Way). We assume that the mass of the parent system enclosed within the pericenter of the orbit of the satellite, $M_{\rm enc}(R_{\rm peri})$, is much larger than the initial mass of the satellite, $M_{\rm enc}(R_{\rm peri})\gg m(t=0)$. The present position of the progenitor is observed to be at heliocentric position ${\bf D}_{\rm p}$ (where any subscript ${\rm p}$ refers to the progenitor). In general, the data for the star and progenitor will have significant uncertainties or missing dimensions at distances typical to the Galactic halo, thus we define ${\bf W}$ and ${\bf W}_{\rm p}$ as the true 6D, heliocentric positions of the star and progenitor and will include these in our model. Since we include these positions as parameters, this method will work even when the star has missing dimensions or the progenitor location is unknown \citep[as in the Orphan stream,][]{belokurov07}.
To model the true, 6D, present-day position of the star, ${\bf W}$, we first transform to cartesian, Galactocentric coordinates where ($\boldsymbol{r}_0,\boldsymbol{v}_0$) and ($\boldsymbol{r}_{{\rm p},0},\boldsymbol{v}_{{\rm p},0}$) are the position and velocity of the star and progenitor today. The star is taken as having been sampled at $t=t_{\rm ub}$ (the unbinding time) from an isotropic Gaussian centered on one of the effective Lagrange points in position, and an isotropic Gaussian centered on the origin in velocity. The present-day phase-space position is the result of integrating this sample forward from $t_{\rm ub}$ in the parent potential, $\Phi$, whose form is parametrized by the vector $\boldsymbol{\Theta}_\Phi$. We assume that once the star becomes ``unbound'' from the satellite the potential of the satellite can be ignored,\footnote{This assumption breaks down for sufficiently large-mass progenitors, somewhere between $\sim10^8-10^9~\mathrm{M}_\odot$.} and thus we treat the star as a test particle. The satellite mass enters this prescription through the tidal radius; a changing satellite mass will skew the positions of the effective Lagrange points and velocity scale. We allow for mass-loss by incorporating the initial satellite mass, $m_0$, and a constant mass-loss term, $\dot{m}$ into the model: $m(t) = m_0 - \dot{m}t$. Finally, we add two additional parameters: a constant, global factor, $\alpha$, that scales the position of the effective Lagrange points relative to the classical tidal radius, and a binary parameter $\beta$ (one for each star), that is either $-1$ or $+1$ depending on whether the star is in the leading or trailing tail. To compress notation, we pack all progenitor parameters into the vector $\boldsymbol{\Theta}_{\rm p} = ({\bf W}_{\rm p}, m_0, \dot{m}, \alpha)$. The likelihood for this model is:
\begin{align}
p({\bf W} \,|\, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi, \beta, t_{\rm ub}) &= p(\boldsymbol{r}_0,\boldsymbol{v}_0, \,|\, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi, \beta, t_{\rm ub})\,\left\vert{\boldsymbol J}\right\vert\label{eq:starlike}\\
p(\boldsymbol{r}_0,\boldsymbol{v}_0, \,|\, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi,\beta,t_{\rm ub}) &= \left[\mathcal{N}(\boldsymbol{r} \,|\, \boldsymbol{\Theta}_{\rm p}, \beta)\,\mathcal{N}(\boldsymbol{v} \,|\, \boldsymbol{\Theta}_{\rm p})\right]_{t_{\rm ub}}
\end{align}
where $r_{\rm tide}$ is given by Eq.~\ref{eq:tidalradius}, and $\sigma_v$ by Eq.~\ref{eq:velscale} and the positions and velocities ($\boldsymbol{r},\boldsymbol{v}, \boldsymbol{r}_{{\rm p}},\boldsymbol{v}_{{\rm p}}$) are evaluated at the unbinding time, $t_{\rm ub}$, by integrating the orbits backwards from ($\boldsymbol{r}_0,\boldsymbol{v}_0, \boldsymbol{r}_{{\rm p},0},\boldsymbol{v}_{{\rm p},0}$). $\left\vert{\boldsymbol J}\right\vert = \left\vert\frac{\partial(x,y,z,v_x,v_y,v_z)}{\partial(l,b,d,\mu_l,\mu_b,v_r)}\right\vert$ is the absolute value of the determinant of the Jacobian that defines the transformation from heliocentric, spherical to Galactocentric, cartesian coordinates. We then marginalize the likelihood over all possible unbinding times, assuming a uniform prior of unbinding times over the entire interaction time:
\begin{align}
p({\bf W} \,|\, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi, \beta) &= \int^{t_{\rm int}}_0 p({\bf W} \,|\, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi, \beta, t_{\rm ub})\,p(t_{\rm ub})\,dt_{\rm ub}
\end{align}
where the interaction time is taken to be $t_{\rm int}=6.2$~Gyr, the total duration of the simulations of Section~\ref{sec:sims}.
The likelihood above is evaluated for each star at all time-steps of the simulation during the marginalization with a uniform prior over all possible unbinding times (no assumptions are made about the orbital phase dependence of mass-loss). We chose this approach because the phase-space position (and hence time) at which a star becomes unbound from a satellite can only be strictly defined for the case of a spherical satellite orbiting on a circular orbit in a spherical potential. Various authors have chosen to measure mass loss in simulations of satellite disruption on eccentric orbits in non-spherical potentials by either looking at the kinetic energy of the particles relative to the satellite's potential energy \citep{johnston95} or looking at the point at which a star crosses the instantaneous tidal radius \citep{kuepper12}. Neither choice is strictly valid and both have been found useful for representing the mass-loss process. Hence we make the simplest choice of using Gaussians to characterize the phase-space distribution of debris as it makes the transition between being bound and unbound. We find that changing this distribution to a log-normal distribution does not affect our conclusions.
We assume that the observational uncertainties are Gaussian in heliocentric, spherical coordinates, such that
\begin{align}
p({\bf D} \,|\, {\bf W}) &= \mathcal{N}({\bf D} \,|\, {\bf W}, {\bf \Sigma})\label{eq:obsstar}\\
p({\bf D}_{\rm p} \,|\, {\bf W}_{\rm p}) &= \mathcal{N}({\bf D}_{\rm p} \,|\, {\bf W}_{\rm p}, {\bf \Sigma}_{\rm p})\label{eq:obsprog}
\end{align}
where $\mathcal{N}$ represents the normal distribution, and the covariance matrices ${\bf \Sigma}$ and ${\bf \Sigma}_{\rm p}$ specify the observational uncertainties on the observed 6D position of the star and progenitor, respectively.
For a single star, the evaluation of the likelihood works as follows: given observed, present-day coordinates for the star and progenitor (${\bf D},{\bf D}_{\rm p}$), potential parameter values ($\boldsymbol{\Theta}_\Phi$), and nuisance parameter values ($\alpha,\beta$),
\begin{enumerate}
\item transform star and progenitor positions in heliocentric coordinates to Galactocentric, cartesian coordinates;
\item integrate star and progenitor orbits backwards as test particles in the potential given by $\boldsymbol{\Theta}_\Phi$ for the total interaction time, $t_{\rm int}$ (in this case, $\sim$$6.2$~Gyr);
\item transform the particle position to the relative, normalized coordinates defined in Section~\ref{sec:debris};
\item compute the likelihood given by Equation~\ref{eq:starlike} at each time-step of the integration and marginalize over the unbinding time, $t_{\rm ub}$.
\end{enumerate}
Assuming each star is an independent tracer, the full likelihood for many stars is then just the product of the individual likelihoods:
\begin{equation}
p(\{{\bf D}^{(i)}\}, {\bf D}_{\rm p} \,|\, \{\boldsymbol{\Theta}^{(i)}\}, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi ) =
p({\bf D}_{\rm p} \,|\, {\bf W}_{\rm p}) \, \prod_i \, p({\bf D}^{(i)} \,|\, {\bf W}^{(i)}) \,
p({\bf W}^{(i)} \,|\, \boldsymbol{\Theta}_{\rm p}, \boldsymbol{\Theta}_\Phi, \beta^{(i)})
\end{equation}
where $\boldsymbol{\Theta}^{(i)} = ({\bf W}^{(i)}, \beta^{(i)})$. All parameters are summarized in Table~\ref{tbl:params}.
\section{Experiments} \label{sec:experiments}
In what follows, we use \texttt{Rewinder}\ to model ``observations'' of a small sample of stars from one of the N-body simulations described in Section~\ref{sec:sims}. We consider this a strong test of the method because the data are generated with N-body simulations and not with anything resembling the likelihood function presented above. For all tests in this article, we use the same functional form for the potential as used in the simulations (Equation~\ref{eq:lm10}). When recovering the potential, we hold fixed the disk and spheroid parameters (see Table~\ref{tbl:params}), along with one of the halo parameters: the scale radius, $r_h$. The priors on the remaining halo parameters are taken to be uniform over a conservative domain of realistic values: for $v_{\rm h}$, 100-200~km/s corresponds to a range in solar circular velocities from $\sim$210-250 km/s (holding other parameters fixed); the range in axis ratios allow for prolate, oblate, and generic triaxiality; and $\phi$ is restricted to $\pm45^\circ$ around the true simulation value, $\phi = 97^\circ$.
For all experiments below, we use a Markov Chain Monte Carlo (MCMC) algorithm to sample from the posterior probability distribution given by our model. Standard MCMC algorithms (e.g., Metropolis-Hastings) update a single chain while exploring parameter space. We instead use an affine-invariant ``ensemble'' sampler \citep{goodman10} that, each step in parameter space, updates the positions of many ``walkers'' (the ensemble). This algorithm is implemented in the \texttt{Python} programming language \citep{foremanmackey13} and runs naturally in a parallelized environment such as the message passing interface (MPI). In each experiment we run the walkers for a large number of steps from starting positions described below, then throw away these samples and take the positions from the final step of this ``burn-in'' phase as a starting point for the samples used for inference. We compute the autocorrelation time for each sampled parameter (using \texttt{ACOR}\footnote{\url{http://www.math.nyu.edu/faculty/goodman/software/acor/}}$^{,}$\footnote{\url{https://github.com/dfm/acor}}) and thin the chains by taking every $\mathrm{max}(t_{\rm acor})$ sample to ensure the samples are close to independent.
\begin{table*}[ht]
\begin{center}
\begin{tabular}{l c c l} \toprule
\multicolumn{4}{l}{{\bf \emph{Milky Way parameters} ($\boldsymbol{\Theta}_\Phi$)}} \\
\toprule
Component & Parameter & Truth & Prior \\\toprule
disk & $M_{\rm disk}$ & $1.0\times10^{11}\,\mathrm{M}_\odot$ & fixed \\
& $a$ & 6.5 kpc & fixed\\
& $b$ & 0.26 kpc & fixed\\
\midrule
spheroid & $M_{\rm spher}$ & $3.4\times10^{10}\,\mathrm{M}_\odot$ & fixed\\
& $c$ & 0.7 kpc & fixed\\
\midrule
halo & $v_{\rm h}$ & 121.858 & $\mathcal{U}(100,200)$ km/s \\
& $q_1$ & 1.38 & $\mathcal{U}(1,2)$\\
& $q_2$ & 1.0 & fixed\\
& $q_z$ & 1.36 & $\mathcal{U}(1,2)$\\
& $\phi$ & 97 & $\mathcal{U}(52,142)$ deg\\
& $r_h$ & 12 kpc & fixed\\
\toprule
\multicolumn{4}{l}{{\bf \emph{Progenitor parameters} ($\boldsymbol{\Theta}_{\rm p}$)}} \\
\toprule
position & $\boldsymbol{r}_{{\rm p},0}$ & -- & $\|\boldsymbol{r}_{{\rm p},0}\|\sim\mathcal{U}(0,200)$~kpc \\
velocity & $\boldsymbol{v}_{{\rm p},0}$ & -- & $\|\boldsymbol{v}_{{\rm p},0}\|\sim\mathcal{U}(0,500)$~km/s\\
Lagrange pt. offset & $\alpha$ & -- & $\mathcal{U}(0.5, 2.5)$\\
initial mass & $m_0$ & $2.5\times10^8\,\mathrm{M}_\odot$ & fixed\\
mass loss & $\dot{m}$ & -- & $3.2\times10^4\,\mathrm{M}_\odot$/Myr (fixed)\\
\toprule
\multicolumn{4}{l}{{\bf \emph{Star parameters} ($\boldsymbol{\Theta}^{(i)}$)}} \\
\toprule
position & $\boldsymbol{r}_0$ & -- & $\|\boldsymbol{r}_0\|\sim\mathcal{U}(0,200)$~kpc \\
velocity & $\boldsymbol{v}_0$ & -- & $\|\boldsymbol{v}_0\|\sim\mathcal{U}(0,500)$~km/s\\
tail assignment & $\beta$ & -- & $\pm1$~equally likely (fixed at truth)\\
\bottomrule
\end{tabular}
\caption{Parameter values used in the experiments of Section~\ref{sec:experiments}. $\mathcal{N}$ is the normal (Gaussian) distribution, and $\mathcal{U}$ the uniform distribution. There are 11 parameters for the Milky Way potential, but only four are left free to vary; some parameters are fixed (denoted by ``(fixed)'') at the true values used in the N-body simulations that generated the fake test data. The progenitor has nine parameters --- the position, $\boldsymbol{r}_{\rm p}$, and velocity, $\boldsymbol{v}_{\rm p}$, vectors each contain three components --- but only five are left free to vary. The sky coordinates (e.g., Galactic $l$, $b$) are assumed to be known with negligible uncertainty. Each star has eight associated parameters, five of which are allowed to vary. Sky coordinates are fixed, along with the tail assignment (whether the star belongs to the leading or trailing tail). For inference with four stars, there are $4+5+4\times4=25$ free parameters. \label{tbl:params}}
\end{center}
\end{table*}
\subsection{Data with negligible uncertainties}\label{sec:exp1}
We test \texttt{Rewinder}\ using only four stars (particles) --- two from the leading tail, and two from the trailing tail --- randomly sampled from the $2.5\times10^8\,\mathrm{M}_\odot$ simulation (Section~\ref{sec:sims}), assuming the observed 6D positions for both the stars and the progenitor have negligible uncertainties. The stars were required to have been stripped after the first pericentric passage and have a present-day distance within $40$ kpc of the Sun. Figure~\ref{fig:sims} (second column from right) shows the randomly chosen stars (black crosses) in Galactic coordinates, over-plotted on all other simulated particles (grey points).
We leave the potential parameters ($q_1,q_z,\phi,v_{\rm h}$) free to vary, along with the Lagrange point offset, $\alpha$, and initialize an ensemble of 64 MCMC walkers by sampling from the priors summarized in Table~\ref{tbl:params}. We run the walkers for 5000 steps to burn-in, then restart the sampler starting from the final position of the burn-in phase and run for another 5000 inference steps. Figure~\ref{fig:trace} shows the walker positions over the 5000 inference steps for each of the five parameters. The autocorrelation time for each parameter is displayed on its corresponding panel. Figure~\ref{fig:exp1_posterior} shows projections of the posterior probability distribution for the parameters; with perfect data, the uncertainties on the potential parameters are all $<1\%$.
\begin{figure}[p]
\begin{center}
\includegraphics[width=0.5\textwidth]{mcmc_trace.pdf}
\caption{ Positions of all 64 MCMC walkers (see Section~\ref{sec:experiments}) for each parameter at each step of the inference (black lines). The MCMC chains do not display any bulk movement or stray walkers, implying, by eye, that the chains have converged. We also compute the autocorrelation functions for each parameter and find that, for this experiment, the autocorrelation times are short (as shown on each panel in this figure). Projections of the binned samples (posterior densities) are shown in Figure~\ref{fig:exp1_posterior}. }\label{fig:trace}
\end{center}
\end{figure}
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp1_posterior.pdf}
\caption{ Projections of the posterior probability distribution over the four potential parameters ($q_1,q_z,\phi,\vhalo$) and Lagrange point offset ($\alpha$) assuming negligible uncertainties on the observed phase-space coordinates of eight stars and the progenitor, visualized as two-dimensional histograms of MCMC samples. Solid contours (black lines) show approximately 1$\sigma$ and 2$\sigma$ levels of the distribution. Vertical and horizontal lines (blue) show the true, input values for the potential parameters used in the N-body simulations. For the potential parameters, the axis ranges are chosen to be the same for this and the potential posterior plots to follow. }\label{fig:exp1_posterior}
\end{center}
\end{figure*}
\subsection{Data with near-future uncertainties}\label{sec:exp2}
We next take the same four stars used in the previous experiment and ``observe'' them with optimistic observational uncertainties. We require these stars to be RR Lyrae variables which are known to be excellent distance indicators via the mid-infrared period-luminosity relation \citep[as shown in, e.g.,][]{madore12}. Nearly 100 RR Lyrae associated with the Sgr stream and $\sim$30 associated with the Orphan stream will be observed with \project{Spitzer}~ as part of the SMASH survey \citep{smashprop} with expected fractional distance uncertainties around $\sim$2\%. These stars will also be included in the \project{Gaia}\, proper motion catalog: at a distance of $\sim$50~kpc, a typical RR Lyrae will have a tangential velocity error around $\sim$20~km/s, though our sample of test stars are all within 40~kpc.
For this experiment, we (1) assume the stars are RR Lyrae stars (e.g., bright distance indicators) such that the fractional distance uncertainty is $2\%$; (2) neglect the uncertainty in angular position, $l,b$ (for a typical RR Lyrae at 50~kpc this is $\sim$$10^{-7}$~deg for \project{Gaia}); (3) assume we can measure radial velocities to these stars with 5~km/s uncertainty; and (4) compute the sky-averaged \project{Gaia}\, proper-motion uncertainty for each star assuming an F0V spectral type using the \texttt{PyGaia}\footnote{\url{https://github.com/agabrown/PyGaia}} code and use this uncertainty for both components of proper motion. We further assume that we know the tail assignment for each star, $\beta$. We observe the position of the progenitor with the same observational uncertainties though in reality some coordinates will have even better measurements.
This experiment samples over the four potential parameters, the Lagrange point offset, four phase-space coordinate parameters for each star (16 total), and four phase-space coordinate parameters for the progenitor --- 25 parameters in total. We use an ensemble of 256 walkers and draw initial conditions for the coordinate parameters by sampling from Gaussian's centered on the ``observed'' value with variances specified by the observational uncertainties. Other parameters --- potential parameters and $\alpha$ --- are initialized by drawing from the priors summarized in Table~\ref{tbl:params}. We burn in the walkers for 50000 steps and run for 100000 inference steps, but thin the chains by taking every $\max(t_{\rm acor})$ sample, where $\max(t_{\rm acor}) \approx 4000$~steps. Figures~\ref{fig:exp2_potential}, \ref{fig:exp2_satellite}, \ref{fig:exp2_particle} show the marginalized posteriors for the potential parameters, satellite parameters, and parameters for a single particle. The uncertainties on the potential parameters are only 5-7\%
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp2_potential.pdf}
\caption{ Projections of the marginal posterior over the triaxial potential parameters for observed stars and progenitor with near-future uncertainties (Section~\ref{sec:exp2}). Axis ranges show the lower and upper bounds on the uniform priors over these parameters. }\label{fig:exp2_potential}
\end{center}
\end{figure*}
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp2_satellite.pdf}
\caption{ Projections of the marginal posterior over the progenitor parameters for observed stars and progenitor with near-future uncertainties (Section~\ref{sec:exp2}). }\label{fig:exp2_satellite}
\end{center}
\end{figure*}
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp2_particle.pdf}
\caption{ Projections of the marginal posterior over parameters for one of the stars for observed stars and progenitor with near-future uncertainties (Section~\ref{sec:exp2}). }\label{fig:exp2_particle}
\end{center}
\end{figure*}
\subsection{Precise distance measurements with missing proper motions}\label{sec:exp3}
The SMASH survey \citep{smashprop} will be completed within a year (early 2015), long before the final \project{Gaia}\, data release. Thus, we will soon have precise distance measurements for stars in the Sgr and Orphan streams, but poor or no proper motion constraints. SMASH also targets several RR Lyrae in the Sgr core, which will enable a high-precision distance measurement of the progenitor. Though proper motions have been measured for the Sgr progenitor, we now consider a case in which we have high-precision (2\%) distances to stream stars and the progenitor, 10~km/s radial velocity uncertainty, but missing proper motions for all stars and progenitor.
As with the previous experiment, this experiment includes 25 model parameters in total. We again use an ensemble of 256 walkers. Other parameters --- potential parameters and $\alpha$ --- are initialized by drawing from the priors summarized in Table~\ref{tbl:params}. We burn in the walkers for 50000 steps and run for 500000 inference steps. We again thin the chains by taking every $\max(t_{\rm acor})$ sample, but here the autocorrelation time is found to be very long, $\max(t_{\rm acor}) \approx 15000$~steps. Figures~\ref{fig:exp3_potential}, \ref{fig:exp3_satellite}, \ref{fig:exp3_particle} show the marginalized posteriors for the potential parameters, satellite parameters, and parameters for a single particle. Even a small sample of stars with precise distance measurements provide an enormous amount of information about the triaxiality of the potential. Fractional uncertainties on the potential parameters are $\sim$15\% and the initially missing proper-motions are recovered with $\sim$25\% uncertainties.
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp3_potential.pdf}
\caption{ Same as Figure~\ref{fig:exp2_potential} but for data with no proper motion measurements. }\label{fig:exp3_potential}
\end{center}
\end{figure*}
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp3_satellite.pdf}
\caption{ Same as Figure~\ref{fig:exp2_satellite} but for data with no proper motion measurements. }\label{fig:exp3_satellite}
\end{center}
\end{figure*}
\begin{figure*}[!h]
\begin{center}
\includegraphics[width=\textwidth]{exp3_particle.pdf}
\caption{ Same as Figure~\ref{fig:exp2_particle} but for data with no proper motion measurements. }\label{fig:exp3_particle}
\end{center}
\end{figure*}
\section{Discussion}\label{sec:discussion}
{\bf 1. Observational uncertainties:} Any method that uses tidal debris as a potential measure must also model the (significant) uncertainties on the kinematic measurements; \texttt{Rewinder}\ handles observational uncertainties and missing data dimensions by including the true 6D positions of stars and the progenitor as model parameters. In this article we have tested \texttt{Rewinder}\ using data of varied quality for a small sample of stars.
{\bf 2. Form of the potential:} No assumptions were made about the the form of the Galactic potential in deriving the likelihood for \texttt{Rewinder}. The simple experiments in this article infer potential parameters from the same static, analytic potential used in the N-body simulations that generated the fake data. However, in principle the potential could be time-dependent, clumpy, or have properties that vary with radius. The orbits of the stars and progenitor are directly integrated and only require a function that evaluates the acceleration due to the potential at a given position (and time). We have shown that with the correct form for the Galactic potential, \emph{precise} measurements of the potential parameters are attainable with near-future data, but we have not discussed the \emph{accuracy} of such measurements due to incorrect assumptions. For example, Bonaca et al. (in prep.) show that using a static potential model to fit the live potential of the \project{Via Lactea} halo with tidal streams can introduce significant biases to measurements of the halo mass, even with perfect knowledge of the 6D coordinates of stars in the streams. In future work, we will (1) try fitting incorrect potential models to see if we can still recover global properties (e.g., flattening); (2) run simulations with smoothly changing potentials \citep[e.g.,][]{buist14} and attempt to model the time dependence; and (3) explore using a generic, non-parametric potential form (e.g., a basis function expansion) as the recovery potential model.
{\bf 3. Multiple debris structures:} Each stream will provide constraints on different properties of the potential. For example, more eccentric streams may better constrain the radial profile \citep[see][who illustrate the power of using multiple streams to simultaneously constrain the potential using orbit fitting]{deg14}. In this article, we only consider a stream on a Sgr-like orbit, which lies nearly in the Galactic $x$-$z$ plane. We find that the stream puts better constraints on the $z$-axis flattening, $q_z$, than on the flattening in the $x$-$y$ plane, $q_1$, as one might expect. The best measurements of the detailed shape of the Galactic potential will come from combining constraints from multiple streams. We have defined above a proper likelihood function for observed stars in a given stream, thus incorporating multiple streams into the inference only requires multiplying the likelihoods computed for each stream and progenitor pair.
{\bf 4. Comparing models to data:} With \texttt{Rewinder}, each star adds constraints on the potential parameters and does not require matching a simulated density to the observed density of stars, thus this method works well even for small samples of well-measured stream stars. It might be that the most relevant stellar samples for inferring the Milky Way potential are small. For instance, stars that produce good distance estimates might be much more valuable than typical stars in the sample so that we may limit to only variable stars, e.g., RR Lyraes. These valuable stars are rare (and their abundances are age and metallicity dependent); there could be many cold structures in the Milky Way halo that are highly constraining on the potential in principle, but which contain only a few good distance-indicating members. It is not yet known what the trade-offs are between having many stars at low precision and a few at high precision, nor is it known how valuable distance information really is, when a structure contains many precisely observed members.
{\bf 5. Computational expense:} One point of concern for \texttt{Rewinder}\ is that the number of parameters scales with eight times the number of stars: each star has six coordinate parameters (the true position), the unbinding time, and the tail assignment. Computational constraints limit the sample size, but even so, the inference is much faster than full N-body modeling because the stars and progenitor are treated as test particles. Each step in parameter space with \texttt{Rewinder}\ requires integrating the orbits of stars and the progenitor: we presently integrate with a fixed time-step using leapfrog integration. Computing the acceleration due to the given potential at each step is implemented in \texttt{Cython} (and approaches \texttt{C}-like speed), but the rest of the code is written in pure-\texttt{Python}. Though already significantly less computationally intensive than running a full N-body simulation for every parameter step, running each MCMC walker for $\gtrsim$100000 steps requires parallelization and many hours of CPU time on a compute cluster. Further optimization of the integration method (e.g., using an adaptive method or implementing in \texttt{Cython}) could speed up the inference significantly.
\section{Conclusions}\label{sec:conclusion}
We have presented a probabilistic model (\texttt{Rewinder}) --- a likelihood function and with priors on the parameters --- for using stars observed in tidal streams to constrain properties of any underlying gravitational potential. \texttt{Rewinder}\ relies on direct orbit integration and not on computing conserved quantities (e.g., actions) and can thus be used with arbitrarily complex (e.g., time-dependent or clumpy) forms of the potential. We have performed several experiments to show that \texttt{Rewinder}\ simultaneously constrains the potential and models a tidal stream given simulated data with a range of realistic uncertainties. We find that with future high-quality data --- that is, high precision distance measurements from RR Lyrae and proper motions from \project{Gaia} --- a small sample of just four stars in a tidal stream modeled with \texttt{Rewinder}\, provide measurements of potential parameters that rival present-day constraints from comparing full N-body simulations to large numbers of stellar tracers with poorly measured kinematics. For this high-quality data, we recover the input potential parameter values with uncertainties of order 5-7 percent. Without proper motion data the uncertainties are around 15 percent. We consider this work to be an encouraging first step towards the goal of recovering the (presumably) much more complex --- time-dependent, clumpy, with axis ratios and orientations that vary with radius --- Milky Way potential with larger data sets.
\acknowledgements
We thank Anthony Brown (Leiden) for insight on the \project{Gaia}\, data quality and for providing the open source \texttt{PyGaia} code for computing predicted uncertainties. We thank the organizers of the Gaia Challenge (2013) and the University of Surrey for hospitality during the workshop. We thank Ana Bonaca (Yale), Jo Bovy (IAS), Dan Foreman-Mackey (NYU), Marla Geha (Yale), Andreas K{\"u}pper (Columbia), and Hans-Walter Rix (MPIA) for useful discussions.
KVJ thanks the Institute of Astronomy at the University of Cambridge for hospitality and Vasily Belokurov (IoA) and his group for discussion of his work while there.
APW is supported by a National Science Foundation Graduate Research Fellowship under Grant No.\ 11-44155. This work was supported in part by the National Science Foundation under Grant No. PHYS-1066293 and AST-1312196.
DWH was partially supported by the NSF (Grant IIS-1124794) and the Moore-Sloan Data Science Environment at NYU.
This research made use of Astropy, a community-developed core \texttt{Python} package for Astronomy \citep{astropy13}. This research has made use of NASA's Astrophysics Data System.
This work additionally relied on Columbia University's \emph{Hotfoot} and \emph{Yeti} compute clusters, and we acknowledge the Columbia HPC support staff for assistance. \\
|
\section*{Supporting Information}
\renewcommand{\thetable}{S.\arabic{table}}
\subsection*{Example template messages used for direct solicitation of phylogenetic datasets}
\subsubsection*{Undergraduate First Request}
Dear Dr. $<$\verb!corresponding author!$>$,
\medskip
\noindent
My name is $<$\verb!solicitor!$>$, and I'm an undergraduate researcher working in the Moore lab at UC Davis. We are doing a meta-analysis of density-dependent rates of lineage diversification. We would like to include your $<$\verb!publication year!$>$ study $<$\verb!study title!$>$ in our meta-analysis, and were hoping that you could send us any tree files that you used to perform the diversification-rate analysis in your study, as well as the alignment used to generate the study tree.
\noindent
Thank you very much for your time and consideration.
\medskip
\noindent
Sincerely,
$<$\verb!solicitor!$>$
\subsubsection*{Undergraduate Second Request}
Dear Dr. $<$\verb!corresponding author!$>$,
\medskip
\noindent
I emailed you a few days ago asking for your tree files and alignments from your $<$\verb!publication year!$>$ study, $<$\verb!study title!$>$, for inclusion in a meta-analysis of density-dependent lineage diversification. Since I have not yet heard back from you, I would like to reiterate my request for the data, which we would very much like to include.
\medskip
\noindent
Sincerely,
$<$\verb!solicitor!$>$
\subsubsection*{Undergraduate Third Request}
Dear Dr. $<$\verb!corresponding author!$>$,
\medskip
\noindent
I emailed you some time ago asking for your tree file and alignment from your $<$\verb!publication year!$>$ study, $<$\verb!study title!$>$, for inclusion in a meta-analysis of density-dependent lineage diversification. I would still appreciate it if you could send me the relevant data.
\medskip
\noindent
Sincerely,
$<$\verb!solicitor!$>$
\newpage
\subsection*{Example R script used to generate messages for direct-solicitation campaign}
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=6in]{R-script.pdf}
\end{center}
\end{figure}
\begin{table}
\caption{\bf{Summary of datasets and models.}} \label{tab:modeltable}
\begin{tabular}{lllp{8.8cm}}
\toprule
Name & Source & Data & Predictor variables \\
\midrule
$x_\text{a,a}$ & archived & alignments only & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding} \\
\rowcolor{gray!25}
$x_\text{a,t}$ & archived & trees only & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding} \\
$x_\text{a,e}$ & archived & alignments or trees & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding} \\
\rowcolor{gray!25}
$x_\text{a,b}$ & archived & alignments and trees & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding} \\
$x_\text{s,a}$ & solicited & alignments only & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership$^*$, NSF funding$^*$, undergraduate student, professor} \\
\rowcolor{gray!25}
$x_\text{s,t}$ & solicited & trees only & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, undergraduate student, professor} \\
$x_\text{s,e}$ & solicited & alignments or trees & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, undergraduate student, professor} \\
\rowcolor{gray!25}
$x_\text{s,b}$ & solicited & alignments and trees & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, undergraduate student, professor} \\
$x_\text{c,a}$ & combined & alignments only & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, solicited} \\
\rowcolor{gray!25}
$x_\text{c,t}$ & combined & trees only & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, solicited} \\
$x_\text{c,e}$ & combined & alignments or trees & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, solicited} \\
\rowcolor{gray!25}
$x_\text{c,b}$ & combined & alignments and trees & \emph{intercept, age, impact factor, no policy, strong policy, JDAP membership, NSF funding, solicited} \\
\end{tabular}\\
$^*$ -- This parameter was initially included in the model, but was later removed because it could not be reliably estimated. \\
\end{table}
\newpage
\subsection*{MCMC diagnosis}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of archived alignments or trees.}} \label{tab:mcmc_archive_any}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7501.00 & 7428.18 & 29931.18 & 0.36 & 0.07 & 0.72 & 0.64 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7770.81 & 7501.00 & 7770.65 & 6917.10 & 29959.56 & 0.74 & 0.37 & 0.98 & 0.64 & 1.00 \\
$\beta_\text{IF}$ & 7236.37 & 7501.00 & 7501.00 & 7501.00 & 29739.37 & 0.59 & 0.47 & 0.11 & 0.06 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 8372.49 & 7489.66 & 30864.15 & 0.47 & 0.01 & 0.68 & 0.32 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7220.13 & 7501.00 & 29723.13 & 0.75 & 0.55 & 0.57 & 0.54 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 7501.00 & 7501.00 & 6773.91 & 29276.91 & 0.16 & 0.18 & 0.83 & 0.84 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7964.70 & 7165.25 & 30131.94 & 0.71 & 0.61 & 0.55 & 0.08 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{a,e}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of archived alignments and trees.}} \label{tab:mcmc_archive_both}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7252.12 & 7501.00 & 29755.12 & 0.29 & 0.08 & 0.77 & 0.21 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7501.00 & 7501.00 & 7821.37 & 30324.37 & 0.26 & 0.11 & 0.14 & 0.86 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 6147.95 & 7501.00 & 7501.00 & 28650.95 & 0.23 & 0.73 & 0.03 & 0.50 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.15 & 0.04 & 0.29 & 0.95 & 1.00 \\
$\beta_\text{strong}$ & 7616.62 & 7501.00 & 7501.00 & 7501.00 & 30119.62 & 0.76 & 0.03 & 0.25 & 0.07 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 7501.00 & 7501.00 & 6887.44 & 29390.44 & 0.45 & 0.08 & 0.16 & 0.20 & 1.00 \\
$\beta_\text{NSF}$ & 7395.85 & 7501.00 & 7501.00 & 7501.00 & 29898.85 & 0.05 & 0.66 & 0.31 & 0.54 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{a,b}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of solicited alignments or trees.}} \label{tab:mcmc_sent_any}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.38 & 0.40 & 0.58 & 0.20 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.48 & 0.57 & 0.68 & 0.51 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.04 & 0.58 & 0.32 & 0.76 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.24 & 0.28 & 0.43 & 0.77 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.63 & 0.84 & 0.82 & 0.73 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 8065.92 & 7501.00 & 7501.00 & 30568.92 & 0.87 & 0.16 & 0.64 & 0.39 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7501.00 & 7383.98 & 29886.98 & 0.56 & 0.87 & 0.97 & 0.96 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{undergrad}$ & 7501.00 & 7009.73 & 7501.00 & 7501.00 & 29512.73 & 0.03 & 0.02 & 0.58 & 0.32 & 1.00 \\
$\beta_\text{prof}$ & 7501.00 & 7235.82 & 7501.00 & 8881.95 & 31119.77 & 0.74 & 0.68 & 0.51 & 0.04 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{s,e}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of solicited alignments and trees.}} \label{tab:mcmc_sent_both}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7221.13 & 8541.62 & 7501.00 & 30764.75 & 0.84 & 0.33 & 0.29 & 0.90 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7223.68 & 7501.00 & 7501.00 & 29726.68 & 0.20 & 0.98 & 0.82 & 0.41 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 8580.70 & 7562.06 & 31144.76 & 0.90 & 0.92 & 0.37 & 0.31 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.09 & 0.23 & 0.56 & 0.17 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 1.00 & 0.77 & 0.62 & 0.58 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.18 & 0.20 & 0.69 & 0.14 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.97 & 0.23 & 0.74 & 0.10 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{undergrad}$ & 7185.12 & 7501.00 & 7501.00 & 7501.00 & 29688.12 & 0.17 & 0.93 & 0.40 & 0.86 & 1.00 \\
$\beta_\text{prof}$ & 7256.96 & 7501.00 & 7501.00 & 7501.00 & 29759.96 & 0.28 & 0.64 & 0.02 & 0.04 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{s,b}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of archived alignments only.}} \label{tab:mcmc_archive_alignments}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7337.49 & 7830.03 & 30169.52 & 0.10 & 0.14 & 0.85 & 0.21 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 6993.23 & 7501.00 & 7501.00 & 7501.00 & 29496.23 & 0.24 & 0.85 & 0.92 & 0.62 & 1.00 \\
$\beta_\text{IF}$ & 8214.77 & 7501.00 & 7501.00 & 7501.00 & 30717.77 & 0.76 & 0.79 & 0.43 & 0.08 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.56 & 0.18 & 0.74 & 0.73 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.94 & 0.24 & 0.77 & 0.29 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 8255.15 & 7855.49 & 7501.00 & 31112.64 & 0.20 & 0.03 & 0.97 & 0.39 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.65 & 0.39 & 0.49 & 0.03 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{a,a}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of archived trees only.}} \label{tab:mcmc_archive_trees}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7501.00 & 7669.34 & 30172.34 & 0.66 & 0.46 & 0.67 & 0.44 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7092.89 & 7501.00 & 7501.00 & 29595.89 & 0.74 & 0.43 & 0.93 & 0.02 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 6660.60 & 7501.00 & 29163.60 & 0.83 & 0.47 & 0.85 & 0.68 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.37 & 0.34 & 0.17 & 0.43 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.42 & 0.73 & 0.68 & 0.34 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 7501.00 & 7501.00 & 7068.63 & 29571.63 & 0.96 & 0.63 & 0.33 & 0.66 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.61 & 0.82 & 0.63 & 0.26 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{a,t}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of solicited alignments only.}} \label{tab:mcmc_sent_alignment}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.95 & 0.31 & 0.85 & 0.98 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.78 & 0.38 & 0.73 & 0.28 & 1.00 \\
$\beta_\text{IF}$ & 7752.87 & 7788.54 & 7501.00 & 7501.00 & 30543.41 & 0.81 & 0.65 & 0.98 & 0.81 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7884.20 & 7501.00 & 7501.00 & 7501.00 & 30387.20 & 0.38 & 0.47 & 0.97 & 0.78 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7757.03 & 7501.00 & 8110.93 & 30869.96 & 0.55 & 0.35 & 0.62 & 0.66 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{undergrad}$ & 7501.00 & 7501.00 & 7501.00 & 8190.22 & 30693.22 & 0.77 & 0.27 & 0.67 & 0.73 & 1.00 \\
$\beta_\text{prof}$ & 7501.00 & 7501.00 & 7501.00 & 7857.31 & 30360.31 & 0.98 & 0.96 & 0.96 & 0.60 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{s,a}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of solicited trees only.}} \label{tab:mcmc_sent_trees}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.68 & 0.76 & 1.00 & 0.33 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.34 & 0.47 & 0.92 & 0.83 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.71 & 0.24 & 0.58 & 0.64 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7229.02 & 7501.00 & 7501.00 & 7806.79 & 30037.81 & 0.46 & 0.34 & 0.66 & 0.29 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.92 & 0.52 & 0.41 & 0.13 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 7501.00 & 7203.03 & 7501.00 & 29706.03 & 0.34 & 0.31 & 0.27 & 0.10 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7231.46 & 7501.00 & 29734.46 & 0.20 & 0.71 & 0.69 & 0.69 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{undergrad}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.60 & 0.98 & 0.58 & 0.84 & 1.00 \\
$\beta_\text{prof}$ & 7429.18 & 7501.00 & 7501.00 & 7501.00 & 29932.18 & 0.52 & 0.97 & 0.72 & 0.38 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{s,t}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of combined alignments only.}} \label{tab:mcmc_combined_alignments}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 8932.81 & 7501.00 & 7501.00 & 7501.00 & 31435.81 & 0.53 & 0.38 & 0.36 & 0.85 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7192.10 & 7501.00 & 7203.17 & 29397.27 & 0.39 & 0.61 & 0.07 & 0.36 & 1.00 \\
$\beta_\text{IF}$ & 7344.84 & 7501.00 & 7501.00 & 7501.00 & 29847.84 & 0.61 & 0.31 & 0.91 & 0.38 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7904.21 & 7501.00 & 7501.00 & 30407.21 & 0.47 & 0.30 & 0.24 & 0.11 & 1.00 \\
$\beta_\text{strong}$ & 8118.81 & 7122.63 & 7501.00 & 7795.53 & 30537.97 & 0.61 & 0.99 & 0.07 & 0.75 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7690.78 & 7501.00 & 7501.00 & 7501.00 & 30193.78 & 0.69 & 0.79 & 0.12 & 0.34 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.53 & 0.66 & 0.00 & 0.76 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{solicited}$ & 7501.00 & 7903.44 & 7501.00 & 7894.54 & 30799.98 & 0.72 & 0.10 & 0.05 & 0.05 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{c,a}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of combined trees only.}} \label{tab:mcmc_combined_trees}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7284.35 & 7501.00 & 7501.00 & 7012.23 & 29298.58 & 1.00 & 0.96 & 0.14 & 0.38 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 6638.97 & 7501.00 & 7501.00 & 7501.00 & 29141.97 & 0.97 & 0.26 & 0.02 & 0.69 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 7716.11 & 7401.50 & 30119.61 & 0.71 & 0.66 & 0.96 & 0.46 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7894.56 & 7501.00 & 7501.00 & 30397.56 & 0.74 & 0.36 & 0.98 & 0.81 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 6726.40 & 7501.00 & 29229.40 & 0.91 & 0.89 & 0.15 & 0.15 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7501.00 & 7501.00 & 8298.20 & 7233.11 & 30533.31 & 0.75 & 0.22 & 0.63 & 0.78 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.63 & 0.37 & 0.39 & 0.12 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{requested}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.27 & 0.66 & 0.23 & 0.14 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{c,t}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of combined alignments or trees.}} \label{tab:mcmc_combined_any}
\centering
\begin{tabular}{lcccccccccc}
\toprule
& \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
Parameter & run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 6261.86 & 7501.00 & 7501.00 & 6837.35 & 28101.21 & 0.25 & 0.42 & 0.85 & 0.45 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.10 & 0.79 & 0.26 & 0.29 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.90 & 0.43 & 0.27 & 0.76 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.09 & 0.39 & 0.71 & 0.36 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.00 & 0.18 & 0.67 & 0.78 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7521.48 & 7501.00 & 7501.00 & 7240.79 & 29764.27 & 0.08 & 0.57 & 0.13 & 0.40 & 1.00 \\
$\beta_\text{NSF}$ & 7730.56 & 7501.00 & 7501.00 & 7271.97 & 30004.53 & 0.34 & 0.85 & 0.40 & 0.76 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{requested}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.11 & 0.74 & 0.35 & 0.85 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{c,e}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\begin{table}[H]
\caption{\bf{MCMC performance for analyses of combined alignments and trees.}} \label{tab:mcmc_combined_both}
\centering
\begin{tabular}{lcccccccccc}
\toprule
Parameter & \multicolumn{5}{c}{Effective Sample Size} & \multicolumn{4}{c}{Geweke's Diagnostic ($p$-value)} & \\
\cmidrule(r){2-6} \cmidrule(r){7-10}
& run 1 & run 2 & run 3 & run 4 & combined & run 1 & run 2 & run 3 & run 4 & PSRF \\
\midrule
$\beta_\text{I}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.44 & 0.38 & 0.62 & 0.75 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.51 & 0.09 & 0.46 & 0.22 & 1.00 \\
$\beta_\text{IF}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.59 & 0.56 & 0.14 & 0.79 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 8623.45 & 7501.00 & 7501.00 & 7841.88 & 31467.33 & 0.77 & 0.72 & 0.44 & 0.91 & 1.00 \\
$\beta_\text{strong}$ & 7501.00 & 7501.00 & 7501.00 & 7501.00 & 30004.00 & 0.18 & 0.47 & 0.34 & 0.27 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 7918.18 & 7501.00 & 7501.00 & 7281.23 & 30201.41 & 0.16 & 0.35 & 0.78 & 0.65 & 1.00 \\
$\beta_\text{NSF}$ & 7501.00 & 7501.00 & 7772.42 & 8128.34 & 30902.75 & 0.71 & 0.29 & 0.28 & 0.23 & 1.00 \\
\rowcolor{gray!25}
$\beta_\text{requested}$ & 7758.05 & 7501.00 & 7240.75 & 7501.00 & 30000.80 & 0.45 & 0.89 & 0.72 & 0.75 & 1.00 \\
\end{tabular}
\begin{flushleft}N.B. This corresponds to analysis $x_{c,a}$ in Tables \ref{tab:modeltable} and \ref{tab:paramtable}.
\end{flushleft}
\end{table}
\newpage
\subsection*{Logistic model parameter estimates}
\begin{table}[H]
\caption{\bf{Parameter estimates based on archived data.}}
\begin{tabular}{l ccc ccc ccc ccc}
\toprule
& \multicolumn{3}{c}{alignments only} & \multicolumn{3}{c}{trees only} & \multicolumn{3}{c}{alignments or trees} & \multicolumn{3}{c}{alignments and trees} \\
\cmidrule(r){2-4} \cmidrule(r){5-7} \cmidrule(r){8-10} \cmidrule(r){11-13}
& & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} \\
\cmidrule(r){3-4} \cmidrule(r){6-7} \cmidrule(r){9-10} \cmidrule(r){12-13}
& mean & lower & upper & mean & lower & upper & mean & lower & upper & mean & lower & upper \\
\midrule
$\beta_\text{I}$ & -3.114 & -5.028 & -1.266 & -0.864 & -2.623 & 0.627 & -0.968 & -1.958 & 0.067 & -3.172 & -5.009 & -1.459 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & -0.033 & -0.069 & 0.002 & -0.080 & -0.139 & -0.026 & -0.038 & -0.060 & -0.016 & -0.019 & -0.043 & 0.003 \\
$\beta_\text{IF}$ & -0.018 & -0.127 & 0.084 & 0.035 & -0.068 & 0.141 & 0.036 & -0.028 & 0.097 & 0.045 & -0.028 & 0.118 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 0.650 & -1.357 & 2.761 & -0.627 & -2.615 & 1.310 & 0.234 & -0.886 & 1.386 & 0.783 & -1.274 & 2.883 \\
$\beta_\text{strong}$ & 1.001 & -1.023 & 3.293 & -0.082 & -2.323 & 1.903 & 1.075 & -0.044 & 2.284 & 1.722 & -0.087 & 3.712 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & 1.685 & -0.114 & 3.594 & 0.066 & -1.529 & 1.791 & 2.370 & 1.382 & 3.528 & 3.048 & 1.289 & 4.854 \\
$\beta_\text{NSF}$ & 1.183 & -0.045 & 2.446 & -0.259 & -1.769 & 1.290 & 0.498 & -0.355 & 1.393 & -0.143 & -1.234 & 0.928 \\
\label{tab:results_archive_alignments}
\end{tabular}
\end{table}
\begin{table}[H]
\caption{\bf{Parameter estimates based on solicited data.}}
\centering
\begin{tabular}{l ccc ccc ccc ccc}
\toprule
& \multicolumn{3}{c}{alignments only} & \multicolumn{3}{c}{trees only} & \multicolumn{3}{c}{alignments or trees} & \multicolumn{3}{c}{alignments and trees} \\
\cmidrule(r){2-4} \cmidrule(r){5-7} \cmidrule(r){8-10} \cmidrule(r){11-13}
& & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} \\
\cmidrule(r){3-4} \cmidrule(r){6-7} \cmidrule(r){9-10} \cmidrule(r){12-13}
& mean & lower & upper & mean & lower & upper & mean & lower & upper & mean & lower & upper \\
\midrule
$\beta_\text{I}$ & -6.248 & -9.587 & -3.130 & -1.577 & -3.096 & -0.248 & -0.328 & -1.225 & 0.548 & -1.708 & -2.780 & -0.619 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & 0.004 & -0.018 & 0.027 & 0.001 & -0.018 & 0.022 & -0.004 & -0.016 & 0.007 & -0.006 & -0.019 & 0.005 \\
$\beta_\text{IF}$ & -0.338 & -0.733 & 0.024 & 0.130 & 0.034 & 0.222 & 0.081 & -0.003 & 0.171 & 0.002 & -0.076 & 0.075 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & 0.653 & -1.530 & 2.787 & 0.578 & -0.953 & 2.062 & 0.281 & -0.599 & 1.127 & -0.089 & -1.013 & 0.805 \\
$\beta_\text{strong}$ & 1.555 & -0.622 & 3.810 & 0.209 & -1.466 & 1.876 & 0.573 & -0.425 & 1.499 & 0.311 & -0.661 & 1.303 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & NA & NA & NA & 0.622 & -1.617 & 2.863 & 0.027 & -1.424 & 1.347 & -0.095 & -1.609 & 1.469 \\
$\beta_\text{NSF}$ & NA & NA & NA & 0.220 & -1.538 & 1.907 & -0.482 & -1.590 & 0.663 & -0.612 & -1.901 & 0.556 \\
\rowcolor{gray!25}
$\beta_\text{undergrad}$ & 2.177 & -0.388 & 4.978 & -2.736 & -4.622 & -0.960 & 0.478 & -0.356 & 1.231 & 1.538 & 0.602 & 2.598 \\
$\beta_\text{prof}$ & 1.507 & -1.153 & 4.508 & -2.334 & -3.991 & -0.819 & 1.501 & 0.653 & 2.378 & 2.598 & 1.649 & 3.645 \\
\label{tab:results_sent_alignments}
\end{tabular}
\end{table}
\begin{table}[H]
\caption{\bf{Parameter estimates based on archived and solicited data.}}
\centering
\begin{tabular}{l ccc ccc ccc ccc}
\toprule
& \multicolumn{3}{c}{alignments only} & \multicolumn{3}{c}{trees only} & \multicolumn{3}{c}{alignments or trees} & \multicolumn{3}{c}{alignments and trees} \\
\cmidrule(r){2-4} \cmidrule(r){5-7} \cmidrule(r){8-10} \cmidrule(r){11-13}
& & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} & & \multicolumn{2}{c}{$95\%$ HPD} \\
\cmidrule(r){3-4} \cmidrule(r){6-7} \cmidrule(r){9-10} \cmidrule(r){12-13}
& mean & lower & upper & mean & lower & upper & mean & lower & upper & mean & lower & upper \\
\midrule
$\beta_\text{I}$ & -0.994 & -1.887 & -0.082 & 0.479 & -1.453 & 2.380 & 2.111 & 1.063 & 3.180 & -0.171 & -1.389 & 0.953 \\
\rowcolor{gray!25}
$\beta_\text{age}$ & -0.008 & -0.019 & 0.002 & -0.038 & -0.078 & 0.002 & -0.007 & -0.018 & 0.002 & 0.006 & -0.009 & 0.021 \\
$\beta_\text{IF}$ & -0.005 & -0.060 & 0.049 & 0.004 & -0.102 & 0.111 & 0.002 & -0.054 & 0.059 & -0.012 & -0.083 & 0.058 \\
\rowcolor{gray!25}
$\beta_\text{none}$ & -0.114 & -0.867 & 0.679 & -0.335 & -2.227 & 1.589 & 0.101 & -0.688 & 0.834 & 0.772 & -0.445 & 1.992 \\
$\beta_\text{strong}$ & 0.289 & -0.548 & 1.084 & -0.609 & -2.517 & 1.313 & 0.338 & -0.477 & 1.171 & 0.353 & -0.814 & 1.611 \\
\rowcolor{gray!25}
$\beta_\text{JDAP}$ & -0.033 & -1.066 & 0.916 & -1.334 & -3.174 & 0.506 & 0.431 & -0.656 & 1.513 & 1.238 & 0.010 & 2.467 \\
$\beta_\text{NSF}$ & 0.232 & -0.593 & 1.074 & -0.858 & -2.429 & 0.663 & -0.342 & -1.235 & 0.520 & 0.172 & -0.880 & 1.211 \\
\rowcolor{gray!25}
$\beta_\text{requested}$ & 0.690 & -0.098 & 1.438 & -3.634 & -5.544 & -1.835 & -2.040 & -2.948 & -1.132 & -3.162 & -4.099 & -2.239 \\
\label{tab:results_combined_alignments}
\end{tabular}
\end{table}
\newpage
\subsection*{Journal Policies}
\begin{table}[!ht]
\caption{\bf{Summary of journal policies.}}
\centering
\resizebox{\textwidth}{!}{%
\begin{tabular}{lll}
\toprule
Journal & Policy rating & Number of studies \\
\midrule
The American Naturalist & \emph{JDAP membership} & 5 \\
\rowcolor{gray!25}
Annals of Botany & \emph{weak policy} & 1 \\
Australian Systematic Botany & \emph{weak policy} & 2 \\
\rowcolor{gray!25}
Biological Journal of the Linnean Society$^\dagger$ & \emph{weak policy} & 4 \\
Biogeography & \emph{weak policy} & 15 \\
\rowcolor{gray!25}
BMC Biology & \emph{weak policy} & 3 \\
BMC Evolutionary Biology$^\dagger$ & \emph{no policy} & 12 \\
\rowcolor{gray!25}
BMC Plant Biology & \emph{weak policy} & 1 \\
Cladistics & \emph{weak policy} & 1 \\
\rowcolor{gray!25}
Copeia & \emph{weak policy} & 1 \\
Ecology$^\dagger$ & \emph{no policy} & 2 \\
\rowcolor{gray!25}
Ecology and Evolution & \emph{strong policy} & 2 \\
Ecology Letters & \emph{no policy} & 4 \\
\rowcolor{gray!25}
Evolution & \emph{JDAP membership} & 35 \\
Evolutionary Biology & \emph{weak policy} & 1 \\
\rowcolor{gray!25}
Hydrobiologia & \emph{no policy} & 1 \\
International Journal of Plant Sciences & \emph{no policy} & 1 \\
\rowcolor{gray!25}
International Journal of Microbial Ecology & \emph{strong policy} & 1 \\
Italian Journal of Zoology & \emph{no policy} & 1 \\
\rowcolor{gray!25}
Journal of Arid Environments & \emph{no policy} & 1 \\
Journal of Evolutionary Biology & \emph{JDAP membership} & 9 \\
\rowcolor{gray!25}
Journal of Heredity$^\dagger$ & \emph{strong policy} & 1 \\
Journal of Mammalogy & \emph{weak policy} & 1 \\
\rowcolor{gray!25}
Journal of Mammalian Evolution & \emph{no policy} & 1 \\
Journal of Theoretical Biology & \emph{no policy} & 1 \\
\rowcolor{gray!25}
Methods in Ecology and Evolution & \emph{strong policy} & 1 \\
Molecular Biology and Evolution$^\dagger$ & \emph{weak policy} & 2 \\
\rowcolor{gray!25}
Molecular Ecology & \emph{JDAP membership} & 8 \\
Molecular Phylogenetics and Evolution & \emph{no policy} & 31 \\
\rowcolor{gray!25}
Nature$^\dagger$ & \emph{weak policy} & 2 \\
New Phytologist & \emph{weak policy} & 1 \\
\rowcolor{gray!25}
PLOS ONE$^\dagger$ & \emph{weak policy} & 12 \\
PLOS Biology$^\dagger$ & \emph{weak policy} & 2 \\
\rowcolor{gray!25}
PeerJ & NA & 1 \\
Plant Systematics and Evolution & \emph{weak policy} & 2 \\
\rowcolor{gray!25}
PNAS & \emph{weak policy} & 11 \\
Perspectives in Plant Ecology, Evolution and Systematics & \emph{no policy} & 1 \\
\rowcolor{gray!25}
Proceedings of the Royal Society B & \emph{strong policy} & 14 \\
Science & \emph{strong policy} & 5 \\
\rowcolor{gray!25}
Systematic Entomology & \emph{no policy} & 2 \\
Systematic Biology & \emph{JDAP membership} & 13 \\
\rowcolor{gray!25}
Trends in Microbiology & \emph{no policy} & 1 \\
Virus Research & \emph{no policy} & 1 \\
\label{tab:journal_policies}
\end{tabular}}
\begin{flushleft} $^\dagger$Data-sharing policy in place at the time of study publication; has since become \emph{JDAP membership.}
\end{flushleft}
\end{table}
\bigskip \noindent \textbf{\emph{The American Naturalist} (strong policy, JDAP member)}
\noindent
The American Naturalist \emph{requires authors to deposit the data associated with accepted papers in a public archive}. For gene sequence data and phylogenetic trees, deposition in GenBank or TreeBASE, respectively, is required. There are many possible archives that may suit a particular data set, including the Dryad repository for ecological and evolutionary biology data (http://datadryad.org). All accession numbers for GenBank, TreeBASE, and Dryad must be included in accepted manuscripts before they go to Production. If the data are deposited somewhere else, please provide a link. If the data are culled from published literature, please deposit the collated data in Dryad for the convenience of your readers. Any impediments to data sharing should be brought to the attention of the editors at the time of submission so that appropriate arrangements can be worked out. For more, see the editorial on data.
\bigskip \noindent \textbf{\emph{Annals of Botany} (weak policy)}
\noindent
Before novel sequences for proteins or nucleotides can be published, authors are required to deposit their data with one of the principal databases comprising the International Nucleotide Sequence Database Collaboration: EMBL Nucleotide Sequence Database, GenBank, or the DNA Data Bank of Japan and to include an accession number in the paper.
\emph{Sequence matrices should only be included if alignment information is critical to the message of the paper}. Such matrices can be in colour but should not occupy more than one printed page. Larger matrices will only be printed by special agreement but may more readily be published electronically as Supplementary Information.
\bigskip \noindent \textbf{\emph{Australian Systematic Botany} (weak policy)}
\noindent
For all papers, whether presenting morphological, cytological or molecular data, voucher specimens must be cited, along with the herbarium where lodged.
\emph{All sequences used as data must be deposited in one of the international nucleotide sequence databases}, preferably GenBank, National Center for Biotechnology Information, 8600 Rockville Pike, Bethesda, MD 20894, USA. Email: <EMAIL>. Request information at <EMAIL>. Post-review final manuscript will not be accepted until sequence database accession numbers are included.
\bigskip \noindent \textbf{\emph{Biological Journal of the Linnean Society} (weak policy$^\dagger$)}
\noindent
\emph{Data that are integral to the paper must be made available in such a way as to enable readers to replicate, verify and build upon the conclusions published in the paper.} Any restriction on the availability of these data must be disclosed at the time of submission. Data may be included as part of the main article where practical. We \emph{recommend that data for which public repositories are widely used, and are accessible to all, should be deposited in such a repository prior to publication.} The appropriate linking details and identifier(s) should then be included in the publication and where possible the repository, to facilitate linking between the journal article and the data. If such a repository does not exist, data should be included as supporting information to the published paper or authors should agree to make their data available upon reasonable request.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{BMC Biology} (weak policy)}
\noindent
Submission of a manuscript to a BioMed Central journal implies that \emph{readily reproducible materials described in the manuscript, including all relevant raw data, will be freely available to any scientist wishing to use them} for non-commercial purposes.
Through a special arrangement with LabArchives, LLC, authors submitting manuscripts to BMC Biology can obtain a complimentary subscription to LabArchives with an allotment of 100MB of storage. LabArchives is an Electronic Laboratory Notebook which will enable scientists to share and publish data files in situ; you can then link your paper to these data. Data files linked to published articles are assigned digital object identifiers (DOIs) and will remain available in perpetuity. Use of LabArchives or similar data publishing services does not replace preexisting data deposition requirements, such as for nucleic acid sequences, protein sequences and atomic coordinates.
\emph{The Accession Numbers of any nucleic acid sequences, protein sequences or atomic coordinates cited in the manuscript should be provided,} in square brackets and include the corresponding database name
\bigskip \noindent \textbf{\emph{BMC Evolutionary Biology} (no policy$^\dagger$)}
\noindent
BMC Evolutionary Biology \emph{encourages} authors to deposit the data set(s) supporting the results reported in submitted manuscripts in a publicly-accessible data repository, when it is not possible to publish them as additional files. This section should only be included when supporting data are available and must include the name of the repository and the permanent identifier or accession number and persistent hyperlink(s) for the data set(s).
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{BMC Plant Biology} (weak policy)}
\noindent
\emph{The Accession Numbers of any nucleic acid sequences, protein sequences or atomic coordinates cited in the manuscript should be provided}, in square brackets and include the corresponding database name$\cdots$
The databases for which we can provide direct links are: EMBL Nucleotide Sequence Database (EMBL), DNA Data Bank of Japan (DDBJ), GenBank at the NCBI (GenBank), Protein Data Bank (PDB), Protein Information Resource (PIR) and the Swiss-Prot Protein Database (Swiss-Prot).
\emph{BMC Plant Biology encourages authors to deposit the data set(s) supporting the results} reported in submitted manuscripts \emph{in a publicly-accessible data repository,} when it is not possible to publish them as additional files. This section should only be included when supporting data are available and must include the name of the repository and the permanent identifier or accession number and persistent hyperlink(s) for the data set(s).
Through a special arrangement with LabArchives, LLC, authors submitting manuscripts to BMC Plant Biology can obtain a complimentary subscription to LabArchives with an allotment of 100MB of storage. LabArchives is an Electronic Laboratory Notebook which will enable scientists to share and publish data files in situ; you can then link your paper to these data. Data files linked to published articles are assigned digital object identifiers (DOIs) and will remain available in perpetuity. Use of LabArchives or similar data publishing services does not replace preexisting data deposition requirements, such as for nucleic acid sequences, protein sequences and atomic coordinates.
\bigskip \noindent \textbf{\emph{Cladistics} (weak policy)}
\noindent
\emph{Cladistics requests the deposition of data matrices} and other material electronically \emph{for publication on the Willi Hennig Society Journal web site.} Please submit these data as e-mail attachments to the Associate Editor Mark Siddall (<EMAIL>) after receiving the tracking number for your manuscript from the Editor. These data will be made available to the referees but not to the community at large until such time as the paper is accepted. If the paper is not found to be acceptable for publication, the data and associated files will be removed from the directory structure and destroyed.
\bigskip \noindent \textbf{\emph{Copeia} (weak policy)}
\noindent
\emph{Analyses based on molecular sequence data must cite the relevant GenBank accession numbers in the text.}
\bigskip \noindent \textbf{\emph{Ecology} (no policy$^\dagger$)}
\noindent
The editors and publisher expect authors to make the data underlying published articles available. Although \emph{public data availability is not strictly a requirement for manuscripts published in Ecology}, Ecological Applications and Ecosphere at this time, any information on materials, methods or data necessary to verify the conclusions of the research reported must be made available to the Subject-matter Editor upon request.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{Ecology and Evolution} (strong policy)}
\noindent
The Journal Ecology and Evolution \emph{requires, as a condition for publication, that data supporting the results in the paper should be archived in an appropriate public archive,} such as GenBank, TreeBASE, Dryad, the Knowledge Network for Biocomplexity or other suitable long-term and stable public repositories. Data are important products of the scientific enterprise, and they should be preserved and usable for decades in the future.
\bigskip \noindent \textbf{\emph{Ecology Letters} (no policy)}
\noindent
It is \emph{recommended that authors deposit the data supporting the results} in the paper in a publically accessible archive, such as Dryad (DataDryad.Org). Data are important products of scientific enterprise, and they should be preserved and remain usable in future decades. DNA sequences published in Ecology Letters should be deposited in the EMBL/GenBank/DDJB Nucleotide Sequence Databases. An accession number for each sequence must be included in the manuscript.
\bigskip \noindent \textbf{\emph{Evolution} (strong policy, JDAP membership)}
\noindent
As a condition for publication, Evolution \emph{requires that data used in the paper are archived. DNA sequence data must be submitted to GenBank and phylogenetic data to TreeBASE. Other types of data must be deposited in an appropriate public archive} such as Dryad, the NCEAS Data Repository, or as supplementary online material associated with the paper published in Evolution. \emph{The data should be given with sufficient detail that, together with the contents of the paper, they allow each result in the published paper to be recreated.} Authors may elect to have the data publicly available at time of publication, or, if the technology of the archive allows, may opt to embargo access to the data for a period up to a year after publication. Exceptions may be granted at the discretion of the Editor-in-Chief, especially for sensitive information such as the location of endangered species. Authors must state their intention to archive their data when they submit their manuscript and must confirm that this has been done before the manuscript is sent to press. If a repository is to be cited, the citation should include the sequence name and accession number, if available. The basic format for citing electronic resources is: Author's Last Name, First initial. Title of data package (e.g., Data from “Article name”). Data Repository Name, Data identifier (or DOI), address/URL. Please include on your title page the location of your data or where you intend to archive your data.
\bigskip \noindent \textbf{\emph{Evolutionary Bioinformatics Online} (weak policy)}
\noindent
Authors publishing in Libertas Academica journals must agree to \emph{make freely available to other academic researchers any of the cells, clones of cells, DNA, antibodies, or other material used in the research reported and not available from commercial suppliers.}
Use DNA Databank of Japan, European Molecular Biology Laboratory, or GenBank.
Please \emph{ensure that accession numbers of any nucleaic acid sequences, protein sequences or atomic coordinates cited in the manuscript are provided} in square brackets with the corresponding database name.
Generally it is possible to provide direct links to data hosted on these databases: EMBL Nucleotide Sequence Database, DNA Data Bank of Japan, GenBank at the NCBI, Protein Data Bank, Protein Information Resource and the SwissProt Protein Database.
\bigskip \noindent \textbf{\emph{Evolutionary Biology} (weak policy)}
\noindent
\emph{DNA sequences must be submitted to GenBank} (NCBI - National Center for Biotechnology Information, Bethesda, USA) \emph{or to the EMBL Nucleotide Sequence Data Base} (EBI - European Institute of Bioinformatics, Cambridge, UK) and \emph{accession numbers must be provided} when the paper is accepted.
\emph{Authors are encouraged to submit other data types to online data sharing resources if such resources are available.}
\bigskip \noindent \textbf{\emph{Hydrobiologia} (no policy)}
\bigskip \noindent \textbf{\emph{International Journal of Plant Sciences} (no policy)}
\noindent
\emph{Material that is not integral to the body of the article and that substantially lengthens the print version of the article} (e.g., genetic and character matrices, \emph{extended cladograms}, extended tables) \emph{should be desgignated as appendixes} (table A1, etc.) and thus appear in the electronic edition of IJPS only. Exception: \emph{If voucher material is presented in your manuscript, this should be listed in the first appendix} (appendix A1), and this will appear in the print version of IJPS. To prepare your accession data, provide an appendix title and a sentence-style row of headings for the data. For each taxon sampled, include specimen voucher information and/or gene accession numbers, separated by commas. To save space, the taxa should be run together in a paragraph.
\bigskip \noindent \textbf{\emph{ISME Journal} (strong policy)}
\noindent
The ISME Journal will \emph{only review and publish manuscripts if the authors agree to make all data that cannot be published in the journal itself} (e.g. novel nucleotide sequences, structural data, or data from large-scale gene expression experiments) \emph{freely available in one of the public databases} (see Submission to public databases below). \emph{Accession codes must be provided} at the time a revised manuscript is returned to the Editorial Office. To avoid delays in publication of the manuscript, \emph{we encourage authors to deposit relevant data in public databases prior to submission.}
\bigskip \noindent \textbf{\emph{Italian Journal of Zoology} (no policy)}
\bigskip \noindent \textbf{\emph{Journal of Arid Environments} (no policy)}
\noindent
Elsevier \emph{accepts electronic supplementary material to support and enhance} your scientific research. Supplementary files offer the author additional possibilities to publish supporting applications, high-resolution images, background datasets, sound clips and more. Supplementary files supplied will be published online alongside the electronic version of your article in Elsevier Web products, including ScienceDirect.
Electronic archiving of supplementary data enables readers to replicate, verify and build upon the conclusions published in your paper. We \emph{recommend that data should be deposited} in the data library PANGAEA
\bigskip \noindent \textbf{\emph{Journal of Biogeography} (weak policy)}
\noindent
Consistent with widely adopted conventions in the field, it is a condition of publication that \emph{papers using new molecular sequences must place the sequences in an appropriate database} (e.g. GenBank). Relevant accession numbers should be provided in the final manuscript. \emph{Accession numbers are required for all sequences used in analyses,} including existing sequences in databases. Museum voucher numbers should also be provided where doing so constitutes the appropriate best practice and/or where this information could be of real value to future researchers. More generally, the journal recognizes that what is considered appropriate best practice regarding data publication/deposition may vary depending on factors such as the nature of the data, the funding sources involved, complexities of prior intellectual ownership issues, etc. We therefore \emph{strongly encourage (where appropriate) but do not require (where it may not be) authors to publish/deposit data sets} in conjunction with papers being published in this journal.
Authors who wish to provide a consolidated statement of how other readers can access the data used in their paper \emph{may wish to refer to outside data repositories} where they have deposited their data, e.g. Dryad, Pangaea, or others. If so, this statement should be included after the Supporting Information section and before the Biosketch entry.
\emph{Additional materials and results (including supporting tables and figures) that are necessary but do not need to be included in the main paper must be compiled into Appendices}, which will be provided to readers as online Supporting Information. No more than three supplementary appendices are permitted (labelled Appendix S1 to Appendix S3).
\bigskip \noindent \textbf{\emph{Journal of Evolutionary Biology} (strong policy, JDAP membership)}
\noindent
Submission of a manuscript to a BioMed Central journal implies that readily reproducible materials described in the manuscript, including \emph{all relevant raw data, will be freely available to any scientist wishing to use them for non-commercial purposes.}
The Journal of Evolutionary Biology requires, as a condition for publication, that \emph{data supporting the results in the paper should be archived in an appropriate public archive,} such as GenBank, TreeBASE, Dryad, the Knowledge Network for Biocomplexity or other suitable long-term and stable public repositories. Data are important products of the scientific enterprise, and they should be preserved and usable for decades in the future.
All accepted papers should \emph{provide accession numbers or DOI} for data underlying the work that have been deposited, so that these can appear in the final accepted article.
\bigskip \noindent \textbf{\emph{Journal of Heredity} (strong policy$^\dagger$)}
\noindent
The primary data underlying the conclusions of an article are critical to the verifiability and transparency of the scientific enterprise, and \emph{should be preserved in usable form for decades in the future.} For this reason, Journal of Heredity has \emph{adopted the Joint Data Archiving Policy} (JDAP)(See Editorial, J Hered (2013) 104 (1):1. doi: 10.1093/jhered/ess137). \emph{The public archiving of all primary data is a requirement} of publication in Journal of Heredity. For data other than nucleotide and protein sequences submitted to GenBank, \emph{suitable archives include Dryad, TreeBASE} or the Knowledge Network for Biocomplexity. \emph{For accepted articles, JHered covers the cost to archive the data in Dryad.}
Authors may elect to have the data publicly available at time of publication, or may opt to embargo access to the data for a period up to a year after publication. Exceptions may be granted at the discretion of the Editor, especially for sensitive information, such as the location of endangered species.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{Journal of Mammalian Evolution} (no policy)}
\noindent
Authors submitting analyses of gene or amino acid sequences should be prepared to supply electronic versions of their sequence alignments \emph{if requested by reviewers.} These are for confidential examination by reviewers only, and reviewers are requested not to use these alignments for any purpose other than the review process.
If the reported research includes examination of \emph{voucher specimens (including tissues), the Museum catalogue number or tissue number if the former is not available must be provided.} If sequences from Genbank are used, the \emph{Museum catalogue number} listed on Genbank must be listed, as well as the Genbank accession number. Genbank accession numbers can be used alone, only in the event that the Museum catalogue number is not available through the Genbank record.
\bigskip \noindent \textbf{\emph{Journal of Mammology} (weak policy)}
\noindent
\emph{All DNA sequences must be submitted to GenBank, and accession numbers provided in the manuscript before publication.}
Supplemental files will be posted online-only and provides information that adds depth to a manuscript but is not essential to a reader’s understanding of the research (e.g., spreadsheets, databases, equations, video or audio files, tables and/or figures).
\bigskip \noindent \textbf{\emph{Journal of Theoretical Biology} (no policy)}
\noindent
Elsevier \emph{encourages authors to connect articles with external databases}, giving their readers one-click access to relevant databases that help to build a better understanding of the described research. Please refer to relevant database identifiers using the following format in your article
Elsevier \emph{accepts electronic supplementary material} to support and enhance your scientific research. Supplementary files offer the author additional possibilities to publish supporting applications, high-resolution images, background datasets, sound clips and more. Supplementary files supplied will be published online alongside the electronic version of your article in Elsevier Web products, including ScienceDirect
\bigskip \noindent \textbf{\emph{Methods in Ecology and Evolution} (strong policy)}
\noindent
Data are important products of the scientific enterprise, and they should be preserved and usable for decades in the future. The British Ecological Society thus \emph{expects that data} (or, for theoretical papers, mathematical and computer models) \emph{supporting the results in Methods in Ecology and Evolution papers will be archived in an appropriate public archive, such as Dryad, TreeBASE,} NERC data centre, GenBank, or another archive of the author's choice that provides comparable access and guarantee of preservation.
\emph{Sufficient details should be archived} so that a third party can reasonably interpret those data correctly, \emph{to allow each result in the published paper to be recreated and the analyses reported in the paper to be replicated} to support the conclusions made. \emph{Authors are welcome to archive more than this, but not less.}
If data have been previously archived then they should not be archived again. \emph{The original archive DOI or reference should be used as the source of the data.}
\bigskip \noindent \textbf{\emph{Molecular Biology and Evolution} (weak policy$^\dagger$)}
\noindent
Among the \emph{requirements for publication} in MBE is that \emph{authors make publicly available, free of charge, any alignment data}, strains, cell lines, or clones used in reported experiments, computer code essential to the analysis, \emph{and any other material or information necessary for the assessment and verification of findings} or interpretations presented in the publication.
\emph{Newly reported sequences must be deposited} in the DDBJ/EMBL/GenBank database (see below). \emph{Accession numbers must be included in the final version of the manuscript} and cannot be added at the proof stage.
Newly reported nucleic acid and amino acid sequences, microarray data, structural coordinates, and all other essential information must be submitted to appropriate public databases (e.g., GenBank; the EMBL Nucleotide Sequence Database; DNA Database of Japan; the Protein Data Bank; Swiss-Prot; GEO; and Array-Express).
\emph{When appropriate, material such as sequence alignments} and large tables \emph{can be published online as supplementary material} permanently linked to an article in the online journal.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{Molecular Ecology} (strong policy, JDAP membership)}
\noindent
Molecular Ecology expects that \emph{data supporting the results in the paper should be archived in an appropriate public archive,} such as \emph{GenBank}, Gene Expression Omnibus, \emph{TreeBASE, Dryad,} the Knowledge Network for Biocomplexity, your own institutional or funder repository, or as Supporting Information on the Molecular Ecology web site. Data are important products of the scientific enterprise, and they should be preserved and usable for decades in the future.
\emph{Authors are expected to archive the data supporting their results and conclusions along with sufficient details so that a third party can interpret them correctly.}
To enable readers to locate archived data from Molecular Ecology papers, \emph{we require that authors include a ‘Data Accessibility’ section} after the references (see below for details). This section must be present at initial submission.
\bigskip \noindent \textbf{\emph{Molecular Phylogenetics and Evolution} (no policy)}
\noindent
Elsevier \emph{encourages authors to connect articles with external databases,} giving their readers one-click access to relevant databases that help to build a better understanding of the described research.
Elsevier \emph{accepts} electronic supplementary \emph{material to support and enhance} your scientific research. Supplementary files offer the author additional possibilities to publish supporting applications, high-resolution images, background datasets, sound clips and more. Supplementary files supplied will be published online alongside the electronic version of your article in Elsevier Web products, including ScienceDirect: http://www.sciencedirect.com.
You can \emph{enrich} your online articles \emph{by providing phylogenetic tree data files (optional)} in Newick or NeXML format, which will be visualized using the interactive tree viewer embedded within the online article. Using the viewer it will be possible to zoom into certain tree areas, change the tree layout, search within the tree, and collapse/expand tree nodes and branches. Submitted tree files will also be available for downloading from your online article on ScienceDirect. Each tree must be contained in an individual data file before being uploaded separately to the online submission system, via the 'phylogenetic tree data' submission category. Newick files must have the extension .new or .nwk (note that a semicolon is needed to end the tree). Please do not enclose comments in Newick files and also delete any artificial line breaks within the tree data because these will stop the tree from showing. For NeXML, the file extension should be .xml. Please do not enclose comments in the file. Tree data submitted with other file extensions will not be processed. Please make sure that you validate your Newick/NeXML files prior to submission. For more information please see http://www.elsevier.com/phylogenetictrees.
\bigskip \noindent \textbf{\emph{Nature} (weak policy$^\dagger$)}
\noindent
Therefore, a condition of publication in a Nature journal is that \emph{authors are required to make materials, data and associated protocols promptly available to readers} without undue qualifications.
The preferred way to share large data sets is via public repositories.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{New Phytologist} (weak policy)}
\bigskip \noindent \textbf{\emph{Perspectives in Plant Ecology, Evolution and Systematics} (no policy)}
\noindent
Elsevier \emph{accepts electronic supplementary material to support and enhance} your scientific research. Supplementary files offer the author additional possibilities to publish supporting applications, high-resolution images, background datasets, sound clips and more.
You can \emph{enrich your online articles} by providing phylogenetic tree data files (optional) in Newick or NeXML format, which will be visualized using the interactive tree viewer embedded within the online article.
Electronic archiving of supplementary data enables readers to replicate, verify and build upon the conclusions published in your paper. We\emph{ recommend that data should be deposited} in the data library PANGAEA (http://www.pangaea.de).
\bigskip \noindent \textbf{\emph{Plant Systematics and Evolution} (weak policy)}
\noindent
Data matrices \emph{including sequence alignments must be made available to the public.} There must be a sentence included in the Materials and methods section that such information is available from the corresponding author. \emph{“DNA or proteine sequences must be deposited in public data bases} (GenBank, EMBL, etc.) before the revised version is sent to the editor.“
\bigskip \noindent \textbf{\emph{PLOS ONE} (weak policy$^\dagger$)}
\noindent
Manuscripts reporting paleontology and archaeology research must include descriptions of methods and specimens in sufficient detail to allow the work to be reproduced. \emph{Data sets supporting statistical and phylogenetic analyses should be provided, preferably in a format that allows easy re-use.}
\emph{Specimen numbers and complete repository information,} including museum name and geographic location, \emph{are required for publication.} Locality information should be provided in the manuscript as legally allowable, or a statement should be included giving details of the availability of such information to qualified researchers.
\emph{Methods sections of papers with data that should be deposited} in a publicly available database \emph{should specify where the data have been deposited} and provide the relevant accession numbers and version numbers, if appropriate. Accession numbers should be provided in parentheses after the entity on first use. If the accession numbers have not yet been obtained at the time of submission, please state that they will be provided during review. They must be provided prior to publication.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{PLOS Biology} (weak policy$^\dagger$)}
\noindent
The results section should provide details of all of the experiments that are required to support the conclusions of the paper.
\emph{Large datasets, including raw data, should be submitted as supplemental files;} these are published online alongside the accepted article.
\noindent
NB: $^\dagger$Data-sharing policy in place at the time of study publication, which has since become \emph{JDAP membership.}
\bigskip \noindent \textbf{\emph{Proceedings of the National Academy of Sciences} (weak policy)}
\noindent
To allow others to replicate and build on work published in PNAS, \emph{authors must make materials, data, and associated protocols available to readers}. Authors must disclose upon submission of the manuscript any restrictions on the availability of materials or information. Data not shown and personal communications cannot be used to support claims in the work. Authors are encouraged to use SI to show all neces- sary data. Authors are encouraged to deposit as much of their data as possible in publicly accessible databases.
Before publication, \emph{authors must deposit large datasets (including microarray data, protein or nucleic acid sequences, and atomic coordinates for macromolecular structures) in an approved database} and provide an accession number for inclusion in the published paper. \emph{When no public repository exists, authors must provide the data as SI online} or, in special circumstances when this is not possible, on the author’s institutional Web site, provided that a copy of the data is provided to PNAS.
\emph{Authors must deposit data in a publicly available database} such as GenBank, EMBL, DNA Data Bank of Japan, UniProtKB/Swiss-Prot, or PRIDE.
\bigskip \noindent \textbf{\emph{Proceedings of the Royal Academy B: Biological Sciences} (strong policy)}
\noindent
To allow others to verify and build on the work published in Royal Society journals \emph{it is a condition of publication that authors make available the data and research materials supporting the results in the article,} as detailed in our Publishing policies; with which our authors are asked to comply.
\emph{Datasets should be deposited in an appropriate, recognized repository and the associated accession number, link or DOI to the datasets must be included in the methods section of the article.} Reference(s) to datasets should also be included in the reference list of the article with DOIs (where available). Where no discipline-specific data repository exists authors should deposit their datasets in a general repository such as Dryad (http://datadryad.org/)
Where possible any other relevant research materials (such as statistical tools, protocols, software etc) should also be made available and details of how they may be obtained should be included in the data accessibility section of the article.
To ensure archived data from Proceedings B articles are available to readers, \emph{authors should include a ‘data accessibility’ section} immediately after the acknowledgements. This should list the database and accession number for all data from the article that has been made publicly available.
\bigskip \noindent \textbf{\emph{Science} (strong policy)}
\noindent
\emph{All data necessary to understand, assess, and extend the conclusions of the manuscript must be available to any reader of Science.} All computer codes involved in the creation or analysis of data must also be available to any reader of Science. \emph{After publication, all reasonable requests for data and materials must be fulfilled.}
Science supports the efforts of databases that aggregate published data for the use of the scientific community. Therefore, \emph{appropriate data sets (including microarray data, protein or DNA sequences, atomic coordinates or electron microscopy maps for macromolecular structures, and climate data) must be deposited in an approved database, and an accession number or a specific access address must be included in the published paper}. We encourage compliance with MIBBI guidelines (Minimum Information for Biological and Biomedical Investigations).
\emph{Large data sets with no appropriate approved repository must be housed as supplementary materials} at Science, or only when this is not possible, on an archived institutional Web site, provided a copy of the data is held in escrow at Science to ensure availability to readers.
\bigskip \noindent \textbf{\emph{Systematic Biology} (strong policy)}
\noindent
\emph{All nucleotide sequence data and alignments must be submitted to GenBank or EMBL} before the paper can be published. In addition, \emph{all data matrices and resulting trees must be submitted to TreeBASE.} GenBank and TreeBASE \emph{reference numbers should be provided in the final version of the paper.}
\bigskip \noindent \textbf{\emph{Systematic Entomology} (no policy)}
\noindent
Paper submissions will be accepted exceptionally, although any relevant data matrices should be electronic.
\bigskip \noindent \textbf{\emph{Trends in Microbiology} (no policy)}
\noindent
A Trends in Microbiology Reviews \emph{must not include unpublished data, simulations or meta-analyses,} or propose a new formal mathematical model.
A Trends in Microbiology Opinion \emph{must not include unpublished data, simulations or meta-analyses,} or propose a new formal mathematical model.
Science \& Society articles \emph{should not include unpublished data, simulations or meta-analyses.}
Letters should \emph{not be used as an opportunity} to promote your own work, \emph{to introduce new data,} to provide an 'update' on a recent review, or to highlight perceived omissions in our articles.
\bigskip \noindent \textbf{\emph{Virus Research} (no policy)}
\noindent
Elsevier \emph{encourages authors to connect articles with external databases,} giving their readers one-click access to relevant databases that help to build a better understanding of the described research.
Elsevier \emph{accepts electronic supplementary material to support and enhance your scientific research.} Supplementary files offer the author additional possibilities to publish supporting applications, high-resolution images, background datasets, sound clips and more. Supplementary files supplied will be published online alongside the electronic version of your article in Elsevier Web products, including ScienceDirec
\bigskip \noindent \textbf{\emph{Zoological Journal of the Linnean Society} (no policy)}
\noindent
\emph{Avoid elaborate tables of original or derived data}, long lists of species, etc.; \emph{if} such data are \emph{absolutely essential,} consider including them as appendices or as online-only supplementary material.
\section*{Abstract}
The scientific enterprise depends critically on the preservation of and open access to published data.
This basic tenet applies acutely to phylogenies (estimates of evolutionary relationships among species). Increasingly, phylogenies are estimated from increasingly large, genome-scale datasets using increasingly complex statistical methods that require increasing levels of expertise and computational investment.
Moreover, the resulting phylogenetic data provide an explicit historical perspective that critically informs research in a vast and growing number of scientific disciplines.
One such use is the study of changes in rates of lineage diversification (speciation -- extinction) through time.
As part of a meta-analysis in this area, we sought to collect phylogenetic data (comprising nucleotide sequence alignment and tree files) from $217$ studies published in $46$ journals over a $13$-year period.
We document our attempts to procure those data (from online archives and by direct request to corresponding authors), and report results of analyses (using Bayesian logistic regression) to assess the impact of various factors on the success of our efforts.
Overall, complete phylogenetic data for $\sim 60\%$ of these studies are effectively lost to science.
Our study indicates that phylogenetic data are more likely to be deposited in online archives and/or shared upon request when: (1) the publishing journal has a strong data-sharing policy; (2) the publishing journal has a higher impact factor, and; (3) the data are requested from faculty rather than students.
Although the situation appears dire, our analyses suggest that it is far from hopeless: recent initiatives by the scientific community---including policy changes by journals and funding agencies---are improving the state of affairs.
\newpage
\section*{Introduction}
Archiving and sharing published data is a social contract that is integral to the scientific enterprise \citep[][]{Vision_10}.
Sharing published data advances the scientific process by:
(1) exposing published results to independent verification (to identify errors and discourage fraud);
(2) providing the pedagogical material for educating students and training future researchers;
(3) acting as a test bed to guide the development of new methods, and;
(4) providing a basis to identify and pursue new questions via synthesis/meta-analysis \citep[][]{Whitlock_11}.
Additionally, archiving published data protects our scientific investment, avoiding
needless costs of data regeneration in terms of time, money, and environmental impact \citep[][]{Piwowar_11b}.
These considerations are particularly germane to phylogenetic data, which include both alignments (estimates of the positional homology of molecular sequences) and phylogenetic trees (estimates of the evolutionary relationships among species).
Phylogenetic trees for individual groups are inherently synthetic---combination of these `twigs' provides a natural approach for elucidating the entire Tree of Life \citep[{\it c.f.},][]{Maddison_07, opentree_14}.
Additionally, phylogenetic data have tremendous potential for reuse, often in ways that were completely unanticipated by the original studies: because they provide an explicit evolutionary perspective, phylogenies have become central to virtually all areas of research in evolutionary biology, ecology, molecular biology and epidemiology \citep[][]{Donoghue_2000, Piwowar_11, Stoltzfus_2012}.
Moreover, the generation of phylogenetic data is an increasingly arduous and technical enterprise.
Clearly, phylogenetic data are a precious scientific resource that must be preserved and shared in order to realize their full potential.
The vast majority of phylogenies are estimated from molecular (primarily nucleotide) sequence data.
Although GenBank and similar public archives provide a robust \citep[albeit imperfect;][]{Noor_06} backstop against the complete loss of the \emph{raw} sequence data, these databases do not safeguard the associated \emph{phylogenetic} data: the alignments estimated from raw sequence data, and the trees inferred from those alignments.
Multiple sequence alignment---the process of estimating the positional homology of each nucleotide site comprising DNA sequences---is a difficult inference problem for which many approaches have been proposed \citep[][]{Notredame_2007, Thompson_2011}.
Different algorithms (or different settings for a given algorithm) may yield dramatically different estimates of the alignment that, in turn, can substantially impact estimates of phylogeny \citep[][]{Wong_2008, Blackburne_2013}.
Moreover, the majority of phylogenetic studies are based on alignments that are subjected to `manual adjustment' after being estimated using formal methods \citep[][]{Morrison_2009}, which effectively destroys the possibility of replicating published alignments from the corresponding raw sequence data.
Even if the alignment could be dependably reproduced, replicating the published phylogeny requires a precise description of how the phylogenetic analysis was performed, details that are typically not provided in phylogenetic studies \citep[][]{Leebens-Mack_06}.
Finally, even if the alignment and details of the analysis were available, re-generating the phylogeny remains a non-trivial proposition: the analysis of a single dataset may require hundreds or thousands of compute hours \citep[][]{Suchard_2009}.
These issues have been appreciated for some time \citep[][]{Sanderson_93}, and motivated the development of a specialized online archive for phylogenetic data, TreeBASE \citep[][]{Sanderson_94}, more than 20 years ago.
Despite such noble efforts, it is increasingly evident that the loss of phylogenetic data is catastrophic:
recent surveys estimate that $\sim 70\%$ of published phylogenetic data are lost forever \citep[][]{Drew_2013a, Drew_2013, Stoltzfus_2012}.
In response to this crisis, several recent community initiatives have been proposed to encourage the preservation and sharing of phylogenetic data.
These include policy initiatives both by funding agencies (the NSF Data Management Plan established in 2011 that requires the preservation of data generated by funded research), and by journals/publishers \citep[the establishment of the Joint Data Archiving Policy, JDAP, by a consortium of prominent journals requiring the submission of data to online archives as a condition of publication;][]{Moore_2010, Whitlock_2010, Rausher_2010, Rieseberg_2010, Uyenoyama_2010},
and the establishment of a new online archive for evolutionary and ecological data, Dryad \citep[][]{Dryad_11}.
We set out to perform a meta-analysis exploring the empirical prevalence of temporal changes in rates of lineage diversification.
To this end, we sought to collect the phylogenetic data from studies using the two most common statistical phylogenetic approaches for detecting temporal shifts in diversification rate; {\it i.e.}, the `gamma' statistic \citep[][`method 1']{Pybus_00} and the `birth-death likelihood' \citep[][`method 2']{rabosky_06} methods.
To be included in our meta-analysis, we required two key data files from each published empirical study: (1) an alignment of nucleotide sequence data, and (2) an ultrametric tree (where the branch lengths are rendered proportional to relative or absolute time).
We document our attempts to procure these data (both via searches of online archives and by direct solicitation from the corresponding authors), and describe results of analyses exploring various factors associated with the availability of phylogenetic data.
We assess a number of correlates---the age of the study, the impact factor and data-sharing policy of the publishing journal, the status of the solicitor, etc.---with a focus on revealing the efficacy of recent community initiatives to ensure the preservation and promote the sharing of published phylogenetic data.
\section*{Methods}
In this section, we document our attempts to procure phylogenetic data from a large and random sample of studies exploring temporal variation in rates of lineage diversification published over a $13$-year period.
We first describe how we sought to collect these data, and then describe the analyses we performed to gauge the success of our efforts.
\subsection*{Data Collection}
During the months of October and November, 2013, we searched for articles citing the two methods papers
using the the ISI Web of Science cited-reference tool.
Our search identified a total of $470$ citing articles ($322$ and $148$ for methods $1$ and $2$, respectively).
Of these, a total of $217$ involved empirical analyses ($165$ and $52$ using methods $1$ and $2$, respectively).
For each study, we captured bibliometric data on authorship, publication month and year, and the name and impact factor of the publishing journal.
We also recorded the data-sharing policy of the publishing journal and whether it was a member of the JDAP initiative at the time of publication.
Specifically, we ascertained the data-sharing policy for each of the $46$ journals from the corresponding `instructions to authors' documentation (see Supporting Information).
Following \citep[][]{Piwowar_10}, we categorized journals that made \emph{no mention} of data sharing as having \emph{no policy}; those that \emph{encouraged} authors to share data upon publication were scored as having a \emph{weak policy}; those that \emph{required} data sharing as a condition of publication were scored as having a \emph{strong policy}; and those that were members of the JDAP initiative were scored as having \emph{JDAP membership}.
Finally, we noted whether the studies acknowledged funding support from the National Science Foundation (NSF).
For each study, we assessed whether data were available online by first searching each article for various keywords (``Dryad'', ``TreeBASE'', etc.), and pursued any links or references to archived data.
If data could not be sourced directly from the article itself, we proceeded to examine any associated Supplemental Material files using a similar strategy.
Articles that did not submit their data to online repositories were targeted for direct solicitation using a semi-automated, multi-step approach (Figure \ref{flow_chart}).
Specifically, we wrote `templates' for three sequential messages comprising an initial, a followup, and a final request for published phylogenetic data.
In the messages, we identified ourselves, provided details of the requested data, and explained the reason for our request;
that is, we explained that we were gathering data for a meta-analysis evaluating the prevalence of temporal changes in diversification rate, and we sought the sequence alignment and ultrametric tree files that were the used to assess temporal changes in diversification rates in the published study.
Each of the three message templates contained `fields' for several variables, including: the name and status of the solicitor; the name and email address of the corresponding author; and the year and title of the published article.
We divided the solicitations evenly (and randomly) between the three of us.
This was intended both to share the burden equably, and also to assess any effect of the solicitor status, which comprised a professor (BRM), a graduate student (MRM) and an undergraduate student (AFM).
We then generated messages using \verb!R! scripts that populated the fields of the templates with the relevant information from the spreadsheet
(the templates and \verb!R! scripts are provided as Supporting Information).
Messages were sent at weekly intervals.
If we received a response, the corresponding author was precluded from receiving subsequent generic email messages, and we corresponded with them on an individual basis.
We recorded various details of each response, including whether the recipient sent the requested alignment file and/or tree file.
Datasets not obtained at the end of this process were deemed unavailable.
A table summarizing information gathered for the $217$ studies is included as Supporting Information.
Following \citep[][]{Wicherts_11}, the data table has been anonymized to protect the identity of corresponding authors ({\it i.e.}, with regard to who did or did not archive and/or share phylogenetic data from published studies).
However, a key is available upon request to allow details of our analyses to be independently verified.
In any case, the issues that we document are general and should not be use to impugn the academic integrity of the individual researchers.
\subsection*{Data Analysis}
We used Bayesian logistic regression to explore correlations between data availability and several variables.
Under this approach, a \emph{trial} is an attempt to recover data for a particular study either from online archives or by direct solicitation, which we deem a \emph{success} if we received data for that study.
The outcomes of a set of $n$ trials are contained in a data vector $\boldsymbol{x} = \{x_1,x_2,\ldots,x_n\}$, where $x_i$ is 1 if we obtained the relevant data for study $i$ and is 0 otherwise.
The outcome of each trial depends on a set of $k$ \emph{predictor variables} that may be continuous (\emph{e.g.}, the journal impact factor) or discrete (\emph{e.g.}, the status of the solicitor).
An $n \times k$ matrix $\mathcal{I}$, the \emph{design matrix}, describes the relationships between trials and predictor variables: $\mathcal{I}_{ij}$ is the value for predictor variable $j$ for trial $i$.
\emph{Parameters} relate the values of each predictor variable to the probability of success of each trial, and are described by the parameter vector $\boldsymbol\beta = \{\beta_1,\beta_2,\ldots,\beta_k\}$, where $\beta_i$ is the contribution of parameter $i$ to the probability of success.
In a Bayesian framework, we are interested in estimating the joint posterior probability distribution of the model parameters $\boldsymbol\beta$ conditional on the data $\boldsymbol{x}$.
According to Bayes' theorem,
\begin{align*}
P(\boldsymbol\beta \mid \boldsymbol x) = \frac{ P( \boldsymbol{x} \mid \boldsymbol\beta ) P(\boldsymbol\beta) }{ \int P( \boldsymbol{x} \mid \boldsymbol\beta ) P(\boldsymbol\beta)\ \mathrm{d}\boldsymbol\beta },
\end{align*}
the \emph{posterior probability} of the model parameters, $P(\boldsymbol\beta \mid \boldsymbol x)$, is equal to \emph{likelihood} of the data given the model parameters, $P(\boldsymbol{x} \mid \boldsymbol\beta)$, multiplied by the \emph{prior probability} of the parameters, $P(\boldsymbol\beta)$, divided by the \emph{marginal likelihood} of the data.
Given the design matrix $\mathcal{I}$, the outcomes of each of the $n$ trials are conditionally independent, so that the likelihood of $\boldsymbol x$ is the product of the likelihoods for each individual trial:
\begin{align*}
P(\boldsymbol{x} \mid \boldsymbol\beta) = \prod_{i=1}^n P(x_i \mid \mathcal{I}, \boldsymbol\beta).
\end{align*}
The likelihood of observing the outcome of a particular trial is
\begin{align*}
P(x_i \mid \mathcal{I}, \boldsymbol\beta) = \begin{cases}
\frac{1}{1+e^{-\omega_i} } & \text{if } x_i=1\\
1-\frac{1}{1+e^{-\omega_i}} & \text{if } x_i=0,
\end{cases}
\end{align*}
where
\begin{align*}
\omega_i = \sum_{j=1}^{k} \mathcal{I}_{ij}\beta_j.
\end{align*}
We specified a multivariate normal prior probability distribution on the $\boldsymbol\beta$ parameters with means $\boldsymbol\mu$ and covariance matrix $\Sigma$.
The complexity of the marginal likelihood precludes an analytical solution to the posterior probability distribution.
Accordingly, we approximated the posterior probability distribution using the Markov chain Monte Carlo algorithm implemented in the \verb!R! package \verb!BayesLogit! \citep[][]{Polson_13,R_13}.
This program uses conjugate prior and posterior probability distributions (via Polya-Gamma-distributed latent variables), which permits use of an efficient Gibbs sampling algorithm to approximate the joint posterior distribution of $\boldsymbol\beta$ conditional on the data.
We defined a set of predictor variables based on the bibliometric metadata captured for each study.
We included an \emph{intercept} predictor variable to describe the background probability of procuring data.
We treated \emph{age} ({\it i.e.}, months since publication) and \emph{journal impact factor} as continuous predictor variables, and \emph{journal policy}, \emph{NSF funding}, and \emph{solicitor status} as discrete predictor variables.
Discrete predictor variables for logistic regression are generally binary, assuming values of 0 or 1.
A few of our discrete bibliometric metadata, however, had more than two possible categories.
We therefore adopted an \emph{indicator-variable} approach in which predictor variables with $p$ categories are discretized into $p$ distinct indicators; each study in a particular predictor category was then assigned a 1 for the corresponding indicator variable.
Under this approach, studies published in journals with no data-sharing policy were assigned a 1 for the \emph{no policy} variable, studies published in journals with a strong policy were assigned a 1 for the \emph{strong policy} variable, and studies published in journals that were members of the JDAP initiative at the time of publication were assigned a 1 for the \emph{JDAP membership} variable.
For the studies included in our direct-solicitation campaign, we also assigned values for solicitor status: datasets solicited by an undergraduate student were scored as 1 for the \emph{undergraduate student} variable, while those solicited by a professor were scored as 1 for the \emph{professor} variable.
In order to avoid overparameteriziation of the logistic model, we did not assign indicator variables for the \emph{weak-policy} or \emph{graduate-student} variables.
Accordingly, the values for \emph{no policy}, \emph{strong policy}, and \emph{JDAP membership} parameters are interpreted as effects relative to weak policies; similarly, the values for \emph{undergraduate student} and \emph{professor} parameters are interpreted as effects relative to a graduate student.
Details of the predictor variables and interpretations of the corresponding parameters are summarized in Table \ref{tab:paramtable}.
We analyzed various subsets of our data table in order to understand the relative importance of the predictor variables on different aspects of data availability.
Specifically, we defined subsets of our data table based on whether study data were sought: (1) by queries to online archives, (2) by direct solicitation from the corresponding author, or (3) either by queries to online archives \emph{or} by direct solicitation.
We further parsed our data table based on whether we successfully procured: (1) \emph{only} trees (\emph{i.e.}, the trial outcome was 1 if we acquired a tree and no alignment, and 0 otherwise); (2) \emph{only} alignments; (3) either alignments \emph{or} trees (\emph{i.e.}, the trial outcome was 0 if we acquired no data, and 1 otherwise), and; (4) both alignments \emph{and} trees (\emph{i.e.}, the trial outcome was 1 if we acquired both an alignment and a tree).
This defined 16 (overlapping) subsets of our data table.
Note that not all predictor variables apply to every subset of our data table; \emph{e.g.} the solicitor-status variable, \emph{undergraduate}, only applies to data that were directly solicited.
Details of the data subsets and their predictor variables are summarized in Table \ref{tab:modeltable}.
We estimated parameters for each data subset by performing four independent MCMC simulations, running each chain for $10^6$ cycles and saving every $100^{th}$ sample to reduce autocorrelation and file size.
We assessed the performance of all MCMC simulations using the \verb!Tracer! \citep[][]{Drummond2012} and \verb!coda! \citep[][]{Plummer_2006} packages.
We monitored convergence of each chain to the stationary distribution by plotting the time series and calculating the Geweke diagnostic \citep[{\it GD}; ][]{Geweke_1992} for every parameter.
We assessed the mixing of each chain over the stationary distribution by calculating both the potential scale reduction factor \citep[{\it PSRF};][]{Gelman_1992} diagnostic and the effective sample size \citep[\emph{ESS};][]{Brooks_1997} for all parameters.
Values of all diagnostics for all parameters in all MCMC simulations indicate reliable approximation of the stationary (joint posterior probability) distributions: {\it e.g.}, $ESS >>1000$; $PSRF \approx 1$; {\it GD} $>>0.05$ (Tables \ref{tab:mcmc_archive_any}$-$\ref{tab:mcmc_combined_both}).
Additionally, we assessed convergence by comparing the four independent estimates of the marginal posterior probability density for each parameter, ensuring that all parameter estimates were effectively identical and SAE compliant \citep[{\it c.f.},][]{Brooks_1997}.
Based on these diagnostic analyses, we discarded the first $25\%$ of samples from each chain as burn-in, and based parameter estimates on the combined stationary samples from each of the four independent chains ($N = 30,000$).
\section*{Results and Discussion}
Overall, our efforts secured complete phylogenetic data for $\sim 40\%$ of the published studies (Figure \ref{pie_bar}).
Accordingly, invaluable phylogenetic data for more than half of these studies are effectively lost to science.
From online archives, we successfully procured \emph{complete} phylogenetic data (both the tree and alignment files) for $11.5\%$ of the studies, and \emph{partial} datasets (either the tree or alignment files) for an additional $13.4\%$ of the studies were archived: $5.5\%$
of these cases had only tree files, $7.9\%$
had only alignment files.
Of these online accessions, $24$ were archived in Dryad, $22$ in TreeBASE, and $8$ as supplemental files on journal websites.
Our (in)ability to recover phylogenetic datasets from online archives over the \emph{entire} 13-year period is comparable to that of recent reports regarding phylogenetic data---where archival rates range from $\sim 4\% - 16.7\%$ \citep[][]{Hughes_2011, Stoltzfus_2012, Drew_2013a}---and also falls within the scope of archival rates for non-phylogenetic data, which range from $\sim 14\%-48\%$ \citep[][]{Piwowar_07, Alsheikh-Ali_11, Vines_13}.
However, our results also reveal a dramatic increase in the archiving of phylogenetic data since 2011; {\it e.g.}, datasets from more than half of the studies published in 2013 were deposited in online archives (Figure \ref{pie_bar}).
Our direct-solicitation campaign entailed the exchange of $786$ emails over the course of four weeks (BRM: $n = 341$; MRM: $n = 212$; AFM: $n =233$).
We received responses to $61.3\%$ of the $163$ messages we sent to corresponding authors ($37\%$, $18\%$, and $7\%$ after the first, second and third message, respectively), $38.7\%$ of the authors never responded to any messages ($28\%$, $46\%$, and $42\%$ for BRM, MRM, and AFM, respectively).
Although $20.2\%$ of the messages were initially undeliverable (owing to invalid/obsolete email addresses), we were able to resolve contact information for all but $3\%$ of the corresponding authors (by performing Internet searches and/or contacting study co-authors).
Our $61\%$ response rate is comparable to that of previous studies.
A recent survey \citep[][]{Drew_2013a} reported a $40\%$ response rate to direct requests for phylogenetic data, which falls within the range for studies involving non-phylogenetic data: {\it e.g.}, $20\%$ for medical/clinical trial data \citep[][]{Savage_09}; $27\%$ for psychological trial data \citep[][]{Wicherts_06}; and $71\%$ for population-genetic data \citep[][]{Vines_13}.
By directly contacting corresponding authors, we successfully procured complete phylogenetic datasets for $29.0\%$
of the published studies, and partial datasets for an additional $12.9\%$
of the studies: $8.8\%$
of corresponding authors sent only tree files, and $4.1\%$
sent only alignment files (Figure \ref{pie_bar}).
Our success in procuring complete $(29\%)$ or some form $(42\%)$ of phylogenetic data by direct solicitation compares favorably to that of a recent study \citep[$16\%$;][]{Drew_2013a}, but again is within the range reported for non-phylogenetic data; {\it e.g.}, $10\%$ for medical/clinical trial data \citep[][]{Savage_09}; $26\%$ for psychological-trial data \citep[][]{Wicherts_06}; $45\%$ for gene-expression data \citep[][]{Piwowar_11a}; $48\%$ for cancer microarray data \citep[][]{Piwowar_07}; $59\%$ for population-genetic data \citep[][]{Vines_13}.
The results of our logistic-regression analysis provide insights into factors associated with the availability of published phylogenetic data (Figure \ref{boxplots}; Tables \ref{tab:results_archive_any_relative_probs}$-$\ref{tab:results_sent_any_relative_probs}).
Studies published in journals with strong data-sharing policies are more likely to archive both complete (tree and alignment files) and incomplete (tree or alignment files) phylogenetic data, and are also more likely to provide complete and incomplete phylogenetic data upon direct request.
Strikingly, the availability of phylogenetic data (via online archives or direct solicitation) from studies published in journals with weak data-sharing policies is comparable to (or slightly worse) than that of studies published in journals with no data-sharing policy \citep[{\it c.f.}, ][]{Piwowar_10, Vines_13}.
This observation substantiates recent calls for establishing strong (and stringently enforced) data-sharing policies \citep[][]{Savage_09, Piwowar_10, Whitlock_11, Drew_2013, Drew_2013a}.
The efficacy of such policies is evident for studies published in JDAP journals.
Surprisingly, there is a \emph{low} probability of directly soliciting data for studies published in JDAP journals.
However, this likely reflects the fact that the data from these studies are so often available in online archives that there is essentially no \emph{need} for direct solicitation.
Our analyses also indicate that corresponding authors are more likely to grant data requests from faculty than from students (Figure \ref{boxplots}).
This may simply reflect the fact that the faculty solicitor (BRM) is acquainted with a larger proportion of the corresponding authors.
However, this does not explain why corresponding authors are more likely to provide data to undergraduate than to graduate students.
An alternative (but not mutually exclusive) explanation involves the perceived risks of data sharing.
Authors may be reluctant to share published data for fear (reasonable or not) that reanalysis may identify errors and/or reach contradictory conclusions
\citep[][]{Ceci_1983, Nature_2006}.
This idea has, in fact, been substantiated by a recent study demonstrating that reluctance to share published data is significantly correlated with weaker evidence and a higher prevalence of apparent errors in the reporting of statistical results \citep[][]{Wicherts_11}.
Accordingly, corresponding authors may perceive requests from undergraduate students to present less potential risk than those from graduate students, whereas the potential risks presented by faculty requests are balanced by their greater familiarity to the authors.
The influence of journal impact factor on data availability might also be interpreted from the perspective of perceived risk.
As for non-phylogenetic data \citep[][]{Piwowar_10, Vines_13}, our analyses indicate that studies published in journals with a higher impact factor are more likely to both deposit their phylogenetic data in online archives and provide these data upon direct request (Figure \ref{recovery_IF}).
If willingness to share published data is correlated with the quality of the research \citep[][]{Wicherts_11}, and if research quality is correlated with the impact factor of the publishing journal, then journal impact factor should positively predict data availability.
An alternative (perhaps less conspiratorial) explanation for the correlation between journal impact factor and data availability invokes an indirect effect of journal impact factor on journal data-sharing policy.
That is, by virtue of their greater prestige, journals with higher impact factors may have greater reign to impose stronger (and more strictly enforced) data-sharing policies on contributing authors \citep[][]{Vines_13}.
As in previous studies \citep[][]{Evangelou_2005}, our results indicate that data availability decreases markedly over time.
Several corresponding authors reported that the requested datasets had been misplaced or had been lost due to hard-drive failures.
As noted above, there appears to be a distinct uptick in the availability of data from studies published since $2011$; this trend was particularly pronounced for archived data (Figure \ref{recovery_age_archive}).
This pattern may simply indicate that the decay of archived phylogenetic data is nonlinear.
Our findings, however, indicate that the recent surge in archived phylogenetic data is attributable to policy changes.
Studies with NSF funding are $\sim 1.4$ times more likely to archive some kind of phylogenetic data (tree or alignment files), but are actually \emph{less} likely to archive complete phylogenetic data (Table \ref{tab:results_archive_any_relative_probs}).
Curiously, the NSF mandate has led to a drastic increase in archiving alignment (but not tree) files (Table \ref{tab:results_archive_alignments}).
By contrast, studies published in journals with JDAP membership are $\sim 2.8$ and $\sim 8.6$ times more likely to archive partial and complete phylogenetic datasets, respectively (Table \ref{tab:results_archive_any_relative_probs}; Figure \ref{recovery_age_archive}).
Paradoxically, the probability of successfully soliciting data from studies with NSF funding and/or published in JDAP journals is \emph{lower} than that for studies without NSF funding and/or published in non-JDAP journals (Figure \ref{recovery_age_solicit}).
However, this likely reflects the decreased demand for these data by direct solicitation.
\section*{Summary}
Phylogenetic data are a precious scientific resource: molecular sequence alignments and phylogenies are expensive to generate, difficult to replicate, and have seemingly infinite potential for synthesis and reuse.
At face value, our results support the conclusion of recent studies \citep[][]{Drew_2013a, Drew_2013, Stoltzfus_2012} that the loss of phylogenetic data is catastrophic:
complete phylogenetic datasets have been lost for $\sim 60\%$ of the studies we surveyed.
Our results also identify factors associated with (phylogenetic) data availability that have been implicated by previous studies:
the probability of procuring phylogenetic data is strongly predicted the age of the study, and the data-sharing policy and impact factor of the publishing journal.
Unlike previous studies, however, our survey of phylogenetic datasets spans important policy initiatives and infrastructural changes, and so provides an opportunity to assess the efficacy of those recent measures.
Overall, the positive impact of these community initiatives has been both substantial and immediate.
Even at this very early stage---spanning the first three years since the introduction of these policies---the archival rate of phylogenetic data has increased dramatically.
Specifically, the proportion of studies that archived partial or complete phylogenetic data since 2011 has increased $4.8$-fold and $2.9$-fold, respectively.
Moreover the proportion of archived phylogenetic data has increased each year since the policy changes, and
deposition rates of phylogenetic data to Dryad have been $4.3$ times that of the more established TreeBASE archive.
The prospects for future progress along these lines appear promising: membership of the JDAP consortium has almost tripled in the three years since its formation.
Although recent policy initiatives have had a clear and welcome effect on the preservation and sharing of phylogenetic data, there nevertheless remains considerable scope for improvement.
The NSF data-management policy, for example, has increased the preservation of alignments but not phylogenetic trees.
This is unfortunate, both because phylogenies are more computationally expensive than alignments, and also because most of the reuse of phylogenetic data entails trees rather than sequence alignments \citep[][]{Piwowar_11, Stoltzfus_2012}.
Moreover, although relative archival rates have increased dramatically, the absolute rate remains low: despite recent policy initiatives, a large proportion of datasets are not being captured in online archives.
Sustaining the momentum of recent initiatives could be achieved via small measures that increase the benefits and decrease the costs of data sharing to data generators.
Although authors who archive data are rewarded with increased citation rates \citep[][]{Piwowar_07, Piwowar_13}, this incentive could be enhanced by rewarding the collection of data as an achievement in its own right.
Journal policies can encourage the direct citation of archived datasets in addition to the studies in which the data were generated,
and funding agencies and academic institutions can recognize alternative metrics that acknowledge the scientific value of data \citep[][]{Piwowar_13a}.
Concordantly, the perceived costs of data sharing could be reduced by implementing more flexible embargo policies that protect the priority access of data generators
\citep[][]{Vision_10, Roche_2014}.
Clearly, we have a long way to go in order to adequately preserve and freely share phylogenetic data, and the road ahead will not be easy.
Nevertheless, our findings suggest that we are moving in the right direction; we are beginning to glimpse the dawn of open access to phylogenetic data.
\section*{Acknowledgments}
We are grateful to Bob Thomson for sharing insights on this work, and to all of the corresponding authors for sharing the requested phylogenetic datasets.
This research was supported by NSF grants DEB-0842181 and DEB-0919529 awarded to BRM.
\newpage
|
\section*{Introduction}
The main goal of the paper is to present explicit formulae for all differentials $\delta_k$ of
the Feigin-Fuchs-Rocha-Carridi-Wallach-resolution (\cite{FeFu}, \cite{RochWall}), a free resolution of a one-dimensional $L_1$-module ${\mathbb C}$. By $L_1$ we denote the positive part of the Witt algebra. The differentials $\delta_k$ are $2\times2$-matrices
whose elements belong to the universal enveloping algebra $U(L_1)$. More precisely, the matrix $\delta_k$ can be expressed by means of singular vectors $S_{p,q}(t)v$ in Verma modules over the Virasoro Lie algebra (Virasoro singular vectors $S_{p,q}(t) \in U(L_1)$), $t$ is a complex parameter:
$$
\delta_1=\left( \begin{matrix} S_{1,1}({-}\frac{3}{2}) , \; S_{1,2}({-}\frac{3}{2})
\end{matrix}\right), \;
\delta_k=\left(
\begin{matrix}
S_{1,3k{+}1}({-}\frac{3}{2}) & S_{2k{+}1,2}({-}\frac{3}{2})\\
{-}S_{2k{+}1,1}({-}\frac{3}{2}) & {-}S_{1,3k{+}2}({-}\frac{3}{2})
\end{matrix}\right),
$$
It is easy to find the first few singular vectors, for instance
$$
S_{1,1}(t)=e_1 , \; S_{1,2}(t)=e_1^2{+}t^{{-}1}e_2, \;
S_{3,1}(t){=}e_1^3{+}4te_2e_1{+}(4t^2{+}2t)
e_3.
$$
However, further computational complexity starts to increase exponentially.
Fuchs and Feigin proposed to look for the operators $S_{p,q}(t)$ as
\begin{equation}
S_{p,q}(t)=e_1^{pq}{+}
\sum_{\begin{array}{c}pq \ge i_1 \ge \dots \ge i_s\ge 1\\i_1+\dots+i_s=pq
\end{array}.}P_{p,q}^{i_1,\dots,i_s}(t)e_{i_1}\dots
e_{i_s},
\end{equation}
where $P_{p,q}^{i_1,\dots,i_s}(t)$ are Laurent polynomials in the complex variable $t$. Feigin, Fuchs and later Astashkevich had found some properties of these Laurent polynomials, however no explicit general formula have been found..
For a long time no one could get to write an explicit formula for
the $S_{p,q}(t)$. At the end of 80s Benoit
and Saint-Aubin \cite{BenSA} found a beautiful explicit expression for one family of singular vectors $S_{p,1}(t)$. Hence three of the four matrix elements of $D_k$ we know (due to the property $S_{p,q}(t)=S_{q,p}(t^{-1})$), but the fourth element $S_{2k{+}1,2}\left({-}\frac{3}{2} \right)$ remained unknown. The key idea of the approach by Benoit and Saint-Aubin was the idea to consider expansions of $S_{p,1}(t)$ in all monomials
$$e_{i_1}e_{i_2}\dots e_{i_s}, i_1{+}i_2{+}\dots{+}i_s{=}pq,$$
which correspond to all unordered partitions of $pq$, not only to ordered partitions $i_1 \ge i_2 \ge \dots \ge i_s\ge 1$, as did Feigin and Fuchs. Of course this extends the formula for $S_{p,q}(t)$, a linear combination of this type is not unique, but in some cases (the calculation of the cohomology with different coefficients) it helps to get interesting combinatorial formulas \cite{Mill}.
Bauer, Di Francesco, Itzykson, Zuber \cite{BDFIZ} have found a very elegant proof of singularity of vectors $S_{p,1}(t)v$. Moreover, using the Benoit-Saint-Aubin formula as a starting point, Bauer, Di Francesco, Itzykson, Zuber \cite{BDFIZ} presented a complete and straightforward algorithm for finding all singular vectors. However, their algorithm encounters technical difficulties for $S_{p,q}(t), p,q \ge 3$
and it is not still clear whether it is possible to get an explicit formula in general case by means of it. The Benoit-Saint-Aubin formula, examples $S_{2,2}(t)$ and $S_{3,2}(t)$
considered in \cite{BDFIZ} has allowed us to guess the general explicit formula for
$S_{2,p}(t)$. We prove the singularity of vectors $S_{2,p}(t)v$ following \cite{BDFIZ} the proof for another family $S_{p,1}(t)v$.
\section{Singular vectors of Verma modules over Virasoro Lie algebra}
\label{Singular_vectors} The Virasoro algebra Vir is infinite
dimensional Lie algebra, defined by its basis $\{z, e_i, i \in
{\mathbb Z}\}$ and commutator relations:
$$
[e_i,z]=0, \; \forall i \in {\mathbb Z}, \quad
[e_i,e_j]=(j-i)e_{i+j}+\frac{j^3-j}{12}\delta_{-i,j}z.
$$
Vir is one-dimensional central extension of the Witt algebra $W$
(the one-dimensional center is spaned by $z$).
\begin{remark}
In \cite{FeFu1, FeFu2, FeFuRe} the symbol $L_1$ denotes the
"positive part" of the Witt algebra $W$ (or the Virasoro
algebra $Vir$), i.e. the algebra of {\it polynomial} vector fields
on the line ${\mathbb R}$ which vanish at the origin together with
their first derivatives. We will use further the symbol $L_1$ for
the notation of both the algebras. This will not cause the
confusion because the cohomology $H^*(L_1)$ and $H^*(W_+)$ are
isomorphic \cite{G, Fu}.
\end{remark}
A $Vir$-module $V(h,c)$ is called a Verma module over the Virasoro
algebra if it is free as a module over the universal enveloping
algebra $U(L_1)$ of the subalgebra $L_1 \subset {\rm Vir}$ and it
is generated by some vector $v$ such that
$$
zv=cv, \; e_0v=hv,\quad e_iv=0, \; i <0,
$$
where $c,h \in {\mathbb C}$. As a vector space $V(h,c)$ can be
defined by its infinite basis
$$
v, \; e_{i_1}\dots e_{i_s}v, \quad i_1\ge i_2 \ge \dots \ge i_s,
\; s \ge 1.
$$
A Verma module $V(h,c)$ is ${\mathbb Z}_+$-graded module:
$$
V(h,c)=\bigoplus_{n=0}^{+\infty} V_n(h,c), \quad V_n(h,c)=\langle
e_{i_1}\dots e_{i_s}v, \; i_1+\dots+i_s=n\rangle.
$$
$V_n(h,c)$ is an eigen-subspace of the operator $e_0$ that
corresponds to the eigenvalue $(h+n)$:
$$
e_0(e_{i_1}\dots e_{i_s}v)=(h+i_1+\dots+i_s)e_{i_1}\dots e_{i_s}v.
$$
In addition to this, $zw=cw$ for all $w \in V(h,c)$.
\begin{definition}
A nontrivial vector $w\in V(h,c)$ is called singular if $e_iw=0$
for all $i <0$.
\end{definition}
\begin{remark}
A subalgebra $Vir^-$ which is spanned by
$\{e_i, i < 0\}$ is multiplicatively generated by two elements $e_{-1}$ and $e_{-2}$. Hence a vector
$w \in V(h,c)$ is singular if and only if
$$
e_{-1}w=e_{-2}=0.
$$
\end{remark}
A homogeneous singular vector $w \in V_n(h,c)$ with the grading
equal to $n$ generates in $V(h,c)$ a submodule that is isomorphic
to $V(h+n,c)$.
It is not difficult to present first examples of singular vectors. At the first level
$n=1$ there is a singular vector $w_1 \in V_1(h,c)$ if and only if $h=0$.
Indeed the subspace $V_1(h,c)$ is one-dimensional and it is spanned by
$e_1v$, but on the another hand
$$
e_{-1}e_1v=2e_0v=2hv, \quad e_{-2}e_1v=0.
$$
A two-dimensional subspace coincides with the span of vectors $e_1^2v$ and $e_2v$. The
condition that the vector $w_2 \in V_2(h,c)$ is annihilated by the operator $e_{-1}$ is equivalent to the condition that (up to multiplication by a constant) the vector $w_2$ is equal to $e_1^2v-\frac{2}{3}(2h+1)e_2v$. On the another hand $e_{-2}$ annihilates $w_2$ if and only if the parameters $h$ and $c$ of a Verma module $V(h,c)$ are related by
$$
\label{V_2_equation}
6h-\frac{2}{3}(2h+1)\left(4h+\frac{c}{2}\right)=0.
$$
It is easy to verify that the set of solutions of this equation can be parametrized in a following way
\begin{equation}
c(t)=13+6t+6t^{-1}, \quad h(t)=-\frac{3}{4}t-\frac{1}{2},
\end{equation}
where $t \ne 0$ runs the complex numbers (or runs the reals, it depends on the task).
The last remark can be reformulated as follows: for any value of $t$ there is a unique (up to multiplication by a constant) singular vector $w_2=e_1^2v+te_2v, w_2 \in V_2(h(t), c(t))$, where
$h(t)$ and $c(t)$ are defined by equations (\ref{V_2_equation}).
\begin{theorem}[\cite{Kac}, \cite{FeFu, FeFu2}]
\label{Kac_theorem} In the Verma module $V(h,c)$ there is a
singular vector $w \in V_n(h,c)$ with the grading not higher than
$n$ if and only if, when two natural numbers $p$ and $q$ can been
found and also a complex number $t$ such that
\begin{equation}
\begin{split}
pq \le n, \quad c=c(t)=13+6t+6t^{-1},\\
h=h_{p,q}(t)=\frac{1-p^2}{4}t+\frac{1-pq}{2}+\frac{1-q^2}{4}t^{-1}.
\end{split}
\end{equation}
\end{theorem}
In particular, the following assertion holds \cite{Fu2}: with
the fixed natural numbers $p$ and $q$ and with an arbitrary
complex number $t$ the Verma module $V(h_{p,q}(t),c(t))$ contains
a singular vector $w_{p,q}(t)$ of degree $pq$, moreover the vector
$w_{p,q}(t)$ is determined unambiguously up to a multiplication by
some scalar:
$$
w_{p,q}(t)=S_{p,q}(t)v=\sum_{|I|=pq}P_{p,q}^I(t)e_Iv=
\sum_{i_1+\dots+i_s=pq}P^{i_1,\dots,i_s}_{p,q}(t)e_{i_1}\dots
e_{i_s}v,
$$
where $S_{p,q}(t)$ denotes some element of the universal
enveloping algebra $U(L_1)$. The coefficients $P_{p,q}^I(t)$
depend polynomially on $t$ and $t^{-1}$. We assume that the
coefficient $P^{1,\dots,1}_{p,q}(t)$ is equal to one. Obviously
that $S_{p,q}(t)=S_{q,p}(t^{-1})$.
\begin{theorem}[Benoit, Saint-Aubin]
\begin{equation}
\label{Benoit_SA}
S_{p,1}(t)=
\sum_{\begin{array}{c}i_1,\dots,i_s\\i_1+\dots+i_s=p
\end{array}.}c_p(i_1,\dots,i_s)t^{p-s}e_{i_1}\dots
e_{i_s}.
\end{equation}
where the sums are all over all partitions of $p$ by positive
numbers without any ordering restriction, and the coefficients
$c_p(i_1,\dots,i_s)$ are defined by the formulas
\begin{equation}
\label{coefficient_c}
c_p(i_1,\dots,i_s)=
\frac{(p{-}1)!^2}{\prod_{l=1}^{s{-}1} \left((\sum_{q=1}^l i_q)(p-\sum_{q=1}^l i_q)\right)}
\end{equation}
\end{theorem}
\begin{example}
$$
S_{1,1}(t)=e_1,\;\;S_{2,1}(t)=e_1^2+te_2,\;\;S_{3,1}(t)=e_1^3+t(2e_1e_2+2e_2e_1)+4t^2e_3.
$$
\end{example}
\begin{theorem}
\label{main_theorem}
Let $V$ be a Verma module over the Virasoro algebra $Vir$. $V$ generated by the vector $v$ and such that $V$
corresponds to the (complex) parameter $t$:
with
$$
c{=}13{+}6t{+}6t^{-1}, h{=}{-}\frac{(p{-}1{+}t)(t^{-1}\left(p{+}1){+}3\right)}{4}.
$$
Let consider an element of universal enveloping algebra $U(L_1)$ definend by the formula
\begin{equation}
\label{new_family}
S_{2,p}(t)=
{\sum_{\begin{array}{c}i_1,\dots,i_s\\i_1{+}{\dots}{+}i_s{=}2p
\end{array}.}}
f_p(i_1,\dots,i_s)e_{i_1}\dots
e_{i_s}.
\end{equation}
where the sums are all over all partitions of $2p$ by positive
numbers without any ordering restriction, and the coefficients (that are rational functions on $t$)
$f_p(i_1,\dots,i_s)$ are defined by the formulas
\begin{equation}
\label{coefficient_f}
f_p(i_1{,}{\dots}{,}i_s)
{=}\frac{(2p{-}1)!^2 (2t)^{s{-}2p} \prod\limits_{r=1}^{2p{-}1}(p{-}t{-}r) \prod\limits_{m=1}^s \left( i_m(2t{+}1){+}2(p{-}t{-}\sum\limits_{n=1}^m i_n)\right)}
{\prod\limits_{l=0}^{2p{-}1} (2p{-}1{-}2l)\prod\limits_{l=1}^{s-1}\left((\sum\limits_{n=1}^l i_n)(2p{-}\sum\limits_{n=1}^l i_n)(p{-}t{-}\sum\limits_{n=1}^l{i_n})\right)}.
\end{equation}
Then $S_{2,p}(t)v$ is a singular vector of $V$:
\begin{equation}
\label{singular_cond}
e_{-k}S_{2,p}v=0, \quad k \in {\mathbb N}.
\end{equation}
\end{theorem}
For instance $S_{2,1}=e_1^2{+}te_2$ and
\begin{equation}
\begin{split}
S_{2,2}(t)=e_1^4{+}4te_1e_2e_1{+}
\frac{(1{-}t^2)}{t}(e_1^2e_2{+}e_2e_1^2){+}\frac{(1{-}t^2)^2}{t^2}e_2^2{+}\\{+}\frac{(1{+}t)(4t{-}1)}{t}e_1e_3{+}\frac{(1{-}t)(4t{+}1)}{t}e_3e_1{+}\frac{3(1{-}t^2)}{t}e_4.
\end{split}
\end{equation}
\begin{proof}
We define the sequence of vectors $v^{(0)}, v^{(1)},\dots, v^{(2p{-}1)}$ by
$$
v^{(0)}=v
$$
and the following recursive relation
for $k=1,\dots, 2p{-}1$:
\begin{equation}
v^{(k)}=\frac{2t\sum\limits_{j{=}1}^k\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}
}{k(2p{-}k)(k{-}p{-}t)}
\end{equation}
For instance
\begin{equation}
\begin{split}
v^{(1)}=\frac{2t(1{-}2p)e_1v^{(0)}}{(2p{-}1)(1{-}p{-}t)}, \\
v^{(2)}=\frac{2t\left(2(t{+}1{-}p)e_2v^{(0)} {+} (3{-}2p)e_1v^{(1)}\right)}{2(2p{-}2)(2{-}p{-}t)}.
\end{split}
\end{equation}
After that we define the vector $w \in V_{2p}$ by the formula:
\begin{equation}
w=2t\sum\limits_{j{=}1}^{2p}\left( (j{-}1)(2t{-}1){+}2p{-}1\right)e_jv^{(2p{-}j)}.
\end{equation}
\begin{lemma}
\begin{equation}
e_{-1}v^{(k)}=-(p{+}2{-}k{+}3t)v^{(k{-}1)}, \quad k=1,\dots, 2p{-}1.
\end{equation}
and the vector $w$ annihilates the operator $e_{-1}$:
$$
e_{-1}w=e_{-1}\left( 2t\sum\limits_{j{=}1}^{2p}\left( (j{-}1)(2t{-}1){+}2p{-}1\right)e_jv^{(2p{-}j)}\right)=0.
$$
\end{lemma}
\begin{proof}
We will prove the following formula for $k=1,\dots,2p$:
\begin{equation}
\begin{split}
e_{-1}\left(2t\sum\limits_{j{=}1}^k\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=-(p{+}2{-}k{+}3t)k(2p{-}k)(k{-}p{-}t)v^{(k{-}1)}.
\end{split}
\end{equation}
We proceed by recursion on $k$. The recursion base is
$$
e_{-1}\left(2t(2{-}2p{-}1)e_1v^{(0)} \right)=2t(1{-}2p)2hv^{(0)}=-(p{+}1{+}3t)(2p{-}1)(1{-}p{-}t)v^{(0)}.
$$
The recursive step is the following calculation
\begin{equation}
\begin{split}
e_{-1}\left(2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)\left((j{+}1)e_{j{-}1}v^{(k{-}j)}
{+}e_{j}e_{-1}v^{(k{-}j)}\right)=\\
=4t(2k{-}2p{-}1)e_0v^{(k{-}1)}{+}2t\sum\limits_{j{=}2}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(j{+}1)e_{j{-}1}v^{(k{-}j)}{+}\\
{-}2t\sum\limits_{j{=}1}^{k{-}1}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(p{+}2{-}k{+}j{+}3t)e_{j}v^{(k{-}j{-}1)}=\\
=t(2k{-}2p{-}1)4(h{+}k{-}1)v^{(k{-}1)}{+}\\
{-}(p{-}k{+}1{+}3t)2t\sum\limits_{j{=}1}^{k{-}1}\left((j{-}1)(2t{-}1){+}2k{-}2p{-}3\right)e_{j}v^{(k{-}j{-}1)}.
\end{split}
\end{equation}
In this chain of equalities we replaced $e_{-1}e_j$ by $(j{+}1)e_{j{-}1}{+}e_je_{-1}$ and used the formula
for $e_{-1}v^{(k{-}j)}$ which we considered true for $j=1,\dots, k$ by the induction hypothesis.
Then we shifted $j{=}j'{+}1$ the summation index in the sum
$$
\sum\limits_{j{=}2}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(j{+}1)e_{j{-}1}v^{(k-j)}
$$
and used the following equality
\begin{equation}
\begin{split}
(j(2t{-}1){+}2k{-}2p{-}1)(j{+}2)-\\
-((j{-}1)(2t{-}1){+}2k{-}2p{-}1)(p{+}2{-}k{+}j{+}3t)=\\
={-}(p{-}k{+}1{+}3t)\left((j{-}1)(2t{-}1){+}2k{-}2p{-}3\right).
\end{split}
\end{equation}
It follows from the definition of $v^{(k{-}1)}$ that
\begin{equation}
\begin{split}
2t\sum\limits_{j{=}1}^{k{-}1}\left((j{-}1)(2t{-}1){+}2k{-}2p{-}3\right)e_{j}v^{(k{-}j{-}1)}=\\
=(k{-}1)(2p{-}k{+}1)(k{-}1{-}p{-}t)v^{(k{-}1)}.
\end{split}
\end{equation}
We remark that
$$
4t(h{+}k{-}1)=\left(-(p{-}1{+}t)(p{+}1{+}3t){+}4kt{-}4t)\right)
$$
and finish our calculations.
\begin{equation}
\label{resulting_equality}
\begin{split}
e_{-1}\left(2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=(2k{-}2p{-}1)\left({-}(p{-}1{+}t)(n{+}1{+}3t){+}4kt{-}4t\right)v^{(k{-}1)}{-}\\
{-}(p{-}k{+}1{+}3t)(k{-}1)(2p{-}k{+}1)(k{-}1{-}p{-}t)v^{(k{-}1)}=\\
=-k(2p{-}k)(k{-}p{-}t)(p{+}2{-}k{+}3t)v^{(k{-}1)}.
\end{split}
\end{equation}
If we divide the resulting equality (\ref{resulting_equality}) by $k(2p{-}k)(k{-}p{-}t)$, we obtain the required formula
$$
e_{-1}v^{(k)}=-(p{+}2{-}k{+}3t)v^{(k-1)}.
$$
Taking $k=2p$ we will get
$$
e_{-1}w=0.
$$
\end{proof}
\begin{lemma}
\begin{equation}
e_{-2}v^{(1)}=0, \quad e_{-2}v^{(k)}=-(p{+}4{-}k{+}5t)v^{(k{-}2)}, \quad k=2,\dots, 2p{-}1.
\end{equation}
and the vector $w$ annihilates the operator $e_{-2}$:
$$
e_{-2}w=e_{-2}\left( 2t\sum\limits_{j{=}1}^{2p}\left( (j{-}1)(2t{-}1){-}1\right)e_jv^{(2p{-}j)}\right)=0.
$$
\end{lemma}
\begin{proof}
We recursively prove this formula for $k=2,\dots,2p$:
\begin{equation}
\begin{split}
e_{-2}\left(2t\sum\limits_{j{=}1}^k\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=-(p{+}4{-}k{+}5t)k(2p{-}k)(k{-}p{-}t)v^{(k{-}2)}.
\end{split}
\end{equation}
The starting point $k{=}1$ is evident
$$
e_{-2}\left(2t(2{-}2p{-}1)e_1v^{(0)} \right)=2t(1{-}2p) (3e_{-1}v^{(0)}+e_1e_{-2}v^{(0)})=0.
$$
Now we take $k{=}2$.
\begin{equation}
\begin{split}
e_{-2}\left(2t(2t{+}2k{-}2p{-}2)e_2v^{(0)} {+} 2t(3{-}2p)e_1v^{(1)}\right)=\\
=2t(3{-}2p)3e_{-1}v^{(1)}{+}2t(2t+2-2p)(4e_0{+}\frac{1}{2}z)v^{(0)}=\\
=2t\left(-3(3{-}2p)(p{+}1{+}3t){+}2(t{+}1{-}p)(4h{+}\frac{13}{2}{+}3t{+}3t^{-1})\right)v^{(0)}=\\
=-(p{+}2{+}5t)2(2p{-}2)(2{-}p{-}t)v^{(0)}.
\end{split}
\end{equation}
Now let consider the recursive step.
\begin{equation}
\begin{split}
e_{-2}\left(2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)\left((j{+}2)e_{j{-}2}v^{(k{-}j)}
{+}e_{j}e_{-2}v^{(k{-}j)}\right)=\\
{=}6t(2k{-}2p{-}1)e_{-1}v^{(k{-}1)}{+}4t(t{+}1{-}p)(4(h{+}k{-}2){+}\frac{13}{2}{+}3t{+}3t^{-1})v^{(k{-}2)}{+}\\
+2t\sum\limits_{j{=}3}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(j{+}2)e_{j{-}2}v^{(k{-}j)}-\\
{-}2t\sum\limits_{j{=}1}^{k{-}2}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(p{+}4{-}k{+}j{+}5t)e_{j}v^{(k{-}j{-}2)}=
\end{split}
\end{equation}
We used the induction assumption
for $e_{-2}v^{(k{-}j)}, j=1,\dots, k{-}2$.
Now we shift the summation index $j'=j{-}2$ in the sum
$$
\sum\limits_{j{=}3}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(j{+}2)e_{j{-}1}v^{(k-j)},
$$
replace $3e_{-1}v^{(k{-}1)}$ by $-(p{+}3{-}k{+}3t)v^{(k{-}2)}$
and we have
\begin{equation}
\begin{split}
e_{-2}\left(2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=2tQv^{(k{-}2)}+
2t\sum\limits_{j{=}1}^{k{-}2}R_je_jv^{(k{-}j{-}2)}
\end{split}
\end{equation}
where
$$
Q{=}{-}3(2k{-}2p{-}1)(p{+}3{-}k{+}3t){+}2(t{+}k{-}p{-}1)\left( 4h{+}4k{-}8{+}\frac{13}{2}{+}3t{+}3t^{-1}\right)
$$
and now we compute the coefficient $R_j$
\begin{equation}
\begin{split}
R_j{=}{-}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)(p{+}4{-}k{+}j{+}5t){+}\\
{+}(4{+}j)\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)=\\
=-(p{-}k{-}2{+}5t)\left((j{-}1)(2t{-}1){+}2k{-}2p{-}5\right)
\end{split}
\end{equation}
The first factor of $R_j$ does not depend on $j$ and moreover
\begin{equation}
\begin{split}
2t\sum\limits_{j{=}1}^{k{-}2}R_je_jv^{(k{-}j{-}2)}=-(p{-}k{-}2{+}5t)(k{-}2)(2p{-}k{+}2)(k{-}p{-}t)v^{(k{-}2)}
\end{split}
\end{equation}
Hence
\begin{equation}
\begin{split}
\label{resultformul}
e_{-2}\left(2t\sum\limits_{j{=}1}^{k}\left( (j{-}1)(2t{-}1){+}2k{-}2p{-}1\right)e_jv^{(k{-}j)}\right)=\\
=\left(2tQ{-}(p{-}k{-}2{+}5t)(k{-}2)(2p{-}k{+}2)(k{-}p{-}t)\right)v^{(k{-}2)}{=}\\
=-k(2p{-}k)(k{-}p{-}t)(p{+}4{-}k{+}5t)v^{(k{-}2)}.
\end{split}
\end{equation}
We divide the resulting equality by $k(2p{-}k)(k{-}p{-}t)$ and obtain the required formula.
Taking $k=2p$ in (\ref{resultformul}) we have
$$
e_{-2}w=0.
$$
This completes the proof of the lemma.
\end{proof}
Proof of the theorem follows from two lemmas. The only one thing we still have to do is to present
explicit formula for $w$.
\begin{equation}
\begin{split}
w=2t\sum\limits_{j_1{=}1}^{2p}\left( (j_1{-}1)(2t{-}1){+}2p{-}1\right)e_{j_1}v^{(2p{-}j_1)},\\
v^{(2p{-}j_1)}{=}\sum\limits_{j_2{=}1}^{2p{-}j_1}\frac{2t \left( (j_2{-}1)(2t{-}1){+}2p{-}2j_1{-}1\right)}{(2p{-}j_1)j_1(p{-}j_1{-}t)}{e_{j_2}}{v^{(2p{-}j_1{-}j_2)}},\\
v^{(2p{-}j_1{-}j_2)}{=}\sum\limits_{j_3{=}1}^{2p{-}j_1{-}j_2}\frac{2t \left( (j_3{-}1)(2t{-}1){+}2p{-}2j_1{-}2j_2{-}1\right)}{(2p{-}j_1{-}j_2)(j_1{+}j_2)(p{-}j_1{-}j_2{-}t)}{e_{j_3}}{v^{(2p{-}j_1{-}j_2{-}j_3)}}.
\end{split}
\end{equation}
Proceeding further step by step we obtain the formula
\begin{equation}
\begin{split}
w={\sum_{\begin{array}{c}j_1,\dots,j_s\\j_1{+}{\dots}{+}j_s{=}2p
\end{array}.}}
\frac{(2t)^s\prod\limits_{r=1}^{s}\left((j_r{-}1)(2t{-}1){+}2p{-}1{-}2\sum\limits_{q=1}^{r{-}1}j_q) \right)}{\prod\limits_{m=1}^{s-1}\left((\sum\limits_{q=1}^{m}j_q)(2p{-}\sum\limits_{q=1}^{m}j_q)(p{-}t{-}\sum\limits_{q=1}^{m}j_q)\right)}e_{j_1}\dots
e_{j_s}v.
\end{split}
\end{equation}
We need to calculate the coefficient facing $e_1^{2p}$ in the expansion of $w$:
$$
w
=\frac{(2t)^{2p}\prod\limits_{k=0}^{2p{-}1}(2p{-}1{-}2k)}{(2p{-}1)!^2\prod\limits_{q{=}1}^{2p{-}1}(p{-}t{-}q)}
e_1^{2p}v+\dots.
$$
Finally we get
$$
S_{2,p}(t)v=\frac{(2p{-}1)!^2\prod\limits_{q{=}1}^{2p{-}1}(p{-}t{-}q)}{(2t)^{2p}\prod\limits_{k=0}^{2p{-}1}(2p{-}1{-}2k)}w.
$$
\end{proof}
\begin{example}
\begin{equation}
\begin{split}
S_{2,3}(t)=e_1^6{+}\frac{(4{-}t^2)}{3t}(e_1^4e_2{+}e_2e_1^4){+}
\frac{8(1{-}t^2)}{3t}(e_1^3e_2e_1{+}e_1e_2e_1^3){+}9te_1^2e_2e_1^2{+}\\
{+}3(4{-}t^2)(e_1^2e_2^2{+}e_2^2e_1^2){+}\frac{64(1{-}t^2)^2}{9t^2}e_1e_2^2e_1{+}\frac{(4{-}t^2)^2}{9t^2}e_2e_1^2e_2{+}\\
{+}\frac{8(1{-}t^2)(4{-}t^2)}{9t^2}(e_1e_2e_1e_2{+}e_2e_1e_2e_1){+}\frac{(4{-}t^2)^2}{t}e_2^3{+}\\
{+}\frac{4(4{-}t^2)(1{-}t^2)(9{-}16t^2)}{9t^4}e_3^2{+}\frac{6(1{+}t)(4t{-}1)}{t}e_1^2e_3e_1{+}
\frac{6(1{-}t)(4t{+}1)}{t}e_1e_3e_1^2{+}\\
{+}\frac{2(1{+}t)(2{+}t)(3{-}4t)}{3t^2}e_1^3e_3{+}\frac{2(1{-}t)(2{-}t)(3{+}4t)}{3t^2}e_3e_1^3{+}\\
{+}\frac{16(1{-}t^2)(1{+}t)(2{+}t)(3{-}4t)}{9t^3}e_1e_2e_3{+}\frac{16(1{-}t^2)(1{-}t)(2{-}t)(3{+}4t)}{9t^3}e_3e_2e_1{+}\\
{+}\frac{2(4{-}t^2)(2{+}t)(1{+}t)(3{-}4t)}{9t^3}e_2e_1e_3{+}\frac{2(4{-}t^2)(2{-}t)(1{-}t)(3{+}4t)}{9t^3}e_3e_1e_2{+}\\
{+}\frac{2(4{-}t^2)(1{-}t)(4t{+}1)}{t^2}e_1e_3e_2{+}\frac{2(4{-}t^2)(1{+}t)(4t{-}1)}{t^2}e_2e_3e_1{+}\\
{+}\frac{6(1{+}t)(2{+}t)(3t{-}1)}{t^2}e_1^2e_4{+}\frac{48(1{-}t^2)}{t}e_1e_4e_1{+}
\frac{6(1{-}t)(2{-}t)(3t{+}1)}{t^2}e_4e_1^2{+}\\
{+}\frac{(4{-}t^2)(2{+}t)(1{+}t)(3t{-}1)}{t^3}e_2e_4{+}\frac{(4{-}t^2)(2{-}t)(1{-}t)(3t{+}1)}{t^3}e_4e_2{+}\\
{+}\frac{4(1{-}t^2)(2{+}t)(8t{-}1)}{t^3}e_1e_5{+}\frac{4(1{-}t^2)(2{-}t)(8t{+}1)}{t^3}e_5e_1{+}\frac{20(1{-}t^2)(4{-}t^2)}{t^3}e_6.
\end{split}
\end{equation}
\end{example}
Obviously, coefficients $f_p(i_1,\dots, i_s)$ facing monomials $e_{i_1}\dots e_{i_s}$ (that correspond to unordered partitions of $2p$) in the expansion of $S_{2,p}(t)$ are not uniquely defined, because these monomials are linearily dependent in $V$. There are, however, two coefficients $f_p(1,\dots,1)$ and $f_p(2,\dots, 2)$ facing $e_1^{2p}$ and $e_2^p$ respectively which, as it is easily seen, are uniquely determined. Let us calculate $f_p(2,\dots,2)$.
$$
f_p(2,\dots, 2)=\frac{\prod\limits_{q=1}^{p}\left(t^2{-}(p{+}1{-}2q)^2\right)}{t^p}.
$$
In particular, in our two previous examples we have
$$
f_2(2,2)=\frac{(t^2{-}1)^2}{t^2}, \quad f_3(2,2,2)=\frac{(t^2{-}4)^2t^2}{t^3}.
$$
It was proved in \cite{AstFu} that
$$
S_{k,l}(t)=(k{-}1)!^{2l}e_k^lt^{(k{-}1)l}+\dots+(l{-}1)!^{2k} e_l^kt^{-(l{-}1)k},
$$
where "$\dots$" denotes intermediate degrees in $t$.
we see that our coefficient $f_p(2,\dots,2)$ has prescribed asymptotic behavior with respect to $t$.
Hence
$$
S_{2,p}(t)=e_2^pt^{p}+\dots
$$
which is consistent with our calculations.
\section{Singular vectors and cohomology}
Now we are going to consider Verma modules over the Virasoro algebra
with $c=0$ that one can consider as Verma modules over the Witt algebra.
\begin{proposition}[Kac \cite{Kac}, Feigin and Fuchs \cite{FeFu, FeFu2}]
There is a singular vector $w_n$ in the homogeneous subspace
$V_n(0,0)$ of the Verma module $V(0,0)$ then and only then when
$n$ is equal to some pentagonal number $n=e_{\pm}(k)=\frac{3k^2\pm
k}{2}$.
\end{proposition}
It follows from the theorem \ref{Kac_theorem} that if a Verma
module $V(h,0)$ (with $c{=}0$) has a singular vector $w_{p,q}(t)$
it implies that $t{=}{-}\frac{3}{2}$ or $t{=}{-}\frac{2}{3}$. We will
fix the value $t=-\frac{3}{2}$ and we will write $S_{p,q}$ instead
of $S_{p,q}\left({-}\frac{3}{2}\right)$ for convenience in the
notations. Let us denote by $V\left(\frac{3k^2\pm k}{2}\right)$ a
submodule in the Verma module $V(0,0)$ generated by a singular
vector with the grading $\frac{3k^2\pm k}{2}$. The submodule
$V\left(\frac{3k^2\pm k}{2}\right)$ is isomorphic to the Verma
module $V\left(\frac{3k^2\pm k}{2},0\right)$.
\begin{proposition}[\cite{RochWall}, \cite{FeFu}]
The system of submodules $V\left(\frac{3k^2\pm k}{2}\right)$ has
the following important properties:
1) the sum $V(1)+V(2)$ is the subspace of codimension one in
$V(0)$;
2) $V\left(\frac{3k^2- k}{2}\right)\cap V\left(\frac{3k^2+
k}{2}\right)=V\left(\frac{3(k{+}1)^2{-} (k{+}1)}{2}\right)+
V\left(\frac{3(k{+}1)^2{+} (k{+}1)}{2}\right),k {\ge} 1$.
\end{proposition}
One can directly verify that the vectors $e_1v$ and
$\left(e_1^2-\frac{2}{3}e_2\right)v$ are singular in the module
$V(0,0)$ with the gradings $1$ and $2$ respectively. Let consider
a submodule $V(1)$ generated by $w_1{=}e_1v$. It is isomorphic to the Verma module
$V(1,0)$ and it containes a singular vector
$S_{1,4}w_1$. A Verma module $V(2,0) = V(2)$ (generated by the vector $w_2{=}S_{1,2}v$)
containes a singular vector
$S_{3,1}w_2.$
Vectors $S_{1,4}w_1$ and $S_{3,1}w_2$ are both singular and they at the level $n{=}5$ in the Verma module $V(0,0)$. Hence they coincide
$$
w_5=S_{1,4}w_1=S_{3,1}w_2.
$$
Similarly, one can check the other equality
$$
w_7=S_{3,2}w_1=S_{1,5}.
$$
We see that singular vectors $w_5, w_7 \in V(1)\cap V(2)$ as well as the sum
$V(5) + V(7)$ of submodules generated by $w_5$ and $w_7$ respectively:
$$
V(5) + V(7) \subset V(1)\cap V(2).
$$
The intersection $V(5) \cap V(7)$ containes two singular vectors
$$
w_{12}=S_{1,7}w_5=S_{5,1}w_7, \quad
w_{15}=S_{5,2}w_5=S_{1,8}w_7.
$$
The inclusions of submodules $V\left(\frac{3k^2\pm k}{2}\right)$
provides us with an exact sequence \cite{RochWall, FeFu, FeFu2}:
\begin{equation}
\label{resolution}
\begin{split}
\begin{CD}
{\dots}{\rightarrow} {V(\frac{3(k{+}1)^2{-}
(k{+}1)}{2}){\oplus}V(\frac{3(k{+}1)^2{+} (k{+}1)}{2})}
@>{\delta_{k{+}1}}>> V(\frac{3k^2{-}
k}{2}){\oplus}V(\frac{3k^2{+} k}{2}){\rightarrow}{\dots}\end{CD}\\
\begin{CD}
{\dots} @>{\delta_3}>> V(5){\oplus}V(7) @>{\delta_2}>>
V(1){\oplus}V(2) @>{\delta_1}>> V(0) @>{\varepsilon}>>
{\mathbb C} {\to} 0
\end{CD},
\end{split}
\end{equation}
where $\delta_k$ are defined with the aid of operators $S_{p,q}
\in U(L_1)$:
\begin{equation}
\begin{split}
\delta_{k+1}\left( \begin{array}{c}x \\y
\end{array}\right)=\left(\begin{array}{cc}S_{1,3k{+}1} & S_{2k{+}1,2}\\
{-}S_{2k+1,1}& {-}S_{1,3k{+}2} \end{array}\right)\left(
\begin{array}{c}x \\y \end{array}
\right), \; k \ge 1;\\
\delta_1\left(
\begin{array}{c}x \\y \end{array}\right)=\left(S_{1,1}, S_{1,2}\right)\left(
\begin{array}{c}x \\y \end{array}\right),
\end{split}
\end{equation}
and $\varepsilon$ is a projection to the one-dimensional ${\mathbb
C}$-submodule generated by the vector $v$.
\begin{theorem}[\cite{RochWall}, \cite{FeFu}]
The exact sequence (\ref{resolution}) considered as a sequence of
$L_1$-modules is a free resolution of the one-dimensional trivial
$L_1$-module ${\mathbb C}$.
\end{theorem}
\begin{corollary}[\cite{FeFu}]
Let $V$ be a $L_1$-module. Then the cohomology $H^*(L_1,V)$ is
isomorphic the cohomology of the following complex:
\begin{equation}
\label{resolution2}
\begin{CD}
{\dots}@<{d_{k{+}1}}<< V{\oplus}V @<{d_k}<< V{\oplus}V
@<{d_{k{-}1}}<< {\dots} @<{d_1}<< V{\oplus}V @<{d_0}<< V
\end{CD},
\end{equation}
with the differentials
\begin{equation}
\label{differential}
\begin{split}
d_k\left(\begin{array}{c}m_1 \\m_2 \end{array}\right)=\left(\begin{array}{cc}S_{1,3k{+}1} & {-}S_{2k+1,1}\\
S_{2k{+}1,2}& {-}S_{1,3k{+}2} \end{array}\right)\left(\begin{array}{c}m_1 \\m_2 \end{array}\right), \; k \ge 1;\\
d_0(m)=\left(\begin{array}{c}S_{1,1}m \\S_{1,2}m
\end{array}\right),\;\; m, m_1, m_2 \in V.
\end{split}
\end{equation}
\end{corollary}
Let consider the trivail one-dimensional module $V={\mathbb C}$. All operators
$d_k$ are trivial and we obtain the famous Goncharova theorem.
\begin{theorem}[\cite{G}] The space of $q$-cohomology
$H^q(L_1,{\mathbb C})$ is two-dimensional for all $q \ge 1$,
moreover it is the direct sum of its one-dimensional subspaces:
$$ H^q(L_1,{\mathbb C})=H_{\frac{3q^2 {-} q}{2}}^q(L_1,{\mathbb C})
\oplus H_{\frac{3q^2{+}q}{2}}^q(L_1,{\mathbb C}). $$
\end{theorem}
The numbers $e_{\pm}(q)=\frac{3q^2 \pm q}{2}$ are
called Euler pentagonal numbers.
\begin{remark}
The original proof of the properties of the free resolution of one-dimensional $L_1$-module
${\mathbb C}$ by Rocha-Carridi and Wallach \cite{RochWall} seriously used the Goncharova
theorem.
\end{remark}
Fuchs and Feigin studied $L_1$-cohomology with coefficients in graded modules $V=\oplus_i V_i$ only with
one-dimensional homogeneous components $V_i$. We will define them
with the aid of the special basis $f_i, V_i=\langle f_{i}\rangle$
($j \in {\mathbb Z}$ in the infinite dimensional case and $j \in
{\mathbb Z}, m\le j\le n$ in the finite dimensional). For a given
graded $L_1$-module $V=\oplus_i V_i$ let us introduce the numbers
$\sigma_{p,q}(j) \in {\mathbb K}$ such that
$$
S_{p,q}f_j=\sigma_{p,q}(j)f_{j{+}pq}.
$$
\begin{example}
\label{F_lambda_mu} The well-known $L_1$-module $F_{\lambda,\mu}$
of tensor densities \cite{Fu}:
$$ e_if_j=\left(j+\mu-\lambda(i+1) \right) f_{i+j}, \forall i \in
{\mathbb N}, j \in {\mathbb Z},
$$
where $\lambda, \mu \in {\mathbb K}$ are two complex parameters.
\end{example}
\begin{corollary}[\cite{FeFu}]
Let $V=\oplus_i V_i$ be a graded $L_1$-module over the field
${\mathbb K}$. Then the one-dimensional cohomology $H^*_s(L_1,V)$
is isomorphic to the cohomology of the following complex:
\begin{equation}
\label{resolution3}
\begin{CD}
{\dots}@<{D_{k{+}1}}<< {\mathbb K}{\oplus}{\mathbb K} @<{D_k}<<
{\mathbb K}{\oplus}{\mathbb K}@<{D_{k{-}1}}<< {\dots} @<{D_1}<<
{\mathbb K}{\oplus}{\mathbb K} @<{D_0}<< {\mathbb K}
\end{CD},
\end{equation}
where the differentials $D_k$ are assigned by the numerical
matrices
\begin{equation}
\label{D_k}
D_k{=}\left(\begin{array}{cc}\sigma_{1,3k{+}1}\left(s{+}\frac{3k^2{-}k}{2}\right)
& {-}\sigma_{2k+1,1}\left(s{+}\frac{3k^2{-}k}{2}\right)\\
\sigma_{2k{+}1,2}\left(s{+}\frac{3k^2{+}k}{2}\right)&
{-}\sigma_{1,3k{+}2}\left(s{+}\frac{3k^2{+}k}{2}\right)
\end{array}\right),
D_0{=}\left(\begin{array}{c}\sigma_{1,1}(s) \\\sigma_{1,2}(s)
\end{array}\right).
\end{equation}
\end{corollary}
Fuchs and Feigin \cite{FeFu} have not found general explicit formulas for singular vectors
of the type $S_{p,1}$ and $S_{p,2}$ but they managed however to find formulae for elements
$\sigma_{p,q}(j)$ for modules $F_{\lambda,\mu}$ (using other arguments). But other graded $L_1$-modules are also can be important for applications \cite{Mill} and require explicit formulas for singular vectors $S_{p,1}$ and $S_{p,2}$. The main result of this article, together with the Benoit-Saint-Aubin theorem provides us with the formulae useful for cohomology calculations.
|
\section{Introduction}
\label{Intro}
The goal of this paper is to develop a new perspective on
quantization, from which some intriguing phenomena in four-dimensional
gauge theory may be naturally understood.
Specifically, the phenomena that we wish to understand are the
following. At low energies, an $\CN = 2$ supersymmetric gauge theory
compactified on a circle $S^1$ is described by a three-dimensional
$\CN = 4$ supersymmetric sigma model \cite{Seiberg:1996nz}. The
target space $\CM$ of the sigma model is a hyperk\"ahler manifold,
which is moreover a complex integrable system in one of the complex
structures \cite{Donagi:1995cf}. On the one hand, it was discovered
by Nekrasov and Shatashvili \cite{Nekrasov:2009rc} that an
$\Omega$-deformation on a two-plane quantizes a real symplectic
submanifold of the complex integrable system. On the other hand, it
was found by Gaiotto, Moore and Neitzke \cite{Gaiotto:2010be} and Ito,
Okuda and Taki \cite{Ito:2011ea} that if the spacetime $\mathbb{R}^3 \times
S^1$ is replaced with a twisted product of $\mathbb{R}^3$ and $S^1$, then
supersymmetric loop operators form a noncommutative deformation of the
algebra of holomorphic functions on $\CM$ in other complex structures.
Despite the similarities between the two phenomena, explanations from
a unified point of view have been lacking. In this paper we provide
such explanations, based on a connection that we establish between a
deformation of $\CN = 4$ supersymmetric sigma model and quantization
of symplectic manifolds.
More precisely, the main result of the paper concerns the
topologically twisted version of the sigma model, known as
Rozansky-Witten theory \cite{Rozansky:1996bq}. We formulate it on a
three-manifold of the form $\mathbb{R} \times \Sigma$, with $\Sigma$ a Riemann
surface, taking the target space to be a hyperk\"ahler manifold $X$.
Given a complex structure on $X$ and a Killing vector field $V$ on
$\Sigma$, we construct a deformation of the theory analogous to the
$\Omega$-deformation in four dimensions. In the case that $\Sigma$ is
a disk $D$, we show that the $\Omega$-deformed Rozansky-Witten theory
is equivalent to a quantum mechanical system whose phase space is a
symplectic submanifold of $X$ determined by the boundary condition.
The phenomena exhibited by the four-dimensional gauge theory then
follow as special cases of this result, applied to low-energy
effective sigma models with target space $X = \CM$. The two cases
differ merely in the choice of complex structure.
While the four-dimensional phenomena are explained with a
three-dimensional theory, the construction of this theory is best
understood from a two-dimensional point of view. To formulate the
$\Omega$-deformation for Rozansky-Witten theory on $\mathbb{R} \times \Sigma$
with target space $X$, we view the theory as a B-twisted
Landau-Ginzburg model \cite{Vafa:1990mu} on $\Sigma$ whose target
space is the space of maps from $\mathbb{R}$ to $X$. For this reason we first
formulate the $\Omega$-deformation for general B-twisted
Landau-Ginzburg models, and then use this formulation to construct
the $\Omega$-deformed Rozansky-Witten theory.
The connection between the $\Omega$-deformed Rozansky-Witten theory
and quantization involves D-branes of novel type, which can be
introduced supersymmetrically in B-twisted Landau-Ginzburg models
only in the presence of the $\Omega$-deformation. An amusing fact is
that in general these branes are similar to A-branes, rather than
B-branes. There is an even more interesting analogy if we specialize
to the $\Omega$-deformed Rozansky-Witten theory. In this case,
relevant branes are much like $(A, B, A)$-branes in $\CN = (4,4)$
supersymmetric sigma models: the support of a brane is the space of
maps from $\mathbb{R}$ to a submanifold $L$ of $X$ that is Lagrangian with
respect to the K\"ahler forms $\omega_I$ and $\omega_K$ associated to
two complex structures $I$ and $K$ of the hyperk\"ahler structure of
$X$, while holomorphic in the third complex structure $J$. This
implies in particular that $L$ is a symplectic manifold with
symplectic form given by the restriction of $\omega_J$.
For $\Sigma = D$ with a brane of this type placed on the boundary, by
localization of the path integral we will derive a formula that
expresses a correlation function in terms of an integral over the
support of the brane. In the context of Rozansky-Witten theory, the
localization formula gives an integration over maps from $\mathbb{R}$ to $L$.
This is nothing but the path integral for a quantum mechanical system
on $L$. We therefore conclude that the $\Omega$-deformed
Rozansky-Witten theory quantizes the symplectic submanifold $(L,
\omega_J)$ of $X$. The algebra of observables is found to be a
noncommutative deformation of the algebra of functions on $L$ that are
restrictions of holomorphic functions on $X$.
The appearance of $(A,B,A)$-like branes in our framework suggests a
relation to another approach to quantization, namely the one using
A-branes, developed by Gukov and Witten~\cite{Gukov:2008ve}. Indeed,
the correspondence between $\Omega$-deformed $\CN = 2$ supersymmetric
gauge theories and quantum integrable systems was explained by
Nekrasov and Witten \cite{Nekrasov:2010ka} from this perspective.
(For explanations from other perspectives, see \cite{Bonelli:2011na,
Aganagic:2011mi}.) Their argument, however, relies on the fact that
the $\Omega$-deformation may be canceled away from the origin of the
two-plane by a redefinition of fields, and this makes the logic a
little involved. Our approach hopefully renders the connection
between the $\Omega$-deformation and quantization more transparent.
In this paper we apply our framework to two specific problems in
four-dimensional gauge theory. It will be interesting to find further
applications. For example, the quantization of Seiberg-Witten curves
proposed in \cite{Fucito:2011pn} may find a natural place in the
present framework. Besides, the framework itself may be generalized.
One direction in this regard would be to consider a gauged version of
Rozansky-Witten theory \cite{Kapustin:2009cd}, which is obtained by
topological twisting of an $\CN = 4$ supersymmetric gauge theory
constructed by Gaiotto and Witten \cite{Gaiotto:2008sd}.
Lastly, the $\Omega$-deformation of B-twisted theories should have a
broader range of applications. The construction can be extended to
include gauge theories (the details of which will appear elsewhere),
and this extension may shed light on the correspondence between $\CN =
2$ superconformal theories in three dimensions and analytically
continued Chern-Simons theory~\cite{Dimofte:2010tz, Terashima:2011qi,
Dimofte:2011jd, Dimofte:2011ju, Dimofte:2011py} via arguments along
the lines of \cite{Yagi:2013fda} (see also \cite{Lee:2013ida,
Cordova:2013cea}). Purely in two dimensions, it may prove fruitful
to study mirror symmetry between $\Omega$-deformed B-twisted theories
and $\Omega$-deformed A-twisted theories \cite{Shadchin:2006yz,
Dimofte:2010tz}.
The rest of the paper is organized as follows. In section~\ref{OBLG},
we formulate the $\Omega$-deformation of B-twisted Landau-Ginzburg
models and derive the localization formula. In section~\ref{QORW}, we
construct the $\Omega$-deformed Rozansky-Witten theory and establish
its connection to quantization. Section~\ref{4d} discusses the
applications to four-dimensional gauge theory. In the appendix we
review the $\Omega$-deformation of twisted $\CN = 2$ supersymmetric
gauge theories in four dimensions.
\section{\texorpdfstring{$\boldsymbol \Omega$}{Omega}-deformation of B-twisted Landau-Ginzburg models}
\label{OBLG}
In this section we formulate the $\Omega$-deformation of B-twisted
Landau-Ginzburg models in two dimensions, based on which the
$\Omega$-deformed Rozansky-Witten theory is constructed.
Furthermore, we study boundary conditions in the presence of the
$\Omega$-deformation, and derive a localization formula for
correlation functions on a disk. The results obtained here will be
essential in our discussion in the next section.
\subsection[\texorpdfstring{$\Omega$}{Omega}-deformed B-twisted Landau-Ginzburg models]{\texorpdfstring{$\boldsymbol \Omega$}{Omega}-deformed B-twisted Landau-Ginzburg models}
Let us recall how the $\Omega$-deformation in four dimensions works
\cite{Nekrasov:2002qd, Nekrasov:2003rj}. A topologically twisted $\CN
= 2$ supersymmetric gauge theory \cite{Witten:1988ze} has a single
scalar supercharge $Q$, satisfying the relation $Q^2 = 0$ modulo a
gauge transformation (as well as a flavor symmetry transformation if
hypermultiplet masses are nonzero \cite{Hyun:1995hz,
Labastida:1996tz}). This is used as a BRST operator, meaning that
physical operators and states are $Q$-cohomology classes. To
introduce an $\Omega$-deformation, one chooses a vector field $V$
generating an isometry of the spacetime four-manifold with respect to
a given metric. With this choice understood, the BRST operator of the
$\Omega$-deformed theory obeys the deformed relation
\begin{equation}
\label{Q2LV}
Q^2 = L_V,
\end{equation}
where $L_V$ is the conserved charge that acts on fields as the Lie
derivative $\CL_V$ by $V$. The construction of the $\Omega$-deformed
theory is reviewed in the appendix.
Similarly, a B-twisted Landau-Ginzburg model in two dimensions
\cite{Vafa:1990mu} has a BRST operator $Q$ satisfying $Q^2 = 0$ up to
a central charge. In order to formulate an $\Omega$-deformation of
this theory, we should therefore pick a Killing vector field $V$ on
the worldsheet $\Sigma$ equipped with a hermitian metric $h$, and
deform the theory so that the modified BRST operator obeys the
deformed relation \eqref{Q2LV}. This is what we are aiming for.%
\footnote{In \cite{Closset:2014pda}, a supergravity background was
found that realizes the $\Omega$-deformation of A-twisted theories
on $S^2$. As mentioned in that paper, one can combine it with a
$\mathbb{Z}_2$-action implementing mirror symmetry to obtain an
$\Omega$-deformation of B-twisted theories. Our formulation
presumably reproduces their results for $\Sigma = S^2$. I would
like to thank Stefano Cremonesi for explaining their work.}
How can we achieve such a deformation? To get the idea, consider the
simplest case $\Sigma = \mathbb{C}$. In this case the theory retains the full
$\CN = (2,2)$ supersymmetry in the twisted form, generated by two
scalar supercharges $\Qb_+$, $\Qb_-$ and a one-form supercharge $G =
G_z \mathrm{d} z + G_\zb \mathrm{d}\zb$. These satisfy $\{\Qb_-, G_z\} = P_z$ and
$\{\Qb_+, G_\zb\} = P_\zb$, where $P = P_z \mathrm{d} z + P_\zb \mathrm{d}\zb$ is
the generator of translations; the other commutators either vanish or
give central charges. The BRST operator of the undeformed theory is
$Q = \Qb_+ + \Qb_-$. If we take a vector field $V = V^z \partial_z +
V^\zb\partial_\zb$ with constant components $V^z$, $V^\zb$ and modify the
BRST operator to $Q = \Qb_+ + \Qb_- + \iota_V G$, then we obtain the
desired relation $Q^2 = \iota_V P$. We are going to generalize this
construction to an arbitrary worldsheet $\Sigma$.
The target space of the theory is a K\"ahler manifold $Y$.
Classically the following construction makes sense for any K\"ahler
target space, but for quantum anomalies to be absent, $Y$ has to be
Calabi-Yau. (This is essentially due to the fact that the axial
$\mathrm{U}(1)$ R-symmetry used in the B-twist is anomalous unless $c_1(Y) =
0$.) We denote the holomorphic and antiholomorphic tangent bundles of
$Y$ by $TY$ and $\overline{TY}$, respectively, and their duals with superscript
$\vee$. We pick a K\"ahler metric $g$ on $Y$.
The bosonic field of the theory is a map $\Phi\colon \Sigma \to Y$.
Given local coordinates on $Y$, we can express $\Phi$ locally by a set
of functions $(\phi^i, \phib^\ib)$. In the standard formulation, the
fermionic fields of the B-twisted theory are scalars $\eta$ with
values in $\overline{TY}$ and $\theta$ with values in $T^\vee Y$, and a
one-form $\rho$ with values in $TY$. In our construction, we use
instead of $\theta$ a two-form $\mu$ with values in $\overline{TY}$; the two
are related by the Hodge duality and the isomorphism between $\overline{TY}$
and $T^\vee Y$ induced by $g$. Thus the fermionic fields of the
theory are
\begin{equation}
\eta \in \Omega^0(\Sigma, \Phi^*\overline{TY}), \quad
\rho \in \Omega^1(\Sigma, \Phi^*TY), \quad
\mu \in \Omega^2(\Sigma, \Phi^*\overline{TY}).
\end{equation}
We also introduce auxiliary bosonic fields. They are two-forms $F$
with values in $TY$ and $\Fb$ with values in $\overline{TY}$:
\begin{equation}
F \in \Omega^2(\Sigma, \Phi^*TY), \quad
\Fb \in \Omega^2(\Sigma, \Phi^*\overline{TY}).
\end{equation}
Starting from $\CN = (2,2)$ supersymmetry transformation laws for
B-twisted chiral multiplets \cite{Labastida:1991qq}, it is
straightforward to write down the $\Omega$-deformed supersymmetry
transformation laws, following the same procedure as in the flat case
described above. After shifting $\Fb$ to absorb dependence on the
worldsheet metric, we have
\begin{equation}
\label{SUSY}
\begin{alignedat}{2}
\delta\phi^i &= \iota_V\rho^i, &
\delta\phib^\ib &= \eta^\ib,
\\
\delta_\nabla\rho^i
&= \mathrm{d}\phi^i + \iota_V F^i, &
\delta_\nabla\eta^\ib &= V(\phib^\ib),
\\
\delta_\nabla F^i
&= \mathrm{d}_\nabla\rho^i
+ \frac12 R^i{}_{j\kb l}\eta^\kb \rho^j \wedge \rho^l, &
\delta_\nabla\mu^\ib &= \Fb^\ib,
\\
& & \quad
\delta_\nabla\Fb^\ib
&= \mathrm{d}_\nabla \iota_V\mu^\ib
+ R^\ib{}_{\jb k\lb} \iota_V\rho^k \eta^\lb \mu^\jb.
\end{alignedat}
\end{equation}
Here $\delta_\nabla$ is the supersymmetry variation coupled to the
pullback of the Levi-Civita connection $\nabla$ of $g$; for example,
$\delta_\nabla \rho^i = \delta\rho^i + \delta\phi^k \Gamma^i_{kj}
\rho^j$. (More generally, $\nabla$ can be any torsion-free connection
on $TY$ whose curvature form $R$ is of type $(1,1)$ and obeys the
first Bianchi identity.) Notice the similarity to the supersymmetry
transformation laws \eqref{SUSY-4d} for $\Omega$-deformed theories in
four dimensions.
One can verify that the ``raw'' supersymmetry variation $\delta$
satisfies $\delta^2 = \CL_V$.%
\footnote{An easy way to see this is to define $G^i = F^i - \frac12
\Gamma^i_{kj} \rho^k \wedge \rho^j$ and $\Gb^\ib = \Fb^\ib -
\Gamma^\ib_{\kb\jb} \eta^\kb\mu^\jb$, in terms of which one has
$\delta\rho^i = \mathrm{d}\phi^i + \iota_V G^i$, \ $\delta G^i =
\mathrm{d}\rho^i$ and $\delta\mu^\ib = \Gb^\ib$, \ $\delta\Gb^\ib =
\mathrm{d}\iota_V\mu^\ib$.}
As usual, we let $Q$ denote the generator of the supersymmetry
variation $\delta$. Then it obeys the deformed relation \eqref{Q2LV},
as desired. Strictly speaking, on the right-hand side of this relation
may appear an extra conserved charge that commutes with fields and
coincides for $V = 0$ with a central charge of the B-twisted $\CN = 2$
supersymmetry algebra. Although such an extra term is important when
one considers the action of $Q$ on states, it plays no role in our
discussion and hence will be ignored.
For the moment we assume that $\Sigma$ has no boundary; we will
discuss boundary effects shortly. Then the $Q$-invariant action $S$
of the $\Omega$-deformed theory consists of two pieces, $S = S_0 +
S_W$. The first piece $S_0$ is $Q$-exact and contains kinetic terms.
The second piece $S_W$ is constructed from a superpotential $W\colon Y
\to \mathbb{C}$, which is a holomorphic function of $\phi^i$. Unlike $S_0$,
this one is not $Q$-exact.
Concretely, we can take $S_0$ to be
\begin{equation}
\label{S0}
S_0
= \delta\int_\Sigma \Bigl(g_{i\jb}\rho^i
\wedge \star\bigl(\mathrm{d}\phib^\jb + \iota_V\Fb^\jb\bigr)
+ g_{i\jb} F^i \wedge \star\mu^\jb\Bigr),
\end{equation}
while $S_W$ is given by
\begin{equation}
\label{SW}
S_W
= i\int_\Sigma\Bigl(
F^i \partial_i W
+ \frac{1}{2} \rho^i \wedge \rho^j \nabla_i\partial_j W
+ \delta\bigl(\mu^\ib \partial_\ib\Wb\bigr)
\Bigr).
\end{equation}
The $Q$-invariance of the action relies on the assumption that $V$ is
a Killing vector field. This ensures that $\CL_V$ commutes with
$\star$, so if we write the $Q$-exact part of the Lagrangian as
$\delta\CV$, then $\delta^2\CV = \CL_V\CV = \mathrm{d}\iota_V\CV$ by the
formula $\CL_V = \mathrm{d}\iota_V + \iota_V\mathrm{d}$.
Computing the supersymmetry variation we find
\begin{multline}
S_0
= \int_\Sigma \Bigl(
g_{i\jb} \bigl(\mathrm{d}\phi^i + \iota_VF^i\bigr)
\wedge \star\bigl(\mathrm{d}\phib^\jb + \iota_V\Fb^\jb\bigr)
+ g_{i\jb} F^i \wedge \star\Fb^\jb
- g_{i\jb} \rho^i \wedge \star\mathrm{d}_\nabla\eta^\jb
+ g_{i\jb} \mathrm{d}_\nabla\rho^i \wedge \star\mu^\jb \\
+ \frac12 R_{\ib j\kb l} \eta^\kb \rho^j \wedge \rho^l \wedge \star\mu^\ib
- \rho^i
\wedge \star\iota_V\bigl(g_{i\jb} \mathrm{d}_\nabla \iota_V\mu^\jb
+ R_{i\jb k\lb} \iota_V\rho^k \eta^\lb \mu^\jb\bigr)
\Bigr).
\end{multline}
For the superpotential terms we get
\begin{equation}
S_W
= i\int_\Sigma\Bigl(
F^i \partial_i W
+ \frac{1}{2} \rho^i \wedge \rho^j \nabla_i\partial_j W
+ \Fb^\ib \partial_\ib\Wb
+ \eta^\ib\mu^\jb\nabla_\ib\partial_\jb\Wb
\Bigr).
\end{equation}
Integrating out the auxiliary fields produces the potential
\begin{equation}
\label{F-term}
\star\|\mathrm{d} W\|^2 + \dotsb
= \star(g^{i\jb} \partial_i W \partial_\jb\Wb) + \dotsb,
\end{equation}
where we have abbreviated the terms involving $V$.
One can add more $Q$-exact terms to the Lagrangian if one wishes, such
as $\delta(g_{\ib j} \eta^\ib V(\phi^j))$ which produces the
$V$-dependent potential $\|V(\phi)\|^2$.
When $V = 0$, the supersymmetry transformation and the action
constructed above reduce to those of the ordinary B-twisted
Landau-Ginzburg model, up to the replacement of $\theta$ with $\mu$
explained above. This construction therefore defines a deformation of
the latter theory for $V \neq 0$.
The worldsheet metric $h$ appears in the action only through the Hodge
duality within the $Q$-exact piece $S_0$. Hence, the
$\Omega$-deformed theory is quasi-topological, in the sense that it is
invariant under deformations of the metric as long as $V$ remains to
be a Killing vector field. In addition, the theory is invariant under
overall rescaling of the target space metric $g$, as this leaves the
supersymmetry transformation invariant and the metric enters the
action through $S_0$.
The observables of the theory are the $Q$-closed operators that are
not $Q$-exact. At the zeros of $V$, any local observables of the
undeformed theory remain to be observables. There is a class of local
observables that are in one-to-one correspondence with the elements of
the ${\bar\partial}$-cohomology $H^{0,\bullet}(Y; \mathbb{C})$. To see this, note that
for $V = 0$, the action of $Q$ on $\phi^i$, $\phib^\ib$ and $\eta$
coincides with that of ${\bar\partial}$ under the identification of $\eta^\ib$
with $\mathrm{d}\phib^\ib$. Thus, if $\omega = \omega_{\ib_1 \dotso \ib_q}
\mathrm{d}\phib^{\ib_1} \wedge \dotsm \wedge \mathrm{d}\phib^{\ib_q}$ is a
${\bar\partial}$-closed $(0,q)$-form on $Y$, then $\omega_{\ib_1 \dotso \ib_q}
\eta^{\ib_1} \dotsm \eta^{\ib_q}$ inserted at a zero of $V$ is a
$Q$-closed local operator, and represents a nonzero $Q$-cohomology
class if and only if $\omega$ represents a nonzero element in
$H^{0,\bullet}(Y;\mathbb{C})$. In particular, holomorphic functions on $Y$
correspond to local observables.
\subsection{Incorporating boundaries}
So far we have assumed that $\Sigma$ has no boundary. Now we consider
the situation that $\Sigma$ has a boundary and the isometry generated
by $V$ restricts to an isometry of $\partial\Sigma$. In this situation
the $Q$-invariance of the action must be reexamined. Also, we have to
ask what sort of boundary conditions are physically sensible. In the
following we will express the $Q$-variation by $Q$-commutator,
reserving $\delta$ for arbitrary variation of fields in order to avoid
possible confusion.
Let us address the issue of $Q$-invariance. The $Q$-exact part of the
action remains to be $Q$-invariant in the presence of boundary. For,
if $\CV$ is a two-form, then
\begin{equation}
\int_\Sigma [Q, \{Q, \CV\}]
= \int_\Sigma (\mathrm{d}\iota_V + \iota_V\mathrm{d})\CV
= \int_{\partial\Sigma} \iota_V \CV
= 0.
\end{equation}
The last equality follows from the assumption that $V$ generates an
isometry of $\partial\Sigma$ and hence is tangent to $\partial\Sigma$. The
potential problem therefore comes from the non-$Q$-exact part.
Indeed, its $Q$-variation gives
\begin{equation}
i\int_{\partial\Sigma} \rho^i \partial_i W,
\end{equation}
breaking the $Q$-invariance by a boundary contribution.
We must somehow eliminate this boundary contribution to recover the
$Q$-invariance. In the case of ordinary B-twisted Landau-Ginzburg
models, one can do this by imposing a B-brane boundary condition.
This condition requires that $\Phi$ map $\partial\Sigma$ to a
submanifold~$\gamma$ of $Y$,
\begin{equation}
\Phi(\partial\Sigma) \subset \gamma,
\end{equation}
and $\mathrm{d} W|_\gamma = 0$, that is, $W$ be locally constant on
$\gamma$. The boundary contribution vanishes then.
In the $\Omega$-deformed case, there is another way of eliminating the
boundary contribution. For simplicity, suppose that there is only one
connected boundary component and it is compact. Let $\vphi$ be a
periodic coordinate on $\partial\Sigma$ such that $h_{\vphi\vphi}$ is
constant. In this coordinate,
\begin{equation}
V|_{\partial\Sigma} = \veps\partial_\vphi
\end{equation}
for some real constant $\veps$. Assuming that $\veps \neq 0$, we can
add to the action the boundary term
\begin{equation}
\label{BT}
-\frac{i}{\veps} \int_{\partial\Sigma} \mathrm{d}\vphi \, (W + W_0),
\end{equation}
where $W_0$ is a locally constant function on $\gamma$. The
$Q$-variation of this term cancels the boundary contribution in
question, recovering the $Q$-invariance of the action.
One interpretation of the above boundary term is that it is the action
for a theory living on the boundary, with $\veps$ being the Planck
constant. The undeformed limit $\veps \to 0$ is the classical limit,
and in this limit $\Phi$ obeys the equation of motion $\mathrm{d} W = 0$ on
the boundary, which reproduces the ordinary B-brane condition on $W$.
This mechanism of recovering the $Q$-invariance is interesting since
it is available only when the $\Omega$-deformation is turned on.
Moreover, it requires a weaker boundary condition on $W$ compared to
the B-brane condition. For the boundary term \eqref{BT} to not spoil
the convergence of the path integral, its real part had better be
nonnegative. For our purposes it is sufficient to consider the
situation that the boundary term is purely imaginary. To ensure this
property, we place on the boundary a brane supported on $\gamma$, and
impose
\begin{equation}
\label{ImdW=0}
\Im\mathrm{d} W|_\gamma = 0.
\end{equation}
Then, the constant imaginary part of $W$ can be absorbed into $W_0$,
and the boundary term can be written as
\begin{equation}
\label{BT-2}
-\frac{i}{\veps} \int_{\partial\Sigma} \mathrm{d}\vphi \, (\Re W + W_0),
\end{equation}
with $W_0$ now chosen to be real.
We would like to write down a set of boundary conditions that defines
this brane. A guiding principle for determining physically sensible
conditions is that in a weak coupling limit, solutions to equations of
motion should be saddle point configurations of the path integral. In
other words, when the fields are varied in that limit, boundary terms
should not arise in the variation of the action. Recalling that in
our case the bulk theory is invariant under rescaling of the target
space metric $g$, we see that there is a natural weak coupling limit,
namely the limit in which $g$ is rescaled by a large factor. This is
also the limit we will consider in the derivation of the localization
formula for correlation functions.%
\footnote{Actually we will consider a slightly different limit which
simplifies the analysis, but the two limits lead to the same
boundary condition.}
We therefore define our brane as the boundary condition
obtained by taking variations in this limit. (A similar choice was
made in \cite{Hori:2013ika} where $\CN = (2,2)$ supersymmetric
theories on a hemisphere were studied.)
Since $S_0$ dominates in the limit under consideration, we can ignore
the remaining part of the action in our analysis. Setting $\delta S_0
= 0$ and using the equations of motion for $F$ and $\Fb$ derived from
$S_0$, we find the constraints
\begin{equation}
g(\delta\Phi, \iota_V\star\mathrm{d}\Phi)|_{\partial\Sigma}
= g\bigl([Q,\delta\Phi], (\iota_V\star\rho, \star\mu)\bigr)|_{\partial\Sigma}
= 0.
\end{equation}
The boundary condition for $\Phi$ implies that any variation of $\Phi$
is tangent to $\gamma$ on $\partial\Sigma$, and we require that the same
be true for the $Q$-variation $[Q,\Phi]$; thus we have
\begin{equation}
\label{QPhi-in-Tgamma}
(\iota_V\rho, \eta) \in T_\mathbb{C}\gamma
\end{equation}
at each point on $\partial\Sigma$. Assuming that the variation of $\Phi$
is not constrained in any other way, we conclude that
\begin{equation}
\iota_V\star\mathrm{d}\Phi \in N_\mathbb{R}\gamma, \quad
(\iota_V\star\rho, \star\mu) \in N_\mathbb{C}\gamma,
\end{equation}
where $N_\mathbb{R}\gamma$ is the normal bundle of $\gamma$ and $N_\mathbb{C}\gamma$
is its complexification. In particular, $\Phi$ obeys the Neumann
boundary condition in the direction normal to the boundary, just as in
the case of ordinary B-branes (with vanishing $B$-field and Chan-Paton
gauge field).
Furthermore, in order for $Q$ to act on the space of allowed field
configurations, the boundary condition itself must be invariant under
the action of $Q$ (or more precisely, the covariant version of it,
coupled to the Levi-Civita connection). This leads to additional
constraints generated by repeated action of $Q$ on the constraints
described above. An example is the constraint \eqref{QPhi-in-Tgamma},
which comes from the D-brane constraint $\Phi(\partial\Sigma) \subset
\gamma$. This procedure generates only a few new constraints since
$Q^2 = L_V$ leaves invariant the space of sections of a vector bundle
over $\partial\Sigma$. It turns out that these additional constraints
follow from the constraints discussed already if we use the equations
of motion derived from the quadratic part of $S_0$.
The above boundary condition is independent of $g$, thanks to the
limit considered here which decouples dependence on the
superpotential. It is also independent of the component of the
worldsheet metric $h$ normal to $\partial\Sigma$. (If $n$ is a coordinate
in the normal direction, we have $\iota_V\star\mathrm{d}\Phi |_{\partial\Sigma}
= \veps \sqrt{h_{\vphi\vphi}/h_{nn}} (\partial_n\phi^i, \partial_n\phib^\ib)$
and $(\iota_V\star\rho, \star\mu) |_{\partial\Sigma} =
\sqrt{h_{\vphi\vphi}/h_{nn}} (\veps \rho^i_n, h^{\vphi\vphi}
\mu^\ib_{n\vphi})$.) Hence, the invariance of the bulk theory under
relevant deformations of the metrics is mostly preserved by the brane,
broken only by the explicit dependence on $h_{\vphi\vphi}$.
\subsection{Localization}
As in ordinary A- and B-twisted theories, the path integral for a
correlation function of $Q$-invariant operators in the
$\Omega$-deformed B-twisted Landau-Ginzburg model reduces to an
integral over a small subspace of the field space. Let us derive a
formula for correlation functions in the case that $\Sigma$ is a disk
$D$, with a brane of the above type placed on the boundary.
We equip $D$ with the metric of the form $h = h_{rr}(r) \mathrm{d} r^2 +
h_{\vphi\vphi}(r)\mathrm{d}\vphi^2$, where $(r, \varphi)$ are polar
coordinates. Then $V = \veps \partial_\varphi$, with $\veps$ constant.
Here $\veps$ is real, but it is also possible to make it complex since
$V$ only needs to satisfy the Killing equation.
Our theory is invariant under rescaling of the target space metric
$g$. In particular, we can rescale it as $g \to t^2 g$ and take the
limit $t \to \infty$. Integrating out the auxiliary fields, we find
that in this limit the action diverges away from the locus where
\begin{equation}
\mathrm{d}\phi^i = 0.
\end{equation}
The path integral therefore localizes to the constant maps, that is to
say, receives contributions solely from an arbitrarily small
neighborhood of the subspace of constant maps in the space of maps
from $D$ to $Y$.
Such a neighborhood may be thought of as a fibration over the space of
constant maps. By the boundary condition, the constant maps are
required to map into $\gamma \subset Y$, the support of the brane.
Thus the base is isomorphic to $\gamma$. The fiber consists of the
bosonic fluctuation $\vphi$ around a constant map. We can extend the
fiber so that it includes the fermionic fields as well. The path
integral is then an integral over the total space of the extended
fibration. What we want to do now is to perform the integration over
the fiber, and reduce the path integral to an integral over the base.
It may be helpful to recall how the fiber integration is done in a
simpler setting where $V = 0$ and $\Sigma$ has no boundary. We
combine $\eta$ and $\mu$ into a single field $\zeta = -\eta + \mu$
which is an even-degree form with values in $\overline{TY}$. To quadratic
order, the part of the action relevant for large $t$ can be written as
\begin{equation}
\label{S-quad}
t^2\bigl(\langle\vphi, \Delta \vphi\rangle
+ \langle (\mathrm{d}_\nabla + \mathrm{d}_\nabla^*)\rho, \zeta\rangle
+ \langle\rho, (\mathrm{d}_\nabla + \mathrm{d}_\nabla^*)\zeta\rangle\bigr).
\end{equation}
Here $\Delta = (\mathrm{d}_\nabla + \mathrm{d}_\nabla^*)^2$, and $\langle\cdot,
\cdot\rangle$ is an inner product defined by the metric on $\Sigma$
and the original metric on $Y$ before the rescaling. We expand
$\vphi$, $\rho$ and $\zeta$ in orthonormal bases of eigenmodes of
$\Delta$, and express the quadratic part of the action in terms of the
expansion coefficients. The integration over the fiber is integration
over these coefficients. We can rescale the coefficients by $1/t$ to
absorb the overall $t^2$ factor in the quadratic part. Provided that
there are no fermion zero modes, after doing so the terms of higher
order are suppressed by inverse powers of $t$. In the limit $t \to
\infty$, the fiber integration produces the ratio of the bosonic and
fermionic one-loop determinants, $\prod_\beta
\sqrt{\lambda'_\beta}/\prod_\alpha \lambda_\alpha$, where
$\lambda_\alpha$ are nonzero eigenvalues for $\vphi$, and
$\lambda'_\beta$ are those for $\rho$ and $\zeta$. The nonzero
eigenvalues for $\rho$ and $\zeta$ agree since their nonzero modes are
related by the action of $\mathrm{d}_\nabla + \mathrm{d}_\nabla^*$, which commutes
with $\Delta$. Similarly, as $\Delta$ and $\star$ commute, the
nonzero modes for zero- and two-forms are related by the Hodge duality
and have the same eigenvalues. This means that the set
$\{\lambda'_\beta\}$ consists of two copies of the set
$\{\lambda_\alpha\}$. Hence, this ratio is equal to $1$, and the
fiber integration is trivial in this case.
We want to carry out a similar computation in the case at hand, where
$V$ generates rotations of $\Sigma = D$. Here the analysis is a bit
more complicated.
One complication is that if $V \neq 0$, the action contains additional
terms and they modify the quadratic part. To simplify the analysis,
we replace the $Q$-exact piece $S_0$ of the action with
\begin{equation}
t^2 \delta\int_\Sigma \Bigl(g_{i\jb}\rho^i
\wedge \star\bigl(\mathrm{d}\phib^\jb + \iota_V\Fb^\jb\bigr)
+ s g_{i\jb} F^i \wedge \star\mu^\jb\Bigr),
\end{equation}
rescale $\mu \to \mu/s$, and take the limit $s \to \infty$.
Integrating out the auxiliary fields and performing integration by
parts using the boundary condition, we find that the relevant part of
the on-shell action now takes the identical form \eqref{S-quad} as in
the case with $V = 0$.
Another complication comes from the presence of boundary, which makes
the analysis of mode expansion more difficult. We can deal with this
problem as follows. Using the freedom of deforming the worldsheet
metric, we can choose $D$ to have the shape of a sphere $S^2$ with a
small hole; in spherical coordinates $(\vtheta, \vphi)$, the metric
takes the form $h = R^2(\mathrm{d}\vtheta^2 + \sin^2\vtheta\mathrm{d}\vphi^2)$,
with $\vtheta$ ranging from $0$ to some value $\vtheta_{\partial D}$ where
the boundary is located. (Our boundary condition depends on
$h_{\vphi\vphi}$. It should be given with respect to the original
metric and fixed throughout the deformation.) Then we take the limit
$\vtheta_{\partial D} \to \pi$. In this limit $D$ becomes the whole
$S^2$, and the boundary state gets mapped to a $Q$-invariant local
operator inserted at $\vtheta = \pi$. If we now expand the fields in
the eigenmodes of $\Delta$ on $S^2$, then in terms of the expansion
coefficients the quadratic part of the action has the same expression
as before.
The conclusion is therefore that the fiber integration is again exact
at one loop and produces a factor similar to the ratio of the bosonic
and fermionic determinants, assuming that there are no fermion zero
modes. The difference is that this time the path integral receives
contributions only from the locus where the expansion coefficients
obey various relations, imposed by the boundary state or the operator
inserted at $\vartheta = \pi$.
The question is whether this one-loop factor depends on the background
constant map $\Phi_0$ around which we are expanding. If it does, the
dependence should come from $g$, $W$ or $\gamma$, since these are the
only objects defined on the target space that enter our setup. The
one-loop computation refers to just the quadratic part of the action
in the limit $t \to \infty$, and this is independent of $W$. It is
also independent of $g$ if we use holomorphic normal coordinates
centered at $\Phi_0$, in which $g_{i\jb}(\Phi_0) = \delta_{ij}$ and
$\partial_k g_{i\jb}(\Phi_0) = \partial_\kb g_{i\jb}(\Phi_0) = 0$. In fact,
in these coordinates the quadratic part takes exactly the same form as
the action for an affine target space $\mathbb{C}^n$ with the standard metric.
This leaves $\gamma$ as the only possible source for nontrivial
dependence on the background. Indeed, the choice of $\gamma$ may
introduce such dependence, since it specifies the boundary condition
which in turn determines the relations among the mode expansion
coefficients. Put another way, the one-loop factor is independent of
the background if we can choose $\gamma$ in such a way that the
boundary condition is the same for all backgrounds. In view of the
fact that the boundary condition in a background $\Phi_0 \in \gamma$
is determined by the tangent and normal spaces of $\gamma$ at
$\Phi_0$, a sufficient condition for background independence is that
the tangent spaces (and hence also the normal spaces) at any two
points of $\gamma$ can be made identical, when regarded as subspaces
of $\mathbb{C}^n$ via a suitable choice of normal coordinate systems centered
at these points. Taking into account the freedom in choosing normal
coordinates, we see that the two spaces need to be identical up to an
action of $\mathrm{U}(n)$.
One way to satisfy this condition is to take $\gamma$ to be a complex
submanifold of $Y$, as in the case for ordinary B-branes. In this
case the superpotential $W$ restricts to a holomorphic function on
$\gamma$. As we require $\Im\mathrm{d} W|_\gamma = 0$, $W$ would then have
to be locally constant on $\gamma$. This is in fact the ordinary
B-brane condition on $W$. However, it is not desirable for our
purposes. We would like to view $W$ as the Lagrangian of a boundary
theory, from which the equation $\mathrm{d} W = 0$ follows as a classical
equation of motion.
A more interesting possibility is to take $\gamma$ to be a Lagrangian
submanifold of $Y$ with respect to the K\"ahler form. The condition
for background independence is then satisfied since $\mathrm{U}(n)$ acts
transitively on the Lagrangian Grassmannian $\mathrm{U}(n)/\mathrm{O}(n)$, the space
of Lagrangian subspaces of $\mathbb{R}^{2n}$. From now on we will consider
this kind of supports.
Finally, we have to make sure that the assumption of absence of
fermion zero modes is actually true. Since the result of the path
integral is independent of the size of the $S^2$, we will show this in
the limit where the $S^2$ is very small. First of all, there are no
harmonic one-forms on $S^2$ and hence no zero modes for $\rho$. The
zero modes of $\eta$ are constants, while those of $\mu$ are their
Hodge duals. For these modes, we have to look at the constraints
imposed by the boundary condition. In the limit we are considering,
the nonzero modes are very massive and decouple, so the fermions can
be replaced by their zero mode parts. Then the boundary condition
forces $(0, \eta) \in T_\mathbb{C} \gamma$ and $(0, \star\mu) \in N_\mathbb{C}
\gamma$. Since $\gamma$ is a Lagrangian submanifold of a K\"ahler
manifold, we have $I(T_\mathbb{R} \gamma) = N_\mathbb{R} \gamma$ and it follows that
$\eta = \mu = 0$ on the boundary. Hence, the zero modes of $\eta$ and
$\mu$ are actually identically zero in this limit.
We have found that the fiber integration just produces an irrelevant
constant. The remaining step in the path integral is to integrate
over all background constant maps. For a constant map $\Phi_0$, the
action is evaluated as
\begin{equation}
S(\Phi_0) = -\frac{2\pi i}{\veps} \bigl(\Re W + W_0\bigr)(\Phi_0).
\end{equation}
Altogether, we conclude that the path integral for the
$\Omega$-deformed B-twisted Landau-Ginzburg model on a disk reduces
to the form
\begin{equation}
\label{LF}
\vev{\CO}
= \int_\gamma \mathrm{d}\Phi_0
\exp\Bigl(\frac{2\pi i}{\veps} \bigl(\Re W + W_0\bigr)(\Phi_0)\Bigr)
\CO(\Phi_0),
\end{equation}
where the operator insertion on the right-hand side is evaluated for
constant maps $\Phi_0 \in \gamma$, with fermions set to zero. In this
expression we have renormalized $W_0$ to absorb the one-loop factor.
\section{Quantization via the
\texorpdfstring{$\boldsymbol\Omega$}{Omega}-deformed Rozansky-Witten theory}
\label{QORW}
Having constructed $\Omega$-deformed B-twisted Landau-Ginzburg
models, we now move up one dimension higher and formulate the
$\Omega$-deformation of Rozansky-Witten theory in three dimensions.
We will then establish, via localization of the path integral, the
connection between the $\Omega$-deformed Rozansky-Witten theory and
quantization of symplectic submanifolds of the hyperk\"ahler target
space.
\subsection{Rozansky-Witten theory}
To begin, let us review the basic aspects of Rozansky-Witten theory.
We refer the reader to the original paper \cite{Rozansky:1996bq} for
more details.
Rozansky-Witten theory is a three-dimensional supersymmetric sigma
model which can be defined on a general three-manifold $M$. The
target space of the theory is a complex symplectic manifold $(X,
\Omega)$. It is a complex manifold $X$ equipped with a nondegenerate
closed holomorphic two-form $\Omega$, called a holomorphic symplectic
form.
Let $\Phi\colon M \to X$ be the bosonic map of the sigma model, and
write $(\phi^i, \phib^\ib)$ for a local expression of $\Phi$. The
theory has two fermionic fields, $\eta$ and $\chi$, and is invariant
under the following supersymmetry:
\begin{equation}
\label{Q-RW}
\begin{alignedat}{2}
\delta\phi^i &= 0, &\qquad
\delta\phib^\ib &= \eta^\ib,
\\
\delta\chi^i &= \mathrm{d}\phi^i, &
\delta\eta^\ib &= 0.
\end{alignedat}
\end{equation}
As can be seen from the transformation laws, $\eta$ is a scalar on $M$
with values in $\overline{TX}$, and $\chi$ is a one-form on $M$ with
values in $TX$. The generator $Q$ of the supersymmetry satisfies $Q^2
= 0$. We use it as a BRST operator, declaring that physical operators
and states are $Q$-cohomology classes.
To construct a $Q$-invariant action we need to make some choices. We
pick a Riemannian metric $h$ on $M$ and a hermitian metric $g$ on $X$.
In addition, we choose a torsion-free connection $\nabla = \mathrm{d} +
\Gamma$ on $TX$, with connection matrices $\Gamma^i_{kj} =
\Gamma^i_{jk}$. The $(1,1)$-part of the curvature form of $\nabla$
(given by the matrix elements $R^i{}_{jk\lb} \mathrm{d}\phi^k \wedge
\mathrm{d}\phib^\lb$ with $R^i{}_{j\lb k} = \partial_\lb\Gamma^i_{kj}$)
represents a ${\bar\partial}$-cohomology class, known as the Atiyah class of
$X$. It is the obstruction to the existence of a holomorphic
connection on $TX$.
The action is the sum of two pieces, $S = S_1 + S_2$. The first piece
is $Q$-exact:
\begin{equation}
\label{S1}
S_1 = \delta\int_M
g_{i\jb} \chi^i \wedge \star\mathrm{d}\phib^\jb
= \int_M\bigl(g_{i\jb} \mathrm{d}\phi^i \wedge \star\mathrm{d}\phib^\jb
- g_{i\jb} \chi^i \wedge \star\mathrm{d}_{\widetilde\nabla}\eta^\jb\bigr).
\end{equation}
Here the connection $\widetilde\nabla = \mathrm{d} + \Gammat$ is defined by
$(\Gammat_\kb)^\ib{}_j = g^{\ib l} \partial_j g_{l\kb}$; if $g$ is
K\"ahler, $\widetilde\nabla$ is the Levi-Civita connection of $g$. The second
piece is $Q$-invariant, but not $Q$-exact:
\begin{equation}
\label{S2}
S_2
= -\frac{i}{4} \int_M\Bigl(
\Omega_{ij} \chi^i \wedge \mathrm{d}_\nabla\chi^j
- \frac{1}{3} \Omega_{ij} R^j{}_{kl\mb}
\chi^i \wedge \chi^k \wedge \chi^l \eta^\mb
+ \frac{1}{3} \nabla_k \Omega_{ij}
\mathrm{d}\phi^i \wedge \chi^j \wedge \chi^k\Bigr).
\end{equation}
The particular normalization is chosen for later convenience.
Since the spacetime metric $h$ appears only in the $Q$-exact part
$S_1$ through the Hodge duality, the theory is topological. Likewise,
the target space metric $g$ appears only in $S_1$, so the theory is
independent of the choice of $g$. It turns out that the theory is
also independent of the choice of the connection $\nabla$, different
choices leading to the same expression for $S_2$ modulo $Q$-exact
terms.
The local observables of Rozansky-Witten theory are in one-to-one
correspondence with the ${\bar\partial}$-cohomology classes of $X$ under the
identification of $\eta^\ib$ with $\mathrm{d}\phib^\ib$. There are also
nonlocal observables. Given a connection $A$ of type $(1,0)$ on any
holomorphic $G$-bundle $E \to X$, we define a $Q$-invariant connection
\begin{equation}
\CA = A_i \mathrm{d}\phi^i + F_{i\jb} \chi^i\eta^\jb,
\end{equation}
where $F$ is the curvature of $A$. Using this connection we can
construct a $Q$-invariant loop operator
\begin{equation}
\mathop{\mathrm{Tr}}\nolimits P\exp\Bigl(\oint_\CC \CA\Bigr),
\end{equation}
by taking the trace of the holonomy of $\CA$ along a closed path $\CC$
in $M$.
A special case of interest is when $X$ admits a hyperk\"ahler
structure $(g, I, J, K)$. In this case, $X$ has a two-sphere $\P^1$
of complex structures, and $g$ is K\"ahler with respect to all of
them. Elements of the $\P^1$ are linear combinations $aI + bJ + cK$,
with $a^2 + b^2 + c^2 = 1$ and $I$, $J$, $K$ satisfying the quaternion
relations
\begin{equation}
I^2 = J^2 = K^2 = IJK = -1.
\end{equation}
If we write $\omega_J$ and $\omega_K$ for the K\"ahler forms
associated to $J$ and $K$, respectively, then
\begin{equation}
\label{Omega_I}
\Omega_I = \omega_J + i\omega_K
\end{equation}
is a holomorphic symplectic form in complex structure $I$. Thus we
can regard $X$ as a complex symplectic manifold with complex
symplectic structure $(I, \Omega_I)$, and construct Rozansky-Witten
theory, choosing $\nabla$ to be the Levi-Civita connection of $g$.
Of course, which complex structure to call $I$ is just a matter of
convention, and any other complex structure in the $\P^1$ gives an
equally good target space. In other words, there is a family of
target spaces parametrized by the $\P^1$. We can continuously change
the theory by moving within this family. Equivalently, we may fix the
target space and vary the BRST operator. In the hyperk\"ahler case
the theory has a second supercharge $\Qb$ which generates the
transformations
\begin{equation}
\label{Qb-RW}
\begin{alignedat}{2}
\deltab\phi^i &= \Omega^{ij} g_{j\kb} \eta^\kb, &\qquad
\deltab\phib^\ib &= 0,
\\
\deltab\chi^i
&= -\Omega^{ij} g_{j\kb} \mathrm{d}\phib^\kb
- \Gamma^i_{kj} \Omega^{kl} g_{l\mb} \eta^\mb \chi^j, &
\deltab\eta^\ib &= 0,
\end{alignedat}
\end{equation}
where $\Omega^{ij} \Omega_{jk} = \delta^i_k$. As $\Qb$ squares to
zero and commutes with $Q$, any linear combination $Q_\zeta \propto Q
+ \zeta\Qb$ with $\zeta \in \P^1$ serves as a BRST operator. (The
$Q$-exact part $S_1$ of the action is also $Q_\zeta$-exact since
$g_{i\jb} \chi^i \wedge \star\mathrm{d}\phib^\jb = \deltab(\Omega_{ij}
\chi^i \wedge \star\chi^j/2)$ and $\delta\deltab = (\delta +
\zeta\deltab) \deltab$. Thus, the topological invariance of the
theory remains to hold.) We see that for $\zeta \neq 0$, $Q_\zeta$
annihilates holomorphic functions on $X$ in a complex structure
different from $I$, so varying the BRST operator indeed amounts to
changing the complex structure of the target space.
\subsection[\texorpdfstring{$\Omega$}{Omega}-deformation of
Rozansky-Witten theory]{\texorpdfstring{$\boldsymbol \Omega$}{Omega}-deformation of
Rozansky-Witten theory}
\label{ORW}
Now let us formulate the $\Omega$-deformation of Rozansky-Witten
theory. Our goal is the following. Consider Rozansky-Witten theory
on $M = \mathbb{R} \times \Sigma$, equipped with a product metric $h = h_\mathbb{R}
\oplus h_\Sigma$. Assume that the target space $X$ is a hyperk\"ahler
manifold with hyperk\"ahler metric $g$. Given a vector field $V$
generating an isometry of $\Sigma$, we wish to construct a deformation
of this theory such that it has a supercharge $Q$ obeying the deformed
relation~\eqref{Q2LV}.
There are a couple of indications that such a deformation does exist.
One is that, as we will explain in section~\ref{RW4d},
Rozansky-Witten theory with hyperk\"ahler target space arises
naturally from $\CN = 2$ supersymmetric gauge theories in four
dimensions by compactification on $S^1$. If we turn on an
$\Omega$-deformation in four dimensions, some deformation should be
induced in three dimensions as well. Another indication is that
reduction of Rozansky-Witten theory on $S^1$ gives the B-model with
the same target space \cite{Thompson:1998vx}. (More generally, if
the target space is not hyperk\"ahler, the dimensional reduction
yields a generalization of the B-model.) As we could construct the
$\Omega$-deformation in two dimensions, it is natural to expect that
there is a corresponding deformation in three dimensions. We take the
second observation as a starting point of our construction.
Our strategy is to describe the $\Omega$-deformed Rozansky-Witten
theory on $\mathbb{R} \times \Sigma$ with target space $X$ as an
$\Omega$-deformed B-twisted Landau-Ginzburg model on $\Sigma$ with
target space $Y = \mathop{\mathrm{Map}}\nolimits(\mathbb{R}, X)$, the space of maps from $\mathbb{R}$ to $X$.
In order to specify the latter theory, we need to pick a complex
structure on $Y$. Such a complex structure is naturally induced from
a complex structure chosen on $X$. This construction therefore
singles out a distinguished element in the $\P^1$ of complex
structures on $X$. We call it $I$.
Roughly speaking, having $\mathop{\mathrm{Map}}\nolimits(\mathbb{R}, X)$ as the target space means that
we should regard a coordinate $t$ of the $\mathbb{R}$ as a continuous index,
putting it on the same footing as coordinate indices of $X$. Hence,
the formula for the action \eqref{S0} now contains an integration over
$t$ in addition to summation over the other indices:
\begin{equation}
\label{S0-RW}
S_0
=
\delta \int_{\mathbb{R} \times \Sigma} \sqrt{h_\mathbb{R}} \mathrm{d} t \wedge
\Bigl(g_{i\jb}\rho^i
\wedge \star_\Sigma\bigl(\mathrm{d}_\Sigma\phib^\jb + \iota_V\star_\Sigma\Fb^\jb\bigr)
+ g_{i\jb} F^i \wedge \star_\Sigma\mu^\jb\Bigr).
\end{equation}
Here $\star_\Sigma$ and $\mathrm{d}_\Sigma$ denote the Hodge star operator
and the exterior derivative (coupled to the Levi-Civita connection) on
$\Sigma$. The supersymmetry transformation laws take the same
form~\eqref{SUSY} as before, the only difference being that the fields
have dependence on~$t$.
The above action lacks terms involving $t$-derivatives, which are
crucial for fully three-dimensional dynamics. These missing terms are
to be provided by a superpotential $W$. In the present context, $W$
is a holomorphic function on $\mathop{\mathrm{Map}}\nolimits(\mathbb{R}, X)$ that respects locality,
namely a holomorphic functional of a map from $\mathbb{R}$ to $X$.
To construct a suitable superpotential, we complete the complex
structure $I$ into a triple $(I, J, K)$ compatible with the
hyperk\"ahler structure, and set $\Omega = \Omega_I$. Since $\Omega$
is a closed holomorphic two-form, we can locally write $\Omega =
\mathrm{d}\Lambda$ with some holomorphic one-form $\Lambda$. Using this
form we define $W$ by
\begin{equation}
\label{W-RW}
W(\Phi) = \frac12 \int_\mathbb{R} \Phi^*\Lambda.
\end{equation}
In this formula we have abused the notation and let $\Phi$ denote a
point on the target space $\mathop{\mathrm{Map}}\nolimits(\mathbb{R}, X)$, as is customary in the
finite-dimensional setting. With this choice of $W$, the
superpotential terms \eqref{SW} are given by
\begin{equation}
\label{SW-RW}
S_W
= \frac{i}{2} \int_{\mathbb{R} \times \Sigma}
\Bigl(\Omega_{ij} F^i \mathrm{d}_\mathbb{R}\phi^j
- \frac{1}{2} \Omega_{ij} \rho^i \wedge \mathrm{d}_\mathbb{R}\rho^j
+ \Omegab_{\ib\jb} \Fb^\ib \mathrm{d}_\mathbb{R}\phib^\jb
+ \Omegab_{\ib\jb} \eta^\ib \mathrm{d}_\mathbb{R}\mu^\jb\Bigr).
\end{equation}
When $V = 0$, the theory described by the action $S_0 + S_W$ reduces
on-shell to Rozansky-Witten theory. Integrating out the auxiliary
fields sets
\begin{equation}
\sqrt{h_\mathbb{R}} \star_\Sigma F^i
= -\frac{i}{2} g^{i\jb} \Omegab_{\jb\kb} \partial_t\phib^\kb,
\quad
\sqrt{h_\mathbb{R}} \star_\Sigma\Fb^\ib
= -\frac{i}{2} g^{\ib j} \Omega_{jk} \partial_t\phi^k.
\end{equation}
From the expression of $\Fb^\ib$, one readily sees that for $V = 0$,
the supersymmetry transformation laws \eqref{SUSY} coincide with the
formula \eqref{Q-RW} under the identification
\begin{equation}
\rho^i_\mu = \chi^i_\mu, \quad
\sqrt{h_\mathbb{R}} \star_\Sigma\mu^\ib
= -\frac{i}{2} g^{\ib j} \Omega_{jk} \chi^k_t .
\end{equation}
One can also check that the action coincides with the Rozansky-Witten
action under this identification.%
\footnote{In checking this, one uses the identity $g^{i\jb}
\Omega_{ik} \Omegab_{\jb\lb} = 4g_{k\lb}$, which follows from the
equation $\Omega = \omega_J + i\omega_K = -g(J + iK)$, and the fact
that $\Omega$ is covariantly constant, which in particular implies
$\Omega_{ik} R^k{}_{jl\mb} + \Omega_{kj} R^k{}_{il\mb} = 0$.}
For example, the F-term potential is
\begin{equation}
\|\delta W\|^2
= \frac14 \int_\mathbb{R} \mathrm{d} t \sqrt{h_\mathbb{R}} h^{tt}
g^{i\jb} \Omega_{ik} \partial_t\phi^k
\Omegab_{\jb\lb} \partial_t\phib^\lb
= \int_\mathbb{R} \mathrm{d} t \sqrt{h_\mathbb{R}}
g_{i\jb} h^{tt} \partial_t\phi^i \partial_t\phib^\jb,
\end{equation}
and this is precisely the kinetic term for the bosonic field along the
$t$-direction. (Matching of the fermionic terms is straightforward.)
Thus, the theory constructed above provides a good definition for the
$\Omega$-deformation of Rozansky-Witten theory on $\mathbb{R} \times \Sigma$
with target space $X$ and complex symplectic structure $(I,
\Omega_I)$.
There is a slight generalization of this construction. It is possible
to modify the definition of $W$ by terms that vanish for $V = 0$.
Locality requires that this is done through a deformation of $\Lambda$
by a locally-defined holomorphic one-form on $X$, which in turn gives
rise to a deformation of $\Omega$ by the equation $\Omega =
\mathrm{d}\Lambda$. The latter is what really matters as far as the effect
on the theory is concerned. After a deformation of this type is
included, generically $\Omega$ would remain nondegenerate and define a
deformed complex symplectic structure. However, it may no longer be a
holomorphic symplectic form associated to some complex structure in a
hyperk\"ahler structure. This generalization will not be considered
in what follows, except that we will briefly discuss its relevance in
applications to $\CN = 2$ supersymmetric gauge theories in four
dimensions.
\subsection{Reduction to quantum mechanics}
Finally we are ready to present the main result of this paper.
Suppose that $\Sigma$ is a disk $D$ and $V$ generates its rotations.
In this situation we can apply the localization formula for
correlation functions obtained in the previous section. Using the
formula, we show that this system is equivalent to a quantum
mechanical system on a real symplectic submanifold of $X$.
First of all, we have to specify the support $\gamma$ of the brane
placed on the boundary of $D$. We recall that $\gamma$ is a
Lagrangian submanifold of $Y = \mathop{\mathrm{Map}}\nolimits(\mathbb{R}, X)$, and $\Im W$ must be
locally constant on $\gamma$. The first condition suggests that we
should take $\gamma = \mathop{\mathrm{Map}}\nolimits(\mathbb{R}, L)$, with $L$ being a Lagrangian
submanifold of $X$ with respect to the K\"ahler form $\omega_I$. The
second condition says that we must have
\begin{equation}
\delta\Im W
= \frac12 \int_\mathbb{R} \Im\Omega_{ij} \delta\phi^i \mathrm{d}\phi^j
= 0
\end{equation}
on the boundary. This is satisfied if $L$ is Lagrangian with respect
to $\Im\Omega = \omega_K$. Then, both $I$ and $K$ give an isomorphism
between $T_\mathbb{R} L$ and $N_\mathbb{R} L$, and $J = KI$ is an endomorphism of
$T_\mathbb{R} L$. Hence, $L$ is a Lagrangian submanifold with respect to
$\omega_I$ and $\omega_K$, while a complex submanifold in $J$. In
particular, it is a symplectic manifold with symplectic form
$\Re\Omega = \omega_J$.
We also need to specify the locally constant function $W_0$ on
$\gamma$, which is part of the boundary term. For this one, we
introduce a $\mathrm{U}(1)$ connection $A$ on a line bundle over $L$ and set
\begin{equation}
\label{W0-RW}
W_0 = \frac12 \int_\mathbb{R} \Phi^*A.
\end{equation}
For $W_0$ to be locally constant, $A$ must be flat.
For simplicity, let us assume for a moment that $\Omega$ is an exact
form so that a holomorphic one-form $\Lambda$ satisfying $\Omega =
\mathrm{d}\Lambda$ exists globally on $X$. Then $W$ is given by the
integral \eqref{W-RW}, and the localization formula \eqref{LF} reads
\begin{equation}
\vev{\CO}
= \int_{\mathop{\mathrm{Map}}\nolimits(\mathbb{R}, L)} \cD\Phi_0
\exp\Bigl(\frac{i}{\hbar} S(\Phi_0)\Bigr) \CO(\Phi_0),
\end{equation}
where the action $S$ and the Planck constant $\hbar$ are given by
\begin{equation}
\label{SQM}
S(\Phi_0) = \int_\mathbb{R} \Phi_0^*(\Re\Lambda + A), \quad
\hbar = \frac{\veps}{\pi}.
\end{equation}
The right-hand side of the formula is the path integral for a quantum
mechanical system with phase space $(L, \Re\Omega)$; in local Darboux
coordinates $(p_a, q^a)$, $a = 1$, $\dotsb$, $\frac12\dim L$ such that
$\Re\Omega|_L = \mathrm{d} p_a \wedge \mathrm{d} q^a$, we have $\Re\Lambda|_L + A
= p_a \mathrm{d} q^a$ up to an exact form, so the Lagrangian is $p_a \dot
q^a$ up to a total derivative. Therefore, this system quantizes the
symplectic manifold $(L, \Re\Omega)$. Notice that the Hamiltonian of
the system is zero, as is consistent with the fact that we started
with a (quasi-)topological field theory.
Observables of the $\Omega$-deformed Rozansky-Witten theory include
local operators that may be regarded as elements of the
${\bar\partial}$-cohomology $H^{0,\bullet}(X; \mathbb{C})$ in complex structure $I$,
inserted at the origin of $D$ and arbitrary points on the $\mathbb{R}$. Among
these observables, those that are nonvanishing after the localization
are holomorphic functions on $X$. The path integral turns these
observables into operators in the quantum mechanical system which form
a noncommutative algebra. They act on the Hilbert space whose
elements are, say, functions of $q^a$ locally. What we have obtained
is thus a noncommutative deformation of the algebra of functions on
$L$ that are restrictions of holomorphic functions on $X$, acting on
the space of sections of a hermitian line bundle over $L$.
It is worth noting that the localization formula derived here is very
similar to one for an $\CN = 4$ supersymmetric sigma model on $\mathbb{R}
\times S^2$, constructed from twisted chiral multiplets of $\CN =
(2,2)$ supersymmetry on $S^2$ \cite{Luo:2013nxa}. In that case, the
action contains the term
\begin{equation}
\frac{2i}{r} \int_{S^2} \Re W,
\end{equation}
where $r$ is the radius of the $S^2$ \cite{Gomis:2012wy}. This term
corresponds to our boundary term \eqref{BT-2}, and eventually becomes
the action of a quantum mechanical system. The localization again
requires the bosonic field to be constant, so the above term is
multiplied by the area of the $S^2$ and the Planck constant is
proportional to $1/r$.
Let us discuss what changes have to be made if we remove the
assumption that $\Omega$ is exact. In this case $\Lambda$ can exist
only locally, so the formula for $W$ is not well-defined. This is not
a problem when we work with a spacetime with no boundary, since what
enters the action then is not $W$ itself, but the derivative of $W$,
which can be expressed in terms of $\Omega$. However, it does cause a
problem in the present setup where $W$ enters the boundary term in the
action.
A better definition for $W$ is the following. First, in each homotopy
class $\CP$ of $\mathop{\mathrm{Map}}\nolimits(\mathbb{R},X)$ we fix a reference map $\Phi_\CP$, so that
any map $\Phi \in \CP$ can be deformed to $\Phi_\CP$ without \linebreak altering
the behavior at infinity. Next, given a map $\Phi \in \CP$, we pick a
homotopy $\Phih\colon [0,1] \to$ $\mathop{\mathrm{Map}}\nolimits(\mathbb{R},X)$ from $\Phi_\CP$ to
$\Phi$; thus $\Phih(0) = \Phi_\CP$ and $\Phih(1) = \Phi$. Finally, we
define
\begin{equation}
W(\Phi) = \frac12 \int_{[0,1] \times \mathbb{R}} \Phih^*\Omega,
\end{equation}
viewing $\Phih$ as a map from $[0,1] \times \mathbb{R}$ to $X$. Changing the
reference maps $\Phi_\CP$ shifts $W$ by a locally constant function,
but such a shift can be absorbed in the definition of $W_0$. If
$\Omega$ is exact, $W$ is given as before by the integral of a
holomorphic one-form $\Lambda$ such that $\Omega = \mathrm{d}\Lambda$.
Since the functional derivative of $W$ can be computed locally, and
locally we can always write $\Omega$ as $\Omega = \mathrm{d}\Lambda$, this
definition of $W$ leads to the same superpotential terms
\eqref{SW-RW}.
The function $W$ so defined is actually not single-valued on $\mathop{\mathrm{Map}}\nolimits(\mathbb{R},
X)$. If we choose a different homotopy $\Phih'$, then the two
homotopies combine into a map $\Delta\Phih\colon S^1 \times \mathbb{R} \to X$,
and $W$ changes by
\begin{equation}
\Delta W = \frac12 \int_{S^1 \times \mathbb{R}} \Delta\Phih^*\Omega.
\end{equation}
By assumption the homotopies leave the behavior at infinity intact, so
$\Delta\Phih$ maps each end of the cylinder $S^1 \times \mathbb{R}$ to a
point. As such, $\Delta\Phih$ may be thought of as really a map from
a two-sphere to $X$. For the path integral with the boundary term
\eqref{BT-2} to be well-defined, the change in the boundary term must be
always an integer multiple of $2\pi i$. This requirement places the
constraint
\begin{equation}
\frac{1}{2\pi\hbar} [\Re\Omega] \in H^2(L; \mathbb{Z}).
\end{equation}
This is nothing but the quantization condition for $(L,\Re\Omega)$.
Now we summarize what we have found. Let $(X, g, I, J, K)$ be a
hyperk\"ahler manifold. Pick a submanifold $L$ of $X$ that is
Lagrangian with respect to $\omega_I$ and $\omega_K$, and holomorphic
in $J$. Then, the $\Omega$-deformed Rozansky-Witten theory with
target space $X$ in complex symplectic structure $(I, \Omega_I)$,
formulated on $\mathbb{R} \times D$ with boundary condition specified by a
brane supported on $L$, is equivalent to a quantum mechanical system
whose phase space is the symplectic manifold $(L, \omega_J)$. The
Planck constant is proportional to the $\Omega$-deformation parameter
$\veps$.
\subsection{Comparison with the A-model approach}
\label{AA}
In \cite{Gukov:2008ve}, Gukov and Witten developed a framework for
quantization of symplectic manifolds using branes in the A-model. It
is illuminating to compare their approach with ours.
In the A-model approach, one first embeds the symplectic manifold
$(L,\omega)$ that one wants to quantize into a complex symplectic
manifold $(X, \Omega)$ of twice the dimension such that $\Re\Omega|_L
= \omega$ and $\Im\Omega|_L = 0$. (We are using the same symbols as
in our approach to emphasize parallels.) One then considers the
A-model whose target space is the symplectic manifold $(X,
\Im\Omega)$. Taking the worldsheet to be a strip, one places two
types of A-branes on its sides. On one side is a Lagrangian A-brane
supported on $L$ and endowed with a complex line bundle with a flat
$\mathrm{U}(1)$ connection $A$. On the other is a canonical coisotropic
A-brane \cite{Kapustin:2001ij} whose support is the whole $X$, endowed
with a line bundle with a connection of curvature $\Re\Omega$. It
turns out that this system quantizes $(L, \omega)$ just like our
system does: the open strings with both ends attached on the canonical
coisotropic brane form a noncommutative deformation of the algebra of
holomorphic functions on $X$, whereas the strings stretched between
the two branes span a Hilbert space on which the deformed algebra
acts.
A particularly nice situation is when $X$ admits a hyperk\"ahler
structure such that $\Omega = \Omega_I$ and $L$ is a complex
submanifold in complex structure $J$, since in this case one can study
the B-model of complex structure $J$ and describe the Hilbert space
explicitly. Since $\omega_K = \Im\Omega$ vanishes on $L$, so does
$\omega_I = -\omega_K J$ then. Thus $L$ is a Lagrangian submanifold
with respect to $\omega_I$ and $\omega_K$, and a complex submanifold
in complex structure $J$; the branes are of type $(A,B,A)$.
Interestingly, this is precisely the property required of the support
of a brane used in our approach. In our case we have a single brane
on the boundary of $D$, and it may be thought of as playing the role
of a combination of the two branes in the A-model approach. For
example, $\Re W$ and $W_0$ correspond to the gauge fields on the
space-filling and middle-dimensional branes, respectively, as can be
seen from their expressions~\eqref{W-RW} and~\eqref{W0-RW}.
In view of these similarities, it may be reasonable to expect that if
we equip $D$ with a cigar metric and regard $D$ as an $S^1$-fibration
over an interval, our system reduces to the A-brane system at low
energies. Nekrasov and Witten \cite{Nekrasov:2010ka} showed that this
is indeed the case under certain circumstances. We will discuss this
point briefly in section \ref{QO4d}
\section{Applications to four-dimensional gauge theory}
\label{4d}
In the final section we discuss applications of our framework to $\CN
= 2$ supersymmetric gauge theories in four dimensions. For each of
these theories, there is a class of physical quantities captured by
Rozansky-Witten theory with hyperk\"ahler target space. Using this
fact and the results obtained in the previous section, we establish
connections between the gauge theory and quantization of objects
associated with the target space.
\subsection{Rozansky-Witten theory from four dimensions}
\label{RW4d}
\enlargethispage{9.7pt}
First we clarify the relation between $\CN = 2$ supersymmetric gauge
theories and Rozansky-Witten theory, and explain some important
features of the geometry of the emergent target spaces. For more
details, see \cite{Seiberg:1996nz, Gaiotto:2008cd}.
Consider an $\CN = 2$ supersymmetric gauge theory, compactified on
$S^1$. On the Coulomb branch, the theory is described at low energies
by a three-dimensional abelian gauge theory with $\CN = 4$
supersymmetry. Dualizing the gauge fields to periodic scalars, we get
a sigma model. Its target space $\CM$ is required to be hyperk\"ahler
by the $\CN = 4$ supersymmetry, and has dimension $4r$, where $r$ is
the rank of the gauge group. For a large class of theories obtained
by compactification of M5-branes on punctured Riemann surfaces, $\CM$
is the Hitchin moduli space of the relevant surface
\cite{Gaiotto:2009we, Gaiotto:2008cd}.
If we instead start from the topologically twisted version of the same
theory, then the resulting sigma model is twisted as well. Placing the
ultraviolet theory on $M \times S^1$, we get Rozansky-Witten theory
on $M$ with target space $\CM$. Recall that when the target space is
hyperk\"ahler, Rozansky-Witten theory has two supercharges $Q$ and
$\Qb$. The first one exists in the more general case of complex
symplectic target spaces, but the second does not. From the
four-dimensional viewpoint, $Q$ is the scalar supercharge of the
twisted theory, while $\Qb$ is the component of the one-form
supercharge along the $S^1$. Any linear combination $Q_\zeta \propto
Q + \zeta\Qb$ with $\zeta \in \P^1$ may be used as a BRST operator.
This corresponds to the fact that $\CM$ has a $\P^1$-worth of complex
structures $J_\zeta$ in which Rozansky-Witten theory can be
formulated. We write $\Omega_\zeta$ for the holomorphic symplectic
form associated to $J_\zeta$.
The above effective description is valid at length scales that are
much larger than the radius of the $S^1$ so that the theory looks
effectively three-dimensional, but much smaller compared to the size
of $M$ so that the effects of the curvature of $M$ are negligible on
the massive modes that are integrated out. This requirement can
always be met by rescaling of the spacetime metric which leaves
physical quantities unaffected. This is clearly true for $\zeta = 0$,
that is when the BRST operator is $Q$, in which case the twisted
theory is well-known to be a topological field theory
\cite{Witten:1988ze}. It is also true for $\zeta \neq 0$. The reason
is that correlation functions of $Q_\zeta$-invariant operators on $M
\times S^1$ are supersymmetric indices and protected under
deformations of the parameters of the theory.%
\footnote{We assume that the $Q_\zeta$-invariant states form a
discrete spectrum, which should be the case if $M$ is compact.
Although the choice $M = \mathbb{R} \times D$ that we will consider is not
compact, we can replace the $\mathbb{R}$ by a finite interval without
altering the conclusions.}
The geometry of $\CM$ is very interesting, in that there is a
distinguished complex structure in which $\CM$ is the phase space of a
complex integrable system \cite{Donagi:1995cf}. This is the complex
structure $J_0$, and usually called $I$. In this complex structure,
$\CM$ is a torus fibration over a complex manifold $\CB$ whose fibers
are complex Lagrangian submanifolds with respect to the holomorphic
symplectic form $\Omega_I$. The base $\CB$ is the Coulomb moduli
space of the ultraviolet theory placed on $\mathbb{R}^4$; it is topologically
an affine space $\mathbb{C}^r$, parametrized by the vacuum expectation values
of the gauge-invariant polynomials in the vector multiplet scalar.
The torus fibers are parametrized by the holonomies of the infrared
abelian gauge fields and their magnetic duals around the $S^1$.
There are particularly nice coordinates on $\CM$ in this context.
$\CN = 2$ supersymmetry requires that $\CB$ admits local holomorphic
coordinates $a^i$, $i = 1$, $\dotsb$, $r$, and their duals $a_{D,i}$
related through a holomorphic function $\CF$ as $a_{D,i} =
\partial\CF/\partial a^i$. The second derivatives $\tau_{ij} = \partial^2\CF/\partial
a^i \partial a^j = \partial a_{D,i}/\partial a^j$ encode the complexified gauge
couplings of the effective abelian gauge theory. If we write the
electric and magnetic holonomies as $\exp(i\theta_e^i)$ and
$\exp(i\theta_{m,i})$, then $z_i = \theta_{m,i} - \tau_{ij}
\theta_e^j$ are holomorphic coordinates on the torus fiber of $\CM$.
Moreover, $(a^i, z_i)$ are complex Darboux coordinates:%
\footnote{Holomorphic objects in complex structure $I$, especially the
structure of complex integrable system, do not receive instanton
corrections coming from BPS particles circling around the $S^1$. This is
because the action for such particles is not holomorphic, but rather
the absolute value of a holomorphic function \cite{Seiberg:1996nz}.}
\begin{equation}
\Omega_I = \mathrm{d} a^i \wedge \mathrm{d} z_i.
\end{equation}
The $a^i$ are conserved charges generating translations in the fiber
directions, and commute with one another with respect to the Poisson
bracket derived from $\Omega_I$. There are $r$ such charges in the
phase space $\CM$ of complex dimension $2r$, reflecting the fact that
the system is completely integrable in the complex sense.
\subsection{Quantization by twisting of the spacetime}
Now we wish to modify the ultraviolet theory in such a way that the
effective theory undergoes an $\Omega$-deformation. For $\zeta \neq
0$, $\infty$, we can achieve this by replacing the spacetime $M \times
S^1$ with a twisted product between $M$ and $S^1$, which is a
nontrivial $M$-fibration over $S^1$. More specifically, we take the
trivial fibration $M \times [0,1]$, and identify the fibers at the two
ends of the interval $[0,1]$ with the action of an isometry of $M$.
Writing this isometry as $\exp(V)$ with $V$ a Killing vector field, we
denote the resulting fibration by $M \times_V S^1$.
Since $Q_\zeta$ are scalars on $M$, it commutes with isometries on $M$
and correlation functions of $Q_\zeta$-invariant operators on $M
\times_V S^1$ are still protected indices. As such, they may be
computed by the effective sigma model. We have $\{Q, \Qb\} \propto
P_4$ and hence $Q_\zeta^2 \propto P_4$ for $\zeta \neq 0$, $\infty$
(up to a central charge), where $P_4$ acts on fields by $\partial_4$. Due
to the isometry twist, at low energies $P_4$ is replaced by $L_V$,%
\footnote{To see this, one can use coordinates $(y^\mu, y^4) =
(\exp(x^4 V) x^\mu, x^4)$, where $x^\mu$ and $x^4$ are coordinates
on $M$ and $S^1$, respectively. In these coordinates the fibration
is ``untwisted,'' $(y^\mu, 0) \sim (y^\mu, 1)$, and at low energies
we simply have $\partial/\partial y^4 = \partial/\partial x^4 - V = 0$ on
functions.}
leading to the deformed relation $Q_\zeta^2 \propto L_V$ in three
dimensions. We thus identify the effective theory for $\zeta \neq 0$,
$\infty$ with Rozansky-Witten theory subject to the
$\Omega$-deformation by a (complex) Killing vector field proportional
to $V$.
Taking $M = \mathbb{R} \times D$ and $V$ to generate rotations of the disk
$D$, we conclude that a twisted $\CN = 2$ supersymmetric gauge theory
on the corresponding fibration quantizes a symplectic submanifold $(L,
\Re\Omega_\zeta)$ of $\CM$, where $L$ is the support of the brane on
the boundary of $D$.
It is interesting to consider $Q_\zeta$-invariant operators. In three
dimensions, relevant operators are holomorphic functions on $\CM$ in
complex structure $J_\zeta$, inserted at the center of $D$ and points
on the $\mathbb{R}$. They form a noncommutative deformation of the algebra of
holomorphic functions on $\CM$. In four dimensions, these local
observables may be represented by line operators wrapped on the $S^1$.
This explains the observation made in \cite{Gaiotto:2010be,
Ito:2011ea} that supersymmetric loop operators realize a deformation
quantization of the algebra of holomorphic functions on $\CM$.
We remark that due to the twisting of the spacetime, there may be
corrections to the holomorphic symplectic form $\Omega_\zeta$ when it
appears in the $\Omega$-deformed Rozansky-Witten theory and hence in
the quantum mechanical system; see the comment at the end of
section~\ref{ORW} for this point.
\subsection[Quantization by the \texorpdfstring{$\Omega$}{Omega}-deformation in four dimensions]{Quantization by the \texorpdfstring{$\boldsymbol \Omega$}{Omega}-deformation in four dimensions}
\label{QO4d}
The above argument does not apply when $\zeta = 0$ or $\infty$. For
$\zeta = 0$, there is a more obvious way to induce an
$\Omega$-deformation in the effective Rozansky-Witten theory. That
is to turn on an $\Omega$-deformation in the ultraviolet.
What we find in this case is that a twisted $\CN = 2$ supersymmetric
gauge theory on $\mathbb{R} \times D \times S^1$, subject to the
$\Omega$-deformation by a rotation generator of $D$, quantizes a real
symplectic submanifold $(L, \Re\Omega_I)$ of the complex integrable
system $(\CM, \Omega_I)$. The commuting Hamiltonians of the quantum
integrable system are the operators corresponding to the special
coordinates $a^i$. In the ultraviolet theory, they are realized
by the gauge-invariant polynomials in the vector multiplet scalar
inserted at the origin of $D$.
Since the twisted theory is topological and its Hamiltonian is
identically zero, the states of the quantum mechanical system
correspond to the vacua of the gauge theory. Let us find where the
vacua are located. To be concrete, we take $L$ to be the locus given
locally by the equations
\begin{equation}
\Im a_{D,i} = \theta_{m,i} = 0.
\end{equation}
This is a good choice; we may choose a hyperk\"ahler metric such that
$\omega_I|_L = 0$ (for example, the semiflat metric obtained by
dimensional reduction of the effective abelian gauge theory), while
$\Omega = \mathrm{d} a^i \wedge \mathrm{d}\theta_{m,i} - \mathrm{d} a_{D,i} \wedge
\mathrm{d}\theta_e^i$ and hence $\Im\Omega_K|_L = 0$, showing that $L$ is a
Lagrangian submanifold with respect to $\omega_I$ and $\omega_K =
\Im\Omega$ as required. $L$ is a real integrable system over the real
submanifold of $\CB$ parametrized by $\Re a_{D,i}$, with the torus
fiber parametrized by $\theta_e^i$. The Lagrangian of the quantum
mechanical system is $-\Re a_{D,i} \mathrm{d}\theta_e^i$. Integrating over
the periodic scalars $\theta_e^i$ imposes the constraints $\Re
a_{D,i}/\hbar \in \mathbb{Z}$. Combining these constraints with the equations
$\Im a_{D,i} = 0$, we obtain
\begin{equation}
\exp\Bigl(\frac{2\pi i}{\hbar} a_{D,i}\Bigr) = 1.
\end{equation}
These are the equations that determine the locations of the vacua.
The above equations are to be identified with the Bethe equations in
the integrable system \cite{Nekrasov:2009rc}. Let $\CF$ be the
($\veps$-corrected) prepotential of the gauge theory, and $\Wt = 2\pi
i\CF/\hbar$. Then we can rewrite the equations as
\begin{equation}
\exp\Bigl(\frac{\partial\Wt}{\partial a^i}\Bigr) = 1.
\end{equation}
If the radius of $S^1$ is much larger than $1/\veps$, then there is a
low-energy regime in which the $\Omega$-deformed theory is effectively
described by a two-dimensional theory with $\CN = (2,2)$
supersymmetry. The function $\Wt$ may be interpreted as the twisted
superpotential for this effective theory on $\mathbb{R} \times S^1$. In the
quantum integrable system, it is interpreted as the Yang-Yang
function.
In the work of Nekrasov and Witten \cite{Nekrasov:2010ka}, the
quantization of the integrable system was explained in the A-brane
framework disscussed in section \ref{AA}. The starting point of their
approach is the same as ours, namely the twisted theory on $\mathbb{R} \times
D \times S^1$, subject to the $\Omega$-deformation on $D$. One puts a
cigar metric on $D$ and thinks of it as an $S^1$-fibration over an
interval. One then reduces the theory to a two-dimensional theory on
a strip. It turns out that away from the tip of the cigar, the effect
of the $\Omega$-deformation can be canceled by a redefinition of
fields. One makes use of this observation and deduces that the
two-dimensional theory is an $\CN = (4,4)$ supersymmetric sigma model
with target space $\CM$, with a space-filling $(A, B, A)$-brane placed
on one side of the strip and a middle-dimensional $(A,B,A)$-brane
placed on the other. This configuration fits in the A-brane
framework, and one concludes that it quantizes a real integrable
system which is the support of the middle-dimensional brane. Here we
have presented an alternative derivation based on the framework
developed in this paper. The relation between the two approaches
should be clear from the discussion in section \ref{AA}.
\section*{Acknowledgments}
I would like to thank Yuan Luo for helpful discussions, and Petr Vasko
for comments on the manuscript. This work is supported by INFN
Postdoctoral Fellowship and INFN Research Project ST\&FI.
|
\section{Introduction}
In 2007, Akutsu et al. \cite{Akutsu(2007)} propose the concept of {\it controllability} of
{\it Boolean control networks} (BCNs), prove that determining the controllability of BCNs is
{\bf NP}-hard\footnote{That is,
there exists no polynomial time algorithm for determining the controllability of BCNs
unless {\bf P=NP}.},
and point out that ``One of the major goals of
systems biology is to develop a control theory for complex biological systems''.
Since then, the study on control-theoretic problems in the areas of Boolean
networks (initiated by Kauffman \cite{Kauffman1969RandomBN} in 1969
to describe
genetic regulatory networks) and Boolean control
networks (initiated in \cite{Ideker(2001)} in 2001) has drawn vast attention (cf.
\cite{Cheng2009bn_ControlObserva,Cheng2011IdentificationBCN,Zhao2010InputStateIncidenceMatrix,Cheng(2011book),Fornasini2013ObservabilityReconstructibilityofBCN,Laschov2013ObservabilityofBN:GraphApproach,Zhang2013ControlObservaBCNTVD,Li2013ObservabilityConditionsofBCN,Zhang2015Invertibility:BCN} etc.).
Controllability and {\it observability} are basic control-theoretic problems.
In 2009, Cheng et al. \cite{Cheng2009bn_ControlObserva}
construct a control-theoretic
framework for BCNs by using a new tool, called the {\it semi-tensor product} (STP) of matrices
proposed in \cite{Cheng2001STP} in 2001,
and give equivalent conditions for controllability
of BCNs and observability of controllable BCNs. Since then, to the best of our knowledge,
how to determine this observability has been open.
This type of observability means that
every initial state can be determined by an input sequence.
Later on, important results on other types of observability
of BCNs came up.
Until now, there are four types of
observability.
Another observability, proposed in
\cite{Zhao2010InputStateIncidenceMatrix} in 2010, stands for that
for every two distinct initial states,
there exists an input sequence which can distinguish them.
There is a sufficient but not necessary condition in \cite{Zhao2010InputStateIncidenceMatrix}.
However, there is no equivalent condition in \cite{Zhao2010InputStateIncidenceMatrix}.
This observability is determined in \cite{Li2015ControlObservaBCN} in 2015
based on an algebraic method.
A third observability stating that there is an input sequence
that determines the initial state, is proposed in \cite{Cheng2011IdentificationBCN}
to study the identifiability problem of BCNs in 2011.
It is proved that determining this observability is {\bf NP}-hard in
\cite{Laschov2013ObservabilityofBN:GraphApproach} in 2013.
Nevertheless, one way is proposed in \cite{Li2013ObservabilityConditionsofBCN}
to determine this observability in 2013.
A fourth observability is determined in
\cite{Fornasini2013ObservabilityReconstructibilityofBCN,Xu2013ObserverFA_STP}\footnote{Note
that the types of observability studied in \cite{Fornasini2013ObservabilityReconstructibilityofBCN,Xu2013ObserverFA_STP}
are the same.} in 2013, which is
essentially the observability of linear control systems,
i.e., every sufficiently long input sequence can
determine the initial state.
Like nonlinear systems,
BCNs are polynomial systems over ${\mathbb F}_2$, the Galois field of two elements \cite{Li2015ControlObservaBCN}.
This explains why
the observability proposed in \cite{Cheng2009bn_ControlObserva,Zhao2010InputStateIncidenceMatrix}
that rely on initial states and inputs are important for BCNs.
The methods of determining the last two types of observability
are not suitable for the first two, mainly because they
are based on the independence of
initial states and/or inputs. Besides, it is not known whether the method for
the second type used in
\cite{Li2015ControlObservaBCN} is suitable for the other three types now.
In this paper, we propose
a unified method based on {\it finite automata}
to determine all the four types of observability regardless of dependence.
To this end, we firstly define
{\it weighted pair graphs} for BCNs, which consist of pairs of states of BCNs producing
the same outputs, and transitions between the pairs.
Secondly, we use the weighed pair graph to transform a BCN to a deterministic finite
automaton. Finally, we use the automaton to determine observability.
The remainder of this paper is organized as follows. Section \ref{sec2}
introduces necessary preliminaries about
STP, BCNs with their algebraic forms, {\it formal languages} and finite automata.
Section \ref{sec3} presents the algorithms to determine all the four types of observability.
Section \ref{sec4} shows the pairwise nonequivalence of
the four types of observability of BCNs.
Section \ref{sec6} ends up with some remarks and challenging open problems.
\section{Preliminaries}\label{sec2}
\subsection{The semi-tensor product of matrices}
We first introduce some related notations in STP.
\begin{itemize}
\item $2^A$: the power set of set $A$
\item $\mathbb{Z}_+$: the set of positive integers
\item $\mathbb{N}$: the set of natural numbers
\item $\mathcal{D}$: the set
$\{0,1\}$
\item $\delta_n^i$: the $i$-th column of the identity matrix $I_n$
\item $\Delta_n$: the set $\{\delta_n^1,\dots,\delta_n^n\}$
($\Delta:=\Delta_2$)
\item $\delta_n[i_1,\dots,i_s]$: the {\it logical matrix}
$[\delta_n^{i_1},\dots,\delta_n^{i_s}]$ ($i_1,\dots,i_s\in\{1,2,\dots, n\}$)
(for the concept of logical matrices, we refer the reader to \cite{Cheng(2011book)}.)
\item $\mathcal{L}_{n\times s}$: the set of
$n\times s$ logical matrices, i.e., $\{\delta_n[i_1,\dots,i_s]|i_1,\dots,i_s\in\{1,2,\dots,n\}\}$
\item $[1,N]$: the first $N$ positive integers
\item $|A|$: the cardinality of set $A$
\end{itemize}
\begin{definition}\cite{Cheng(2011book)}
Let $A\in \mathbb{R}_{m\times n}$, $B\in \mathbb{R}
_{p\times q}$, and $\alpha=\mbox{lcm}
(n,p)$ be the least common multiple of $n$ and $p$. The STP of $A$
and $B$ is defined as
$A\ltimes B = (A\otimes I_{\frac{\alpha}{n}})(B\otimes I_{\frac
{\alpha}{p}})$,
where $\otimes$ denotes the Kronecker product.
\end{definition}
From this definition, it is easy to see that the conventional
product of matrices is a particular case of STP.
Since STP keeps most properties of the conventional product \cite{Cheng(2011book)},
e.g., the associative law, the distributive law,
etc.,
we usually omit the symbol ``$\ltimes$'' hereinafter.
\subsection{Boolean control networks and their algebraic forms}
In this paper, we investigate the following BCN
with $n$ state nodes, $m$ input nodes and $q$ output nodes:
\begin{equation}\label{BCN1}
\begin{split}
&x (t + 1) = f (u (t),x (t)), \\
&y(t)=h(x (t)),\\
\end{split}
\end{equation}
where $x\in\mathcal{D}^n$; $u\in\mathcal{D}^m$; $y\in\mathcal{D}^q$;
$t=0,1,\dots$; $f:\mathcal{D}^{n+m}\to\mathcal{D}^n$ and $h:\mathcal{D}^n\to\mathcal{D}^q$ are logical
functions.
Using the STP of
matrices, (\ref{BCN1}) can be equivalently represented in
the following algebraic form \cite{Cheng2009bn_ControlObserva}
\begin{equation}\label{BCN2}
\begin{split}
&x(t + 1) = Lu(t)x(t),\\
&y(t) = Hx(t),
\end{split}
\end{equation}
where $x\in\Delta_{N}$, $u\in\Delta_{M}$ and $y\in\Delta_{Q}$ denote
states, inputs and outputs, respectively;
$t=0,1,\dots$; $L\in\mathcal{L}_{N\times (NM)}$; $H\in\mathcal{L}_{Q\times N}$;
hereinafter, $N:=2^n$, $M:=2^m$ and $Q:=2^q$.
For more details on properties of STP,
and how to transform a BCN into its equivalent algebraic form,
we refer the reader to \cite{Cheng2009bn_ControlObserva}.
\subsection{Formal languages and finite automata}
The theories of formal languages and finite automata are
among the mathematical foundations of theoretical computer
science \cite{Kari2013LectureNoteonAFL}.
Let $\Sigma$ be a finite nonempty set (called {\it alphabet}).
We use $\Sigma^*$ to denote the set of all finite sequences (called {\it words}) of
elements (called {\it letters}) of $\Sigma$.
The empty word is denoted by $\epsilon$.
$|u|$ denotes the length of word $u$. For example, $|abc|=3$ for
the alphabet $\{a,b,c\}$, $|\epsilon|=0$. The set of all words of length $p$
is denoted by $\Sigma^p$. Notice that $\Sigma^{0}=\{\epsilon\}$.
Then $\Sigma^{*}=\cup_{p=0}^{\infty}\Sigma^{p}$.
A formal language (or language for short) is a subset of $\Sigma^*$.
A deterministic finite automaton (DFA) is defined as
5-tuple $A=(S,\Sigma,\sigma,s_0,F)$, where $S$ denotes the finite state set,
$\Sigma$ the finite alphabet,
$s_0\in S$ the initial state, $F\subset S$ the final state set,
and $\sigma:S\times\Sigma\to S$ the transition partial function,
i.e., a function defined on a fixed subset of $S\times\Sigma$,
which can naturally be
extended to $\sigma:S\times \Sigma^*\to S$.
We call a
DFA {\it complete} if $\sigma$ is a function from $S\times\Sigma^*$ to $S$.
A language $L$ over alphabet $\Sigma$ is called {\it regular}, if it is {\it recognized} by
a DFA $A=(S,\Sigma,\sigma,s_0,F)$, i.e., $L=\{w\in\Sigma^*|\sigma(s_0,w)\in F\}$.
A word $u\in\Sigma^*$ such that $\sigma(s_0,u)\in F$ is called {\it accepted} by DFA $A$.
A DFA accepts the empty word $\epsilon$ iff its initial state is final.
In order to represent a DFA,
we introduce the transition graph of DFA
$A=(S,\Sigma,\sigma,s_0,F)$.
Let $V,E$ and $W$ be the vertex set, the edge set and the weight function of
a weighted directed graph $G=(V,E,W)$.
$G$ is called the transition graph of DFA $A$,
if $V=S$, $E=\{(s_i,s_j)\in V\times V|\text{there is }a\in\Sigma\text{ such that }
\sigma(s_i,a)=s_j\}\subset V\times V$, and
$W:E\to 2^\Sigma$, $(s_i,s_j)\mapsto \{a\in\Sigma|\sigma(s_i,a)=s_j\}$.
In the transition graph of a DFA, we add a ``start'' input arrow to the vertex of
the initial state, and use double circles to denote
final states. We omit the curly bracket ``$\{\}$''
in the weights of edges. See Fig. \ref{fig4:observability} for an example.
Now we give a proposition on finite automata that will be used in
the main results.
\begin{proposition}\label{prop6_observability}
Given a DFA $A=(S,\Sigma,\sigma,s_0,F)$. Assume that $F=S$ and for each
$s\in S$, there is a word $u\in\Sigma^*$ such that $\sigma(s_0,u)=s$.
Then $L(A)=\Sigma^*$ iff $A$ is complete.
\end{proposition}
\begin{proof}
``if'': If $A$ is complete and $F=S$, then $\epsilon\in L(A)$ and
for any nonempty word $w\in\Sigma^*$, $\sigma(s_0,w)\in F$, i.e., $w\in
L(A)$. Hence $L(A)=\Sigma^*$.
``only if'': Assume that $F=S$ and $A$ is not complete.
Choose an $s\in S$ such that $\sigma$ is not well defined at $(s,a)$
for some $a\in\Sigma$.
Choose word $w\in \Sigma^*$ such that $\sigma(s_0,w)=s$,
then
word $wa\notin L(A)$, for $A$ is deterministic.
That is, $L(A)\ne \Sigma^*$.
\end{proof}
\section{Determining the observability of BCNs}\label{sec3}
\subsection{Weighted pair graph}
In this subsection, we define a weighted directed graph for BCN \eqref{BCN2},
named weighted pair graph.
Based on the weighted pair graph, in the following subsections,
we construct a DFA in the sense of each observability, and then
use the obtained DFA and Proposition \ref{prop6_observability}
to determine observability.
\begin{definition}\label{pairgraph}
Consider BCN \eqref{BCN2}.
Let $\mathcal{V},\mathcal{E}$ and $\mathcal{W}$ be the vertex set, the edge set and the weight function of
a weighted directed graph $\mathcal{G}=(\mathcal{V},\mathcal{E},\mathcal{W})$. $\mathcal{G}$ is
called the weighted pair graph of the BCN, if
$\mathcal{V} = \{(x,x')\in\Delta_N\times\Delta_N|Hx=Hx'\}$\footnote{Here
$(x,x')$ is an unordered pair, i.e., $(x,x')=(x',x)$.},
$\mathcal{E}=\{((x_1,x_1'),(x_2,x_2'))\in\mathcal{V}\times\mathcal{V}|
\text{there exists }u\in\Delta_M\text{ such that }Lux_1=x_2\text{ and }Lux_1'=x_2',
\text{ or, }Lux_1=x_2'\text{ and }Lux_1'=x_2\}\subset \mathcal{V}\times\mathcal{V}$, and
$\mathcal{W}:\mathcal{E}\to 2^{\Delta_M}$, $((x_1,x_1'),(x_2,x_2'))\mapsto
\{u\in\Delta_M|Lu_1x_1=x_2\text{ and }Lu_1x_1'=x_2'\text{, or, }
Lux_1=x_2'\text{ and }Lux_1'=x_2\}$.
\end{definition}
Intuitively, there is an edge from a vertex $v$ to another one $v'$, iff there is
an input $u$ driving one state in $v$ to one state in $v'$ and driving the other
state in $v$ to the other state in $v'$.
Similar to the transition graph of a DFA, we omit the curly
bracket ``$\{\}$'' in the weights of edges.
Hereinafter, we call each vertex $(x,x)\in\Delta_N\times\Delta_N$ a diagonal
vertex.
From Definition \ref{pairgraph}, the weighted pair graph
consists of every state pair producing the same output.
In fact, to test whether a BCN is observable,
is to test whether these states can be distinguished by
input sequences.
Let $(\mathcal{V},\mathcal{E},\mathcal{W})$ be a weighted pair graph. For a subset $V$ of $\mathcal{V}$, the subgraph
generated by $V$ is the graph $(V,\mathcal{E}\cap (V\times V),\mathcal{W}|_{\mathcal{E}\cap (V\times V)})$,
where $\mathcal{W}|_{\mathcal{E}\cap (V\times V)}$ is the restriction of $\mathcal{W}$ to $\mathcal{E}\cap (V\times V)$.
The weighted pair graph of the following BCN \eqref{eqn2_observability} is depicted in
Fig. \ref{fig3:observability}.
\begin{equation}
\begin{split}
x(t+1) &= \delta_4[1,1,2,1,2,4,1,1]x(t)u(t),\\
y(t) &= \delta_2[1,2,2,2]x(t),
\end{split}
\label{eqn2_observability}
\end{equation}
where $t\in\mathbb{N}$, $x\in\Delta_4$, $y,u\in\Delta$.
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=1.5 cm, scale = 0.8, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\node[state] (11) {$11$};
\node[state] (22) [right of = 11] {$22$};
\node[state] (42) [left of = 11] {$24$};
\node[state] (32) [right of = 22] {$23$};
\node[state] (44) [below of = 11] {$44$};
\node[state] (33) [below of = 22] {$33$};
\node[state] (34) [left of = 44] {$34$};
\path [->] (11) edge [loop above] node {$1,2$} (11)
(22) edge [loop above] node {$1$} (22)
(22) edge node {$2$} (11)
(42) edge node {$2$} (11)
(32) edge node {$1$} (22)
(44) edge node {$1,2$} (11)
(33) edge node {$1$} (22)
(33) edge node {$2$} (44)
;
\end{tikzpicture}
\caption{The weighted pair graph of BCN \eqref{eqn2_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, and the weight $k_1,k_2,\dots$ beside each edge denotes the
weight $\{\delta_2^{k_1},\delta_2^{k_2},\dots\}$ of the edge.}
\label{fig3:observability}
\end{figure}
\subsection{Notations}
\begin{figure
\centering
\begin{tikzpicture}[->,>=stealth,node distance=1.0cm]
\tikzstyle{state}=[shape=rectangle,fill=blue!80,draw=none,text=white,inner sep=2pt]
\tikzstyle{input}=[shape=rectangle,fill=red!80,draw=none,text=white,inner sep=2pt]
\tikzstyle{output}=[shape=rectangle,fill=red!50!blue,draw=none,text=white,inner sep=2pt]
\tikzstyle{disturbance}=[shape=rectangle,fill=red!20!blue,draw=none,text=white,inner sep=2pt]
\tikzstyle{time}=[inner sep=0pt,minimum size=5mm]
\node [state] (x0) {$x_0$};
\node [state] (x1) [right of=x0] {$x_1$};
\node [state] (x2) [right of=x1] {$x_2$};
\node [input] (u0) [above of=x1] {$u_0$};
\node [input] (u1) [right of=u0] {$u_1$};
\node [time] (x3) [right of=x2] {$\cdots$};
\node [time] (u2) [above of=x3] {$\cdots$};
\node [state] (x4) [right of=x3] {$x_p$};
\node [time] (u2) [above of=x3] {$\cdots$};
\node [input] (u3) [above of=x4] {$u_{p-1}$};
\node [time] (x5) [right of=x4] {$\cdots$};
\node [time] (u4) [above of=x5] {$\cdots$};
\node [output] (y0) [below of=x0] {$y_0$};
\node [output] (y1) [below of=x1] {$y_1$};
\node [output] (y2) [below of=x2] {$y_2$};
\node [output] (y4) [below of=x4] {$y_p$};
\node [time] (y3) [below of=x3] {$\cdots$};
\node [time] (y5) [below of=x5] {$\cdots$};
\path (x0) edge (x1)
(x1) edge (x2)
(u0) edge (x1)
(u1) edge (x2)
(x2) edge (x3)
(x3) edge (x4)
(u3) edge (x4)
(x4) edge (x5)
(x0) edge (y0)
(x1) edge (y1)
(x2) edge (y2)
(x4) edge (y4);
\end{tikzpicture}
\caption{The input-state-output-time transfer graph of BCN
\eqref{BCN2}, where
subscripts stand for time steps,
$x_0,x_1,\dots$ states, $u_0,u_1,\dots$ inputs,
$y_1,y_2,\dots$ outputs, and arrows infer dependence.}
\label{fig:input-state-output-time-graph}
\end{figure}
The input-state-output-time transfer graph
of BCN \eqref{BCN2} is drawn in Fig.
\ref{fig:input-state-output-time-graph}.
In order to define these observability, we
define the following mappings:
Let $\Delta_M,\Delta_N,\Delta_Q$ be three alphabets.
For all $x_0\in\Delta_N$ and all $p\in\mathbb{Z}_{+}$,
\begin{enumerate}
\item
\begin{equation}\label{lx0n}
\begin{split}
&L_{x_0}^p:(\Delta_M)^{p}\to(\Delta_N)^{p},u_0
\dots u_{p-1}\mapsto x_1\dots x_p,\\
&L_{x_0}^{\mathbb{N}}:(\Delta_M)^{\mathbb{N}}\to(\Delta_N)^{\mathbb{N}},u_0u_1\dots
\mapsto x_1x_2\dots.
\end{split}
\end{equation}
\item
\begin{equation}\label{hx0n}
\begin{split}
&(HL)_{x_0}^p:(\Delta_M)^{p}\to(\Delta_Q)^{p},u_0
\dots u_{p-1}\mapsto y_1\dots y_p,\\
&(HL)_{x_0}^\mathbb{N}: (\Delta_{M})^{\mathbb{N}}\to(\Delta_{Q})^{\mathbb{N}},
u_0u_1\dots\mapsto y_1y_2\dots.
\end{split}
\end{equation}
\end{enumerate}
For all $p\in\mathbb{Z}_{+}$,
all $U=u_1\dots u_p\in(\Delta_M)^{p}$,
and all $1\le i\le j\le|U|$, we use $U[i,j]$ to denote the word $u_i\dots u_j$.
In particular, $U[i]$ (or $U(i)$) is short for $U[i,i]$.
Given $U\in(\Delta_M)^*$, $U^{\infty}$ denotes the {\it concatenation} of infinite copies
of $U$, i.e., $UU\dots$. For all input sequences $U=u_0u_1\dots$$\in(\Delta_M)^{\mathbb{N}}$, and
all $0\le i\le j\in\mathbb{N}$, we use $U[i,j]$ to denote the word $u_i\dots u_j$.
\subsection{Determining the observability in \cite{Cheng2009bn_ControlObserva}}
\begin{definition}[\cite{Cheng2009bn_ControlObserva}]\label{def1_observability}
BCN \eqref{BCN2} is called observable, if for every initial state $x_0
\in\Delta_N$, there exists an input sequence such that the initial state
can be determined by the output sequence.
\end{definition}
Definition \ref{def1_observability} can be expressed equivalently as follows:
\begin{definition}\label{def3_observability}
BCN \eqref{BCN2} is called observable, if for every initial state $x_0
\in\Delta_N$, there exists an input sequence $U\in(\Delta_M)^{p}$ for some
$p\in\mathbb{Z}_{+}$ such
that for all states $x_0\ne\bar x_0\in\Delta_N$,
$Hx_0=H\bar x_0$ implies $(HL)_{x_0}^{p}(U)\ne (HL)_{\bar
x_0}^{p}(U)$.
\end{definition}
In this subsection, the observability of BCN \eqref{BCN2} refers to
Definition \ref{def3_observability}.
According to Definition \ref{def3_observability},
BCN \eqref{BCN2} is not observable iff there is a state $\delta_N^i$
in a non-diagonal vertex of its weighted pair graph $\mathcal{G}=(\mathcal{V},\mathcal{E},\mathcal{W})$
such that
for all $p\in\mathbb{Z}_{+}$, all $U\in(\Delta_M)^p$, there is a state $\delta_N^j$
with $j\ne i$, $(\delta_N^i,\delta_N^j)\in \mathcal{V}$ and $(HL)_{\delta_N^i}^p(U)=(HL)_{\delta_N^j}^p(U)$.
For fixed $\delta_N^i$, we design an algorithm to construct a DFA
for BCN \eqref{BCN2} according to its weighted pair graph $(\mathcal{V},\mathcal{E},\mathcal{W})$.
The new DFA is denoted by $A_{\delta_N^i}$, and accepts exactly all finite input
sequences that do not determine $\delta_N^i$.
The states of DFA $A_{\delta_N^i}$ are subsets of $\mathcal{V}$.
\begin{algo}\label{alg1_observability}
\begin{enumerate}
\item\label{item1_obser1}
Set $\Delta_M$ to be the alphabet of the DFA.
Set the subset of $\mathcal{V}$ consisting of all the
non-diagonal vertices of $\mathcal{V}$
that contain $\delta_N^i$ to be the initial state of the DFA.
That is, the set $s_0:=\{(\delta_N^k,\delta_N^l)|k,l\in[1,N],
H\delta_N^k=H\delta_N^l,
k\ne l, k\text{ or }l=i\}$ is the initial state
of the DFA.
\item\label{item2_obser1}
For each letter $\delta_M^j$, $j\in[1,M]$, find the value
for the transition partial
function of the DFA at $(s_0,\delta_M^j)$.
The specific procedure is as follows:
Fix $j\in[1,M]$. Set $s_j:=\{v\in\mathcal{V}|\text{there is }v'\in s_0
\text{ such that }(v',v)\in\mathcal{E},\text{ and }\delta_M^j\in\mathcal{W}((v',v))\}$.
If $s_j\ne\emptyset$, add $s_j$ to the state
set of the DFA and set $s_j$ to be the value of the transition
partial function at $(s_0,\delta_M^j)$;
otherise, the transition partial function
is not well defined at
$(s_0,\delta_M^j)$.
\item\label{item3_obser1}
For each new state $s$ of the DFA found in the previous step,
and each letter $\delta_M^j$, $j\in[1,M]$, find the value
for the transition partial
function at $(s,\delta_M^j)$ according to
Step \ref{item2_obser1}.
\item\label{item4_obser1}
Repeat Step \ref{item3_obser1} until
no new state of the DFA occurs. (Since $\mathcal{V}$
is a finite set, so is its power set,
this repetition will stop.)
\item\label{item5_obser1}
Set all the states of the DFA to be final states.
\end{enumerate}
\end{algo}
Take BCN \eqref{eqn2_observability} for example. Choose state $\delta_4^2$.
Then the DFA $A_{\delta_4^2}$ generated by Algorithm \ref{alg1_observability}
is as shown in Fig. \ref{fig4:observability}.
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=2.0 cm, scale = 1.0, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\node[accepting,initial,state] (2*) {$23,24$};
\node[accepting,state] [above right of = 2*] (11) {$11$};
\node[accepting,state] [below right of = 2*] (22) {$22$};
\path [->] (2*) edge node {$2$} (11)
(2*) edge node {$1$} (22)
(22) edge [bend right] node {$2$} (11)
(22) edge [loop right] node {$1$} (22)
(11) edge [loop right] node {$1,2$} (11)
;
\end{tikzpicture}
\caption{The DFA $A_{\delta_4^2}$ with respect to BCN
\eqref{eqn2_observability} generated by Algorithm
\ref{alg1_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, and the weight $k$ beside each edge denotes the input $\delta_2^k$.
}
\label{fig4:observability}
\end{figure}
Now we give a necessary and sufficient condition for this observability.
\begin{theorem}\label{thm3_observability}
BCN \eqref{BCN2} is not observable in the sense of
Definition \ref{def3_observability} iff there is a state
$\delta_N^i$ in a non-diagonal
vertex of its weighted pair graph such that
the DFA $A_{\delta_N^i}$ generated by Algorithm \ref{alg1_observability}
recognizes language $(\Delta_M)^*$.
\end{theorem}
\begin{proof}
``only if'':
Assume that BCN \eqref{BCN2} is not observable, then there is
a state $\delta_N^i$ such that for all $p\in\mathbb{Z}_{+}$, all
$U\in(\Delta_M)^p$, there is a state $\delta_N^j$ satisfying $i\ne j$,
$H\delta_N^i=H\delta_N^j$, and $(HL)_{\delta_N^i}^p(U)=(HL)_{\delta_N^j}^p(U)$.
According to Algorithm \ref{alg1_observability}, $v_0:=(\delta_N^i,\delta_N^j)$
is in the initial state of DFA $A_{\delta_N^i}$. Denote the weighted pair graph
of BCN \eqref{BCN2} by $\mathcal{G}=(\mathcal{V},\mathcal{E},\mathcal{W})$.
Then there exist vertices $v_k:=(\delta_N^{i_k},\delta_N^{j_k})\in\mathcal{V}$
such that
$U[k]\in \mathcal{W}( (v_{k-1},v_{k}))$, $k=1,\dots,p$.
That is, for all $p\in\mathbb{Z}_{+}$, each $U$ in $(\Delta_M)^p$ is accepted by
DFA $A_{\delta_N^i}$.
It is obvious that $\epsilon\in L (A_{\delta_N^i})$.
Then $L (A_{\delta_N^i})=(\Delta_M)^*$.
``if'':
Note that the DFA $A_{\delta_N^i}$ accepts exactly all finite input
sequences that do not determine $\delta_N^i$. Then $L (A_{\delta_N^i})=(\Delta_M)^*$
implies that for all $p\in\mathbb{Z}_{+}$, all
$U\in(\Delta_M)^p$, there is a state $\delta_N^j$ such that $i\ne j$,
$H\delta_N^i=H\delta_N^j$, and $(HL)_{\delta_N^i}^p(U)=(HL)_{\delta_N^j}^p(U)$.
That is, BCN \eqref{BCN2} is not observable.
\end{proof}
Proposition \ref{prop6_observability},
Theorem \ref{thm3_observability} and Algorithm
\ref{alg1_observability} directly imply the following result that
can be used to check whether a given BCN is observable.
\begin{theorem}\label{alg2_observability}
BCN \eqref{BCN2} is not observable in the sense of
Definition \ref{def3_observability} iff there is a state
$\delta_N^i$ in a non-diagonal
vertex of its weighted pair graph such that
the DFA $A_{\delta_N^i}$ generated by Algorithm \ref{alg1_observability}
is complete.
\end{theorem}
\begin{example}\label{exam2_observability}
Check whether BCN \eqref{eqn2_observability} is observable.
According to Theorem \ref{alg2_observability}, we should check
$\delta_4^2,\delta_4^3,\delta_4^4$ one by one.
First we check $\delta_4^2$. According to Algorithm
\ref{alg1_observability}, we calculate DFA $A_{\delta_4^2}$, and
derive the transition graph of this DFA as shown in Fig.
\ref{fig4:observability}. This DFA is complete, by
Theorem \ref{alg2_observability}, BCN \eqref{eqn2_observability} is
not observable.
\end{example}
\subsection{Determining the observability in \cite{Zhao2010InputStateIncidenceMatrix}}
\begin{definition}\label{def4_observability}
BCN \eqref{BCN2} is called observable, if for any distinct states
$x_0,\bar x_0\in\Delta_N$, there is an input sequence $U\in(\Delta_M)^p$
for some $p\in\mathbb{Z}_{+}$, such that $Hx_0=H\bar x_0$ implies
$(HL)_{x_0}^{p}(U)\ne (HL)_{\bar x_0}^{p}(U)$.\footnote{Actually,
after removing ``$Hx_0=H\bar x_0$ implies'' in Definition \ref{def4_observability},
Definition \ref{def4_observability} becomes
the observability studied in \cite{Zhao2010InputStateIncidenceMatrix}.
In order to make the observability studied in \cite{Zhao2010InputStateIncidenceMatrix}
exactly the widely accepted one for nonlinear control systems, we modify it in
Definition \ref{def4_observability}.}
\end{definition}
In this subsection, the observability of BCN \eqref{BCN2} means Definition \ref{def4_observability}.
According to Definition \ref{def4_observability},
BCN \eqref{BCN2} is not observable iff there is a non-diagonal vertex
$(\delta_N^i,\delta_N^j)$ in its weighted pair graph such that
for all $p\in\mathbb{Z}_{+}$, and $U\in(\Delta_M)^p$,
$(HL)_{\delta_N^i}^p(U)=(HL)_{\delta_N^j}^p(U)$.
For a fixed non-diagonal vertex $(\delta_N^i,\delta_N^j)$,
we design an algorithm to construct a DFA
for BCN \eqref{BCN2} according to its weighted pair graph.
The new DFA is denoted by $A_{(\delta_N^i,\delta_N^j)}$,
and accepts exactly all finite input sequences that do not
distinguish $\delta_N^i$ and $\delta_N^j$.
\begin{algo}\label{alg3_observability}
\begin{enumerate}
\item Set $\Delta_M$ to be the alphabet of the DFA.
Set vertex $(\delta_N^i,\delta_N^j)$
to be the initial state of the DFA.
\item Find each vertex $v$ such that there is a path
from $(\delta_N^i,\delta_N^j)$ to $v$.
Keep the subgraph generated by
$(\delta_N^i,\delta_N^j)$ and those vertices, and remove all vertices and
edges outside of the subgraph.
\item Set each remainder vertex to be a final state of the DFA.
\end{enumerate}
\end{algo}
Again take BCN \eqref{eqn2_observability} as an example.
The DFA of each non-diagonal vertex of the weighted pair
graph generated by Algorithm \ref{alg3_observability}
is shown in Fig. \ref{fig5:observability}.
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=1.1 cm, scale = 0.9, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\tikzstyle{time}=[inner sep=0pt,minimum size=5mm]
\node[accepting,initial,state] (241) {$24$};
\node[accepting, state] (111) [right of = 241] {$11$};
\path [->] (241) edge node {$2$} (111)
(111) edge [loop right] node {$1,2$} (111)
;
\node[time] (1) [right of = 111] {};
\node[time] (2) [right of = 1] {};
\node[accepting,initial,state] (232) [right of = 2] {$23$};
\node[accepting, state] (222) [right of = 232] {$22$};
\node[accepting, state] (112) [right of = 222] {$11$};
\path [->] (232) edge node {$1$} (222)
(222) edge [loop below] node {$1$} (222)
(222) edge node {$2$} (112)
(112) edge [loop right] node {$1,2$} (112)
;
\node[accepting,initial,state] (345) [below of = 241] {$34$};
\end{tikzpicture}
\caption{The DFA of each non-diagonal vertex of the weighted pair
graph of BCN \eqref{eqn2_observability}
generated by Algorithm \ref{alg3_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, and the weight $k$ beside each edge denotes the input $\delta_2^k$.
}
\label{fig5:observability}
\end{figure}
The following is a necessary and sufficient condition for this observability.
\begin{theorem}\label{thm4_observability}
BCN \eqref{BCN2} is not observable in the sense of
Definition \ref{def4_observability} iff there is a non-diagonal vertex
$(\delta_N^i,\delta_N^j)$ in its weighted pair graph such that
the DFA $A_{(\delta_N^i,\delta_N^j)}$ generated by Algorithm
\ref{alg3_observability}
recognizes language $(\Delta_M)^*$.
\end{theorem}
\begin{proof}
``only if'':
Assume that BCN \eqref{BCN2} is not observable, then there is
a non-diagonal vertex $(\delta_N^i,\delta_N^j)$ in the weighted pair graph of
BCN \eqref{BCN2} such that for all $p\in\mathbb{Z}_{+}$, all $U\in(\Delta_M)^p$,
$(HL)_{\delta_N^i}^p(U)=(HL)_{\delta_N^j}^p(U)$.
Then for all $p\in\mathbb{Z}_{+}$, each $U$ in $(\Delta_M)^p$ is accepted by DFA
$A_{(\delta_N^i,\delta_N^j)}$.
It is obvious that $\epsilon\in L (A_{(\delta_N^i,\delta_N^j)})$.
Then $L (A_{(\delta_N^i,\delta_N^j)})=(\Delta_M)^*$.
``if'':
Obvious by Definition \ref{def4_observability}.
\end{proof}
From Proposition \ref{prop6_observability},
Theorem \ref{thm4_observability} and Algorithm
\ref{alg3_observability}, the following result which follows
can be used to determine whether a given
BCN is observable.
\begin{theorem}\label{alg4_observability}
BCN \eqref{BCN2} is not observable in the sense of
Definition \ref{def4_observability} iff there is a non-diagonal vertex
$(\delta_N^i,\delta_N^j)$ in its weighted pair graph such that
the DFA $A_{(\delta_N^i,\delta_N^j)}$ generated by Algorithm
\ref{alg3_observability} is complete.
\end{theorem}
\begin{example}\label{exam3_observability}
Check whether BCN \eqref{eqn2_observability} is observable.
According to Theorem \ref{alg4_observability}, one should check
$(\delta_4^2,\delta_4^3),(\delta_4^2,\delta_4^4),(\delta_4^3,\delta_4^4)$
one by one.
From Fig. \ref{fig5:observability}, one sees that
$\delta_2^2\notin L(A_{(\delta_4^2,\delta_4^3)})$,
$\delta_2^1\notin L(A_{(\delta_4^2,\delta_4^4)})$, and
$\delta_2^1,\delta_2^2\notin L(A_{(\delta_4^3,\delta_4^4)})$.
Then by Theorem \ref{alg4_observability}, BCN \eqref{eqn2_observability}
is observable.
\end{example}
At the end of this subsection,
using the concept of weighted pair graphs, we give a further result on this
observability.
\begin{theorem}
Consider BCN \eqref{BCN2}. Denote
the number of non-diagonal vertices
of its weighted pair graph by $N_{nd}$.
The following two items are equivalent.
\begin{enumerate}[(i)]
\item \label{item2_observability_length}
The BCN is observable in the sense of Definition \ref{def4_observability}.
\item \label{item1_observability_length}
$N_{nd}=0$ or
for all distinct states $x_0,\bar x_0\in\Delta_N$, there is an input sequence
$U\in(\Delta_M)^{N_{nd}}$ such that $Hx_0=H\bar x_0$ implies
$(HL)_{x_0}^{N_{nd}}(U)\ne(HL)_{\bar x_0}^{N_{nd}}(U)$.
\end{enumerate}
\label{thm:length:NewObservability}
\end{theorem}
\begin{proof}
\eqref{item1_observability_length} $\Rightarrow$
\eqref{item2_observability_length}:
Obvious by Definition
\ref{def4_observability}.
\eqref{item2_observability_length} $\Rightarrow$ \eqref{item1_observability_length}:
Assume that \eqref{item1_observability_length} does not hold.
That is, $N_{nd}>0$ and there are distinct $x,x'\in\Delta_N$,
for all $U\in(\Delta_M)^{N_{nd}}$,
$Hx=Hx'$ and $(HL)_{x}^{N_{nd}}(U)=(HL)_{x'}^{N_{nd}}(U)$.
Use Algorithm \ref{alg3_observability} to generate DFA $A_{(x,x')}=(S,\Delta_M,\sigma,
(x,x'),S)$. Then $\cup_{i=0}^{N_{nd}}(\Delta_M)^i\subset L(A_{(x,x')})$.
We claim that $A_{(x,x')}$ is complete. Suppose the contrary: there is a state
$v$ of $A_{(x,x')}$ and an input $u\in\Delta_M$ such that $\sigma$ is not well defined
at $(v,u)$. Then $v$ is a non-diagonal vertex of the weighed pair graph,
because for all diagonal vertices $v'$ (if exist), for all inputs $u'\in\Delta_M$,
$\sigma$ is well defined at $(v',u')$.
There are exactly $N_{nd}$ non-diagonal vertices, then there exists an input sequence
$U_1$ of length less than $N_{nd}$ such that $\sigma((x,x'),U_1)=v$.
We get a contradiction $U_1u\in\cup_{i=0}^{N_{nd}}(\Delta_M)^i\setminus L(A_{(x,x')}) $.
By Theorem \ref{alg4_observability}, the BCN is not observable.
\end{proof}
\subsection{Determining the observability in \cite{Cheng2011IdentificationBCN}}
\begin{definition}[\cite{Cheng2011IdentificationBCN}]
BCN \eqref{BCN2} is called observable, if
there exists an input sequence $U\in(\Delta_M)^p$
for some $p\in\mathbb{Z}_{+}$, such that for any distinct states
$x_0,\bar x_0\in\Delta_N$, $Hx_0=H\bar x_0$ implies
$(HL)_{x_0}^{p}(U)\ne (HL)_{\bar x_0}^{p}(U)$.
\label{def7_observability}
\end{definition}
In this subsection, the observability of BCN \eqref{BCN2} refers to Definition \ref{def7_observability}.
According to Definition \ref{def7_observability},
to judge whether BCN \eqref{BCN2}
is observable, we need to
check the set $\mathcal{V}_n$ of all non-diagonal vertices
of its weighted pair graph
$(\mathcal{V},\mathcal{E},\mathcal{W})$.
Now we design an algorithm to construct a DFA
for BCN \eqref{BCN2} according to its weighted pair graph $(\mathcal{V},\mathcal{E},\mathcal{W})$.
The new DFA is denoted by $A_{\mathcal{V}_n}$, and accepts exactly every finite
input sequence by which not all non-diagonal state pairs can
be distinguished. The states of the DFA $A_{\mathcal{V}_n}$ are subsets of $\mathcal{V}$.
\begin{algo}\label{alg5_observability}
\begin{enumerate}
\item\label{item1_obser2}
Set $\Delta_M$ to be the alphabet of the DFA.
Set the set $\mathcal{V}_n$ of all non-diagonal
vertices of $\mathcal{V}$ to be
the initial state of the DFA.
\item\label{item2_obser2}
For each letter $\delta_M^j$, $j\in[1,M]$, find the value
for the transition partial
function of the DFA at $(\mathcal{V}_n,\delta_M^j)$.
The specific procedure is as follows:
For each $j\in[1,M]$, let $s_j:=\{v\in\mathcal{V}|\text{there is }v'\in \mathcal{V}_n
\text{ such that }(v',v)\in\mathcal{E},\text{ and }\delta_M^j\in\mathcal{W}((v',v))\}$.
If $s_j\ne\emptyset$, add $s_j$ to the state
set of the DFA and set $s_j$ to be the value of the transition
partial function at $(\mathcal{V}_n,\delta_M^j)$;
otherwise, the transition partial function
of the DFA is not well defined at
$(\mathcal{V}_n,\delta_M^j)$.
\item\label{item3_obser2}
For each new state $s$ of the DFA found in the previous step,
for each letter $\delta_M^j$, $j\in[1,M]$, find the value
for the transition partial
function of the DFA at $(s,\delta_M^j)$ according to
Step \ref{item2_obser2}.
\item\label{item4_obser2}
Repeat Step \ref{item3_obser2} until
no new state of the DFA occurs. (Since $\mathcal{V}$
is a finite set, so is its power set,
this repetition will stop.)
\item\label{item5_obser5}
Set all the states of the obtained DFA to be final states.
\end{enumerate}
\end{algo}
According to Algorithm \ref{alg5_observability}, the following
theorem holds.
\begin{theorem}
BCN \eqref{BCN2} is not observable in the sense of Definition
\ref{def7_observability} iff the DFA $A_{\mathcal{V}_n}$ generated by
Algorithm \ref{alg5_observability} recognizes language $(\Delta_M)^*$.
\label{thm5_observability}
\end{theorem}
\begin{proof}
Notice that BCN \eqref{BCN2} is not observable iff none of finite input sequences
can distinguish all state pairs of $\mathcal{V}_n$, that is, $L(A_{\mathcal{V}_n})=
(\Delta_M)^*$.
\end{proof}
From Proposition \ref{prop6_observability},
Theorem \ref{thm5_observability} and Algorithm \ref{alg5_observability},
the following result which follows can be used to judge
whether BCN \eqref{BCN2} is observable.
\begin{theorem}\label{alg6_observability}
BCN \eqref{BCN2} is not observable in the sense of Definition
\ref{def7_observability} iff the DFA $A_{\mathcal{V}_n}$ generated by
Algorithm \ref{alg5_observability} is complete.
\end{theorem}
\begin{example}\label{exam4_observability}
Check whether BCN \eqref{eqn2_observability} is observable.
According to Theorem \ref{alg6_observability}, we should check
whether DFA
$A_{\{(\delta_4^2,\delta_4^3),(\delta_4^2,\delta_4^4),
(\delta_4^3,\delta_4^4)\}}$ is complete.
From Fig. \ref{fig6:observability}, one sees that this DFA is complete.
Then by Theorem \ref{alg6_observability}, BCN \eqref{eqn2_observability}
is not observable.
\end{example}
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=2.0 cm, scale = 1.0, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\node[accepting,initial,state] (2*) {$23,24,34$};
\node[accepting,state] [above right of = 2*] (11) {$11$};
\node[accepting,state] [below right of = 2*] (22) {$22$};
\path [->] (2*) edge node {$2$} (11)
(2*) edge node {$1$} (22)
(22) edge [bend right] node {$2$} (11)
(22) edge [loop right] node {$1$} (22)
(11) edge [loop right] node {$1,2$} (11)
;
\end{tikzpicture}
\caption{The DFA $A_{\mathcal{V}_{
\{(\delta_4^2,\delta_4^3),(\delta_4^2,\delta_4^4),
(\delta_4^3,\delta_4^4)\} }}$ with respect to BCN
\eqref{eqn2_observability} generated by Algorithm
\ref{alg5_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, and the weight $k$ beside each edge denotes the input $\delta_2^k$.
}
\label{fig6:observability}
\end{figure}
\begin{remark}
In \cite{Laschov2013ObservabilityofBN:GraphApproach}, it is proved that
determining this observability is {\bf NP}-hard. Actually, the results of
\cite{Laschov2013ObservabilityofBN:GraphApproach} show that determining each of
the four types of observability is {\bf NP}-hard.
How to determine this observability has been
solved in \cite{Li2013ObservabilityConditionsofBCN}
by enumerating all possible input sequences of a common finite length.
However, one can use our method to find any input sequence that determines
the initial state.
Due to the independence of initial states, their method cannot
be applied to deal with
Definitions \ref{def3_observability} or \ref{def4_observability}.
\end{remark}
\subsection{Determining the observability in
\cite{Fornasini2013ObservabilityReconstructibilityofBCN}}
\begin{definition}[\cite{Fornasini2013ObservabilityReconstructibilityofBCN}]\label{def8_observability}
BCN \eqref{BCN2} is called observable, if for any distinct states
$x_0,\bar x_0\in\Delta_N$, for any input sequence $U\in(\Delta_M)^{\mathbb{N}}$,
$Hx_0=H\bar x_0$ implies
$(HL)_{x_0}^{\mathbb{N}}(U)\ne (HL)_{\bar x_0}^{\mathbb{N}}(U)$.
\end{definition}
In this subsection, the observability of BCN \eqref{BCN2}
means Definition \ref{def8_observability}.
According to Definition \ref{def8_observability}, BCN \eqref{BCN2} is not
observable iff there are two distinct states $\delta_N^i,\delta_N^j$ and an input
sequence $U\in(\Delta_M)^{\mathbb{N}}$ such that $H\delta_N^i=H\delta_N^j$ and $(HL)_{\delta_N^i}
^{\mathbb{N}}(U)=(HL)_{\delta_N^j}^{\mathbb{N}}(U)$.
Then the following theorem can be used to determine this observability.
\begin{theorem}
BCN \eqref{BCN2} is not observable in the sense of Definition
\ref{def8_observability} iff there is a non-diagonal vertex
$(\delta_N^i,\delta_N^j)$ of the weighted pair graph of BCN \eqref{BCN2}
such that the transition graph of the DFA
$A_{(\delta_N^i,\delta_N^j)}$ generated by Algorithm \ref{alg3_observability}
has a cycle.
\label{thm6_observability}
\end{theorem}
\begin{proof}
Since the transition graph has a finite number of vertices,
the graph has a cycle iff there is an input sequence
$U\in(\Delta_M)^{\mathbb{N}}$ such that $(HL)_{\delta_N^i}^{\mathbb{N}}(U)=
(HL)_{\delta_N^j}^{\mathbb{N}}(U)$.
\end{proof}
In fact, one can determine the observability directly from the weighted pair
graph of BCN \eqref{BCN2}. Theorem \ref{thm6_observability} directly implies
the following result.
\begin{theorem}
BCN \eqref{BCN2} is not observable in the sense of Definition
\ref{def8_observability} iff there is a cycle in its weighted pair graph,
and either the cycle contains a non-diagonal vertex, or there is a path
from a non-diagonal vertex to the cycle.
\label{thm11_observability}
\end{theorem}
\begin{example}\label{exam5_observability}
Check whether BCN \eqref{eqn2_observability} is observable.
By Theorem \ref{thm11_observability} and Fig. \ref{fig3:observability},
BCN \eqref{eqn2_observability} is not observable.
\end{example}
\begin{remark}
An equivalent condition for this observability is given
in \cite{Fornasini2013ObservabilityReconstructibilityofBCN} by
checking each pair of distinct periodic state-input trajectories of the
same minimal period and same length.
In addition, a specific critical length is given in \cite{Fornasini2013ObservabilityReconstructibilityofBCN}
such that if none of the input sequences of
that specific length can determine the initial
states, nor can input sequences of any other length.
Due to the independence of initial states and inputs, their method cannot
be used to deal with
Definitions \ref{def3_observability} or \ref{def4_observability} either.
\end{remark}
By the end of this subsection, we give a further result on this observability.
\begin{theorem}
Consider BCN \eqref{BCN2}. Denote
the number of non-diagonal vertices
of its weighted pair graph by $N_{nd}$.
The BCN is observable in the sense of Definition \ref{def8_observability}, iff
$N_{nd}=0$ or,
for all distinct states $x_0,\bar x_0\in\Delta_N$, for all input sequences
$U\in(\Delta_M)^{N_{nd}}$, $Hx_0=H\bar x_0$ implies
$(HL)_{x_0}^{N_{nd}}(U)\ne(HL)_{\bar x_0}^{N_{nd}}(U)$.
\label{thm1:length:NewObservability}
\end{theorem}
\begin{proof}
``if'':
Obvious by Definition
\ref{def8_observability}.
``only if'':
Assume that
$N_{nd}>0$ and there are distinct $x,x'\in\Delta_N$ and an input sequence
$U\in(\Delta_M)^{N_{nd}}$ such that
$Hx=Hx'$ and $(HL)_{x}^{N_{nd}}(U)=(HL)_{x'}^{N_{nd}}(U)$.
Use Algorithm \ref{alg3_observability} to generate DFA $A_{(x,x')}=(S,\Delta_M,\sigma,
(x,x'),S)$. Then $U\in L(A_{(x,x')})$. Denote $\sigma( (x,x'),U)$ by $v_U$.
If $v_U$ is diagonal, then $(HL)_{x}^{\mathbb{N}}(U(\delta_M^1)^{\infty})=
(HL)_{x'}^{\mathbb{N}}(U(\delta_M^1)^{\infty})$, and the BCN is not observable.
If $v_U$ is not diagonal, there are distinct $i,j\in[1,N_{nd}]$ such that either
$\sigma( (x,x'),U[1,i])=\sigma( (x,x'),U[1,j])$ or $(x,x')=\sigma( (x,x'),U[1,j])$,
for there are exactly $N_{nd}$ non-diagonal vertices.
By Theorem \ref{thm11_observability}, the BCN is not observable.
\end{proof}
\section{Pairwise nonequivalence of the four types of observability of Boolean control networks}
\label{sec4}
In this section, we prove that no pairs of the four types of observability of BCNs
are equivalent, which reveals the essence of nonlinearity of BCNs (shown in Fig.
\ref{fig10:observability}).
\begin{theorem}
If BCN \eqref{BCN2} is observable in the sense of Definition
\ref{def3_observability}, then it is also observable in the sense of
Definition \ref{def4_observability}. The converse is not true.
\label{thm2_observability}
\end{theorem}
\begin{proof}
The first part naturally follows from Definitions
\ref{def3_observability} and \ref{def4_observability}.
We use BCN \eqref{eqn2_observability} to prove the second part.
First, we prove that BCN \eqref{eqn2_observability} is not observable
in the sense of Definition \ref{def3_observability}.
Denote $M:=\delta_4[1,1,2,1,2,4,1,1]W_{[2,4]}{\bf1}_2
=\delta_2[1,2,2,1,1,1,4,1]{\bf1}_2=\left[
\begin{smallmatrix}
2 & 1 & 0 & 2\\ 0 & 1 & 1 & 0\\ 0 & 0 & 0 & 0\\
0 & 0 & 1 & 0
\end{smallmatrix}\right]$. Then for all $k\in\mathbb{Z}_{+}$, $M^k=\left[
\begin{smallmatrix}
* & * & * & *\\ * & * & * & *\\0 & 0 & 0 & 0\\0 & 0 & * & 0
\end{smallmatrix}\right].$ By
\cite[Theorem 3.3]{Zhao2010InputStateIncidenceMatrix},
BCN \eqref{eqn2_observability} is not controllable.
So one cannot use the test criteria proposed in
\cite{Cheng2009bn_ControlObserva}
to check whether BCN \eqref{eqn2_observability} is observable.
Next we prove that BCN \eqref{eqn2_observability} is not observable
by showing that for state $\delta_4^2$, there is no input sequence
such that the corresponding output sequence can determine it.
We only need to consider states $\delta_4^3,\delta_4^4$,
as $H\delta_4^1\ne H\delta_4^2$.
Arbitrarily given an input sequence $U\in(\Delta)^{\mathbb{N}}$. If $U(0)=\delta_2^1$,
then $L_{\delta_4^2}^1(\delta_2^1)=L_{\delta_4^3}^1(\delta_2^1)=\delta_4^2$.
Then for each such $U$, $(HL)_{\delta_4^2}^{\mathbb{N}}(U)=
(HL)_{\delta_4^3}^{\mathbb{N}}(U)$. Else if $U(0)=\delta_2^2$,
then $L_{\delta_4^2}^1(\delta_2^2)=L_{\delta_4^4}^1(\delta_2^2)=\delta_4^1$.
Then for each such $U$, $(HL)_{\delta_4^2}^{\mathbb{N}}(U)=
(HL)_{\delta_4^4}^{\mathbb{N}}(U)$. Then
BCN \eqref{eqn2_observability}
is not observable in the sense of Definition \ref{def3_observability}.
Second, we prove that BCN \eqref{eqn2_observability} is observable
in the sense of Definition \ref{def4_observability}.
We only need to check the state pairs $(\delta_4^2,\delta_4^3)$,
$(\delta_4^2,\delta_4^4)$ and $(\delta_4^3,\delta_4^4)$.
For $(\delta_4^2,\delta_4^3)$, $(HL)_{\delta_4^2}^1(\delta_2^2)=\delta_2^1\ne
(HL)_{\delta_4^3}^1(\delta_2^2)=\delta_2^2$.
For $(\delta_4^2,\delta_4^4)$, $(HL)_{\delta_4^2}^1(\delta_2^1)=\delta_2^2\ne
(HL)_{\delta_4^4}^1(\delta_2^1)=\delta_2^1$.
For $(\delta_4^3,\delta_4^4)$, $(HL)_{\delta_4^3}^1(\delta_2^1)=\delta_2^2\ne
(HL)_{\delta_4^4}^1(\delta_2^1)=\delta_2^1$.
Thus, BCN \eqref{eqn2_observability} is observable in the sense of
Definition \ref{def4_observability}.
\end{proof}
\begin{theorem}
If BCN \eqref{BCN2} is observable in the sense of Definition
\ref{def8_observability}, then it is also observable in the sense of
Definition \ref{def4_observability}. The converse is not true.
\label{thm7_observability}
\end{theorem}
\begin{proof}
The first part follows from Definitions
\ref{def4_observability} and \ref{def8_observability}.
We also use BCN \eqref{eqn2_observability} to prove the second part.
We have proved that BCN \eqref{eqn2_observability} is observable
in the sense of Definition \ref{def4_observability} in Theorem
\ref{thm2_observability}.
BCN \eqref{eqn2_observability} is not observable
in the sense of Definition \ref{def8_observability}, because
$H\delta_4^2=H\delta_4^4=\delta_2[1,2,2,2]\delta_4^2=\delta_2^2$ and
$(HL)_{\delta_4^2}^{\mathbb{N}}(\delta_2^2(\delta_2^1)^{\infty})=
(HL)_{\delta_4^4}^{\mathbb{N}}(\delta_2^2(\delta_2^1)^{\infty})$.
\end{proof}
\begin{theorem}
If BCN \eqref{BCN2} is observable in the sense of Definition
\ref{def8_observability}, then it is also observable in the sense of
Definition \ref{def7_observability}. The converse is not true.
\label{thm10_observability}
\end{theorem}
\begin{proof}
Assume that a given BCN \eqref{BCN2} is observable in the sense of
Definition \ref{def8_observability}, then arbitrarily given
$U\in(\Delta_M)^{\mathbb{N}}$,
for any distinct $\delta_N^i,\delta_N^j$, $H\delta_N^i=H\delta_N^j$ implies
$(HL)_{\delta_N^i}^{\mathbb{N}}(U)\ne(HL)_{\delta_N^j}^{\mathbb{N}}(U)$.
Since $N<+\infty$, there is $p\in\mathbb{Z}_{+}$ such that for any distinct
$\delta_N^i,\delta_N^j$, $H\delta_N^i=H\delta_N^j$ implies
$(HL)_{\delta_N^i}^{p}(U[0,p-1])\ne(HL)_{\delta_N^j}^{p}(U[0,p-1])$.
That is, the BCN is observable in the sense of
Definition \ref{def7_observability}.
To prove the second part, consider the following BCN:
\begin{equation}
\begin{split}
x(t+1) &= \delta_4[1,1,3,3,1,2,3,2]x(t)u(t),\\
y(t) &= \delta_2[1,1,2,2]x(t),
\end{split}
\label{eqn4_observability}
\end{equation}
where $t\in\mathbb{N}$, $x\in\Delta_4$, $y,u\in\Delta$.
Choose $U=\delta_2^1\in(\Delta)^1$. $H\delta_4^1=H\delta_4^2=\delta_2^1$,
$(HL)_{\delta_4^1}^1(U)=\delta_2^1\ne(HL)_{\delta_4^2}^1(U)=\delta_2^2$.
$H\delta_4^3=H\delta_4^4=\delta_2^2$,
$(HL)_{\delta_4^3}^1(U)=\delta_2^1\ne(HL)_{\delta_4^4}^1(U)=\delta_2^2$.
Then BCN \eqref{eqn4_observability} is observable in the sense of
Definition \ref{def7_observability}.
Consider any $U\in(\Delta)^{\mathbb{N}}$ such that $U(0)=\delta_2^2$.
Then $L_{\delta_4^3}^{\mathbb{N}}(U)=L_{\delta_4^4}^{\mathbb{N}}(U)$ and
$(HL)_{\delta_4^3}^{\mathbb{N}}(U)=(HL)_{\delta_4^4}^{\mathbb{N}}(U)$. That is,
BCN \eqref{eqn4_observability} is not observable in the sense of
Definition \ref{def8_observability}.
\end{proof}
\begin{theorem}
If BCN \eqref{BCN2} is observable in the sense of Definition
\ref{def7_observability}, then it is also observable in the sense of
Definition \ref{def3_observability}. The converse is not true.
\label{thm8_observability}
\end{theorem}
\begin{proof}
The first part holds naturally. To prove the second part,
consider the following BCN:
\begin{equation}
\begin{split}
x(t+1) &= \delta_4[1,1,1,3,1,2,3,2]x(t)u(t),\\
y(t) &= \delta_2[1,1,2,2]x(t),
\end{split}
\label{eqn3_observability}
\end{equation}
where $t\in\mathbb{N}$, $x\in\Delta_4$, $y,u\in\Delta$.
The weighted pair graph of BCN \eqref{eqn3_observability} is as shown
in Fig. \ref{fig7:observability}.
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=2.0 cm, scale = 1.0, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\node[state] (11) {$11$};
\node[state] (22) [right of = 11] {$22$};
\node[state] (12) [left of = 11] {$12$};
\node[state] (34) [right of = 22] {$34$};
\node[state] (33) [below of = 11] {$33$};
\node[state] (44) [below of = 22] {$44$};
\path [->] (11) edge [loop above] node {$1,2$} (11)
(22) edge node {$1$} (11)
(22) edge node {$2$} (33)
(12) edge node {$1$} (11)
(34) edge node {$2$} (22)
(44) edge node {$1$} (33)
(44) edge node {$2$} (22)
(33) edge node {$1$} (11)
(33) edge node {$2$} (22)
;
\end{tikzpicture}
\caption{The weighted pair graph of BCN \eqref{eqn3_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, the weight $k_1,k_2,\dots$ beside each edge denotes the
weight $\{\delta_2^{k_1},\delta_2^{k_2},\dots\}$ of the edge.}
\label{fig7:observability}
\end{figure}
The DFA $A_{\{(\delta_4^1,\delta_4^2),(\delta_4^3,\delta_4^4)\}}$ generated by Algorithm
\ref{alg5_observability} (see Fig. \ref{fig8:observability}) is complete.
Then by Theorem \ref{thm5_observability}, BCN \eqref{eqn2_observability} is
not observable in the sense of Definition \ref{def7_observability}.
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=2.0 cm, scale = 1.0, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\node[accepting,initial,state] (2*) {$12,34$};
\node[accepting,state] [above right of = 2*] (11) {$11$};
\node[accepting,state] [below right of = 2*] (22) {$22$};
\node[accepting,state] [right of = 22] (33) {$33$};
\path [->] (2*) edge node {$1$} (11)
(2*) edge node {$2$} (22)
(11) edge [loop above] node {$1,2$} (11)
(22) edge node {$1$} (11)
(22) edge node {$2$} (33)
(33) edge node {$2$} (22)
(33) edge node {$1$} (11)
;
\end{tikzpicture}
\caption{The DFA $A_{\mathcal{V}_n
}$ with respect to BCN
\eqref{eqn3_observability} generated by Algorithm
\ref{alg5_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, the weight $k$ of each edge denotes the input $\delta_2^k$.
}
\label{fig8:observability}
\end{figure}
The DFAs $A_{\delta_4^1}$ and $A_{\delta_4^3}$ generated by Algorithm
\ref{alg1_observability} (see Fig. \ref{fig9:observability})
satisfy $\delta_2^2\notin L(A_{\delta_4^1})$ and
$\delta_2^1\notin L(A_{\delta_4^3})$. Then by Theorem \ref{thm3_observability},
BCN \eqref{eqn3_observability} is observable in the sense of Definition
\ref{def3_observability}.
\begin{figure
\centering
\begin{tikzpicture}[>=stealth',shorten >=1pt,auto,node distance=1.5 cm, scale = 0.7, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\tikzstyle{time}=[inner sep=0pt,minimum size=5mm]
\node[accepting,initial,state] (2*) {$12$};
\node[accepting,state] [above right of = 2*] (11) {$11$};
\path [->] (2*) edge node {$1$} (11)
(11) edge [loop right] node {$1,2$} (11)
;
\node[time] [above right of = 2*] (*) {};
\node[time] [right of = *] (**) {};
\node[time] [right of = **] (***) {};
\node[accepting,initial,state] [below right of = ***] (3*) {$34$};
\node[accepting, state] [above right of = 3*] (22*) {$22$};
\node[accepting, state] [right of = 22*] (11*) {$11$};
\node[accepting, state] [below of = 11*] (33*) {$33$};
\path [->] (3*) edge node {$2$} (22*)
(22*) edge node {$2$} (33*)
(33*) edge node {$2$} (22*)
(22*) edge node {$1$} (11*)
(33*) edge node {$1$} (11*)
(11*) edge [loop right] node {$1,2$} (11*)
;
\end{tikzpicture}
\caption{The DFAs $A_{\delta_4^1}$ and $A_{\delta_4^3}$
with respect to BCN
\eqref{eqn3_observability} generated by Algorithm
\ref{alg1_observability},
where the number $ij$ in each circle denotes the state pair $(\delta_4^i,
\delta_4^j)$, the weight $k$ beside each edge denotes the input $\delta_2^k$.
}
\label{fig9:observability}
\end{figure}
\end{proof}
\begin{theorem}
If BCN \eqref{BCN2} is observable in the sense of Definition
\ref{def8_observability}, then it is also observable in the sense of
Definition \ref{def3_observability}. The converse is not true.
\label{thm9_observability}
\end{theorem}
\begin{proof}
The first part holds naturally. To prove the second part,
consider BCN \eqref{eqn3_observability} again.
We have proved that BCN \eqref{eqn3_observability} is observable in the sense of
Definition \ref{def3_observability} in Theorem
\ref{thm8_observability}. Note that in Fig. \ref{fig9:observability},
the DFAs are just the corresponding ones
generated by Algorithm \ref{alg3_observability}.
Then By Theorem
\ref{thm6_observability},
BCN \eqref{eqn3_observability} is not
observable in the sense of Definition \ref{def8_observability}.
\end{proof}
\section{Concluding remarks}\label{sec6}
In this paper, we solved the problem on determining the observability
of Boolean control networks (BCNs) completely
by using techniques in finite automata.
Also, we showed that no pairs of all the four types of observability notions are equivalent
by counterexamples, which
reveals the essence of nonlinearity of BCNs (shown in Fig. \ref{fig10:observability}).
\begin{figure
\centering
\begin{tikzpicture} [>=stealth',shorten >=1pt,auto,node distance=2.5 cm, scale = 1.0, transform shape,
->,>=stealth,inner sep=2pt,state/.style={shape=circle,draw,top color=red!10,bottom color=blue!30},
point/.style={circle,inner sep=0pt,minimum size=2pt,fill=},
skip loop/.style={to path={-- ++(0,#1) -| (\tikztotarget)}}]
\node[state] (7) {Def. \ref{def7_observability}};
\node[state] [right of =7] (4) {Def. \ref{def3_observability}};
\node[state] [below of =4] (5) {Def. \ref{def4_observability}};
\node[state] [left of =5] (8) {Def. \ref{def8_observability}};
\draw [->] ([yshift=2pt] 7.east) -- node {$+$} ([yshift=2pt] 4.west);
\draw [->] ([yshift=-2pt] 4.west) -- node {$-$} ([yshift=-2pt] 7.east);
\draw [->] ([xshift=2pt] 4.south) -- node {$+$} ([xshift=2pt] 5.north);
\draw [->] ([xshift=-2pt] 5.north) -- node {$-$} ([xshift=-2pt] 4.south);
\draw [->] ([yshift=2pt] 8.east) -- node {$+$} ([yshift=2pt] 5.west);
\draw [->] ([yshift=-2pt] 5.west) -- node {$-$} ([yshift=-2pt] 8.east);
\draw [->] (8) to [out=22.5, in=249.5] node {$+$} (4);
\draw [->] (4) to [out=204.5, in=67.5] node {$-$} (8);
\draw [->] ([xshift=2pt] 8.north) -- node {$-$} ([xshift=2pt] 7.south);
\draw [->] ([xshift=-2pt] 7.south) -- node {$+$} ([xshift=-2pt] 8.north);
\end{tikzpicture}
\caption{The implication relationships between Definitions
\ref{def3_observability}, \ref{def4_observability},
\ref{def7_observability} and \ref{def8_observability},
where ``$+$'' means ``implies'' and ``$-$'' means ``does not imply''.
}
\label{fig10:observability}
\end{figure}
Note that the computational complexity of algorithms for determining the first and
fourth types of observability is in exponential time, and the algorithms for the other two types
are in doubly exponential time.
How to reduce the computational complexity effectively is
a challenging and urgent problem, and we are naturally concerned with
``Is there a nondeterministic polynomial time algorithm for determining
the observability of BCNs?'' Furthermore, we conjecture that
``Determining the observablity of BCNs is {\bf PSPACE}-hard.''
\section*{Acknowledgment}
The first author is in debt to Prof. Jarkko Kari, Drs. Charalampos Zinoviadis,
Ville Salo and Ilkka T\"{o}rm\"{a} at the
University of Turku, Finland for fruitful discussions while visiting
the same university in 2013.
Both authors thank Mr. Chuang Xu
at the University of Alberta, Canada, the
anonymous referees and the associate editor for their
valuable comments that highly improve the presentation of this paper.
\ifCLASSOPTIONcaptionsoff
\newpage
\fi
|
\section*{The UK financial mathematics M.Sc.}
The last three decades have seen an unprecedented explosion of activity in a new sub-discipline of mathematics: {\it financial mathematics}. The emergence of this field of research and study has paralleled the manifold expansion of the financial services industry and, accordingly, has transformed the role that mathematics departments now play in feeding the financial services industry.
\medskip
\noindent For many university mathematics departments, the volume of demand for taught courses that have been set up in the direction of financial mathematics, in particular at the postgraduate level, has proved to be a gift from the gods. In the UK, the country to which we henceforth restrict the majority of our forthcoming discussion, the {\it Master of Science (M.Sc.) in financial mathematics} has been pioneered for the best part of the last twenty years. Such masters programmes are marketed as premium postgraduate education, with the allure of a well-paid career in the banking industry thereafter. Moreover, they are often populated by large numbers of foreign students unable to access this type of specialised education in their country of origin. Accordingly, these study programmes are generally priced with some of the highest annual tuition fees in the UK. (To give a little perspective, quite a few universities are now charging well in excess of what one would pay for a top-end, brand new family-sized car.) For some universities, the fees are found to be double, treble or even quadruple those of many other masters programmes on the same campus, in the same faculty or even in the same department. A brief glance at the table in the Appendix A reveals some eye-watering statistics. Appendix B illustrates further the severity of the differential between the cost of undertaking a regular M.Sc. in Mathematics (or a related course with a very strong mathematical component) compared to an M.Sc. in Financial Mathematics. For home/EU students, the situation is considerably worse than for overseas students who, traditionally, have always paid significantly higher fees to attend UK universities. Indeed one might even question from the data presented whether home/EU students are being priced out as there is a deliberate focus on the overseas market where more money is to be made. Nonetheless, even for overseas students, there is a clear and significant pricing discrepancy with `nearest-neighbour' mathematics masters programmes.
\medskip
\noindent For many years, the setting of fees for postgraduate taught courses has been left to the discretion of the awarding institution. It is assumed that the principle of the free market should come into play. Universities are expected to set fees that represent a balanced view on reputation, expertise, running costs and demand. As British academia moves into an age where the commercialisation of education is becoming ever more the norm, we examine whether things have gone too far in the case of the M.Sc. in financial mathematics, thereby highlighting some of the incongruities and future dangers of free-market education in the UK.
\section*{Niche mathematics}
\noindent In 1997, Robert Merton and Myron Scholes were awarded the Nobel Prize for Economics for their work on the use of probability theory and its connections to partial differential equations, in order to show how it was possible to explicitly hedge and hence price a whole family of option contracts in an idealised market. One of the key innovations in their work was the translation of rational behaviour of market agents into a rigorous mathematical framework.
Although their model of a financial market was an idealised version of reality, their calculations involved a remarkable and stunning application of one of the most important developments in the theory of probability: {\it stochastic calculus}. This was a truly exciting development that provided a rare example of where a highly non-trivial mathematical theory could be used to address a real-world application. Moreover, it offered the possibility of much more sophisticated computations and, eventually, opened up the sub-discipline that we now call {\it financial mathematics}.
\medskip
\noindent This Nobel prize-winning work dated back to the mid-seventies, and pertained to an original publication that Scholes had co-authored with Fisher Black in 1973. (Black died in 1995 and could not be awarded the prize posthumously.) Black, Scholes and Merton were all financial economists and, whilst Black and Scholes had some initial difficulties publishing their work, momentum in the development of the so-called Black-Scholes-Merton theory developed through the late seventies. By the early 1990s this theory had captured the attention of the mathematical mainstream. Thanks to a general sense of awareness, in part brought about by the award of the Nobel Prize, a much broader perspective on the original contributions of the financial economist trio had been realised by the late 1990s and the amount of research being published in mathematical journals and books blossomed dramatically. By the turn of the Millennium, financial mathematics was well and truly established as a field of research. Today, it claims the right to be recognised as a mainstream sub-discipline of mathematics and is supported by a huge international community of researchers.
\medskip
\noindent In parallel to (as well as stimulating) these academic developments, the global derivatives markets had significantly swollen in size thanks to a process of deregulation that began over 30 years ago, predominantly in the US and UK. By the 1980s, traders were selling in large volume complex contracts `over the counter', which were generally known as {\it financial derivatives}. For an up-front premium, a financial derivative would allow investors to collect potential future financial returns, or perhaps no return at all, depending on how future markets played out. Option contracts, such as those originally studied by Black, Scholes and Merton, were relatively straightforward examples of such derivatives. By the mid 1990s, for good or for bad, the volume and variety of financial derivatives that were being traded reflected the degree of sophistication with which traders were using these products to build complex investment portfolios.
As markets are unpredictable, each financial derivative carries an individual risk and, accordingly, portfolios of derivatives carry compounded and interactive risks. With such large-scale and complicated trading taking place, financial institutions sensed the increasing need for in-house mathematicians to quantify the risks involved.
And so, the number of positions for so-called {\it quantitative analysts}, or {\it quants} for short, along with other supporting roles of a numerate nature, has exploded in number over the last 20 years or more.
\medskip
\noindent Together with the US, UK academia was quick to respond to this rapidly expanding job market and to identify a rather unique opportunity to set up specialist postgraduate mathematics teaching programmes that, in some sense, also served as high-level vocational training for the quantitative financial services industry. Here was a self-reinforcing niche market. As more financial mathematics graduates proliferated into various quantitative support roles in the financial sector, the more the next generation of students perceived the necessity to engage with such training. The M.Sc. in financial mathematics has become the iconic intermediate goal of many young women and men, particularly those with a first degree in mathematics, statistics or physics, whose long-term aspiration is to `work in banking'.
\section*{Cash cow}
\noindent The very first UK-based M.Sc. in financial mathematics made its debut at the University of Warwick in 1996. A similar programme was launched the following year jointly by the University of Edinburgh and Heriot--Watt University in the Scottish capital. According to their webpages, both programmes still enrol very healthy numbers today: approximately 50 in Warwick every year and 25-35 in Edinburgh every year. This is even more impressive when one takes account of the number of competing courses that have sprung up in the UK in the meantime, not to mention in other countries around the world. As can be seen from Appendix A, there are currently over 30 M.Sc. courses on the topic of financial mathematics, that can be found the length and breadth of the UK. The majority of annual fees for these masters programmes exceed the upper cap on undergraduate fees (\pounds 9,000). Around two thirds of them exceed \pounds 15,000, going as high as \pounds 31,000, for home and EU students. Whilst it is a long-standing practice in the UK higher education system that a higher rate is charged for non-EU students, the majority of masters in financial mathematics do not discriminate between whether a student is from the EU or not. A back-of-the-envelope calculation shows that a department with a well maintained M.Sc. degree of this kind should easily manage to bring in well over half a million pounds a year from it. Some programmes currently in existence are more than capable of bringing in well over a million pounds per year.
\medskip
\noindent The lure of this kind of income for mathematics (and partner) departments has proved to be far too attractive to ignore in light of the financial strain that British academia is under. Over the last 20 years, encouraged by the government, UK universities have experienced a massive programme of expansion, opening their doors to many more home and international students.
Naturally, state funding has spread thinner and become seemingly less generous. Student grants have been replaced by student loans, study fees for undergraduate home students have been passed from the state to the student, national research funding council scholarships for taught postgraduate courses have all but disappeared and research grant capture has become increasingly more competitive. Universities have found themselves in a position where, in order to survive in the modern academic climate, it has become necessary to explore a whole variety of new sources of income.
\medskip
\noindent Laying on low-cost, easy-to-populate taught postgraduate degrees with premium rate fees is a straightforward way of generating income. Pursuing the ever-expanding, higher fee-paying international student market is another.
Running an M.Sc. in financial mathematics has proved to be very successful in both respects for many universities.
\section*{Value for money}
\medskip
\noindent How is value for money of premium-fee M.Sc. programmes justified? {\it Specialism}, {\it opportunity} and {\it demand} are three of the words that many offering such education cite.
\medskip
\noindent For the large majority of financial mathematics masters programmes on offer, it is not unusual that a proportion of what is taught is standard material which can be found in third or fourth year undergraduate programmes. Indeed, it is not uncommon that, in the interests of the economy of scale, masters students will find themselves sharing the lecture theatre with undergraduates and postgraduates of other masters programmes for some of their modules.
Specialist material, exclusive to the masters programme itself, is also to be found.
For example, it is normal in the UK that the successful completion of an M.Sc. requires the submission and satisfactory assessment of a summer dissertation. Here the student has the opportunity, over a period of around 3-4 months, to be guided through an extensive programme of reading and research, addressing considerably more advanced theory and application. However, most of what is offered up in this respect is considered `specialist' because it addresses current trends in academic research rather than what is actually used or needed in the industrial roles that the majority of these masters graduates go on to occupy. This discrepancy is an important point that is worth emphasising.
\medskip
\noindent
A surprising fact concerning modern research in financial mathematics is that, although presenting substantial technical challenges, it does not find its way into industry as often as one might expect. Many of the underlying models which are studied in theory are simply inappropriate or impractical to work with. It is not uncommon for the academic financial mathematician to work from the safety of an axiomatic treatment of a financial scenario in order to investigate phenomenological behaviour or to develop a mathematical method within that context. The discrepancy with reality comes about because, unlike for example physical systems, it is difficult to reasonably model the mechanisms that drive the randomness in generic financial markets by appealing to fundamental universal laws.
\medskip
\noindent It is not true, however, that academic research in financial mathematics has no impact at all. There are some very important successful examples attributed to quite exceptional mathematicians. These individuals have carved out relatively rare careers, interacting closely and consistently with banks and influencing in-house banking activity, sometimes from within an academic position, sometimes as an employee of a bank and sometimes both. But such individuals are few and far between and the level of mathematical experience involved goes far beyond what can be established on the back of an M.Sc. in financial mathematics.
Moreover, it is usual that their interaction takes place with some of the bigger banks or specialist consultancies that have a unique inner core of `super-quants' who are educated and experienced as academic researchers well beyond the level of a Ph.D.
\medskip
\noindent
As far as masters-level vocational training is concerned, the reality for most academics is that many are not up to date on the kind of challenges and routines that are undertaken en masse by industrial quantitative analysts on a daily basis. Moreover, few can claim to have recently seen the inside of a bank and interacted in a mathematical capacity with its quants.
Therefore, one has to question the extent to which the academic specialism that is injected into postgraduate financial mathematics taught courses should be valued {\it over and above} any other numerate discipline taught at this level, in terms of vocational training.
\medskip
\noindent That said, there is still the issue of what employers want. Webpages, posters and brochures for postgraduate education in financial mathematics often advocate the value of the course in question as useful in preparing for a job in the financial industry and point towards some of the many employers that their graduates have gone on to work for. But, realistically, to what extent are banks interested in hiring individuals with such specific masters qualifications? It would be more appropriate if marketing for these programmes could use hard data to indicate the true demand for such education {\it from within the industry itself}. In the experience of the author of this article, an academic who {\it has} spent a modest amount of time inside a number of banks and financial institutions engaging with their quantitative analysts, very few banks seem to care whether or not a student is in possession of an M.Sc. in financial mathematics.
\medskip
\noindent Indeed, hiring strategies for quants (the industrial job whose technical requirements most closely match the training on the typical Financial Mathematics M.Sc.) are based around looking for sharp problem solvers who can deal with the fast pace of work and learn and develop in-house technical procedure. The familiarity with certain concepts and academic specialisms that an M.Sc. in financial mathematics offers, some of which are not directly relevant to working practice, is not necessarily deemed to be of huge value to the quant employer, who would expect that the smart hire can pick up what is needed anyway once in post.
Often, but not always, this means that it is those who are trained to the level of a Ph.D. in a numerate discipline (not necessarily financial mathematics) that have a competitive advantage as far as quant jobs are concerned. If nothing else, this is because of a self-selecting bias through their ability to undertake a Ph.D. in the first place.
\medskip
\noindent This does not contradict the observation made by vendors of financial mathematics postgraduate education that many of their graduates do go on to work in the finance industry. It just means that the technical needs of the roles they are hired into (for example careers in risk management or trading, which have relatively soft graduate technical numerical requirements) are not specific to an M.Sc. in financial mathematics. Moreover, self-selection again plays a role here; commitment to an expensive one-year postgraduate certificate is a strong affirmation of the desire to end up working in the financial services industry.
\medskip
\noindent Another justification for the value for money that can be given, but only for a handful of M.Sc. programmes in financial mathematics (see Appendix A), is that they offer the opportunity for a summer placement in industry or a summer research project in collaboration with industry. However, with the exception of one university, it is not entirely clear from publicly available on-line course prospectuses exactly how many students on each programme have this experience accessible to them during the summer months. Certainly, if mentioned at all, the choice of language one tends to find regarding the exact nature of such opportunities within the degree is rather vague and non-committal.
\medskip
\noindent More generally, participating in an M.Sc. in financial mathematics will almost certainly give students the opportunity to enter job interviews in the financial services industry at large (not just quant jobs) with more confidence. But it seems somewhat anecdotal to claim that this would {\it drastically} improve one's chances of employment over and above any other masters level education in mathematics given that the core theory that supports such masters programmes is relatively easy to access in most mathematics undergraduate and taught postgraduate programmes anyway. Certainly there is no apparent data (at least not supplied by vendors of financial mathematics masters programmes) to support such a claim. In particular, the kind of data that is desired here would certainly need to take account of the phenomenon of self-selection as noted above.
\medskip
\noindent All of the above arguments aside, one still cannot discount the fact that there is ongoing demand for these masters courses from the students themselves, not to mention many more financial mathematics bachelor programmes that are also in existence (priced in line with standard undergraduate fees). More broadly, there is at least one commercially operated programme of courses which could be argued to be in competition with university education. A Certificate in Quantitative Finance is provided by Fitch Learning in collaboration with the Wilmott Forum, which currently charges \pounds 11,950 for a six-month programme with six specialist modules.
\medskip
\noindent As long as students are keen to pay the current costs involved in postgraduate training in financial mathematics, then one could assert that the product speaks for itself. But, realistically speaking, is this an acceptable justification for the extent of the pricing differential observed in universities?
\medskip
\medskip
\noindent Amidst the prevailing culture shift in UK academia to a more corporate approach, one has to ask whether `opportunistic education' has started to appear. The packaging, marketing and selling of education is part and parcel of the many challenges that universities are now confronted with as a matter of survival. Against this backdrop, one has to question the motivation behind the scramble in recent years by institutions wanting to offer postgraduate education in financial mathematics whilst imposing relatively (in some cases, exceptionally) high fees, when, given the arguments above, there isn't necessarily clear evidence of value for money at that tariff. There are grounds to feel concern that there is an element of implicit cynicism around the running of such courses. Said another way, there is a danger that the answer to the question {\it `why are fees for financial mathematics postgraduate taught courses so high in relative terms?'} may simply be: {\it `because universities can get away with it'.}
\medskip
\noindent Would any other public service industry tolerate or be tolerated for such an approach to its pricing policy?
\section*{Responsible education}
Financial mathematics is beautiful. It has proved itself again and again, showing amazing robustness in the way that financial and economic concepts and problems can be reformulated in an abstract context and addressed using a wide variety of different mathematical tools. Moreover, there can be no doubt that heightened interest in financial mathematics has helped to motivate some of the major waves of development in probability theory over the last 30 years.
Nonetheless, financial mathematics has its critics. There are those who feel that the intelligentsia should not promote a false sense of security in the use of mathematics where there are known to be inaccuracies, especially within an economic system that has led to an extreme, and ultimately devastating, financial breakdown.
\medskip
\noindent But financial mathematics is not about making money and it is not about promoting capitalism (or any other socio-economic ideal for that matter). Financial mathematics is a science, it is a theory that aims to quantify risk and randomness within economic scenarios, with a view to characterising {\it rational} behaviour, thereby promoting {\it fairness}. In the experience of this author, most mathematicians who advocate the use of this theory (both within academia and within the financial services industry) have a sincere interest in seeing it used to improve the quality of decision making within financial markets, leading to a more accountable and responsible system.
\medskip
\noindent In this respect, financial mathematics {\it is} an important intellectual discipline and deserves to be taught at depth in an academic environment. There is a lot to share and a lot to think about. And like all other scientific disciplines, it deserves to be delivered in a responsible way. In this respect most of the masters programmes in existence are extremely rewarding in their intellectual content. Indeed, some of the programmes listed in Appendix A are exemplary in their academic delivery, for example, in turning their attention to the mathematics of post-crash finance. But, whilst this is no doubt relevant to the issue of value for money, there is a much bigger picture to consider here.
\medskip
\noindent
There is a real danger that the M.Sc. in financial mathematics has become an iconic symbol within UK academia of an overzealous attitude towards the commercialisation of education.
Prospective students are ill informed about the relationship between research in financial mathematics, which guides what is taught on masters programmes devoted to this theme, and what is actually needed in the banking industry for the different types of numerate jobs. Marketing for masters programmes in financial mathematics often confuses what is needed by an elite core of quantitative analysts in large banks and consultancies (who typically look far beyond the training that a UK M.Sc. in financial mathematics has to offer) with other softer quantitative roles in banking (for which a whole array of other quantitative undergraduate and postgraduate qualifications are equally relevant).
\medskip
\noindent The fee structure of the UK financial mathematics M.Sc. sets an undesirable precedent for other degrees in the future. Take, for example, postgraduate taught degrees in statistics. Unlike the case of an M.Sc. in financial mathematics, one can make a much stronger case that there {\it is} directly relevant high-level vocational training taking place in a statistics M.Sc. Moreover, such a qualification genuinely {\it is} considered to be needed by a whole variety of industries that {\it do} hire masters-level students into well-paid jobs as statisticians. What is now stopping universities charging significantly higher fees than the average for these degrees? Indeed, there is plenty of demand in this field too.
Do we want our academic system in the UK to fall into the trap in which academic excellence is confused with costly branding? As a quick google will confirm, there is already much debate of an emerging bubble in the education market, in particular in the US. If this is the case, would it not be a sad irony if {\it financial mathematics} found itself at the head of such a market in the UK?
\medskip
\noindent
\medskip
\noindent Arguing the case of a sellers' market for this kind of education seems the most dangerous justification of all for premium fees. Given the significant differential in fees with nearest-neighbour mathematics postgraduate degrees for home-EU students at a large number of institutions and given the large proportion of non-EU students that are enrolled every year, universities could potentially risk exposing themselves to an accusation of `predatory education' directed at wealthy foreign students. If the principle of appealing to `demand' to justify the setting of exceptionally high fees becomes a more common practice, would the system not attract the uncomfortable prospect of external regulation?
\medskip
\noindent For the specific case of the M.Sc. in financial mathematics, justifying premium fees on the basis of there being a sellers' market also opens up another genuine concern. On completion of their studies, some overseas students are returning to developing and less regulated economies where it is unclear to what extent a qualification of this kind, marketed and sold to students the way that it is, is being put to use. This aspect alone may prove to be the most frightening consequence of all in the financial mathematics postgraduate education frenzy. Who is checking these things?
\medskip
\noindent Twenty years ago, when the profession of the quantitative analyst was still relatively young, the M.Sc. in financial mathematics meant something completely different to what it does today. The financial services industry has matured from where it was and, at least in the main financial hubs (London, New York, Frankfurt, Paris, Hong Kong, Singapore, Tokyo, etc.), it has a much clearer sense of identity and a clearer understanding of how it is prepared to interact with high-end mathematics.
The role that universities can play in vocational training for this profession has accordingly changed.
As academics, we need to take a closer look at the data available to correctly understand, quantify and qualify the true value of what is currently on offer by way of taught postgraduate education in financial mathematics.
\medskip
\noindent Conversely, {\it if} UK university education in this field is to be subjected to the full force of the free market ideal, then is it clear that the students, i.e. the consumers, are correctly informed when choosing to enroll for a M.Sc. in financial mathematics? Students should demand greater justification when premium fees are involved.
How do these fees break down? Can the university really quantify the value it will have in their future career on demand (other than a fancy poster, brochure or webpage with quotes from satisfied students)? Can a university's claims be independently verified? What is the relative additional value for money over and above nearest-neighbour masters programmes whose graduates also feed the financial services industry, and whose fees already take account of the fact that graduate education has value to employers at large?
\medskip
\noindent
After all,
a fundamental concept that students will learn on an M.Sc. in financial mathematics is that rational pricing forbids the occurrence of arbitrage.
\section*{Acknowledgements}
\noindent The motivation for this document comes from
more than 15 years of the author's experience in teaching, managing and externally examining M.Sc. programmes in financial mathematics across a wide spectrum of universities, both nationally and internationally.
Parts of this document are inspired by many discussions with colleagues over the aforesaid period in light of the growing commercialisation of education in the UK. More recently, whilst writing this article, the author has had a number of extensive, candid and extremely helpful discussions with stakeholders in postgraduate financial mathematics education, in both academia and industry. I am most grateful for their input.
\newpage
\section*{Appendix A}
{\footnotesize The following table lists current postgraduate taught courses in the UK with emphasis on financial mathematics, ranked in decreasing order of fees for home/EU students. This data has been collected from on-line, publicly available information provided by each of the institutions. Fees indicated are for the academic year 2014-2015 or 2013-2014 where information for the former is not yet available. Figures, expressed in k\pounds, have been rounded to one decimal place. The list of postgraduate courses that relate to financial mathematics is more extensive than the one presented below, but we have chosen to include only the ones which present significant mathematical content.
The fourth column, `P', indicates whether there is mention on the programme webpages of the possible availability of of an industrial placement or industrially oriented research project during the summer dissertation period. In this column, `N' means there is no mention, `V' means there is vague mention, `E' means there is explicit mention of opportunities (without quantifying how many students will have access to this opportunity or whether there are any guarantees) and `Y' means that a strong commitment to engaging the majority of students with an industrial placement or industrially oriented research project has been indicated.
\[
\hspace{-0.4cm}\begin{footnotesize}
\begin{array}{| l | c | c |c|l|}
\hline
\text{
University} & \text{ home/EU (k\pounds)} & \text{ non-EU(k\pounds)} & \text{P} & \text{M.Sc. Course name}\\
\hline
\hline
\text{Warwick} & 31&31 &\text{E}& \text{Financial Mathematics}\\
\hline
\text{Oxford} & 27.5 & 31.1 &\text{N}^{1}& \text{Mathematical Finance (part time)}
\\
\hline
\text{Imperial} & 27 & 27 &\text{Y}& \text{Mathematics and Finance}
\\
\hline
\text{Oxford} & 24.9 & 24.9 &\text{N}& \text{Mathematical and Computational Finance
}
\\
\hline
\text{City} & 23 & 23 &\text{V}& \text{Financial Mathematics}
\\
\hline
\text{LSE} & 23 & 23 &\text{N}& \text{Financial Mathematics}
\\
\hline
\text{LSE} & 23 & 23 &\text{N}& \text{Risk and Stochastics}
\\
\hline
\text{King's} & 22.7 & 22.7 &\text{N}& \text{Financial Mathematics}
\\
\hline
\text{Manchester} & 21.4 & 21.4 &\text{E}& \text{Quantitative Finance: Financial Engineering }
\\
\hline
\text{Reading} &19.8 & 19.8&\text{N}& \text{Financial Engineering}
\\
\hline
\text{UCL} &19.6 &21.7 &\text{N}& \text{Financial Mathematics}
\\
\hline
\text{Birkbeck} & 19 & 20 &\text{N}& \text{Financial Engineering}
\\
\hline
\text{Edinburgh/Heriot--Watt} &19 &22 &\text{E}& \text{Financial Mathematics}\\
\hline
\text{Brunel} & 17.5 & 17.5 &\text{V}& \text{Financial Mathematics}
\\
\hline
\text{York} & 17.4&22.2 &\text{N}& \text{Financial Engineering}
\\
\hline
\text{York} & 16.5&21 &\text{N}& \text{Financial Mathematics}
\\
\hline
\text{Queen Mary} &15.5 & 18&\text{N}& \text{Mathematical Finance}
\\
\hline
\text{Birmingham} & 15.5 & 15.5 &\text{N}& \text{Mathematical Finance}
\\
\hline
\text{Liverpool} & 15.4 & 15.4 &\text{N}& \text{Financial Mathematics}
\\
\hline
\text{Exeter} &11.9 &19.5 &\text{N}& \text{Financial Mathematics}\\
\hline
\text{Manchester} & 10.7 & 18.4 &\text{V}& \text{Mathematical Finance}
\\
\hline
\text{Leeds} & 9 & 19 &\text{N}& \text{Financial Mathematics}
\\
\hline
\text{Leicester} &8.5 &16 &\text{E}& \text{Financial Mathematics and Computation}
\\
\hline
\text{Glasgow} & 8.3& 17.2 &\text{N}& \text{Financial Modelling}\\
\hline
\text{Strathcylde} & 7&13 &\text{N}& \text{Quantitative Finance}
\\
\hline
\text{Loughborough} & 6.3& 13.8&\text{N}& \text{Mathematical Finance}
\\
\hline
\text{Birmingham} & 6 & 13.7 &\text{N}& \text{Financial Engineering}
\\
\hline
\text{Sheffield} & 6 &15.4 &\text{N}& \text{Statistics with Financial Mathematics}
\\
\hline
\text{Nottingham} & 6& 13.7 &\text{N}& \text{Numerical Techniques for Finance}
\\
\hline
\text{Swansea} & 5&12.5 &\text{N}& \text{Mathematics and Computing for Finance}\\
\hline
\end{array}
\end{footnotesize}\]
$^1$ This programme is designed for individuals who are already employed in the financial services industry.
\pagebreak
\section*{Appendix B}
{\footnotesize For comparison with the table in Appendix A, we list fees for other postgraduate taught courses in mathematics or programmes with a very strong mathematical content (what we refer to as `nearest-neighbour' programmes in the main text).
Some of the departments listed in Appendix A are not listed here as they do not offer such programmes. Some universities not listed in Appendix A are also to be found below. For universities that are found in both tables, and for convenient cross-referencing, we have transferred the figures from the first table {\color{blue}(in blue)} into the second table and included the relative multiplicative factor {\color{red}[in red]}. As with the previous table, this data has been collected from on-line, publicly available information provided by the respective institutions. Fees indicated are for the academic year 2014-2015 or 2013-2014 where information for the former is not yet available. Figures, expressed in k\pounds, have been rounded to one decimal place and courses are listed in location alphabetical order.}
\[
\hspace{-0.6cm}\begin{footnotesize}
\begin{array}{| l | l | l |l|}
\hline
\text{
University} & \text{Fee (k\pounds) home/EU} & \text{Fee (k\pounds) non-EU} & \text{M.Sc. Course name}\\
\hline
\hline
\text{Aberdeen} & 3.4 & 12.6 & \text{Math.}
\\
\hline
\text{Bath} & 6 & 18.6 & \text{Math. Sci.}
\\
\hline
\text{Bath} & 6 & 18.6 & \text{Modern Appl. Math.}
\\
\hline
\text{Bath} & 6 & 18.6 & \text{Math. Biol.}
\\
\hline
\text{Birkbeck} & 4 {\color{blue}\mbox{ } (19)} {\color{red}\mbox{ } [4.5]}& 7.4 {\color{blue}\mbox{ } (20)} {\color{red}\mbox{ } [2.70]}& \text{Math.}
\\
\hline
\text{Birmingham} & 5.9 {\color{blue}\mbox{ } (15.5)} {\color{red}\mbox{ } [2.63]}& 13.7 {\color{blue}\mbox{ } (15.5)} {\color{red}\mbox{ } [1.13]}& \text{Math. Model./MORSE}
\\
\hline
\text{Bristol} & 8& 17.5 & \text{Math. Sci.}
\\
\hline
\text{Cambridge (univ. + coll.)} & 10.7 & 23.2 & \text{Part III of the Math. Tripos}
\\
\hline
\text{Dundee} & 3.8& 13 & \text{Math. Biol.}
\\
\hline
\text{Durham} & 5.7& 14 & \text{Math. Sci.}
\\
\hline
\text{Edinburgh} &9.3 {\color{blue}\mbox{ } (19)} {\color{red}\mbox{ } [2.04]} &17.4 {\color{blue}\mbox{ } (22)} {\color{red}\mbox{ } [1.26]}& \text{OR}
\\
\hline
\text{Exeter} & 7.5 {\color{blue}\mbox{ } (11.9)} {\color{red}\mbox{ } [1.59]}&17.5{\color{blue}\mbox{ } (19.5)} {\color{red}\mbox{ } [1.11]}& \text{Adv. Math.}\\
\hline
\text{Glasgow} & 5.4 {\color{blue}\mbox{ } (8.3)} {\color{red}\mbox{ } [1.54]}& 17.3 {\color{blue}\mbox{ } (17.3)} {\color{red}\mbox{ } [1]}& \text{Math./Appl. Math.}\\
\hline
\text{Heriot--Watt} & 5.9 {\color{blue}\mbox{ } (19)} {\color{red}\mbox{ } [3.22]}& 13 {\color{blue}\mbox{ } (22)} {\color{red}\mbox{ } [1.69]}& \text{Math.}
\\
\hline
\text{Heriot--Watt} & 5.9 {\color{blue}\mbox{ } (19)}{\color{red}\mbox{ } [3.22]}& 13 {\color{blue}\mbox{ } (22)} {\color{red}\mbox{ } [1.69]}& \text{Appl. Math. Sci.}
\\
\hline
\text{Heriot--Watt} & 5.9 {\color{blue}\mbox{ } (19)}{\color{red}\mbox{ } [3.22]}& 13 {\color{blue}\mbox{ } (22)} {\color{red}\mbox{ } [1.69]}& \text{Comp. Math.}
\\
\hline
\text{Imperial} & 8.3 {\color{blue}\mbox{ } (27)} {\color{red}\mbox{ } [3.25]}& 23 {\color{blue}\mbox{ } (27)} {\color{red}\mbox{ } [1.17]}& \text{Pure Math./Appl. Math.}
\\
\hline
\text{King's} & 8.3 {\color{blue}\mbox{ } (22.7)} {\color{red}\mbox{ } [2.73]}& 16.5 {\color{blue}\mbox{ } (22.7)} {\color{red}\mbox{ } [1.37]}& \text{Math.}
\\
\hline
\text{Leeds} & 5.1 {\color{blue}\mbox{ } (9)} {\color{red}\mbox{ } [1.76]}& 13.3 {\color{blue}\mbox{ } (19)} {\color{red}\mbox{ } [1.43]}& \text{Math.}
\\
\hline
\text{Leicester} &8.5 {\color{blue}\mbox{ } (8.5)} {\color{red}\mbox{ } [1]}& 16 {\color{blue}\mbox{ } (16)} {\color{red}\mbox{ } [1]}& \text{Math. Model. Biol.}
\\
\hline
\text{Leicester} &8.5 {\color{blue}\mbox{ } (8.5)} {\color{red}\mbox{ } [1]}& 16 {\color{blue}\mbox{ } (16)} {\color{red}\mbox{ } [1]}& \text{Appl. Comp. Num. Model.}
\\
\hline
\text{Liverpool} & 5.3 {\color{blue}\mbox{ } (15.4)} {\color{red}\mbox{ } [2.90]}& 12.2 {\color{blue}\mbox{ } (15.4)} {\color{red}\mbox{ } [1.26]}& \text{Math. Sci.}
\\
\hline
\text{Loughborough} & 6.3 {\color{blue}\mbox{ } (6.3)} {\color{red}\mbox{ } [1]}& 13.8 {\color{blue}\mbox{ } (13.8)} {\color{red}\mbox{ } [1]}& \text{Indust. Math. Model.}
\\
\hline
\text{LSE} & 11.6 {\color{blue}\mbox{ } (23)} {\color{red}\mbox{ } [1.98]}& 17.9 {\color{blue}\mbox{ } (23)} {\color{red}\mbox{ } [1.28]} & \text{Appl. Math.}
\\
\hline
\text{Manchester} & 8.4 {\color{blue}\mbox{ } (10.7, 21.4)} {\color{red}\mbox{ } [1.27, 2.55]}& 14 {\color{blue}\mbox{ } (18.4, 21.4)} {\color{red}\mbox{ } [1.31, 1.53]}& \text{Appl. Math.}
\\
\hline
\text{Manchester} & 8.4 {\color{blue}\mbox{ } (10.7, 21.4)} {\color{red}\mbox{ } [1.27, 2.55]}& 14 {\color{blue}\mbox{ } (18.4, 21.4)} {\color{red}\mbox{ } [1.31, 1.53]}& \text{Pure Math. Math. Logic}
\\
\hline
\text{Nottingham} & 6 {\color{blue}\mbox{ } (6)} {\color{red}\mbox{ } [1]}& 13.7 {\color{blue}\mbox{ } (13.7)} {\color{red}\mbox{ } [1]}& \text{Pure Math.}
\\
\hline
\text{Oxford} & 5.6 {\color{blue}\mbox{ } (24.9, 27.5)} {\color{red}\mbox{ } [4.45, 4.91]}& 15.3 {\color{blue}\mbox{ } (24.9, 31.1)} {\color{red}\mbox{ } [1.62, 2.03]}& \text{Math. Model. Sci. Comp.}
\\
\hline
\text{Queen Mary} &6.3 {\color{blue}\mbox{ } (15.5)} {\color{red}\mbox{ } [2.46]}& 13.5 {\color{blue}\mbox{ } (18)} {\color{red}\mbox{ } [1.33]}& \text{Math.}
\\
\hline
\text{Sheffield} & 6 {\color{blue}\mbox{ } (6)} {\color{red}\mbox{ } [1]}&15.4 {\color{blue}\mbox{ } (15.4)} {\color{red}\mbox{ } [1]}& \text{Math.}
\\
\hline
\text{Southampton} & 7.3 & 15 & \text{OR}
\\
\hline
\text{Surrey} & 6.3 & 16.3 & \text{Math.}
\\
\hline
\text{Sussex} & 5.5 & 13 & \text{Math.}
\\
\hline
\text{UCL} &8.5 {\color{blue}\mbox{ } (19.6)} {\color{red}\mbox{ } [2.30]}&16.8 {\color{blue}\mbox{ } (21.7)} {\color{red}\mbox{ } [1.29]}& \text{Math. Model.}
\\
\hline
\text{Warwick} & 7{\color{blue}\mbox{ } (31)} {\color{red}\mbox{ } [4.42]}& 15.9 {\color{blue}\mbox{ } (31)} {\color{red}\mbox{ } [1.94]}& \text{Math.}\\
\hline
\text{York} & 6.2 {\color{blue}\mbox{ } (17.4, 16.5)} {\color{red}\mbox{ } [2.80, 2.66]}& 14.3 {\color{blue}\mbox{ } (22.2, 21)} {\color{red}\mbox{ } [1.55, 1.47]}& \text{Adv. Math. Biol.}
\\
\hline
\end{array}
\end{footnotesize}\]
\newpage
\noindent {\footnotesize Based on the figures in red above, we group universities according to the differential between the presented M.Sc. in Financial Mathematics and other M.Sc. programmes in a strongly mathematical subject area on the same campus.}
\[
\begin{footnotesize}
\begin{array}{| l | c | c| c| c| c| }
\hline
\text{Differential} & 0-10\% & 10-20\% & 20-50 \% & 50-100\% & > 100\% \\
\hline
\hline
\text{Home-EU}
&
\begin{array}{l}
\text{Leicester}\\
\text{Loughborough} \\
\text{Nottingham}\\
\text{Sheffield}
\end{array}
&
&
\begin{array}{l}
\text{Glasgow} \\
\text{Manchester}
\end{array}
&
\begin{array}{l}
\text{Exeter} \\
\text{Leeds} \\
\text{LSE}
\end{array}
&
\begin{array}{l}
\text{ }\\
\text{Birkbeck}\\
\text{Birmingham}\\
\text{Heriot--Watt}\\
\text{Imperial}\\
\text{King's}\\
\text{Liverpool}\\
\text{Manchester}\\
\text{Queen Mary}\\
\text{UCL}\\
\text{Warwick}\\
\text{York}\\
\text{ }
\end{array} \\
\hline
\text{Overseas} &
\begin{array}{l}
\text{Glasgow}\\
\text{Leicester}\\
\text{Loughborough}\\
\text{Nottingham}\\
\text{Sheffield}
\end{array}
&
\begin{array}{l}
\text{Birmingham}\\
\text{Exeter}\\
\text{Imperial}
\end{array}
&
\begin{array}{l}
\text{ }\\
\text{Edinburgh}\\
\text{King's}\\
\text{Leeds}\\
\text{Liverpool}\\
\text{LSE}\\
\text{Manchester}\\
\text{Queen Mary}\\
\text{UCL}\\
\text{York}\\
\text{ }
\end{array}
&
\begin{array}{l}
\text{Heriot-Watt}\\
\text{Oxford}\\
\text{Warwick}\\
\text{York}
\end{array}
&
\begin{array}{l}
\text{Birkbeck}\\
\text{Oxford}
\end{array}
\\
\hline
\end{array}
\end{footnotesize}
\]
\end{document}
|
\section{Introduction} \label{sec:intro}
Multiphase flows represent the cornerstone of many fields
of science and technology ranging from micro-scale devices to the large scale
cyclonic separators of industrial plants. In the context of micro/nano
technologies, the transport of small particles or bubbles by a carrier fluid
is fundamental in designing micro-devices where particles must be separated,
mixed or advected towards the sensible regions of the apparatus for detection
purposes, see e.g \cite{stone2004}. Concerning larger scale devices, the turbulent
transport of a disperse phase is relevant for the dynamics of small fuel
droplets in combustion chambers, \cite{post}, or in the spatial evolution of
sprays employed for surface coating, \cite{pawlowski2008}.
Important aspects of multiphase flows are related to the intrinsic coupling
between the motion of the disperse phase and the carrier fluid which involves
mass, momentum and energy exchange between the two phases. Hydrodynamic
interactions among the particles or inter-particles collisions might also
occur. The regime where all these interactions take place is known as four-way
coupling regime, see e.g. \cite{balachandar_rev,elgo_map}.
The straightforward method to capture
such complex physics is represented by numerical simulations where the
fluid flow around each particle is fully resolved. This means that the actual
particle boundary has to be resolved on the computational grid and the
coupling with the fluid occurs via the non slip boundary conditions imposed on
the particle surface. The hydrodynamic force on each particle can be directly
computed by integrating the pressure and shear stress distribution on the
boundary. Even though this approach captures entirely the physics, it is
computationally demanding and limited to the simulation of a relatively
small number of ``large'' particles. The adjective large means that the particle
typical size, the diameter $d_p$, is larger than the smallest physically
active hydrodynamical scale $\eta$. For instance $\eta$ could be either
the Kolmogorov dissipative scale in a turbulent flow or the smallest spatial
scale in a micro-fluidic apparatus. In the context of the so called
resolved particles simulations many approaches are available ranging from
finite volume techniques, \cite{burton}, immersed boundary methods,
\cite{lucci_jfm},
or approaches based on the Lattice-Boltzman equations, \cite{ten_cate,gao2011}.
Alternative approaches are however available. For instance, the PHYSALIS technique,
see e.g. \cite{zhang} and references therein, has been recently adopted to address
the interaction of solid particle and a turbulent flow, \cite{naso}.
\cite{homann_bec} adapted the pseudo-penalization spectral method proposed by
\cite{pasquetti} to account for the coupled dynamics of neutrally buoyant
particles in a turbulent flow. The Force Coupling Method (FCM) proposed by Maxey
and coworkers, see e.g. among many others the papers by \cite{maxeypat,lomholt},
is certainly worth mentioning in detail. In the FCM the effect that each particle
exerts on the fluid is approximated by a multipole expansion of a regularized
steady Stokes solution where the concentrated delta-function forces are mollified
to a Gaussian.
The basic method has been continuously improved by
including several physical effects such as lubrication forces for closely packed
particles \cite{dance} or the effects of elongated particles \cite{liu}.
Recently a numerical simulation of homogeneous isotropic turbulence laden with
thousands of relatively large particles $(d_p/\eta=6\div12)$ has been reported
by \cite{yeo}.
The opposite limit of particles much smaller than the smallest
hydrodynamical scale is also relevant in many applications. For instance
the mixing and combustion of a turbulent spray after that the primary
atomization phase has occurred, takes place in presence of significant
momentum coupling among the carrier fluid and the fuel droplets, see
e.g. the recent review by \cite{jenny}. In fact,
in dilute suspensions the volume fraction of the particles is small enough
to neglect hydrodynamic interactions and collisions among particles.
However, for large values of the particle-to-fluid density ratio,
significant mass loads (ratio between the mass of the disperse phase and
the fluid) may occur. In such regime, the so called the two-way coupling regime,
the momentum exchange between the two phases is significant and must be
accounted for. The Particle In Cell (PIC) method, \cite{crowe}, is still a
valuable tool to model the momentum coupling. Such approach needs substantial
care, however, due to technical issues associated with the injection of the point-wise
forcing of the particles on the computational grid where the continuous fluid
phase is resolved. Indeed, the force that the particles exert on the fluid is
regularized by averaging on the volume of the computational cell.
Hence, the coupling term results strongly grid dependent unless the number
of particles per cell $N_p/N_c$ exceeds a certain threshold, see e.g. the
numerical results in \cite{jfm_2way}, the discussion by \cite{balachandar2009}
and the comments in \cite{jenny}.
Alternative to the PIC approach, other methods which are able to work
irrespective of particle number density do indeed exist. For instance \cite{pan}
modeled the disturbance flow produced by each point particle in terms of
the steady Stokeslet. Though interesting, this approach has several potential
shortcomings. The disturbance
flow decays in space away from the particle as slow as the inverse distance and
the perturbation induced by a single particle affects the whole domain.
In these conditions, any truncation is undoubtably bound to deeply alter the
dynamics. Additionally, the disturbance flow presents the singularity associated
with the steady Stokeslet. Moreover, the steady Stokes solution used to model
the fluid-particle interaction is not uniformly valid and fails away from
the particle. The Oseen correction consistently accounts for the unavoidable
far field convective effects, see classical textbooks like \cite{lamb,batchelor}.
Numerical approaches based on this improved modeling can be found, e.g.,
in \cite{subramanian2008,pignatel2011}.
In the present paper we propose a new approach able to provide a physically
consistent and numerically convergent solution for the flow disturbance
produced by a huge number of small, massive particles coupled to a generic,
possibly turbulent, carrier flow. Hereafter this new formulation will be
referred to as the Exact Regularized Point Particle (ERPP) method.
As it will be shown in detail, this approach presents several advantages.
The most significant one is related to the physical accuracy of the momentum
coupling modeling. In a nutshell, in the relative motion with respect to
the fluid, the particle generates a vortical field.
Even though the relative Reynolds number is small, the local flow is dominated by
unsteady viscous effects as discussed by \cite{bruno2008}. Vorticity production is
a localized process that takes a finite elapsed time $\epsilon_R$ since generation
to reach the relevant hydrodynamic scales of the flow. It is indeed this transient
process of localized generation and finite time diffusion that introduces the
actual momentum coupling with the carrier flow. Indeed, the model here envisaged
reproduces this physical process by addressing the velocity field,
rather than vorticity. The finite time delay $\epsilon_R$ automatically provides
the regularization of the disturbance field. Instead of being a purely
mathematical or numerical ingredient, the regularization featured by ERPP is
intrinsically associated with the actual physical process of vorticity generation
and viscous diffusion. A distinguishing aspect of ERPP is that all the vorticity
generated by the particle is properly transferred to the fluid phase, entailing
momentum conservation. A crucial concern is the small scale component of the
disturbance field associated with the instantaneously generated vorticity not yet
diffused up-to the hydrodynamic scales. This localized, inner scale part of the
disturbance exhibits a $1/r$ local singularity and vanishes altogether at
the relevant hydrodynamic scales. Although, in principle, this field
should contribute locally to the convective term of the Navier-Stokes equations,
its effect is proportional to the (small) particle Reynolds number based on
the slip velocity. Consistently, it negligibly contributes to the dynamics of
the relevant hydrodynamic scales.
Concerning the hydrodynamic force acting on the particles in the
two way-coupling regime, the expression provided by \cite{maxril} is easily
adapted to the present context. A crucial issue is the fluid-to-particle slip
velocity appearing in the expression of the Stokes drag that should be
understood as the undisturbed fluid velocity (i.e. the relative fluid-particle
velocity in absence of the particle). In the ERPP the self-induced velocity
disturbance can be evaluated in a closed form, allowing to explicitly remove its
contribution. It follows a consistent evaluation of slip velocity and
hydrodynamic force.
Despite the underlying theoretical aspects may look complicated at first sight,
the practical implementation of the ensuing algorithm is remarkably simple
and efficient. In principle, the coupling algorithm can indeed be embedded in
any available discretization scheme as implemented in one's favorite
Navier-Stokes solver. This flexibility allows to easily handle hundred thousands
particles at affordable computational cost.
The paper is organized as follows. The next section \S \ref{sec:methodology}
forms the main theoretical body of the paper. Along with its subsections, it
introduces the physical model and discusses the inter-phase momentum coupling.
In section \S \ref{sec:validation} the proposed approach is validated
against available analytical results. Section \S \ref{sec:turb_flow} reports
preliminary results concerning a turbulent particle-laden shear flow.
Finally, the last section \S \ref{sec:final} summarizes the main findings.
To smooth out the reading, several appendices are devoted to lengthy technical
issues whose description inside the main text would have hampered a clear
exposition of the main material.
\section{Methodology} \label{sec:methodology}
In this section we present the physical model used to achieve the momentum coupling
between the carrier fluid and the disperse phase in view of describing the
algorithm for the simulation of particle laden flows in the two way coupling regime.
In doing so, we assume to know the state of the system at time $t$ and propagate
the solution for one time step $Dt$. Clearly, reiteration of the procedure allows
to proceed in time, as in standard time integration algorithms. During the
generic time step of length $Dt=t_{n+1}-t_n$ the state of the system
will propagate from $t_n$ to $t_{n+1}$. For the sake of simplicity hereafter
we shall often address the generic step as the step $n = 0$. In this case the
running time will be $0 \le t \le Dt$ in all the differential equations to
be addressed. In the discussion, a quantity $\epsilon_R$ with dimension of time
and the role of a regularization parameter will play a central role.
In this case, having dubbed the current time as instant $t_0 = 0$, it could well
happen that certain time-delayed variables (i.e. $t-\epsilon_R$) could be
negative: we like to assure the reader before hand that this will be no harm.
Integral representation of the solution may represent an exception to this rule.
Indeed, such integral representation will be used to derive a systematic
regularization procedure from which we obtain by differentiation the regularized
pde's to be solved in the algorithm. In such cases the time extrema
will typically range in the interval $\left[0,t\right]$, with $t$ arbitrary, and
typically larger than $\epsilon_R$.
In this framework, the short time evolution ($Dt \ll 1$) of the overall flow
(fluid and particles) is conceptually split into a (modified) Navier-Stokes
evolution of the carrier fluid and a superimposed disturbance flow produced by
the relative motion of the particles, here assumed spherical, with respect to
the fluid. Relying on the small Reynolds number of the particle-fluid relative
motion, the disturbance flow is described by the linear unsteady Stokes equations.
In fact, we will rearrange the equation in such a way that the exact solution
of the particle disturbance field is consistently embedded into the carrier
phase Navier-Stokes solver allowing to reconstruct the actual fluid-particle
coupled solution in the limit of vanishing time step and grid spacing for
small particle Reynolds number.
The detailed derivation of the coupling model needs a gradual illustration better
achieved starting from a schematic description divided in five conceptual steps:
\begin{itemize}
\item[$i)$] Carrier flow-disperse phase interaction and disturbance flow
equation (subsection \S \ref{sec:interaction})
\item[$ii)$] Solution of the disturbance flow equation
(subsection \S \ref{sec:disturbance})
\item[$iii)$]Regularization (subsection \S \ref{sec:sing_reg})
\item[$iv)$] Embedding the disturbance flow into the Navier-Stokes equations
(subsection \S \ref{sec:coupling_phases})
\item[$v)$] Evaluation of the hydrodynamic force on the particles in the
two-way coupling regime and removal of the self-induced velocity disturbance
(subsection \S \ref{sec:prtcl_motion_force}).
\end{itemize}
\subsection{Interaction between the two phases} \label{sec:interaction}
In presence of a disperse phase, the carrier fluid fills the domain
${\cal D}\backslash \Omega$ where $\cal D$ is the flow domain and
$\Omega(t) = \cup_p \Omega_p(t)$ denotes the region occupied by the collection of
$N_p$ rigid particles, with $\Omega_p(t)$ the time dependent domain occupied by
the $p$th particle, see the sketch in figure \ref{fig:domain_sketch}.
The set theoretic notation $\cup_p$ denotes the union of
sets indexed by $p$ and $A\backslash B$ denotes the complement in $A$ of set $B$.
The motion of the carrier fluid is assumed to be described by the standard
incompressible Navier-Stokes equations endowed with the no-slip condition at
the particle boundaries
\begin{equation}
\label{eqn:ns_resolved}
\begin{array}{l}
\left.
\begin{array}{l}
\displaystyle \nabla \cdot \vu = 0 \\ \\
\displaystyle \frac{\partial \vu}{\partial t} + \vu \cdot \nabla \vu =
-\frac {1}{\rho_f} \nabla {\rm p} + \nu \nabla^2 \vu
\end{array} \right\} \qquad \vx \in {\cal D}\backslash \Omega(t)
\\
~\\
\displaystyle \vu\lvert_{\partial \Omega_p(t)} = \vV_p(\vx)\lvert_{\partial \Omega_p(t)} \qquad \qquad p =1,\ldots,N_p \\
\displaystyle \vu\lvert_{\partial {\cal D}} = \vu_{wall} \\
~\\
\displaystyle \vu(\vx,0)=\vu_0(\vx) \qquad \qquad \vx \in {\cal D}\backslash \Omega(0)\ .
\end{array}
\end{equation}
In equations (\ref{eqn:ns_resolved}), $\vu_0(\vx)$ is the velocity field at time
$t=0$, $\rho_f$ denotes the fluid density, $\nu$ is the kinematic viscosity,
$\partial \Omega_p$ is the boundary of the $p$th particle and $\partial {\cal D}$
is the boundary of the overall flow domain, see figure \ref{fig:domain_sketch}.
In this microscopic description, the
particles affect the carrier fluid through the no-slip condition at the moving
particle surface $\partial \Omega_p(t)$ where the fluid matches the local rigid
body velocity of the particle
$\vV_p(\vx) = \vv_p + \vomega \times \left(\vx-\vx_p \right)$, with $\vv_p$
the velocity of the particle geometric center $\vx_p(t)$
and $\vomega(t)$ the angular velocity. The equations of rigid body dynamics need
be coupled to the equation for the fluid velocity field to determine the
particle motions, where the fluid tension acting at the particle boundary provide
the relevant forces and moments.
In principle the system (\ref{eqn:ns_resolved}) can be numerically integrated at
the price of resolving all the particle boundaries on the computational grid.
When the suspension is formed by a huge number of small particles their direct
solution is unaffordable. In any case, equations (\ref{eqn:ns_resolved}) still
provide the basic description of the flow in terms of the interaction between the
two phases. Purpose of the present subsection is to manipulate and approximate the
basic equations to derive a viable model for the suspension.
\begin{figure}
\centerline{
\includegraphics[scale=0.35,angle=-90]{./sketch_laden_fluid.pdf}}
\caption{Sketch of the flow domain. The fluid fills the domain
${\cal D}\backslash \Omega$ with $\Omega(t) = \cup_p \Omega_p(t)$ the region
occupied by the $N_p$ rigid particles and $\Omega_p(t)$ the time
dependent domain of the $p$th particle. $\partial {\cal D}$ denotes
the boundary of $\cal D$. The fluid velocity at the generic point $\vx \in {\cal D}\backslash \Omega$ is decomposed as
$\vu=\vw+\vv$, to be understood as the definition of $\vw$ given the fluid velocity $\vu$ and the solution $\vv$ of the linear, unsteady Stokes problem
(\ref{eqn:unsteady_stokes}).
\label{fig:domain_sketch}}
\end{figure}
As a starting point, for small time intervals $0 \le t \le Dt \ll1$, the carrier flow velocity is decomposed
into two parts, $\vu(\vx,t)=\vw+\vv$. The field $\vw(\vx,t)$ is assumed to satisfy the equations
\begin{equation}
\label{eqn:ns_background}
\begin{array}{l}
\begin{array}{l}
\displaystyle \nabla \cdot \vw = 0 \\ \\
\displaystyle \frac{\partial \vw}{\partial t} + \vF =
-\frac {1}{\rho_f} \nabla \pi + \nu \nabla^2 \vw
\end{array}
\\
\\
\displaystyle \vw\lvert_{\partial {\cal D}} = \vu_{wall} -
\vv_{\partial {\cal D}} \\
\displaystyle \vw(\vx,0)={\bar \vu}_0(\vx)\, ,
\end{array}
\end{equation}
where $\vx \in {\cal D}$ and
\begin{equation}
\label{eqn:ns_F}
\vF = \left\{
\begin{array}{ll}
\vu \cdot \nabla \vu & \qquad \mbox{for} \,\, \vx \in {\cal D}\backslash\Omega(t)\\ \\
\vV_p \cdot \nabla \vV_p & \qquad \mbox{for} \,\, \vx \in \Omega(t)
\end{array}
\right.
\end{equation}
is a field reproducing the complete convective term of the Navier-Stokes equation
in the carrier fluid domain ${\cal D}\backslash \Omega$ which is prolonged inside
$\Omega$ using the solid particle velocity field. Other choices are possible, but
the actual shape of the field inside the particle domains is irrelevant to our
present purposes: under this respect, the solid body motion provides an elegant
example given the continuity of the field $\vF$ at the particle boundaries.
In problem (\ref{eqn:ns_background}), a part from the prolongation of the
field $\vF$, the particles disappeared altogether from the domain and the
convective term, retaining its complete nonlinear nature in the fluid domain,
is treated as a prescribed forcing term. The initial field $\bar \vu_0$ is
prolonged inside the particle domains by the same rule, i.e. as the solid body
motion of relevant particle.
The field $\vv(\vx,t)$ exactly satisfies the \emph{linear} unsteady Stokes
problem (the complete non-linear term has been retained in the equation for $\vw$)
\begin{equation}
\label{eqn:unsteady_stokes}
\begin{array}{l}
\left.
\begin{array}{l}
\displaystyle \nabla \cdot \vv = 0 \\ \\
\displaystyle \frac{\partial \vv}{\partial t}
=-\frac{1}{\rho_f} \nabla {\rm q} +\nu \nabla^{2}{\vv}
\end{array}
\right\} \qquad \vx \in {\cal D}\backslash \Omega(t)
\\
~\\
\displaystyle \vv\lvert_{\partial \Omega_p(t)}=\vV_p(\vx)\
\lvert_{\partial \Omega_p(t)} -
\vw\lvert_{\partial \Omega_p(t)} \qquad p = 1, \ldots N_p \\
~\\
\displaystyle \vv(\vx,0)=0 \qquad \vx \in {\cal D}\backslash \Omega(0)\, ,
\end{array}
\end{equation}
where boundary conditions are applied at the particle surfaces. It should be
observed that no boundary condition are applied to the field $\vv$ at the
flow domain boundary $\partial {\cal D}$. In other words, the field $\vv$ can
be regarded as a free space solution in the whole $\Rset^3$ restricted the actual
flow domain $\cal D$. Indeed the value of $\vv$ at the domain boundary is used
to correct the boundary condition for $\vw$. It is worth calling the reader's
attention to the initial conditions for the two complementary problems: the
initial velocity field is assigned as initial condition for $\vw$, leaving
homogenous initial data for $\vv$. As shown in a later section, the homogeneous
initial conditions for the perturbation field $\vv$ will turn out to be a crucial
feature of the decomposition.
The solution of equations (\ref{eqn:unsteady_stokes}) can be expressed
in terms of the boundary integral representation of the unsteady Stokes equations
that involves the unsteady Stokeslet $G_{ij}(\vx,\vxi,t,\tau)$, a second order
Cartesian tensor, and the associated stresses in the form of the third order
tensor
${\cal T}_{ijk} (\vx,\vxi,t,\tau)$, see appendix \ref{app:unsteady_stokeslet}
and classical textbooks, \cite{zapryanov,kim_book}.
The unsteady Stokeslet $G_{ij}(\vx,\vxi,t,\tau)$ is readily interpreted as
the fluid velocity ($i$th direction) at position $\vx$ and time $t$ due to the
singular forcing $\delta(\vx-\vxi)\delta(t-\tau)$ ($j$th direction) applied at
$\vxi$ at time $\tau$. Exploiting the vanishing initial condition,
the solution of equations (\ref{eqn:unsteady_stokes}) is recast in the boundary
integral representation
\begin{equation}
\label{eqn:unsteady_stokes_solution}
v_i(\vx,t)=\int_0^t d\tau \int_{\partial \Omega} t_j(\vxi,\tau)
G_{ij}(\vx,\vxi,t,\tau)
-v_j(\vxi,\tau) {\cal T}_{ijk} (\vx,\vxi,t,\tau) n_k(\vxi) \, dS_{\vxi}.
\end{equation}
Equation (\ref{eqn:unsteady_stokes_solution}) expresses $\vv(\vx,t)$ in terms of
a boundary integral on $\partial \Omega = \cup_p \partial \Omega_p$ involving
the (physical) tension $t_j(\vxi,\tau)$ and the boundary condition on the
perturbation velocity $v_j(\vxi,\tau)$ at each particle boundary.
In principle, the tension $t_j(\vxi,\tau)$
can be determined by solving the boundary integral equation (indeed a system of
coupled boundary integral equations, one for each particle) associated with
representation (\ref{eqn:unsteady_stokes_solution}). Once the tension is known at
each particle boundary, representation (\ref{eqn:unsteady_stokes_solution})
provides the perturbation field everywhere in the flow domain.
Moreover the boundary integral of the tension $t_j$ would provide the forces
acting on the particles.
Since the present aim is capturing the effects of many small particles of
diameter $d_p$, the interest is focused on the far field particle
disturbance that can be approximated by a multipole expansion of equation
(\ref{eqn:unsteady_stokes_solution}). Substituting in equation
(\ref{eqn:unsteady_stokes_solution}) the first order truncation of the Taylor
series of $G_{ij}(\vx,\vxi,t,\tau)$ and ${\cal T}_{ijk} (\vx,\vxi,t,\tau)$,
centered at the particle position $\vx_p$, leads to the far field expression for
large $r_p/d_p$, where $r_p = |\vx -\vx_p |$,
\begin{equation}
\label{eqn:unsteady_stokes_far_field}
v_i(\vx,t)=-\sum_p \int_0^t D^p_j(\tau) G_{ij}(\vx,\vx_p,t,\tau) \, d\tau \, ,
\end{equation}
showing that the far field disturbance depends only on the hydrodynamic force
$\vD_p(\tau)$, with Cartesian components $D^p_j$, which acts on the generic
particle. Given the physical interpretation of the unsteady Stokeslet $G_{ij}$,
the partial differential equation whose solution is given by
(\ref{eqn:unsteady_stokes_far_field}) follows as
\begin{equation}
\label{eqn:unsteady_stokes_singular}
\frac {\partial \vv}{\partial t} - \nu \nabla^2 \vv +
\frac{1}{\rho_f} \nabla {\rm q} = - \frac{1}{\rho_f} \sum_p \vD_p(t)
\, \delta\left[ \vx - \vx_p(t) \right] \, ; \qquad \vv(\vx,0)=0 \, ,
\end{equation}
as is directly verified by combining the time derivative of equation
(\ref{eqn:unsteady_stokes_far_field}) with its Laplacian.
In equation (\ref{eqn:unsteady_stokes_singular}) the boundary condition at the
particle surfaces disappear altogether and the fluid-particle coupling
occurs via the (singular) forcing term in the unsteady Stokes problem.
Given the linearity, hereafter we shall explicitly consider the single
contribution of the generic particle $p$, keeping in mind that a final summation
all over the particles is required.
It is also clear that as the particle diameter gets smaller and smaller, the
term $\vF$ in equations (\ref{eqn:ns_background}) uniformly fills the
entire domain $\cal D$ and reduces everywhere to the standard convective term
of the Navier-Stokes equation $\vu \cdot \nabla \vu$, where $\vu = \vw + \vv$.
\subsection{Disturbance flow due to a small particle} \label{sec:disturbance}
The vorticity equation associated with (\ref{eqn:unsteady_stokes_singular})
is
\begin{equation}
\label{eqn:vort_stokes}
\frac {\partial \vzeta}{\partial t} - \nu \nabla^2 \vzeta
= \frac{1}{\rho_f} \vD_p(t) \times \nabla
\delta\left[\vx-\vx_p(t) \right]; \, \quad \vzeta(\vx,0) = 0 \, ,
\end{equation}
where $\vzeta = \nabla \times \vv$. The solution can be expressed as
a convolution with the fundamental solution of the diffusion equation
$g(\vx-\vxi,t-\tau)$, given by (see appendix \ref{app:diffusion_eq})
\begin{equation}
\label{eqn:fundamental_fourier}
g(\vx-\vxi,t-\tau) = \frac{1}{\left[ 4 \pi \, \nu (t-\tau)\right]^{3/2}}
\exp\left[-\frac{\lVert \vx -\vxi \rVert ^2}{4 \nu (t-\tau)} \right] \, ,
\end{equation}
that is a Gaussian with time dependent variance
$\sigma(t-\tau) = \sqrt{2 \nu (t-\tau)}$. Observe that $g$ is the fundamental
solution of the diffusion equation in free-space, since $\vv$ is itself a free-space
field, as noted when discussing eq.~(\ref{eqn:unsteady_stokes}).
By rearranging the forcing on the right hand side of equation
(\ref{eqn:vort_stokes}) as a time-convolution,
\begin{equation}
\label{eqn:forcing_conv}
\vD_p(t) \times \nabla \delta\left[\vx-\vx_p(t) \right] =
\int_0^{t^+}
\vD_p(\tau) \times \nabla \delta\left[\vx-\vx_p(\tau) \right]
\delta(t-\tau) d\tau \,
\footnote{with $\displaystyle \int_0^{t^+} f(\tau) \, d \tau$ we intend
$\displaystyle \lim_{\epsilon \to 0}\int_0^{t+\epsilon} f(\tau) \, d \tau$},
\end{equation}
the solution of equation (\ref{eqn:vort_stokes}) follows at once as
\begin{equation}
\label{eqn:sol_vort_forced}
\vzeta(\vx,t) = \frac{1}{\rho_f} \int_0^{t^+}
\vD_p(\tau) \times \nabla g\left[ \vx-\vx_p(\tau),t-\tau\right] d\tau \ .
\end{equation}
The original fluid velocity $\vv(\vx,t)$ can be reconstructed from the vorticity
using the non-canonical decomposition
\begin{eqnarray}
\label{eqn:disturbance_decomp}
\vv(\vx,t) = \vv_{\vzeta}(\vx,t) + \nabla \phi(\vx,t) \, ,
\end{eqnarray}
where
\begin{eqnarray}
\label{eqn:pseudo_velo}
\vv_{\vzeta}(\vx,t) = - \frac{1}{\rho_f} \int_0^{t^+} \vD_p(\tau)
g\left[ \vx-\vx_p(\tau),t-\tau\right] d\tau
\end{eqnarray}
is a pseudo-velocity, such that its curl equals the vorticity,
$\nabla \times \vv_{\vzeta} = \vzeta$, and the gradient term is
added to make the field solenoidal, as appropriate for incompressible flows,
\begin{eqnarray}
\label{eqn:velo_proj}
\nabla^2 \phi(\vx,t) = - \frac{1}{\rho_f}
\int_0^{t^+} \vD_p(\tau) \cdot \nabla g\left[ \vx-\vx_p(\tau),t-\tau\right]
d\tau \ .
\end{eqnarray}
The pseudo-velocity $\vv_{\vzeta}$ obeys the equation
\begin{eqnarray}
\label{eqn:v_zeta_forced}
\frac{\partial \vv_{\vzeta}}{\partial t} - \nu \nabla^2 \vv_{\vzeta} =
- \frac{1}{\rho_f} \vD_p(t) \delta\left[\vx - \vx_p(t) \right]\, , \qquad
\vv_{\vzeta}(\vx,0) = 0 \ .
\end{eqnarray}
\subsection{Regularization of the disturbance field due to a small particle}
\label{sec:sing_reg}
Both the velocity $\vv$ and the vorticity $\vzeta$ are apparently singular,
with singularity arising from the contribution to the integral near the upper
integration limit, $\tau \simeq t$, where $g(\vx-\vxi,t-\tau)$ tends to
behave as ``badly'' as the Dirac delta function. On the contrary away from
the upper integration limit the integrand is nicely behaved since it involves a
Gaussian or its gradient.
In this paragraph we define a regularization procedure based on a temporal cut-off
$\epsilon_R$ such that the fields are additively split into a regular and a
singular component. For instance the decomposition of the vorticity reads
\begin{eqnarray}
\label{eqn:vort_forced_decomp}
\vzeta(\vx,t) = \vzeta_R(\vx,t;\epsilon_R) + \vzeta_S(\vx,t;\epsilon_R) \, ,
\end{eqnarray}
with smooth and singular part respectively given by
\begin{eqnarray}
\label{eqn:vort_regular}
\vzeta_R(\vx,t) = \frac{1}{\rho_f} \int_0^{t - \epsilon_R}
\vD_p(\tau) \times \nabla g\left[ \vx-\vx_p(\tau),t-\tau\right] d\tau \, ,
\end{eqnarray}
and by
\begin{eqnarray}
\label{eqn:vort_singular}
\vzeta_S(\vx,t) = \frac{1}{\rho_f} \int_{t - \epsilon_R}^{t^+}
\vD_p(\tau) \times \nabla g\left[ \vx-\vx_p(\tau),t-\tau\right] d\tau \ .
\end{eqnarray}
As implied by the fundamental solution of the diffusion equation,
the regular part of the vorticity field is everywhere smooth and
characterized by the smallest spatial scale
$\sigma_R=\sigma(\epsilon_R) = \sqrt{2 \nu \epsilon_R}$.
Thanks to the semigroup property of solutions of the diffusion equation,
the regular field $\vzeta_R(\vx,t)$ can be interpreted as the free
diffusion from time $t-\epsilon_R$ to time $t$ of the complete
field at time $t-\epsilon_R$, $\vzeta(\vx,t-\epsilon_R)$, namely
\begin{eqnarray}
\label{eqn:vort_regular_diffused}
\vzeta_R(\vx,t) = \int \vzeta(\vxi,t-\epsilon_R)
g\left( \vx-\vxi,\epsilon_R\right) d\vxi \, ,
\end{eqnarray}
where the spatial convolution integral propagates the field from $t-\epsilon_R$
to $t$. Although physically obvious, equation~(\ref{eqn:vort_regular_diffused})
can be directly proved using the result
\begin{eqnarray}
\label{eqn:group_green}
g(\vx,t) = \int g(\vxi,t-\epsilon_R) g\left( \vx-\vxi,\epsilon_R\right)
d\vxi \, ,
\end{eqnarray}
that is nothing more that a re-expression of the semigroup property for the
free-space diffusion equation applied to the fundamental solution $g$.
Actually, using the property (\ref{eqn:group_green}) and introducing
eq.~(\ref{eqn:sol_vort_forced}) at time $t-\epsilon_R$
into eq.~(\ref{eqn:vort_regular_diffused}), after integration by parts, one
readily gets
\begin{eqnarray}
\label{eqn:proof_vort_reg}
\vzeta_R(\vx,t) = \frac{1}{\rho_f}
\int \left\{
\int_0^{t-\epsilon_R}
\vD_p(\tau) \times \nabla_{\vxi} g\left[ \vxi-\vx_p(\tau),t-\epsilon_R-\tau\right]
d\tau \right\} g\left( \vx-\vxi,\epsilon_R\right) d\vxi = \nonumber \\
\frac{1}{\rho_f} \int_0^{t-\epsilon_R}
\vD_p(\tau) \times \int \nabla_{\vxi}
g\left[ \vxi-\vx_p(\tau),t-\epsilon_R-\tau\right]
g\left( \vx-\vxi,\epsilon_R\right) d\vxi \,
d\tau = \nonumber \\
\frac{1}{\rho_f} \int_0^{t-\epsilon_R}
\vD_p(\tau) \times \nabla \int
g\left[ \vxi-\vx_p(\tau),t-\epsilon_R-\tau\right]
g\left( \vx-\vxi,\epsilon_R\right) d\vxi \,
d\tau = \nonumber \\
\frac{1}{\rho_f} \int_0^{t-\epsilon_R}
\vD_p(\tau) \times \nabla g\left[ \vx-\vx_p(\tau),t-\tau\right] \,
d\tau \nonumber \, ,
\end{eqnarray}
which is indeed equation (\ref{eqn:vort_regular}).
The corresponding vorticity field $\vzeta_R$ at time $t$ obeys a forced diffusion
equation where the forcing is applied at the slightly earlier time $t-\epsilon_R$,
\begin{eqnarray}
\label{eqn:pde_vort_regular}
\frac{\partial \vzeta_R}{\partial t} - \nu \nabla^2 \vzeta_R = - \frac{1}{\rho_f}
\nabla \times \vD_p(t-\epsilon_R) g\left[\vx -\vx_p(t-\epsilon_R),\epsilon_R
\right] \, ; \quad
\vzeta_R(\vx,0) = 0 \, ,
\end{eqnarray}
see appendix \ref{app:pde_vort_regular} for the detailed calculation.
The velocity field $\vv_R$ associated with the regularized vorticity field
$\vzeta_R$ can be expressed though the general decomposition
(\ref{eqn:disturbance_decomp}),
\begin{equation}
\label{eqn:v_regular_decomp}
\vv_R(\vx,t) = \vv_{{\vzeta}_R} + \nabla \Phi_R \, ,
\end{equation}
where, by analogy with eq.~(\ref{eqn:v_zeta_forced}), the regularized
pseudo-velocity $\vv_{{\vzeta}_R}$ is
\begin{eqnarray}
\label{eqn:v_zeta_regular}
\frac{\partial \vv_{{\vzeta}_R}}{\partial t} - \nu \nabla^2 \vv_{{\vzeta}_R} =
- \frac{1}{\rho_f} \vD_p(t-\epsilon_R)
g\left[\vx - \vx_p(t-\epsilon_R), \epsilon_R \right]\, ; \quad
\vv_{\vzeta_R}(\vx,0) = 0 \, ,
\end{eqnarray}
and the potential correction follows from the equation
\begin{eqnarray}
\label{eqn:poisson_regular}
\nabla^2 \Phi_R = - \nabla \cdot \vv_{{\vzeta}_R} \ .
\end{eqnarray}
It is worth noticing that the complete regularized field obeys instead the
forced unsteady Stokes equation
\begin{eqnarray}
\label{eqn:velo_reg_eq}
\frac{\partial \vv_R}{\partial t} - \nu \nabla^2 \vv_R
+\frac{1}{\rho_f}\nabla {\rm q}_R = - \frac{1}{\rho_f}
\vD_p(t-\epsilon_R) \, g\left[ \vx-\vx_p(t-\epsilon_R),\epsilon_R \right]
\end{eqnarray}
for the solenoidal field $\vv_R$. The crucial point to observe here is that the
regularized component of the velocity disturbance $\vv_R(\vx,t)$
evolves according to a diffusion equation forced by the anticipating Stokes drag
(i.e. evaluated at $t-\epsilon_R$) times the regular spatial distribution
$g\left[ \vx-\vx_p(t-\epsilon_R),\epsilon_R \right]$. Equation~(\ref{eqn:v_zeta_regular})
can in principle be straightforwardly solved
on a discrete grid, once the spatial scale $\sigma_R$ of the forcing is properly
resolved by the grid. Once $\vv_{{\vzeta}_R}$ is known, the correction needed to
make the field solenoidal calls for the solution of the Poisson
equation~(\ref{eqn:poisson_regular}).
For the future application to the full solver for the carrier phase in presence
of the suspension, it is also worth mentioning that the field $\vv_{{\vzeta}_R}$ is
rapidly decaying in space as far as the observation time $t$ is small, since it
implies the short-time diffusion of a rapidly decaying forcing.
All the long-range effects are indeed confined to the potential
correction $\nabla \Phi_R$. As will be discussed in the forthcoming sections,
the field $\vv_R$ does not need to be separately evaluated, since it will be
embedded in the solution procedure for the single field $\vu$
which accounts for both the undisturbed carrier flows and the particle
perturbation.
At variance with $\vv_R$, the singular contribution $\vv_S$ cannot be represented
on a discrete grid. It can be decomposed as well into a vorticity related
component plus a potential correction, according to the general
representation~(\ref{eqn:disturbance_decomp}). The vortical component
$\vv_{{\vzeta}_S}$ is an extremely fast decaying function of distance from the
actual position of the particle, while its potential correction $\nabla \Phi_S$
is not. In order to address the error propagation of the algorithm that
will be illustrated in the next section, it is instrumental to explicitly provide
an estimate on the order of magnitude of the field $\nabla \Phi_S$.
The singular part of the pseudo-velocity is given by
\begin{eqnarray}
\label{eqn:pseudo_velo_S}
\vv_{\vzeta_S}(\vx,t) = \frac{1}{\rho_f} \int_{t-\epsilon_R}^{t^+} \vD_p(\tau)
g\left[ \vx-\vx_p(\tau),t-\tau\right] d\tau \, ,
\end{eqnarray}
see eq.~(\ref{eqn:pseudo_velo}) for comparison. The equation for the potential
correction is then
\begin{eqnarray}
\label{eqn:potential_S}
\nabla^2 \Phi_S = - \nabla \cdot \vv_{\vzeta_S} =
\frac{1}{\rho_f} \int_{t-\epsilon_R}^{t^+} \vD_p(\tau) \cdot
\nabla g\left[ \vx-\vx_p(\tau),t-\tau\right] d\tau \ .
\end{eqnarray}
It follows
\begin{eqnarray}
\label{eqn:potential_S_solution}
\Phi_S = - \frac{1}{\rho_f} \int_{t-\epsilon_R}^{t^+} d\tau \vD_p(\tau) \cdot
\nabla \int_{\Rset^3}
\frac{g\left[ \vy-\vx_p(\tau),t-\tau\right]}{4 \pi |\vx-\vy|} d^3\vy \, ,
\end{eqnarray}
where $-1/(4 \pi |\vx - \vy|)$ is the fundamental solution of the Laplace equation.
From the solution (\ref{eqn:potential_S_solution}) a rough estimate for the
correction field $\nabla \Phi_S$ is immediately obtained as
\begin{eqnarray}
\label{eqn:estimate}
| \nabla \Phi_S | \le \frac{1}{\rho_f} \sup_{t-\epsilon_R \le \tau \le t^+}
|\vD_p |
\Bigg| \nabla \otimes \nabla \int_{\Rset^3} \frac{\displaystyle
\int_{t-\epsilon_R}^{t^+} g\left[ \vy-\vx_p(\tau),t-\tau\right] d \tau}
{4 \pi |\vx-\vy|} d^3\vy \Bigg| \ .
\end{eqnarray}
Sufficiently away from the particle, i.e. $|\vx - \vx_p|/d_p \gg 1$, the above
estimate is asymptotically expressed as
\begin{eqnarray}
\label{eqn:estimate_asy}
| \nabla \Phi_S | \le \frac{1}{\rho_f} \int_{\Rset^3} \int_{t-\epsilon_R}^{t^+}
g\left[ \vy-\vx_p(\tau),t-\tau\right] d\tau d^3\vy
\sup_{t-\epsilon_R \le \tau \le t^+} |\vD_p | \Bigg| \nabla \otimes \nabla
\frac{1}{4 \pi |\vx-\vx_p^*|} \Bigg| \, ,
\nonumber \\
\end{eqnarray}
where $\vx_p^* = \vx_p(\tau^*)$, $t-\epsilon_R \le \tau^* \le t^+$, is the
position along the portion of the particle trajectory closest to the point $\vx$.
Given the known integral
\begin{eqnarray}
\label{eqn:known_integral}
\int_{\Rset^3} \frac{1}{\left(2 \pi \sigma^2 \right)^{3/2} }
e^{- r^2/(2 \sigma^2)} d^3 \vr = 1\, ,
\end{eqnarray}
one ends up with
\begin{eqnarray}
\label{eqn:estimate_asy2}
| \nabla \Phi_S | \le \sup_{t-\epsilon_R \le \tau \le t^+} |\vD_p |
\frac{\epsilon_R}{4 \pi \rho_f |\vx-\vx_p^*|^3} \, ,
\end{eqnarray}
where the norm of the double tensor
$\nabla \otimes \nabla \left[1/(4 \pi |\vx -\vx_p^*)\right]$ is given by
\begin{eqnarray}
\label{eqn:tensor_norm}
\Bigg| \nabla \otimes \nabla \frac{1}{4 \pi |\vx - \vx_p^*|} \Bigg| =
\sup_{|\bf{ \hat e} | = 1} \left[ \left(\bf{ \hat e} \cdot \nabla\right)
\nabla
\frac{1}{4 \pi |\vx - \vx_p^* |} \right] = \frac{1}{4 \pi |\vx -\vx_p^*|^3} \ .
\end{eqnarray}
From the expression of the singular component of the pseudo-velocity
(\ref{eqn:pseudo_velo_S}) it is clear that, far from the particle,
$\vv_{\vzeta_S}$ decays exponentially fast, hence the far field dominating
component of $\vv_S$ is provided by the long range correction $\nabla \Phi_S$
that is order $\epsilon_R/r^3$. It is also clear that close to the particle the
singular contribution is unbound. This singular near-field is however unessential
as far as the relevant length scales of the system, either the smallest
hydrodynamic scale $\eta$ or the inter-particle distance, are larger than
$\sigma_R = \sqrt{2 \nu \epsilon_R}$. For this reason, it will be neglected when
advancing the solution of one time step in the actual algorithm illustrated in the
following sections. However, this highly localized field will eventually diffuse
to larger scales at later times. Hence, the singular contribution that is
neglected
during a single time step is successively reintroduced in the field as soon as
it reaches the smallest physically relevant scales of the system. This procedure
guarantees that the error does not accumulate in time, thereby maintaining the
accuracy of the calculation.
Figure~\ref{fig:vorticity_injection} sketches the decomposition into regular
and singular fields, using the vorticity field to describe the process which
is easier to visualize than the velocity field. The sketch highlight the
singular production of vorticity by the particle, its diffusion, associated
to momentum transfer the fluid, and the regularizing effects of viscosity.
A crucial point is that the singular component of the field, which cannot be
represented on a discrete mesh, is fully recovered at a successive time instant
when its characteristic length-scale reaches the grid size.
In the following section the convective effect of the
singular field will be dealt with in more detail, to show that it is indeed
negligible when the dynamics is observed at the relevant hydrodynamical scale.
\begin{figure}
\centerline{
\includegraphics[scale=0.225,angle=-90]{./vorticity_all.pdf}
\includegraphics[scale=0.225,angle=-90]{./vorticity_regular.pdf}
\includegraphics[scale=0.225,angle=-90]{./vorticity_regular_late.pdf}}
\caption{Coupling mechanism and regularization procedure. The green curves
sketch the vorticity field. Left panel: the complete vorticity field generated by
the particle at the current time $t$ is split into the regular $\vzeta_R(\vx,t)$
(solid green line) and singular $\vzeta_S(\vx,t)$ (dashed green line) components
respectively. Central panel: only the regular component $\vzeta_R(\vx,t)$ can be
represented by the computational grid with mesh size $Dx$ at the generic time $t$.
Right panel: after the
elapsed time $\epsilon_R$ (time $t+\epsilon_R$) the singular component of
the vorticity field diffuses to scales large enough to be captured by the
discrete grid. The momentum transfer towards the fluid occurs via viscous
diffusion of the vorticity generated by the particle. When only the regularized
field is considered a small error is incurred in the exchanged momentum. However,
the successive diffusion of the singular field fully recovers the correct
amount of vorticity at a successive time step. Thus the error does
not accumulate in time and remains under control along the simulation.
\label{fig:vorticity_injection}}
\end{figure}
\subsection{Coupling with the carrier flow} \label{sec:coupling_phases}
The regularized fluid velocity of the carrier flow in presence of the perturbing
particles is obtained by aggregating the two contributions of the velocity
decomposition $\vu_R=\vw+\vv_R$ described in subsection \S \ref{sec:interaction}.
The resulting field obeys the equations
\begin{equation}
\label{eqn:ns_regularized}
\begin{array}{l}
\displaystyle \nabla \cdot \vu_R = 0 \\ \\
\displaystyle \frac{\partial \vu_R}{\partial t} + \vu_R \cdot \nabla \vu_R +
\left\{ \vv_S \cdot \nabla \vu_R + \vu_R \cdot \nabla \vv_S + \vv_S \cdot \nabla
\vv_S \right\} = \\
\displaystyle
-\frac {1}{\rho_f} \nabla p + \nu \nabla^2 \vu_R
- \frac{1}{\rho_f} \sum_p^{N_p}
\vD_p(t-\epsilon_R) \, g\left[ \vx-\vx_p(t-\epsilon_R),\epsilon_R \right] \\
\end{array}
\end{equation}
with boundary and initial conditions given by
\begin{equation}
\label{eqn:ns_regularized_bc}
\displaystyle \vu_R\lvert_{\partial {\cal D}} = \vu_{wall} -
\vv_S\vert_{\partial {\cal D}} \, , \qquad
\displaystyle \vu_R(\vx,0)=\vu_0(\vx)\, ,
\end{equation}
where we have added the contributions arising from all the $N_p$ particles
transported by the fluid. It should be stressed that the boundary condition for the
regularized velocity $\vu_R$ at $\partial {\cal D}$ needs taking the singular
contribution $\vv_S$ into account.
An interpretation of equation (\ref{eqn:ns_regularized}) could now be helpful.
Along its motion the particle experiences the hydrodynamic force. In the
formulation here proposed, the force is naturally regularized by viscous
diffusion,
hence the mollified Dirac delta functions takes the form of the fundamental
solution of the diffusion equation. The effect of the hydrodynamic force is the
generation of the regularized vorticity, (\ref{eqn:vort_regular}), that is
characterized by the smallest length-scale $\sigma_R=\sqrt{2 \nu \epsilon_R}$
where $\epsilon_R$ is the regularization diffusion time scale.
A crucial point to be stressed again is that the hydrodynamic forcing acting on
the regularized solution at time $t$ is the one experienced by the particles at
a slightly previous time $t-\epsilon_R$ when their position were
$\vx_p(t-\epsilon_R)$. The net effect of the disperse phase on the regularized
carrier flow field is then accounted for by the extra forcing term corresponding
to the time-delayed hydrodynamic force expressed as the Gaussian
$g\left[ \vx-\vx_p(t-\epsilon_R),\epsilon_R \right]$ with variance $\sigma_R$.
The total field $\vu$ will involve a singular part that is concentrated on scales
smaller than the physically relevant ones. As such the singular contribution is
actually neglected since only the regularized field needs to be considered.
However contributions from the singular disturbance field $\vv_S$ appear
in the term in curly brackets in eq.~(\ref{eqn:ns_regularized}).
Indeed in the far field of the particles $\vv_S$ was already shown to be of the
order of $\epsilon_R/r^3$, which is negligible in comparison with the other terms
in the equation. It is worth reminding that, at the successive time step, the
corresponding contribution is re-introduced in the field, giving rise to no
error accumulation in the long run. The crucial point here is that
eq.~(\ref{eqn:ns_regularized}) is taken to hold everywhere in ${\cal D}$, also
close and inside the domains which are actually occupied by the particles.
In this near field the curly bracket term needs to be treated with some care.
In fact, due to the scale separation between $\vu_R$ and $\vv_S$, the
filtering of the fields on a scale $\Delta$ which is order of the smallest
hydrodynamic scale does not alter $\vu_R$, i.e. denoting by ${\hat \vu}_R$ the
filtered field one has $\vu_R = {\hat \vu}_R$. In such conditions the equations
for the regularized field follows by applying the filter to
system~(\ref{eqn:ns_regularized}). As a result
of scale separation, the filter is actually acting only on the terms in curly
brackets which involve the singular contribution $\vv_S$. As explicitly shown in
appendix \ref{app:singular_velocity}, a few detailed calculations show that
the filtered terms give contribution of the order
\begin{eqnarray}
\label{eqn:h_estimate_text}
\widehat{\vv_S \cdot \nabla \vv_S} \sim
Re_p \left(\frac{\sigma_R}{\Delta}\right)^3 \frac{\vD_p}{\rho_f}g_{max}
\nonumber \\
\\
\widehat{\vv_S \cdot \nabla \vu_R} \sim \widehat{\vu_R \cdot \nabla \vv_S} \sim
Re_p \left(\frac{\sigma_R}{\Delta}\right)^2 \frac{\vD_p}{\rho_f}g_{max} \, ,
\nonumber
\end{eqnarray}
where $g_{max} = 1/(2 \pi \sigma_R^2)^{3/2}$ is the maximum of the mollified delta
function. Clearly the above filtered convective terms are an order $Re_p$ smaller
than the forcing term on the right hand side of eq.~(\ref{eqn:ns_regularized}).
Under the assumption of small particle Reynolds number they can be safely
neglected in the evolution equation of the regularized field.
We like to stress the simplicity of the final equations that have to
be solved,
\begin{equation}
\label{eqn:ns_regularized_filtered}
\begin{array}{l}
\displaystyle \nabla \cdot \vu_R = 0 \\ \\
\displaystyle \frac{\partial \vu_R}{\partial t} + \vu_R \cdot \nabla \vu_R
=
\displaystyle
-\frac {1}{\rho_f} \nabla p + \nu \nabla^2 \vu_R
- \frac{1}{\rho_f} \sum_p^{N_p}
\vD_p(t-\epsilon_R) \, g\left[ \vx-\vx_p(t-\epsilon_R),\epsilon_R \right] \ .
\end{array}
\end{equation}
The effects of the disperse phase on the carrier fluid is taken
into account by an extra term in the Navier-Stokes
equations. Under this point of view, any standard Navier-Stokes solver
can be easily equipped with such extra term which is known in closed form.
Furthermore each particle will produce an active
forcing on the fluid localized in a sphere of radius order $\sigma_R$
centered at the particle position. In presence of many particles only the
few grid points in the sphere of influence of the particle will receive the
disturbance produced by the particle itself. Finally, the forcing term is
grid independent in the sense that, once the grid spacing is refined
($Dx$ progressively getting smaller at fixed $\sigma_R$), any successively finer
grid will only provide a better numerical approximation of the same forcing.
\subsection{Evaluation of the hydrodynamic force \& removal of self-interaction}
\label{sec:prtcl_motion_force}
The dynamics of a point particle of mass $m_p$ in the relative motion with
respect to a Newtonian fluid is described by the equation of motion,
\begin{equation}
\label{eqn:particle_motion}
\frac{d \vx_p}{dt}= \vv_p(t), \quad m_p \frac{d \vv_p}{dt} = \vD_p(t)
+\left( m_p -m_f\right)\vg,
\end{equation}
where $m_f$ is the displaced mass of fluid, $\vD_p(t)$ is the hydrodynamic force,
and $\vg$ the acceleration due to gravity.
Clearly, for the accurate evaluation of the particle trajectories and of the
inter-phase momentum coupling, an accurate and efficient expression for the hydrodynamic
force is mandatory. To obtain such expression one should reconsider the equation
for the perturbation field $\vv$ addressed in \S~\ref{sec:interaction}.
As shown there, the perturbation due to the presence of a particle obeys the
unsteady Stokes equation~\eqref{eqn:unsteady_stokes} where, noteworthy, the
initial condition for the perturbation field $\vv$ is homogeneous. Indeed, in our
scheme, the solution of the unsteady Stokes equation for $\vv$ at the generic time
step provides the stress at the fluid-particle interface and ultimately yields the
drag force. Luckily there is no need to work out the details, since \cite{maxril}
already provided the expression for the general unsteady drag force of a spherical
particle when the field has homogenous initial condition, which is the case of
interest here. Their solution can in fact be fully exploited to provide
the drag force for the perturbation flow that is asymptotically expressed
as \eqref{eqn:unsteady_stokes_far_field} during the generic time step.
Following \cite{maxril} the force $\vD_p(t)$ can be evaluated as
\begin{eqnarray}
\label{eqn:force_particle}
\vD_p(t) &=& 6\pi\mu a_p \left[\tilde \vu(\vx_p,t)
+\frac{a_p^2}{6}\nabla^2 \tilde \vu(\vx_p,t)-\vv_p(t) \right] \nonumber \\
&+&m_f \frac{D \tilde \vu}{Dt}\bigg|_{x_p} +\frac{1}{2} m_f\frac{d}{dt}
\left[\tilde \vu(\vx_p,t) + \frac{a_p^2}{10}\nabla^2 \tilde \vu(\vx_p,t)
-\vv_p(t) \right] \nonumber \\
&+&6\pi\mu a_p^2 \int_0^t d\tau \,
\frac{1}{\left[\pi\nu\left(t-\tau\right)\right]^{1/2}}
\frac{d}{d\tau}\left[ \tilde \vu(\vx_p,\tau)+
\frac{a_p^2}{6}\nabla^2 \tilde \vu(\vx_p,\tau)-\vv_p(\tau)\right]
\end{eqnarray}
where $a_p=d_p/2$ is the particle radius. Expression (\ref{eqn:force_particle})
involves the steady Stokes drag (first line), the added mass terms (second line),
and the Basset history force (third line). In all terms the Faxen correction
associated with spatial non-uniformity of the flow is included and,
following the original derivation by \cite{maxril}, the velocity
$\tilde \vu(\vx_p,t)$ must be interpreted as the fluid velocity, at the particle
position, in absence of the particle self-interaction, i.e. $\tilde \vu_p$ should
account for the background -- possibly turbulent -- flow and for the disturbance
generated by all the other particles except the $p$th one, see also \cite{boivin}
and \cite{gatignol}. In the regime of
our interest, where the particle back-reaction modifies the carrier flow, the
correct calculation of $\tilde \vu_p$ is crucial and calls for an effective
procedure to deprive the field $\vu(\vx,t)$ evaluated at the particle position
from the particle self-interaction contribution.
Luckily, the (regularized) disturbance flow generated by each particle is known in
closed form and can thus be easily removed from the complete field in computing
the hydrodynamic force, at least for numerical algorithms using explicit time
integration schemes. As an illustration, let us consider the simple case of
heavy small particles, $\rho_p \gg \rho_f$, where the
hydrodynamic force (\ref{eqn:force_particle}) reduces to the Stokes drag
\begin{eqnarray}
\label{eqn:force_stokes_drag}
\vD_p(t) &=& 6\pi\mu a_p \left[\tilde \vu(\vx_p,t)
-\vv_p(t) \right].
\end{eqnarray}
The explicit calculation of the velocity $\vv_R(\vx-\vx_0,t)$ induced at time $t$
and position $\vx$ by a particle located at $\vx_0$ is provided
in appendix \ref{app:self_velocity}. This result can be exploited to remove the
self-interaction term in the illustrative case of an explicit Euler time
advancement algorithm. Indeed, in this case, to correctly evaluate the right
hand side of equation~(\ref{eqn:force_stokes_drag}) it suffices to subtract from
$\vu(\vx_p,t)$ the value $\vv_R[\vx_p(t)- \vx_p(t-Dt), Dt]$ induced at time $t$
at the current particle position $\vx_p(t)$ by the same particle when
it was placed at $\vx_p(t -Dt)$. The same kind of reasoning can be
straightforwardly extended to other explicit time integration schemes, e.g.
to each intermediate step of a Runke-Kutta algorithm and to the different
contributions in the general expression of the force
(\ref{eqn:force_particle}), e.g. bubbly flows \cite{climent2006dynamics}.
Before closing this section devoted to force evaluation, a final note is in
order concerning the Basset force: it represents the effects on the
force due to the particle-fluid interaction during the particle previous motion
before the actual time $t$. In cases where the particle do not modify the
carrier flow, see the derivation by \cite{maxril}, this interaction is modeled
by a memory convolution integral which mimics the particle vorticity production
and its viscous diffusion occurring from the initial time $t=0$ up to the actual
time $t$. In our case, the carrier fluid is perturbed step by step by the
particle motion (two-way coupling regime) and the diffusion of the
vorticity produced by the particle during the past motion before the actual time $t$
is captured without any modeling by equations (\ref{eqn:ns_regularized_filtered}).
Hence, the time integral must model the only vorticity production occurring
during the last time step $Dt$, i.e. the memory integral is limited to a single
time step of the eventual integration algorithm. Actually, the effects of the
previous history come in through the boundary condition of equations
(\ref{eqn:unsteady_stokes}) where the field $\vw$ must be interpret as the
background velocity acting on the particle, i.e. as the carrier flow velocity field
that would occur at the particle boundary during the last time step
in absence of the particle. For small particles such field reduces to the
value at particle center plus a Faxen-like correction accounting for spatial flow
variations on the scale of the particle.
\section{Algorithm validation} \label{sec:validation}
The methodology illustrated in the previous sections needs to be validated.
We will address several test cases where analytical data can be employed for
comparison. To better focus our attention on the interaction between the fluid
and the disperse phase, we will consider a periodic box ${\cal D}$ free from solid
boundaries which may hinder the analysis. The numerical solution of equation
(\ref{eqn:velo_reg_eq}) and (\ref{eqn:ns_regularized}) for the carrier fluid is based
on a pseudo-spectral Fourier-based spatial discretization where the non linear terms are
calculated by the standard $3/2$ dealiasing procedure. Time advancement is achieved
by a low-storage semi-implicit Runge-Kutta method with a fourth-order Adams-Bashforth
formulation for the convective terms and an implicit Crank-Nicholson formula for the
diffusive terms. The details on the implementation are described elsewhere, see
\cite{pof_shear}.
\subsection{Response to a localized force}
\label{sec:spatio_temporal_convergence}
We start by addressing a simple case where a known small amplitude constant force
$\vF_0$ is applied at a fixed point $\vx_p$ to the fluid which is initially at rest in
the domain ${\cal D}$. Due to the small amplitude of the forcing, the flow is assumed to
obey the linear, unsteady Stokes equations. Equation (\ref{eqn:velo_reg_eq}) is
suitable of an analytical solution which, in terms of vorticity, is provided by
equation (\ref{eqn:vort_regular}). This reference solution allows to verify that the
algorithm correctly transfers the proper impulse to the fluid, a crucial aspect in view
of simulations in the two-way coupling regime.
Due to periodicity, the exact impulse,
$\vI_E = \vF_0 t = \int_{\cal D} \rho_f \vu(\vx,t) d^3 \vx$, can be expressed in terms
of the vorticity moment (\cite{saffman1992vortex}),
\begin{equation}
\label{eqn:impulse_def}
\vI_E(t)=\frac{1}{2} \rho_f \int_\Omega \vx \times \vzeta(\vx,t) \, d^3 \vx.
\end{equation}
The error $E_I = \lvert \vI_E(t)-\vI_N(t) \rvert$, where
$\vI_N(t)$ is the estimate of (\ref{eqn:impulse_def}) from the numerical solution,
is shown in semi-logarithmic scale in the left panel of figure \ref{fig:conv_imp} as
a function of the normalized time $t/\epsilon_R$ for a fixed value of the
regularization timescale $\epsilon_R=0.01$ and different spatial resolutions, namely
the ratio $\sigma_R/Dx$. In the unresolved cases ($\sigma_R/Dx < 1$),
the error is order one and increases in time. In contrast, when a proper spatial
resolution is adopted, i.e. $\sigma_R/Dx > 1$, the error becomes progressively
smaller as the resolution is increased and stays constant in time.
In other words, as the simulation advances in time
$E_I$ does not accumulate. The right panel of figure \ref{fig:conv_imp} reports the
supreme $\sup_{t \ge 0} \lvert \vI_E(t)-\vI_N(t) \rvert$ as a function of the
ratio $\sigma_R/Dx$. This plot emphasized the convergence rate of the impulse
against the spatial resolution at a fixed value of $\epsilon_R$. The inset
shows $\sup_{t \ge 0} E_I$ in a different manner. Here the spatial resolution
is fixed, $\sigma_R/Dx=1$, and the regularization timescale is progressively
reduced denoting convergence also with respect to the parameter $\epsilon_R$.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./err_imp_vs_time.pdf}
\includegraphics[scale=0.30]{./Sup_E_I_vs_resolution_inset.pdf}}
\caption{Convergence study in the case of a constant force $\vF_0=(1,0,0)$
applied in a fixed point $\vx_p$ to the fluid initially at rest in a periodic box
$\cal D$.
Left: the error $E_I$ of the total impulse, see text for definition,
is plotted against the normalized time $t/\epsilon_R$ for a fixed regularization
timescale $\epsilon_R=0.01$ and different values of the spatial resolution $\sigma_R/Dx$.
Right: the supreme $\sup_{t \ge 0} E_I$ is plotted against the ratio $\sigma_R/Dx$ for
$\sigma_R=0.01$. The inset reports the supreme $\sup_{t \ge 0} E_I$ for
different values of $\epsilon_R$ and a fixed spatial resolution
$\sigma_R/Dx=1$.
\label{fig:conv_imp} }
\end{figure}
The impulse, though a fundamental quantity, does not retain any information
concerning the spatial structure of the fluid field. To go more in depth into
the convergence analysis we have addressed the vorticity field.
The error is now defined by using the standard $L^2$ norm
as $E_{\vzeta}=\lVert \vzeta_E-\vzeta_N \rVert_2$ where the subscripts refer to the
exact regularized solution (\ref{eqn:vort_regular}) and its numerical counterpart.
The error $E_{\vzeta}$ is shown in the left panel of figure \ref{fig:conv_vort}
as a function of time for $\epsilon_R=0.01$ and several values of the ratio
$\sigma_R/Dx$. When a proper spatial resolution is adopted, also the error $E_{\vzeta}$
stays constant in time or decreases. Note that the largest error is
achieved at the early stages of the simulation when the force is applied to
the fluid at rest. In any case, the supreme $\sup_{t \ge 0} E_{\vzeta}$
converges with respect to the refinement of the spatial resolution as
shown in the right panel of the figure. The inset
reports $\sup_{t \ge 0} E_{\vzeta}$ against the regularization timescale
$\epsilon_R$ for a fixed spatial resolution documenting the convergence of
$\sup_{t \ge 0} E_{\vzeta}$ also with respect to $\epsilon_R$.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./err_vort_vs_time.pdf}
\includegraphics[scale=0.30]{./Sup_E_vort_vs_resolution_inset.pdf}}
\caption{Convergence study in the case of a constant force $\vF_0=(1,0,0)$
applied in a fixed point $\vx_p$ to the fluid initially at rest in a periodic box
$\cal D$.
Left: the error $E_{\vzeta}$, see text for definition,
is plotted against the normalized time $t/\epsilon_R$ for a fixed regularization timescale
$\epsilon_R=0.01$ and different values of the spatial resolution $\sigma_R/Dx$.
Right: the supreme $\sup_{t \ge 0} E_{\vzeta}$ is shown as a function of
the ratio $\sigma_R/Dx$ for $\sigma_R=0.01$. The inset reports the supreme
$\sup_{t \ge 0} E_{\vzeta}$ versus the regularization timescale
$\epsilon_R$ for a fixed spatial resolution $\sigma_R/Dx=1$.
\label{fig:conv_vort} }
\end{figure}
A more detailed insight concerning the ensuing fluid motion generated by the fixed
force is achieved by a direct inspection of the flow field.
Figure \ref{fig:u_vs_rx_sigmaR_f=1} reports the fluid velocity in a one dimensional
cut across the complete three-dimensional field. It is
useful to fix the notation: the constant force has cartesian components
$\vF_0=(F_x,F_y,F_z)$, the corresponding velocity vector is $\vu=(u,v,w)$ and the
distance from $\vx_p$ is measured by the vector $\vr=\vx-\vx_p$ whose cartesian
components are $\vr=(r_x,r_y,r_z)$.
In the plots of figure \ref{fig:u_vs_rx_sigmaR_f=1} the force is $\vF_0=(1,0,0)$.
The figure presents the velocity disturbance
$u$ in the direction of the force as a function of the separation $r_x$ along a one dimensional cut aligned
with the x-axis passing through the point where the force is applied.
The different profiles pertain to simulations which share the same regularization
timescale $\epsilon_R=0.01$ but differ for the spatial resolution. Namely,
a typical unresolved case $\sigma_R/Dx=0.38$ and a resolved simulation
$\sigma_R/Dx=1$ are compared against the exact solution at two different times
along the simulation. In fact, when a fixed constant force is applied to the
fluid, an exact solution can be easily determined in closed form
by evaluating the time convolution integral between the unsteady Green tensor, see
appendix equation (\ref{eqn:green_tensor}), and the force $\vF_0$. After some algebra,
the fluid velocity disturbance in the direction of the force reads
\begin{equation}
\label{eqn:ref_velo_f_fixed}
u(\vr,t)=\frac{1}{4 \pi \mu r}
\left[\frac{1}{2 \eta_t^2}\mbox{erf}\left(\eta_t\right) -
\mbox{erf}\left(\eta_t\right)-
\frac{1}{\sqrt{\pi} \eta_t}\exp\left(-\eta_t^2\right)+1\right]
\end{equation}
where $\eta_t=r/\sqrt{4 \nu t}$ and $r=\sqrt{r_k \, r_k}$.
As shown in the figure \ref{fig:u_vs_rx_sigmaR_f=1},
when $\sigma_R/Dx \ge 1$ the present algorithm well reproduces
the exact solution. Note that insufficient spatial resolution results in a
clear underestimate of the fluid velocity disturbance. This is emphasized
by the plots reported in the insets figure \ref{fig:u_vs_rx_sigmaR_f=1}
where the data are represented in a semi-logarithmic scale.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./u_vs_rx_t=005.pdf}
\includegraphics[scale=0.30]{./u_vs_rx_t=1.pdf}}
\caption{Fluid velocity disturbance generated by a fixed constant force
$\vF_0=(1,0,0)$ on an initially motionless fluid contained in a
periodic box $L_B=4\pi$. The $1D$ profile representing the fluid velocity
in the direction of the force (symbols) is compared against the exact solution
(solid line). The velocity component $u(\vr,t)$ in the $x-$direction is plotted
against the separation $r_x$ for $\epsilon_R=0.01$ and two spatial resolutions,
namely $\sigma_R/Dx=0.38$ ($\square$) and
$\sigma_R/Dx=1$ ($\bigcirc$). Left: velocity disturbance at $t=0.05$.
Right: velocity disturbance at $t=1$. In the insets of the two panels
the data are plotted in a semi-logarithmic scale.
\label{fig:u_vs_rx_sigmaR_f=1} }
\end{figure}
Figure \ref{fig:u_vs_rx_epsR_f=1} documents the behavior of the ERPP method
when the spatial resolution $\sigma_R/Dx$ is kept fixed and the regularization
timescale is progressively reduced. In fact, as $\epsilon_R$ is decreased,
the numerical solution describes a progressively wider range of the exact solution
avoiding in all cases the occurrence of the singularity at the point $\vx_p$
where the force is applied, $r_x=0$ in the plot.
The different cases share the same far field behavior
away from $\vx_p$ irrespective of the value of $\sigma_R$ as
emphasized by the plots in the insets of figure \ref{fig:u_vs_rx_epsR_f=1}
where the velocity disturbance is represented in a semi-logarithmic scale.
In summary, the solution provided by the ERPP retains the relevant features
of the exact solution and avoids the occurrence of the singularity at $\vx_p$
which is clearly an unwanted trait in any numerical solution.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./u_vs_rx_vari_epsR_t=005.pdf}
\includegraphics[scale=0.30]{./u_vs_rx_vari_epsR_t=1.pdf}}
\caption{Fluid velocity disturbance generated by a fixed constant force
$\vF_0=(1,0,0)$ on an initially motionless fluid contained in a
periodic box $L_B=4\pi$. The $1D$ profile representing the fluid velocity
in the direction of the force (symbols) is compared against the exact solution
(solid line). The velocity component $u(\vr,t)$ in the $x-$direction is plotted
against the separation $r_x$ for several values of the regularization
timescale, $\epsilon_R=0.01$ ($\square$); $\epsilon_R=0.02$ ($\triangle$);
$\epsilon_R=0.005$ ($\diamond$); $\epsilon_R=0.0025$ ($\bigcirc$), at a fixed
spatial resolution $\sigma_R/Dx=1$.
Left: velocity disturbance at $t=0.05$.
Right: velocity disturbance at $t=1$. The insets of the two panels show
the data plotted in a semi-logarithmic scale.
\label{fig:u_vs_rx_epsR_f=1} }
\end{figure}
Figure \ref{fig:u_vs_ry} reinforces the conclusion of the previous analysis by showing the fluid velocity
component in the direction of the force $u(r_y)$ as a function of the distance $r_y$ in a transversal one-dimensional
cut through the point of application of the force.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./u_vs_ry_vari_t.pdf}
\includegraphics[scale=0.30]{./u_vs_ry_vari_epsR.pdf}}
\caption{Fluid velocity disturbance generated by a fixed constant force
$\vF_0=(1,0,0)$ on an initially motionless fluid contained in a
periodic box $L_B=4\pi$. The $1D$ profile representing the fluid velocity
component $u(\vr,t)$ in the $x-$direction (symbols)
is plotted against the separation $r_y$. The exact solution (solid line)
is reported for comparison.
Left: fluid velocity disturbance at $t=0.05$ (main panel) and
$t=1$ (inset) plotted for a fixed value of $\epsilon_R=0.01$ and two spatial resolutions
namely $\sigma_R/Dx=0.38$ and $\sigma_R/Dx=1$.
Right: fluid velocity disturbance at $t=0.05$ (main panel) and
$t=1$ (inset) plotted for several values of $\epsilon_R$ and a fixed spatial
resolution $\sigma_R/Dx=1$.
\label{fig:u_vs_ry}}
\end{figure}
Finally figure \ref{fig:u_vs_r_normalized} reports the fluid velocity component
in the direction of the force plotted against the normalized distance
$r_x/\sigma_R$ and $r_y/\sigma_R$. The discussion of these plots requires some care.
The ERPP model has been conceived to describe the far field effect produced on the
fluid by a point-like particle avoiding the occurrence of singularities in the
point $\vx_p$ where the particle is located. In fact, the disturbance produced by the
particle is described by retaining only the first term in the multipole
expansion of the general solution of the unsteady Stokes flow and the regularization
timescale $\epsilon_R$ accounts for the viscous diffusion process which naturally
regularizes the solution. Hence, the solution provided by the ERPP has un intrinsic
inner cut-off provided by $\sigma_R$ and is expected to reproduce the disturbance flow
generated by a point-particle in the far field. This is indeed what happens and what is
documented by the plots in figure \ref{fig:u_vs_r_normalized}. The regularized
solution stays on top of the exact solution everywhere, see e.g. the main panels of
figure \ref{fig:u_vs_r_normalized} which, on the scale of the complete computational
domain $\cal D$, reports the fluid velocity disturbance produced by the particle. The
insets of figure \ref{fig:u_vs_r_normalized} show the same data in proximity of the
origin where the particle is placed. Such representation emphasizes that,
after a distance of a few $\sigma_R$, the ERPP solution falls on top of the exact solution.
The threshold $3\sigma_R$ can be safely assumed as an inner cut-off for the
disturbance flow field produced by the particle.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./u_vs_rx_sigma_R_t=1.pdf}
\includegraphics[scale=0.30]{./u_vs_ry_sigma_R_t=1.pdf}}
\caption{Fluid velocity disturbance generated by a fixed constant force
$\vF_0=(1,0,0)$ on an initially motionless fluid contained in a
periodic box $L_B=4\pi$. The velocity component $u(\vr,t)$ in the $x-$direction
(direction of the force, symbols) is plotted against the normalized
separation $r_x/\sigma_R$ (left panel) and $r_y/\sigma_R$ (right panel) for
$\epsilon_R=0.0025$ and $\sigma_R/Dx=1$. The inset of the two panels
provides a close-up view of the solution at $\vx_p=0$.
\label{fig:u_vs_r_normalized}}
\end{figure}
\subsection{Unsteady motion of an isolated particle}
\label{sec:unsteady_motion}
The following subsections address more realistic cases where the point $\vx_p$
is allowed to move according to equations (\ref{eqn:particle_motion}) with
the initial conditions $\vx_p(t=0)=\vx_p^0; \, \vv_p(t=0)=0$.
In order to proceed gradually we will first discuss a series of tests where the
particle motion is not affected by the fluid velocity disturbance that the particle
generates. In such decoupled cases the fluid disturbance field is still amenable
of exact solutions which can be employed for further comparisons.
Successively, we will consider the fully coupled case where the dynamics of the
particles and of the fluid are intertwined.
\subsubsection{Imposed particle motion}
We consider the motion of a small particle subjected to an external force,
e.g. gravity, and to the Stokes drag. The particle velocity is given by the
solution of the equation
\begin{equation}
\label{eqn:Drag_one_way}
m_p \frac{d \vv_p}{dt}= m_p\vg - 6\pi\mu a_p \vv_p(t) \, ,
\end{equation}
namely
\begin{equation}
\label{eqn:uncoupled_solution}
v_p(t)=v_t\left(1-e^{-t/\tau_p}\right)
\end{equation}
where $v_p(t)$ denotes the particle velocity in the direction of the gravity
acceleration, say the $x-$direction, $v_t=\tau_p\,g$ is the particle terminal
velocity and $\tau_p=\rho_p d_p^2/{18 \mu}$ the Stokes relaxation timescale.
In this framework the motion of the particle is imposed and its dynamics
is decoupled from the dynamics of the carrier fluid.
The left panel of figure \ref{fig:sigmaR_Re} reports
the fluid velocity disturbance produced by the moving particle
at $t/\tau_p=20$ when the particle has reached its terminal velocity $v_t$.
The fluid velocity profile is plotted for two cases which differs for
the value of the regularization timescale $\epsilon_R$. Once again
the value of $\epsilon_R$ controls the regularized near field and does not
affect the far field, see the inset of the figure.
In this case the fluid velocity has an explicit solution given by
the time convolution integral of
the unsteady Stokeslet, equation (\ref{eqn:green_tensor}), and the
the hydrodynamic force $D_p(t)$ where the Stokeslet is placed at the instantaneous position of
the particle evolving as specified by equation (\ref{eqn:uncoupled_solution}). Performing the time convolution integral is now a little more
tricky than in the previous example since $\vr = \vx-\vx_p(t)$ where $\dot{x}_p=v_p(t)$.
The integration of the expression
\begin{equation}
\label{eqn:ref_velo_disturbance}
u_i(\vx,t)=\int_0^t G_{ik}\left[\vx-\vx_p(\tau),t-\tau\right] D^p_k(\tau)
\quad d\tau \,
\end{equation}
in a closed form becomes cumbersome even though the integral can be evaluated
numerically by a quadrature formula. Indeed, such numerical approximation
of the exact solution can be still used for useful comparisons.
The data produced by the ERPP algorithm are compared with this reference
solution in figure \ref{fig:sigmaR_Re}.
The plots show that the ERPP approach is able to capture
the expected solution and provides a consistent regularization of the
singularity which occurs at $\vx_p$. The large fluid domain, note that the box size
is $L_B=4\pi$, allows a good approximation of the unbounded domain where the reference
solution (\ref{eqn:ref_velo_disturbance}) holds. The right panel of figure
\ref{fig:sigmaR_Re} provides the fluid velocity disturbance produced by
particles with different terminal velocities $v_t$.
The effect of increasing the terminal velocity is worth discussing. As $v_t$ is increased the
front-aft symmetry in the disturbance flow is progressively broken.
This symmetry breaking is indeed easily explained in terms of vorticity released along the path of
the moving particle. In fact in the body-fixed frame, the convection term
$\vv_p\cdot \nabla \vu$ is responsible of the constant velocity advection of the
vorticity even in a Stokes regime. In order to capture the tails of the disturbance flow in the far field and
to avoid confinement effects introduced by the periodic boundary conditions, the computational box is now $L_B=8\pi$ wide.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./u_vs_x_t=10_uncoupled_vari_epsR.pdf}
\includegraphics[scale=0.30]{./u_vs_x_t=10_uncoupled_vari_Re.pdf}}
\caption{Normalized velocity disturbance produced by a particle moving with
velocity $v_p(t)=v_t\left(1-e^{-t/\tau_p}\right)$ in the $x-$direction in a
periodic box $L_B=4\pi$.
The normalized velocity disturbance $u/v_t$ in the direction of
the particle motion, is plotted against the separation $r_x$ at time
$t/\tau_p=20$ when the particle has reached the terminal velocity $v_t$.
Left panel: data obtained for different values of the regularization timescale
$\epsilon_R=0.006$ ($\triangle$); $\epsilon_R=0.001$ ($\bigcirc$)
at a fixed spatial resolution $\sigma_R/Dx=1$, are
compared against the exact solution (solid line) given by equation
(\ref{eqn:ref_velo_disturbance}). The inset reports
the data of the main panel plotted in a semi-logarithmic scale.
Right panel: data pertaining particles with
with different terminal velocities, $v_t^*= v_t d_p/\nu=10^{-5}$ ($\square$);
$v_t^*=10^{-3}$ ($\triangle$), are compared against the corresponding exact
solution (\ref{eqn:ref_velo_disturbance}), dashed and solid line respectively,
in a periodic box $L_B=8\pi$. The inset reports the date of the main panel
plotted in a semi-logarithmic scale.
\label{fig:sigmaR_Re} }
\end{figure}
We conclude the discussion by presenting in figure \ref{fig:cfr_PIC_uncoupled}
the comparison between the ERPP solution and what one would obtain by using the
classical Particle In Cell (PIC) approach. As expected, the solution
provided by the PIC method is grid dependent, as demonstrated by comparing the disturbance
velocity profile of two simulations which share the same physical parameters but
differ for the grid resolution, namely $N=192^3$ and $N=384^3$ Fourier modes.
As the grid is refined
a singular-like behavior occurs at $\vx_p$ and the field is characterized by
numerical aliasing, see e.g. the top right inset where the velocity profile
is plotted in a semi-logarithmic scale. In contrast, once the regularization timescale
$\epsilon_R$ is fixed, the ERPP approach provides a numerically convergent, asymptotically
grid-independent solution. This behavior
can be better appreciated in the top left inset where a close-up view of the
velocity disturbance is reported. In a nutshell the ERPP retains all the features
of the physical solution produced by a small point-like particle except for the (undesired)
singularity which unavoidably occurs at $\vx_p$. The regularization
of the solution is controlled by the timescale $\epsilon_R$ which is
related to a diffusive lengthscale $\sigma_R$. Indeed $\sigma_R$ is naturally
introduced by the process of vorticity diffusion and can be fixed on a physical ground.
For instance, in a
turbulent flow, velocity fluctuations are physically irrelevant below the Kolmogorov
lengthscale $\eta$. At the same time the effects that a swarm of point-like particles generates on length scales larger
than $\eta$ is physically relevant. In such framework the regularization lengthscale $\sigma_R$
is naturally selected as $\sigma_R=\eta$.
\begin{figure}
\centerline{
\includegraphics[scale=0.50]{./u_vs_x_t=10_uncoupled_cfr_PP_2inset.pdf}}
\caption{Normalized velocity disturbance at $t/\tau_p=20$ produced by a particle
moving with velocity $v_p(t)=v_t\left(1-e^{-t/\tau_p}\right)$ in the $x-$direction
in a periodic box $L_B=4\pi$. The velocity profile calculated with
the ERPP for different values of the regularization timescale
$\epsilon_R=0.006$, $N=192^3$ Fourier modes, $\sigma_R/Dx=1$ ($\square$);
$\epsilon_R=0.006$, $N=384^3$ Fourier modes, $\sigma_R/Dx=2$, ($\triangle$);
$\epsilon_R=0.001$, $N=384^3$ Fourier modes, $\sigma_R/Dx=1$, ($\bigcirc$);
is compared against the exact solution (solid line) and corresponding results obtained
by the PIC approach, $N=192^3$ Fourier modes, ($\triangledown$);
$N=384^3$ Fourier modes, ($\diamond$).
Top left panel: close-up view of the velocity disturbance near
the singular point $\vx_p$. Top right inset: velocity profile plotted in
semi-logarithmic scale.
\label{fig:cfr_PIC_uncoupled} }
\end{figure}
\subsubsection{Particle motion in the coupled regime}
This subsection addresses the unsteady motion of a particle which
settles from rest under the action of gravity in the coupled regime
where the particle induces a disturbance in the surrounding fluid and such
disturbance enters the expression of the hydrodynamic force.
For simplicity we will consider small particles much heavier than the surrounding
fluid, i.e. $\rho_p\gg\rho_f$, where the only relevant force is the Stokes drag.
The general expression of the force \eqref{eqn:force_particle} simplifies to
\begin{equation}
\label{eqn:drag_coupled}
\vD_p(t)=m_p\vg + 6\pi\mu a_p \left[\tilde\vu(\vx_p,t)
-\vv_p(t) \right] \,.
\end{equation}
Following the discussion of section \ref{sec:prtcl_motion_force} the velocity
$\tilde\vu(\vx_p,t)$ must be interpreted as the background fluid velocity in absence of the
p$th$ particle, e.g. turbulent fluctuations plus the disturbance flow generated by all
the other particles. This makes the calculation of the
hydrodynamic force particularly challenging in the two-way coupling regime.
In the particular case where only one particle is considered the value
$\tilde\vu(\vx_p,t)$ should be set to zero. However this way of proceeding is unfeasible
in the general case where many particles are present since the value of
$\tilde\vu(\vx_p,t)$ must also account for the velocity disturbance generated by all the
other particles and the background flow. This conundrum can be disentangled in
the context of the ERPP approach since the disturbance flow produced by the $p$th
particle on itself is known in a closed form and thus can be removed from
the background fluid velocity $\vu(\vx_p,t)$ even in presence of many other particles.
The two panels of figure \ref{fig:velo_vs_t} provide evidence to the above
considerations.
The plots report the particle velocity normalized with the settling velocity
$v_t$ as a function of the dimensionless time $t/\tau_p$ both for the ERPP
calculation and for the PIC approach. The particle trajectory should be compared with
the reference solution given by equation (\ref{eqn:uncoupled_solution}).
The left panel shows the particle velocity calculated by
the ERPP method for different values of the ratio $d_p/\sigma_R$. We recall
that our approach is designed to model the disturbance flow produced by point-like
particles, i.e. particles whose diameter $d_p$ is much smaller that any other
lengthscale in the system. In the ERPP approach, in absence of any other length-scales
introduced by the background flow, the only significant lengthscale is the
diffusive scale $\sigma_R$. Hence, the nominal diameter of the particle should be
smaller than $\sigma_R$. Indeed, as the ratio $d_p/\sigma_R$ decreases the particle
velocity rapidly approaches the reference curve provided by
equation (\ref{eqn:uncoupled_solution}). When the scale $\sigma_R$ has been fixed
the ERPP gives a grid-independent solution as can be appreciated in
figure \ref{fig:velo_vs_t} where two trajectories which share the same $\sigma_R$
but have different grids, namely $N=192^3$ and $N=24^3$ Fourier modes,
give practically undistinguishable results. It's worth noting that the error in the
particle velocity is already below $10\%$ for the relatively large ratio
$d_p/\sigma_R=0.5$ we have considered. The right panel of figure \ref{fig:velo_vs_t}
reports the particle velocity calculated with the PIC method. The solution
presents now larger deviations from the exact
result (\ref{eqn:uncoupled_solution}). This is due to a poor estimate of the
hydrodynamic force. In fact, in the PIC approach the self-induced
disturbance produced by the $p$th particle is unknown or, if eventually modeled by
the steady Stokeslet, is singular at the particle position $\vx_p$. In both cases
it can not be removed from the particle-to-fluid slip velocity resulting in
an inaccurate prediction of the hydrodynamic force and, consequently, of the
particle trajectory. For instance, for $d_p/Dx=0.5$ the error on the terminal velocity is $50\%$ for the PIC approach
compared to a much lower $10\%$ for the ERPP. Clearly the error reduces as
the ratio $d_p/Dx \to 0$.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./velo_vs_t_erpp.pdf}
\includegraphics[scale=0.30]{./velo_vs_t_pp.pdf}}
\caption{Normalized particle velocity $v_p/v_t$ as a function of the dimensionless
time $t/\tau_p$. Left panel: particle velocity in the ERPP simulations for
different values of the ratio; $d_p/\sigma_R=0.5$, ($\square$);
$d_p/\sigma_R=0.25$, ($\triangle$); $d_p/\sigma_R=0.125$, ($\diamond$);
$d_p/\sigma_R=0.0625$, ($\bigcirc$). The fluid field is discretized with $N=192^3$
Fourier modes (open symbols) in a periodic box $L_B=2\pi$. In the case
$d_p/\sigma_R=0.0625$, ($\blacksquare$) the fluid field is discretized with $N=24^3$
Fourier modes to check grid-independence.
Right panel: particle velocity provided by the PIC approach in comparable conditions
as in the ERPP. $d_p/Dx=0.5$, ($\square$); $d_p/Dx=0.25$, ($\triangle$);
$d_p/Dx=0.125$, ($\diamond$); $d_p/Dx=0.0625$, ($\bigcirc$).
In both panels the reference solution (\ref{eqn:uncoupled_solution}) is reported for
comparison (solid line).
\label{fig:velo_vs_t} }
\end{figure}
A more direct comparison between the two approaches is provided in the
left panel of figure \ref{fig:velo_terminal} where we plot the particle velocity
for the largest ratio $d_p/\sigma_R=0.5$ and the smallest one $d_p/\sigma_R=0.0625$
both for the ERPP and the PIC. For comparison in the ERPP calculation we
have reported the particle velocity in a case where we did not subtract
from $\vu(\vx_p,t)$ the self-induced disturbance. The right panel of the figure
presents the relative error committed in the estimate of the terminal velocity
as a function of $d_p/\sigma_R$. Although the error scaling with grid resolution is comparable (see inset),
the error pertaining to the ERPP approach is substantially smaller.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./velo_vs_t_erpp_vs_pp.pdf}
\includegraphics[scale=0.30]{./Err_vt_inset.pdf}}
\caption{Normalized particle velocity $v_p/v_t$ as a function of the dimensionless
time $t/\tau_p$. Left panel: direct comparison of the ERPP results
against the PIC approach and the reference solution (solid line)
for different values of the ratio $d_p/\sigma_R$ or equivalently $d_p/Dx$.
$d_p/\sigma_R \div d_p/Dx =0.5$ (squares), $d_p/\sigma_R \div d_p/Dx =0.0625$
(circles). Black filled symbols refer to the PIC approach, open symbols to the ERPP
method. The grey square refers to an ERPP calculation where intentionally we
did not remove the self-induced velocity disturbance.
Right panel: relative error committed on the evaluation of the
terminal velocity as a function of the dimensionless parameter $d_p^*=d_p/\sigma_R$
for the ERPP and the PIC. In the inset data plotted in log-log scale.
\label{fig:velo_terminal} }
\end{figure}
A last issue concerns the sensitivity of the ERPP
method in poorly resolved cases where $\sigma_R < Dx$. In figure
\ref{fig:velo_unresolved} we compare the particle velocity for
three different resolutions at fixed $d_p/\sigma_R$. As expected, the method loses accuracy as the
regularization kernel is not resolved on the computational grid.
\begin{figure}
\centerline{
\includegraphics[scale=0.30]{./velo_vs_t_erpp_unresolved.pdf}}
\caption{Normalized particle velocity $v_p/v_t$ as a function of the dimensionless
time $t/\tau_p$. The trajectories obtained in two unresolved cases, namely
$\sigma_R/Dx=0.5$ and $\sigma_R/Dx=0.25$ are compared against
a resolved case at $\sigma_R/Dx=1$ for a given value of the ratio
$d_p/\sigma_R=0.0625$.
\label{fig:velo_unresolved}}
\end{figure}
\section{Application to turbulent flows} \label{sec:turb_flow}
In order to discuss the feasibility of turbulent, particle-laden flow simulations,
in this section we present preliminary results obtained by the ERPP method
for a homogeneous shear flow at moderate Reynolds number. The mean velocity profile
in the $x$-direction (streamwise direction) is imposed and is characterized
by a constant velocity gradient $S$ along the $y$-direction (shear direction).
The third coordinate is denoted with $z$ (spanwise direction). The Reynolds
decomposition $\vu = S y \ve_x + \vu^\prime$ allows to write the Navier-Stokes
equations for the turbulent fluctuating component $\vu^\prime$ which
are solved in a reference frame advected by the mean flow,
see e.g. \cite{rogallo}. The Rogallo's transformation allows to restore
the spatial homogeneity of the fluctuations in the convected frame. A sketch
of the flow domain is reported in figure \ref{fig:flow_config}.
In the homogeneous shear flow the turbulent fluctuations are sustained by the
off-diagonal component of the Reynolds shear stress $- \langle u \, v\rangle$
resulting in a neat turbulent kinetic energy production rate
${\cal P}=-S \langle u \, v \rangle$.
The large scale anisotropic forcing feeds the energy cascade operated by the
non-linear terms of the Navier-Stokes equations which eventually restore
isotropy at smaller scales. The so called shear scale $L_S$ ideally
separates the production range $L_S < \ell < L_0$ ($L_0$ is the
integral scale), from the isotropy recovery range $\eta < \ell < L_S$.
It follows that the nature of turbulent fluctuations is parametrized
by two dimensionless parameter,
the shear intensity $S^*=\left(L_0/L_S\right)^{2/3}$ and the Corsin parameter
$S_c=\left(\eta/L_S\right)^{2/3}$. The latter can be recast in terms of the
inverse of the classical turbulent Reynolds number $Re_\lambda$ based on the
Taylor length-scale.
In the conditions discussed above the transport of inertial particles is non
trivial. In fact, the disperse phase is characterized by small scales
aggregates (clusters) which preserve a spatial preferential orientation induced
by the large scale motions, up to the smallest scales where, in contrast,
turbulent fluctuations recover isotropy, see e.g. \cite{jfm_part,jfm_2way}.
\begin{figure}
\begin{center}
\includegraphics[scale=0.35]{./3D_view.jpg}
\end{center}
\caption{Sketch of the flow configuration. The flow domain is represented by
a periodic box of length $L_x=4\pi$, $L_y=2\pi$ and $L_z=2\pi$ in the
steamwise, shear and span wise directions respectively. The mean flow $S y \ve_x$
is in the $x$-direction and linearly changes at a rate $S$ along the
$y$-direction. The contour plot shows the intensity of the velocity fluctuations
in selected coordinate planes.
\label{fig:flow_config}}
\end{figure}
The first result concerns a flow at a Taylor based Reynolds number of
$Re_\lambda=60$ and shear intensity $S^*=7$. The carrier phase is resolved by
using $N_x\times N_y\times N_z = 256 \times 256 \times 128$ Fourier modes in a
$4\pi \times 2\pi \times 2\pi$ periodic box. Such spectral-based
discretization corresponds to $384 \times 384 \times 192$ collocation
points in physical space due to the $3/2$ dealiasing procedure required for
the calculation of the non linear terms. The spatial discretization
fully resolves the Kolmogorov length scale with $Dx/\eta \sim 1.07$.
Time integration is performed
by the low-storage Runge-Kutta method already mentioned in section
\S \ref{sec:validation}. The carrier fluid is laden with
$N_p=2200000$ inertial particles. The particle to fluid density ratio is
$\rho_p/\rho_f=1800$ corresponding to a Stokes number
$St_\eta=\tau_p/\tau_\eta=1$ where
$\tau_p=\left(\rho_p/\rho_f\right) d_p^2/18 \nu$ is the Stokes relaxation
time and $\tau_\eta$ is the Kolmogorov time scale. In such conditions
the particle diameter $d_p$ is much smaller than the Kolmogorov length,
namely $d_p/\eta=0.1$. The mass load $\Phi$ defined as the
ratio between the mass of the disperse phase and the carrier fluid
is $\Phi=0.4$.
In figure \ref{fig:snapshots} we present a snapshot of the particle position
in a $xy$ plane containing the mean flow (from left to right).
As expected, particles with unitary Stokes number
are characterized by small scale clusters, i.e. the particles concentrate in
narrow regions, the clusters, separated by voids where neither a particle
can be found. The preferential alignment of the aggregates
along the principal strain direction of the mean flow is also evident from
the snapshot. This is the signature
of the persistent anisotropy of the clusters at small scales.
In the context of the ERPP methodology we are able to compute
in a closed form the forcing operated by the particles on the
fluid. In the middle panel of figure \ref{fig:snapshots} we report
the intensity of the forcing term of equations
(\ref{eqn:ns_regularized}) which accounts for the back-reaction on the
fluid. The pattern of the back-reaction field is strongly correlated to the
cluster structure and inherit from the latter its characteristic
multi-scale nature. The forcing is actually active in a broad range
of scales up to the smallest scales where intense peaks occur. Note
however, that the forcing field is everywhere smooth and can be successfully
represented on the discrete grid by virtue of the regularization naturally
operated by the viscosity. The highest forcing intensity is
localized in the spatial regions where the particles concentrate
while in the void regions the forcing vanishes.
The correlation between the instantaneous particles spatial
configuration and the corresponding back-reaction on the fluid can be
visually appreciated in the bottom panel of the figure where the two fields are
superimposed.
\begin{figure}
\begin{center}
\includegraphics[scale=0.45]{./clusters_2D.jpg}
\includegraphics[scale=0.45]{./forcing_2D.jpg}
\includegraphics[scale=0.45]{./clusters_and_forcing_2D.jpg}
\end{center}
\caption{
Snapshot of the instantaneous particle configuration (top) and corresponding
intensity of the forcing on the fluid (center) in a thin slice along the
$xy$ plane. The mean flow $S\, y$ is in the $x$ direction from left to right.
The two top panels are superimposed in the bottom panel to provide a visual
correlation between the instantaneous particles configuration and the corresponding
forcing field.
\label{fig:snapshots}}
\end{figure}
As anticipated, this short paragraph is aimed at the clear illustration of
the potential of the ERPP in dealing with
actual turbulent flows laden with millions of particles. Clearly a complete
analysis of the turbulence modulation in the two-way coupling regime
would require a more complete statistical analysis which is however beyond
the intents of the present work and is postponed to future investigation.
\section{Final remarks} \label{sec:final}
In this paper we have presented a new methodology, dubbed the ERPP method,
able to capture the
inter-phase momentum exchange between a carrier flow and a disperse phase modeled
as lumped massive points. The coupling mechanism is designed on the physical ground
provided by the unsteady Stokes flow around a small sphere. In short, along its
trajectory the particle continuously generates a highly localized vorticity field
that can be evaluated in a closed form. Successively, because of viscous
diffusion, the vorticity field reaches the physically significant length-scales of
the flow field. When this occurs, the newly generated vorticity can be injected
on the computational grid
where the Navier-Stokes equations for the carrier flow are solved thus
achieving the inter-phase momentum coupling. Under this respect
the viscous diffusion naturally regularizes the disturbance flow produced by
each particle without requiring any ``ad hoc'' numerical artifact.
The proposed approach can be implemented in a highly efficient computational
algorithm since the disturbance produced by a particle is strongly
compact in space and localized around the actual particle position.
This means that at each time step only few grid points perceive the particle
disturbance which decays exponentially fast in space. As a consequence the ERPP
method can handle millions of particles at an affordable computational cost as
proved by preliminary results of a particle laden turbulent homogeneous shear flow
in the two-way coupling regime.
The ERPP overcomes several drawbacks of established methods, like the Particle In
Cell method. Indeed, the regularization of the back-reaction field provided by
the viscous diffusion allows numerically convergent solutions, preventing the
strong grid-dependence which spoils singularity-based approximations.
Even more important, the ERPP method solves the intrinsic difficulty of
numerical simulations
in the two-way coupling regime associated with the calculation of the correct
particle-to-fluid slip velocity. Actually, in the ERPP method the disturbance flow
produced by the particles at each time step can be evaluated in a closed form.
This allows to remove from the particle-to-fluid slip velocity the
spurious self-induced velocity disturbance allowing for a correct evaluation of
the hydrodynamic force.
The preliminary results concerning a turbulent particle-laden shear flow presented
in the last section demonstrate the potential of the ERPP method in the
simulation of turbulent flows in the two-way coupling regime. It is known that the
dynamics of the two phase system is fully characterized by a given set of
dimensionless parameters, namely
$\left\{Re_0, \, St_\eta, \, \rho_p/\rho_f, \, d_p/\eta, \, \Phi, \, N_p \right\}$.
To comment on the effectiveness of the ERPP method in modeling turbulent suspensions,
let us assume that the turbulence characteristics are prescribed i.e. the
the turbulent Reynolds number $Re_0$, the integral scale $L_0$, the
Kolmogorov scale $\eta$ or timescale $\tau_\eta$ are fixed, and let us
consider small particles, i.e. $d_p/\eta \ll 1$. In these conditions
the Stokes number $St_\eta=\displaystyle \tau_p / \tau_\eta$
controls the dynamics of the disperse phase in terms of its preferential spatial
accumulation, either small scale clustering in homogeneous
flows
\cite{toschi2009lagrangian,meneguz2011statistical,jfm_bec,monchaux2010preferential,reade2000effect}
or turbophoresis in wall bounded flows \cite{pica,marsol}.
Once the Stokes number is fixed, the mass load $\Phi$ follows as
\begin{equation}
\label{eqn:phi_constraint}
\Phi=\frac{\pi}{6} \, N_p \, \frac{\left(18 \, St_\eta\right)^{3/2}}
{\left(\rho_p/\rho_f\right)^{1/2} \, Re_0^{9/4}} \, .
\end{equation}
The value of $\Phi$ can be adjusted by means of the density ratio $\rho_p/\rho_f$
and the number of particles $N_p$.
However, in an actual experiment the ratio $\rho_p/\rho_f$ must fall in the
range of the
available materials and the most straightforward way to achieve the desired mass
load consists in adjusting the number of particles $N_p$. Although rather
easy in experiments, adjusting the number of particles turns out to be a big
issue in numerical simulations since, most often, the momentum coupling model
is unable to handle an arbitrary number of particles while providing
grid-independent and physically consistent results. At variance with most
available methods, both these requirements are fulfilled by the ERPP approach.
Indeed, the number of particles can be freely changed since disturbance flow
and back-reaction of each particle are smooth fields. This implies that the
solution is correctly reproduced also in flow regions where the particles are
extremely dilute, like it happens for the exterior region of
turbulent jets or spatially evolving boundary layers.
For comparison, classical approaches like the PIC method intrinsically
suffer of spurious numerical oscillations in the back-reaction field
when too few particles per computational cell are available, leading to
strong limitations in the achievable mass load $\Phi$.
Indeed, in any Direct Numerical Simulation of a turbulent flow the number of
computational cells scales with the Reynolds number like $N_c \sim Re_0^{9/4}$
suggesting through equation (\ref{eqn:phi_constraint}) that no room is available
to adjust the mass load if the additional constraint $N_p/N_c \sim 1$
needs to be enforced. This limitation is overcome in the approach proposed here
by relaxing the request on the particle density to allow for the
modeling freedom needed to reproduce any physically relevant condition.
\section{Ackonwledgements}
The results of this research have been achieved using the
PRACE-2IP project (FP7 RI-283493) resource {\em Zeus} based in Poland at
Krakow. The authors are gratefull to the COST Action MP0806 Particles in
Turbulence.
|
\section{Introduction}
NGC~7538 is a well studied optically visible H\,{\tiny II}\ region, located at a distance of 2.65~kpc \citep{moscadelli09}, surrounded by very massive star--forming complexes \citep{Minn75, Werner79, Read80, Rots81, Campbell84, Reid05}. NGC~7538 IRS~1--3 form a cluster of embedded massive protostars with luminosities of $\sim 10^4$~L$_{\odot}$\ located 2 arcmin (1.5~pc) southeast of the center of the optical H\,{\tiny II}\ region \citep{Wynn74}. IRS~1, 2 and 3 have luminosities of 8, 5 and $0.6\times10^4$~L$_{\odot}$, respectively and all of them have ultra compact (UC) \ion{H}{ii} regions \citep{Campbell84, Campbell88}.
IRS~1, whose luminosity is equivalent to a O7.5 main sequence star, is surrounded by an UC \ion{H}{ii} region with a double-lobed structure along the North--South direction with a size of 0\asec2 ($\simeq 500$~AU) \citep{Campbell84a, Gaume95}. At these scales, the radio emission is dominated by a collimated, ionized wind that exhibits time variability \citep{Franco04, Sandell09}. This is supported by the extremely broad line widths of radio recombination lines \citep{Gaume95,Keto08}. A bipolar molecular outflow extending in the NW--SE direction is detected in CO \citep[][hereafter QZM11]{Kayema89, davis98, qiu11}. The outflow has created a cavity that is well detected in the near and mid-IR at scales of a few arcseconds north of the protostar \citep{Buizer05, Kraus06}. Despite of the existence of the UC \ion{H}{ii} region, the protostar is probably still actively accreting gas with a high infall rate (mm inverse P--Cygni profiles yield infall rates of $\sim 10^{-3}$~$M_{\odot}$~yr$^{-1}$) and it probably has a massive circumstellar disk \citep[QZM11; ][]{Klaassen09, Sandell09, Beuther12}. This source is associated with a rich variety of masers, most of them arising from the interaction zone of the dense molecular gas with the ionized gas and from the outflow \citep{Rots81,Dickel82, Johnston89, Schilke91, Minier98, Hoffman03, Hutawarakorn03,Pestalozzi06, Galvan10, Surcis11, Hoffman11}.
IRS~2 and IRS~3 are not associated with a dusty envelope and are probably in a more evolved phase. IRS~2 is associated with an O5 star and it is probably the most evolved source. \citet{bloomer98} propose that its stellar winds are shocking the surrounding material, generating a ``stellar wind bow shock'' visible as a shell in H\,{\tiny II}\ and Fe$_{\rm II}$. IRS~3 is associated with an O6--O9 star that might power one or more CO outflows \citep[QZM11;][]{ojha04}.
The dense molecular gas around IRS~1--3 appears to have a filamentary morphology, with an arc--like shape southeast of IRS~1 \citep[QZM11; ][]{Pratap90}. \citet{kawabe92} interpret this dense molecular gas filament as part of an expanding ring--like structure with a radius of 0.25~pc and a mass of 230~$M_{\odot}$, created or piled up by the strong protostellar winds. QZM11 presented a study of the NGC~7538~IRS~1--3 region using the Submillimeter Array (SMA) in the 1.3~mm wave band. They found star--forming cores deeply embedded within the filamentary, dense molecular gas cloud of IRS~1--3 and multiple molecular outflows.
Previous submm polarization observations by \citet{Momose01} were done with single--dish bolometers toward IRS~1--3 and IRS~11 (a younger source, located $1\farcm5$ south of IRS~1). The two sources show striking differences in the polarization properties. Thus, while IRS 11 exhibits an extremely well-ordered magnetic field and a high degrees of polarization, in IRS~1 the field appears locally disturbed, and the degrees of polarization are much lower than those of IRS~11. They interpreted this as an evolutionary effect (more ordered fields are observed in younger sources), which has been also observed in other high mass star forming regions \citep{girart09, girart13,Tang09a, Tang09b}.
Polarimetric observations allow us to study the magnetic fields at the relevant scales (100--$10^4$~AU) where the star formation takes place \citep{Girart99, girart06, Rao09, Rao14, Hull13,liu13,qiu13}. Recently, \citet{zhang14} presented a large sample of massive, clustered, star forming clumps and showed that magnetic fields play an important role during the formation of dense cores at spatial scales of 0.01--0.1~pc. In this paper, we present SMA observations carried out at 345 GHz toward the massive cluster NGC~7538~IRS~1. Section~\ref{sec-obs} briefly describes observations and data reduction. Section~\ref{sec-results} presents the results of the observations. Sections~\ref{sec-analysis} and \ref{sec-disc} contain the analysis and discussion. Finally, in Section~\ref{sec-concl} we draw our main conclusions.
\begin{figure*}
\centering
\includegraphics[width=.7\textwidth,angle=0]{n7538i1-selfcal3.pdf}
\caption{Contour map of the SMA dust continuum emission map towards NGC~7538~IRS~1\ at 878~$\mu$m overlapped with the grey image of the polarized intensity. The blue and red segments show the magnetic field direction obtained from the polarization data with cutoff at rms of 2 and 3--$\sigma$, respectively (see Section~\ref{sec-pol}). Solid contours go from 3--$\sigma$ to 33--$\sigma$ in steps of 3--$\sigma$, where $\sigma=0.017$~Jy~beam$^{-1}$ is the rms noise of dust continuum. The dashed contours around IRS~1 (MM1) go from 50--$\sigma$ to 300--$\sigma$ in steps of 50--$\sigma$. The scale of the polarized intensity image is shown in the right--hand side of the figure (the units are Jy~beam$^{-1}$). The black and white crosses show the peak positions of the dust continuum sources (Table~\ref{tab-cont}). The name of the dust continuum sources are also shown. Red circles mark the positions of IRS~1--3. The solid circle shows the FWHM of the SMA primary beam at the observed frequency. The physical scale of the map and synthesized beam are shown in the bottom left corner of the panel.}
\label{fig-pol}
\end{figure*}
\begin{table*}
\caption{
878~$\mu$m continuum emission parameters.
}
\begin{tabular}{cccccccccc}
\hline
&
\multicolumn{1}{c}{$\alpha$(J2000)} &
\multicolumn{1}{c}{$\delta$(J2000)} &
\multicolumn{1}{c}{$I_\nu^{\textrm{\tiny{Peak}}}$} &
\multicolumn{1}{c}{$S_{\nu}$} &
\multicolumn{1}{c}{$\diameter_{\rm Peak/2}$ $^b$} &
\multicolumn{1}{c}{$T_{\rm dust}$ $^a$} &
\multicolumn{1}{c}{$N_{\rm H_2}$ $^c$} &
\multicolumn{1}{c}{$n_{\rm H_2}$ $^c$} &
\multicolumn{1}{c}{Mass $^c$}
\\
\multicolumn{1}{c}{Source} &
\multicolumn{1}{c}{h:m:s} &
\multicolumn{1}{c}{$^\circ$:$'$:$''$} &
\multicolumn{1}{c}{(Jy~beam$^{-1}$)} &
\multicolumn{1}{c}{(Jy)} &
\multicolumn{1}{c}{(10$^{3}$AU)} &
\multicolumn{1}{c}{(K)} &
\multicolumn{1}{c}{(10$^{23}$cm$^{-2}$)} &
\multicolumn{1}{c}{(10$^{6}$cm$^{-3}$)} &
\multicolumn{1}{c}{($M_{\odot}$)}
\\
\hline
MM1 &23:13:45.349 &61:28:10.47 &5.24 &\valerr{6.07}{0.05} &4.93 &245 &14 &23 &17 \\
MM2 &23:13:44.698 &61:28:12.30 &0.563 &\valerr{1.87}{0.08} &9.38 &40 &8.4 &7.5 &37 \\
MM3 &23:13:46.397 &61:28:14.05 &0.358 &\valerr{0.52}{0.05} &6.04 &40 &5.6 &7.8 &10 \\
MM3b &23:13:45.976 &61:28:13.52 &0.214 &\valerr{0.57}{0.06} &9.54 &40 &2.5 &2.2 &11 \\
MM4 &23:13:43.774 &61:28:11.74 &0.419 &\valerr{1.30}{0.06} &8.36 &40 &7.4 &7.4 &26 \\
MM5 &23:13:45.207 &61:28:16.83 &0.149 &\valerr{0.18}{0.04} &6.28 &40 &1.8 &2.4 &3.5 \\
MM6 &23:13:46.849 &61:28:10.12 &0.251 &\valerr{0.39}{0.05} &6.75 &43 &3.1 &3.8 &7.0 \\
MM7 &23:13:45.140 &61:27:58.44 &0.404 &\valerr{0.97}{0.07} &7.60 &58 &4.3 &4.7 &12 \\
MM7b &23:13:44.622 &61:27:58.30 &0.224 &\valerr{0.37}{0.05} &6.75 &58 &2.1 &2.6 &4.7 \\
MM9 &23:13:43.770 &61:28:02.32 &0.153 &\valerr{0.31}{0.05} &8.36 &54 &1.2 &1.2 &4.2 \\
MM10 &23:13:46.439 &61:28:05.30 &0.229 &\valerr{0.36}{0.05} &6.52 &43 &3.1 &3.9 &6.5 \\
MM11 &23:13:44.421 &61:28:25.76 &0.194 &\valerr{0.50}{0.05} &8.17 &40 &2.9 &3.0 &9.6 \\
MM12 &23:13:46.019 &61:28:01.47 &0.137 &\valerr{0.32}{0.04} &9.54 &43 &1.3 &1.1 &5.6 \\
MM13 &23:13:45.092 &61:28:26.36 &0.105 &\valerr{0.18}{0.04} &8.71 &40 &0.94 &0.90 &3.5 \\
\hline
\end{tabular}
\smallskip\\
$^a$ For the sources detected by QZM11 we use the temperature they use. For the rest of the sources we adopt 40~K.\\
$^b$ Diameter of the circle with area equal to the source area satisfying $I_\nu$$>$$I_\nu^{\rm Peak}/2$\\
$^c$ Assuming $\kappa_{\rm 341.4~GHz}$$=$0.015~cm$^2$~g$^{-1}$ \citep{ossenkopf94}. See Appendix~1 in Frau et al. (2010) for details on the calculation. The uncertainties are approximately 50\% of the computed values. \\
\label{tab-cont}
\end{table*}
\section{Observations and data reduction\label{sec-obs}}
The SMA observations were undertaken on 2005 July 7 with seven antennas in the compact configuration. The weather was good during the observations, with system temperatures in the range of 200--300~K. The phase center used for NGC~7538~IRS~1\ was $\alpha$(J2000.0)$= 23^{\rm h}13^{\rm m}43\fs359$, $\delta$(J2000.0)$= 61\degr 28' 10\farcs60$ \citep{davis98}. A single receiver was used and tuned to a Local Oscillator frequency of 341.6 GHz (878~$\mu$m), with a total bandwidth of 2~GHz per sideband, covering a frequency range of 335.51--337.49~GHz in the lower sideband and about 345.51--347.89~GHz in the upper sideband. At these frequencies, the full width at half maximum (FWHM) of the antennae's primary beam is $32\farcs4$. The 2$\times$2~GHz correlator was configured to sample the aforementioned frequency ranges with a uniform spectral resolution of 0.81 MHz ($\simeq0.7$~km~s$^{-1}$). The standard reduction of the data was done by using the IDL MIR package and selecting the QSO 3C454.3 for the bandpass calibration and BLLAC for the gain calibration. The polarization calibration was performed by observing 3C454.3 over a large parallactic angle range ($-70\degr$ to $+70\degr$). This allowed us to correct for the instrumental polarization using the MIRIAD software package at an accuracy of 0.1\% \citep{Marrone08}. We used the MIRIAD task UVLIN to separate the continuum and the line data in the $u,v$ domain.
The line-free channels were selected by inspecting MM1 because it is the core with the most molecular lines. These channels were used for the final maps, minimizing the possible contamination from molecular line emission. Self--calibration on NGC~7538~IRS~1\ was performed using the Stokes $I$ continuum data for each baseline independently. The derived gain solutions were applied to the molecular line data.
The final maps were obtained using Natural weighting, which yielded a synthesized beam of $2\farcs33\times2\farcs01$ with a position angle of $34\degr$. The continuum map sensitivity is $\sigma_{I}$=0.017~Jy~beam$^{-1}$\ for Stokes~$I$ and $\sigma_{\rm pol}$=0.010~Jy~beam$^{-1}$\ for Stokes~$Q$ and $U$. For the molecular line data, we made channel maps with a spectral resolution of 1.4~km~s$^{-1}$\ that resulted in a sensitivity of $\sigma$=0.25~Jy~beam$^{-1}$\ per channel.
\begin{figure*}
\centering
\includegraphics[width=.9\textwidth]{n7538i1-moments-4-gamma3.pdf}
\caption{Moment maps of molecular transitions with extended emission. Rows: moments 0, 1, and 2, in descending order, labeled on the left-hand side of the figure. Columns: from left to right, C$^{17}$O~3--2, H$^{13}$CO$^{+}$~4--3, SO~8--7, and C$^{34}$S~7--6, respectively, labeled on the top of the figure. Color map: moment maps. Common scale is shown on the right-hand side of the figure. Contours: 3$\sigma$, 9$\sigma$, and 27$\sigma$ continuum emission levels, where $\sigma$=0.017~Jy~beam$^{-1}$.}
\label{fig-momnt}%
\end{figure*}
\begin{figure*}
\centering
\includegraphics[angle=-90,width=.9\textwidth]{spectrum.pdf}
\caption[MM1 spectrum USB 1 of 2]{Spectrum towards the MM1 peak. Top panel: lower side band. Bottom panel: upper side band. A common temperature range from -2 to 14~K is displayed for better visualization. Additional temperature ranges are shown for the upper (-2 to -20~K) and lower (14 to 42~K) sidebands to display the entire spectrum. A dashed lines marks the 0~K level. For both sidebands, the frequencies of the main molecular transitions are noted by a dotted vertical line and the transition is specified.}
\label{fig-spec}
\end{figure*}
\section{Results\label{sec-results}}
\subsection{Continuum emission}
Figure~\ref{fig-pol} presents the 878~$\mu$m continuum SMA map of NGC~7538~IRS~1. A total of fourteen sources are detected, including IRS~1 (MM1 in this figure). MM1 is the brightest source at 878~$\mu$m, with an intensity of 5.24~Jy~beam$^{-1}$. The rest of the continuum sources have intensities in the 0.11--0.56~Jy~beam$^{-1}$\ range. Our mass sensitivity at a 3$\sigma$ level for a typical $T_{\rm dust}=40$~K is $\sim$0.9~$M_{\odot}$~beam$^{-1}$, twice as better as that of QMZ11 (our second contour in Fig.~\ref{fig-pol} roughly compares to their first contour in Fig.~1). Hence, only nine sources were previously detected by QMZ11. MM8, detected by QZM11, is not detected at 878~$\mu$m because it falls outside the FWHM of the primary beam at this wavelength. The 878~$\mu$m position of MM9 is offset by $4\farcs5$ with respect to the value given at 1.3~mm by QZM11. This is probably due to the relative weakness of this source at the two wavelengths and that it is close to the primary beam edge. MM3 and MM7 show a secondary peak at 878~$\mu$m, MM3b and MM7b, which were not detected previously. Both new sources are located westward of the main component and have peak fluxes of $\sim$55\% that of the main component. In addition, four new sources are detected at 878~$\mu$m, named MM10--MM13 following the convention of QZM11. All the dust continuum sources except MM11 and MM13 appear to be located in a diffuse arc--like filamentary structure, resembling a ``spiral arm'' which encloses IRS~1 (hereafter we refer to this filament as ``spiral--arm''). This ``spiral arm'' was first reported by \citet{kawabe92} from CS~2--1 observations. At the center of the complex, MM1 and MM2 sources seem to be closer to each other and embedded in a dense environment. The other sources, located mainly east and south of IRS~1, are embedded in more diffuse medium and form a C-shaped structure that seems to arise from MM1-MM2.
The measured and derived physical parameters of the sources are listed in Table~\ref{tab-cont}. Temperatures are assumed to be those used by QZM11: 245~K for MM1 and 40--58~K for the rest of the sample. For the new sources (MM10--13), we adopted a temperature of 40~K (these sources are weaker and relatively far from IRS~1). The diameter of the sources (FWHM of the dust emission) are relatively homogeneous with an average value of 3\asec0$\pm$0\asec5 ($\simeq 8 \times 10^3$~AU), slightly larger than the beam, and thus are poorly resolved. The densest object is MM1 (\tenpow{2.8}{7}~cm$^{-3}$) but the most massive one is MM2 ($\simeq 37$~$M_{\odot}$). The total mass is $\simeq$160~$M_{\odot}$, where 50~\% is contained in the MM1-MM2-MM4 region ($\simeq 80$~$M_{\odot}$) and 40~\% in the C-shaped filament SE to MM1 ($\simeq 60$~$M_{\odot}$). In addition, the central sources appear to be denser than those in the C-shaped filament, with the exception of MM3.
\subsection{Linearly polarized continuum emission\label{sec-pol}}
The polarized emission is broadly detected along the dust filamentary structure, with a polarized intensity between $\simeq 0.017$ and 0.059~Jy~beam$^{-1}$. Two sets of polarization segments were computed. Firstly, the high--significance set (red segments in Fig.~\ref{fig-pol}) is computed using 3.0--$\sigma_{\rm pol}$ cutoff in the Stokes $Q$ and $U$ maps and 6--$\sigma_{\rm I}$ cutoff in the Stokes $I$ maps. Secondly, the low--significance set (blue segments in Fig.~\ref{fig-pol}) is computed using 2--$\sigma_{\rm pol}$ and 3--$\sigma_{\rm I}$ cutoffs, respectively. The agreement in the magnetic field direction is remarkable between both sets, thus the low cutoff values deliver realistic information (Section~\ref{ssec-houde09}). The overall morphology of the magnetic field segment directions\footnote{The so called magnetic field segments, represent the angle of the line-of-sight (LOS) integrated linearly polarized emission flipped by 90$^\circ$, which is assumed to roughly trace the magnetic field direction.} is not uniform across the region, unlike the ordered directions detected in other regions \citep{girart06,girart09}, and seem to roughly follow the arc--like filament direction as traced by the continuum emission.
\begin{figure*}
\centering
\includegraphics[angle=0,width=.24\textwidth]{data-vs-model-paper-1.pdf}
\hfill
\includegraphics[angle=0,width=.74\textwidth]{data-vs-model-paper-2.pdf}
\caption{Velocity structure of NGC~7538~IRS~1. {\it Contours}: observed dust continuum map, contours are 3$\sigma$ to 21$\sigma$ in steps of 6$\sigma$, where $\sigma$=0.025~Jy~beam$^{-1}$. {\it First column}: order~1 moment maps, i.e. velocity structure. The common scale is shown at the top of the column. {\it Second to fifth columns}: channel maps with the velocity labeled at the top. {\it Rows a and b}: observed H$^{13}$CO$^{+}$~4--3 (a) and C$^{17}$O~3--2 (b) maps. The gray-scale for the channel maps is shown at the right-hand side of the figure. {\it Rows c and d}: synthetic maps generated with \texttt{RATPACKS} (Section~\ref{ssec-toy-model}) for a logarithmic spiral with radial expanding motions (c), and with rotational motions (d).}
\label{fig-toymodel}
\end{figure*}
\subsection{Molecular line emission}
In this section we present the molecular line data towards NGC~7538~IRS~1\ at 878~$\mu$m that can be summarized in two different behaviors: ({\it i}) four molecular transitions that have extended emission arising from the diffuse material, and ({\it ii}) many spatially unresolved hot-core lines only present towards the chemically differentiated MM1 source.
The four molecular transitions that exhibit extended emission are shown in Figure~\ref{fig-momnt}, which presents the zero, first and second order moment maps (integrated intensity, velocity and velocity dispersion maps, respectively) of the emission. The different transitions are ordered by increasing critical density\footnote{$n_{\rm crit}=A/\gamma$ with the Einstein coefficient $A$ and the collisional rate $\gamma$ were taken from LAMBDA \citep{schoier05} for a temperature of 40--50~K.}: $\sim$10$^{5}$~cm$^{-3}$\ for C$^{17}$O~3--2, a few 10$^{6}$~cm$^{-3}$\ for H$^{13}$CO$^{+}$~4--3 and SO~8--7, and $\sim$10$^{7}$~cm$^{-3}$\ for C$^{34}$S~7--6. These moment maps cover the velocity range between $-66$ and $-60$~km~s$^{-1}$, where the emission is detected. The two most extended molecular transitions, C$^{17}$O~3--2 and H$^{13}$CO$^{+}$~4--3 (first and second column, respectively) trace the dust emission with high fidelity. For both transitions, the first order moment maps show an almost equivalent velocity pattern, which strongly suggest that the complex velocity structure is real. MM1-MM2 sources have a velocity around $-64.5$~km~s$^{-1}$, while the rest of the sources seem to be closer to $-61$~km~s$^{-1}$. These features are not so well observed in the C$^{17}$O\ because it does not trace very well MM1 and MM4. The ``spiral arm'' possibly starting in the MM1-MM2 region and ending up in MM9, is very well traced by the C$^{17}$O\ and H$^{13}$CO$^{+}$\ emission. The projected velocity pattern shows an increase in velocity from MM1 to MM6, and then a slight decrease along the filament down to MM7. An interesting feature of this filament is that appears to be a marginal velocity gradient across the filament, with the inner edge of the arc--like structure blueshifted with respect to the outer edge.
The SO~8--7 and C$^{34}$S~7--6 emission (third and fourth column, respectively, of Fig.\ref{fig-momnt}) appears to be less extended than that of the C$^{17}$O\ and H$^{13}$CO$^{+}$\ lines, tracing only partially the dusty arc--like filament. This can be due to their higher critical density, i.e. they trace the densest parts of the filament. Both are present around MM1 (including MM2, MM3, MM3b and MM5) and MM7. The SO line is also detected toward MM6 and MM11. MM1 is particularly bright in these two molecular transition maps and shows a large velocity dispersion of 2--3~km~s$^{-1}$, in contrast with the rest of the filament that hardly achieves a dispersion of 1.5~km~s$^{-1}$. The velocity pattern of the SO and C$^{34}$S\ is compatible with that of the more extended, C$^{17}$O\ and H$^{13}$CO$^{+}$\ lines.
The set of molecules detected only towards MM1 continuum emission peak are all spatially unresolved. Figure~\ref{fig-spec} presents the spectrum towards the peak of MM1 with most of the lines labeled. This spectra is typical of an evolved hot-core (see \S~\ref{sec-disc}): it contains a few transitions of SO, SO$_{2}$, $^{34}$SO$_{2}$, NS, HC$_{3}$N, H$_{2}$CO, and NH$_{2}$CHO; as well as many transitions of CH$_{3}$OH and CH$_{3}$OCHO. The upper energies of these transitions span an order of magnitude, ranging from $\sim$50~K to $\sim$500~K. Several unidentified lines were detected as well.
\begin{table}
\caption{Velocity parameters of the kinematic models}
\begin{center}
\begin{tabular}{c ccccc}
\hline
&
\multicolumn{4}{c}{Velocity}
\\
Fig.~\ref{fig-toymodel}&
Type & $v(r_{\rm out})$ & $r_{\rm out}$ & $\alpha$ & $\theta_{\rm LOS}^{\rm spir}$
\\
(row)&
& km~s$^{-1}$ & AU & & deg\\
\hline
(c) &
Expansion & 9 & \tenpow{1.5}{4} & 0 & 70 \\
(d) &
Rotation & 2 & \tenpow{1.5}{4} & 1 & 45 \\
\hline
\end{tabular}
\end{center}
\label{tab-toymodel}
\end{table}
\section{Analysis\label{sec-analysis}}
\subsection{The kinematics of the ``spiral arm'' around NGC~7538~IRS~1\ \label{ssec-kinetic}}
The velocity pattern found towards the ``spiral arm'' of NGC~7538~IRS~1\ is consistent among the tracers and shows very smooth variations (Fig.~\ref{fig-momnt}). This pattern is present towards the south and east of the UC H\,{\tiny II}\ region and includes MM3, MM3b, MM5, MM6, MM7, MM7b, MM10, and MM12. The total amount of mass in the ``spiral arm'' from dust continuum is $\simeq 60$~$M_{\odot}$.
\subsubsection{Kinematic models\label{ssec-toy-model}}
We used a set of simplified models with different geometries and different velocity structures. For the geometrical structure, we used two-dimensional Archimedean spirals $r = a + b \theta $ and logarithmic spirals $r = a e^{\theta/b} $, both expressed in polar coordinates. A certain thickness was applied to the spiral to form a tubular three-dimensional structure.
For the kinematics, we took into account radial and rotational motions, both with speed following a potential law of the form $v(r)_{rad/rot} = v_{rad/rot}(r_{\rm out}) \left[r/r_{\rm out}\right]^{\alpha}$. At any point in the space, the radial velocity vector is tangential to the radial direction with respect to the source center, while the rotation velocity vector is simultaneously perpendicular to the rotation axis and the radial direction. Positive $v(r_{\rm out})$ mean expansion and counter-clockwise rotation for the cases of radial and rotational motions, respectively. However, changing the signs of the angle with respect to the LOS and of the velocity would produce the same map and, therefore, there is an uncertainty on the direction of the gas flow. The parameter $\alpha$ can be used in radial motions to accelerate or decelerate the gas as a function of the radius. For rotating motions, $\alpha$ can be used to simulate rigid body rotation, constant speed rotation, and keplerian rotation with $\alpha=1$, $\alpha=0$, and $\alpha=-1/2$, respectively.
Finally, we developed a simple RAdiative Transfer Package for Adaptable Construction of Kinematical and Structural models (\texttt{RATPACKS}) to generate synthetic velocity cubes using any combination of geometric and kinematic input as a synthetic source. The synthetic source can be rotated around any axis allowing any orientation in the three-dimensional space. The velocities are projected on the plane-of-the-sky (POS) according to the velocity pattern chosen, and to the three-dimensional orientation given to the synthetic source. Then, a simple radiative transfer routine assuming optically thin emission is used to derive noiseless synthetic channel maps, and order~0, 1, and 2 moment maps.
\subsubsection{Application to NGC~7538~IRS~1\label{sssec-toy-model-ngc7538}}
We explore the parameter space in order to reproduce best the H$^{13}$CO$^{+}$\ and C$^{17}$O\ maps of the entire ``spiral arm''. We adopt a depth of the spiral arm equal to the average observed width: $5''$ (64~mpc). Since we are only interested in the general kinematics, we assume uniform gas density, constant molecular abundance and optically thin emission. In Fig.~\ref{fig-toymodel} we present the models with the best fitted parameters together with the H$^{13}$CO$^{+}$\ and C$^{17}$O\ data for comparison. Logarithmic spirals fit best the data than Archimedean spirals and are the ones presented in Fig.~\ref{fig-toymodel}. For the models presented we used $a=$\tenpow{1.4}{4}~AU and $b=42$~rad. The parameters used for the kinematics are listed in Table~\ref{tab-toymodel}. For both velocity cases we set $v_{\rm LSR}=-63$~km~s$^{-1}$\ and $r_{\rm out}=$\tenpow{1.5}{4}~AU ($\sim$5\asec7). $\theta_{\rm LOS}^{\rm spir}$ represents the angle between the plane containing the spiral and the LOS, where 0$^\circ$\ and 90$^\circ$\ mean edge-on and face-on, respectively.
Row $c$ in Fig.~\ref{fig-toymodel} shows a synthetic source with radial expanding motions, while row $d$ shows counter-clockwise rotating motions. The radial expanding motions seem to reproduce best the channel maps in both molecules. This velocity pattern produces channel maps with extended emission. The best fit is achieved using $\alpha=0$ with a constant radial velocity of 9~km~s$^{-1}$\ and $\theta_{\rm LOS}^{\rm spir}$=70$^\circ$, close to face-on. The rotational pattern produces channel maps with concentrated emission, not seen in the data. The best fit is achieved with a rigid-body rotation of 2~km~s$^{-1}$\ at $r_{\rm out}$ and $\theta_{\rm LOS}^{\rm spir}$=45$^\circ$. The values of the velocities depend strongly on the orientation angles of the synthetic source and, therefore, velocities should be taken as upper limits.
\subsection{Statistical derivation of the magnetic field strength\label{ssec-houde09}}
\begin{figure*}[ht]
\centering
\includegraphics[angle=270,width=.9\textwidth]{houde09-n7538i1-split.pdf}
\caption{Angular dispersion function of the magnetic field segments detected towards the spiral arm (panels a and b) and the central region (panels c and d). {\it Top sub-panels (a and c):} dots represent the data with uncertainty bars, dashed line marks the zero value, dotted vertical line notes the beam size, dotted horizontal line shows the expected value for a randomic magnetic field, red line shows the best fit to the large scale magnetic field (summation in Eq.~\ref{eq-houde09}), and blue line shows the best fit to the data (Eq.~\ref{eq-houde09}). {\it Bottom sub-panels (b and d):} dots represent the correlated component of the best fit to the data, dashed line marks the zero value, dotted vertical line notes the beam size, red line shows the correlation due to the beam, and blue line shows the correlation due to the beam and the turbulent component of the magnetic field. \label{fig-houde09}}
\end{figure*}
\subsubsection{Formalism}
Based on observational data, a widely used method to estimate the magnetic field strength of the POS component of the large-scale magnetic field is the Chandrasekhar-Fermi \citep[hereafter CF;][]{chandrasekhar53} equation
\begin{equation}
\frac{\delta B}{B_0} \simeq \frac{\sigma_v}{V_{\rm A}},
\label{eq-chand-fermi}
\end{equation}
\noindent where $B_0=\left|\vec{B}_0\right|$ is the large-scale component of the magnetic field, $\delta B$ is the variation about $B_0$, $V_{\rm A}$=$B_0/\sqrt{4\pi\rho}$ is the Alfv\'en speed at density $\rho$, and $\sigma_{\rm v}$ is the velocity dispersion of an appropriate spectral line. Recently, different statistical methods have been developed to avoid some of the CF method caveats \citep{hildebrand09,houde09,houde11,koch10}. These methods rely on the study of the extended polarized emission in observational maps.
\begin{comment}
One can define the difference between the angles of the polarization segments of two positions separated by a displacement $\vec{l}$ as
\begin{equation}
\Delta\Phi \left(\,l\,\right) angular dispersion function
\equiv
\Phi\left(\,\vec{x}\,\right)-\Phi\left(\,\vec{x}+\vec{l}\,\right) ,
\end{equation}
\noindent where $\Phi(\vec{x})$ is the angle of the polarization segment at position $\vec{x}$. Then, the structure function (SF) of second order can be defined as
\begin{equation}
\langle \Delta\Phi^2 \left(\,l\,\right) \rangle
\equiv
\frac{1}{N\left(\,l\,\right)}
\sum^{N\,\left(l\,\right)}_{i=1}
\left[\Phi\left(\,\vec{x}\,\right)-\Phi\left(\,\vec{x}+\vec{l}\,\right)\right]^2,
\label{eq-sf}
\end{equation}
\noindent where $\langle\ldots\rangle$ denotes an average, $l=|\vec{l}|$, and $N(l)$ is the number of pairs of segments separated by a displacement $\vec{l}$.
\end{comment}
\citet{houde09} assume two statistically independent components of $\vec{B}$, the large-scale magnetic field $\vec{B}_0(\vec{x})$, and the turbulent magnetic field $\vec{B}_{\rm t}(\vec{x})$. Then, they derive the turbulent to large-scale magnetic field strength ratio from the angular dispersion function that accounts for the polarization angle differences as a function of the distance between the measured positions. The analysis is based in an analytical derivation for a turbulent cloud (see their Eq.~4) including the effect of beam and LOS averaging. They further assume a stationary, homogeneous, and isotropic magnetic field strength, an isotropic but negligible turbulent polarized emission, and a magnetic field turbulent correlation length $\delta$ much smaller than the thickness of the cloud $\Delta'$ ($\delta\ll\Delta'$). Applying all these simplifications, the angular dispersion function can be written as
\begin{eqnarray}
1-\langle \cos\left[\Delta\Phi\left(\,l\,\right)\right] \rangle
&\simeq&
\frac{\langle B^2_{\rm t} \rangle}{\langle B^2_0 \rangle}
\,\frac{1}{N}\,
\left[1-e^{-l^2/2(\delta^2+2W^2)}\right] \nonumber \\
&&+\sum^{\infty}_{j=1}a'_{2j}\,l^{\,2j} ,
\label{eq-houde09}
\end{eqnarray}
\noindent where
\begin{equation}
N=\left[\frac{(\delta^2+2\,W^2)\Delta'}{\sqrt{2\,\pi}\,\delta^3}\right]
\label{eq-nturb}
\end{equation}
\noindent is the number of independent turbulent cells along the LOS, $W$ is the standard deviation ($\sigma={\rm FWHM}/\sqrt{8\ln{2}}$) of the gaussian beam, and the summation is a Taylor expansion representing the large-scale magnetic field component that does not involve turbulence. The first term in the square brackets contains the integrated turbulent magnetic field contribution, while the exponential term represents the correlation by the combined effect of the beam ($W$) and the turbulent magnetic field ($\delta$). The intercept of the fit to the data of the non-correlated part at $l=0$, $f_{{\rm NC}}(0)$, together with the assumption of a cloud thickness $\Delta'$, allow us to estimate the turbulent to large-scale magnetic field strength ratio as
\begin{equation}
\frac{\langle B^2_{\rm t} \rangle}{\langle B^2_0 \rangle}
=
N\,f_{{\rm NC}}(0).
\label{eq-btb0}
\end{equation}
\noindent Finally, identifying $\langle B^2_{\rm t} \rangle \equiv \delta B^2$, one can apply the CF equation (Eq.~\ref{eq-chand-fermi}) to derive the large-scale component of the magnetic field as
\begin{equation}
\langle B^2_0 \rangle^{1/2}=
\sqrt{4\,\pi\,\rho}\,\sigma_v\,
\left[\frac{\langle B^2_{\rm t} \rangle}{\langle B^2_0 \rangle}\right]^{-1/2}.
\label{eq-b0}
\end{equation}
\noindent We note that the magnetic field component labeled as ``turbulent'' describes, more generally, any contribution to the total magnetic field other than the uniform large-scale one. Therefore, when we refer in the next sections to the ``turbulent'' magnetic field we are discussing the non-uniform magnetic field contribution.
\begin{table}
\caption{Derived magnetic field strength~$^a$}
\begin{center}
\begin{tabular}{ l cc}
\hline
&Spiral~$^{b}$ & Central~$^{c}$ \\
\hline
\smallskip
$\delta$ (\arcsec, mpc) & 2.6$\pm$0.3 (33$\pm$4) & 1.0$\pm$0.6 (13$\pm$8) \\
$f_{{\rm NC}}(0)$ & 0.61$\pm$0.04 & 0.356$\pm$0.017 \\
$a'_2$ (\arcsec$^{-2}$) & \tenpow{(-8.5$\pm$1.3)}{-4} & -- \\
$N$ & 0.9$\pm$0.3 & 6$\pm$8 \\
$\langle B^2_{\rm t} \rangle / \langle B^2_0 \rangle$ & 1.47$\pm$0.16 & 3.2$\pm$2.8 \\
$\langle B^2_0 \rangle^{1/2}$~$^{c}$ (mG) & 2.64$\pm$0.14 & 2.3$\pm$1.0 \\
\hline
\end{tabular}
\end{center}
$^a$ Following \citet{houde09} method. We assumed $\Delta'$=5\arcsec$\pm$1\arcsec\ (64$\pm$13~mpc), roughly the width of the filament and the central structure. \\
$^b$ Assuming $\rho$=4$\times$10$^6$~cm$^{-3}$\ and $\sigma_{v}$=2.3~km~s$^{-1}$\ (H$^{13}$CO$^{+}$~4--3).\\
$^c$ Assuming $\rho$=8$\times$10$^6$~cm$^{-3}$\ and $\sigma_{v}$=2.2~km~s$^{-1}$\ (H$^{13}$CO$^{+}$~4--3).
\label{tab-houde09}
\end{table}
\subsubsection{Application to NGC~7538~IRS~1\label{sssec-houde09-ngc7538}}
The magnetic field segments in this complex region do not follow a defined homogeneous pattern along the observed field (see Fig.~\ref{fig-pol}) such as, e.g., the hour-glass shape reported and modeled in simpler sources \citep{girart06,girart09,frau11,padovani12}. No analytical models are available that can be compared to this complex source. Thus, in order to extract physical information, a statistical approach seems to be the best strategy. The ``spiral arm'' and the central sources seem to have different kinematics and different directional patterns of the segments, probably related to the YSO embedded in the central region (Sec.~\ref{ssec-kinetic} and Fig.~\ref{fig-pol}). Consequently, we analyze the magnetic field for each of the regions independently. Figure~\ref{fig-houde09} shows the angular dispersion function for both structures. Bins are equally spaced by 1\arcsec. Data points represent the mean within the bin, with uncertainties that are smaller than the point size. We used the nonlinear least-squares Marquardt-Levenberg algorithm to fit Eq.~\ref{eq-houde09} to the data. The best fit is shown in Fig.~\ref{fig-houde09} and the parameters are listed in Table~\ref{tab-houde09}.
Panels (a) and (b) show the results for the spiral arm. The uncorrelated large scale component is fitted with a $j=1$ polynomial following Eq.~\ref{eq-houde09}. The correlated component dominates at small distances ($\sim$5\arcsec--\,6\arcsec\ or $\sim$64\,--\,77~mpc at 2.65~kpc). The turbulent magnetic field effect in correlating the segments is significantly more important than the beam effect. The correlation length is $\delta$=2\asec6$\pm$0\asec3 (33$\pm$4~mpc at 2.65~kpc), almost three times larger than the beam correlation distance.
Panels (c) and (d) show the results for the central sources. The correlated component is only important in a distance half that of the spiral arm, and is mostly due to the effect of the beam. For larger distances, the data flattens to a value compatible with a random magnetic field \citep[$(1-\cos(\simeq52^\circ)\simeq0.384$:][see also \citealt{girart13}]{poidevin10}. Therefore, the summation in Eq.~\ref{eq-houde09} (large-scale magnetic field) was dropped in our analysis and only the correlated component was used (see Table~\ref{tab-houde09}). The best fit leads to a turbulent magnetic field correlation length of $\delta$=1\asec0$\pm$0\asec6 (13$\pm$8~mpc at 2.65~kpc), roughly equal to $W\sim$0\asec92 (the telescope beam correlation length, Eq.~\ref{eq-houde09}).
To estimate the magnetic field strength, one has to assume either a certain cloud thickness $\Delta'$ or a certain number of turbulent cells $N$ (see Eq.~\ref{eq-nturb}). Since the magnetic field that we are tracing mostly comes from a filamentary structure, an educated guess is to assume that the thickness is that of the filament width, $5''$ (64~mpc). Under this assumption, we found that the spiral arm contains one turbulent cell along the LOS, while the result for the central sources is poorly constrained to 6$\pm$8 cells. In both cases, the local turbulent field is more important than the large scale, ordered field. However, the central sources have a magnetic turbulence two times more important than the spiral arm. The strength of the field is $\sim$2.5~mG, similar for both regions.
\begin{figure}
\centering
\includegraphics[angle=0,width=8.5cm]{pol-dust-histo-statIndep.pdf}
\caption{Comparison of the dusty filament orientation to the magnetic field segments orientation towards NGC~7538~IRS~1. Pixels have been resized to the beam size to ensure statistical independence (see Fig.~\ref{fig-pol} for Nyquist sampling). {\it Top panel}: contours are 3, 9, and 27 times 0.02~Jy~beam$^{-1}$ continuum emission levels. Blue segments are derived as for Fig.\ref{fig-pol}. Red thick line is the axis of the filament (see Section~\ref{ssec-dust-vs-pol}). Red segments show the orientation of the filament corresponding to each pixel with polarization detection. {\it Bottom panel}: histogram of the angle difference between the polarization segment and the filament axis for each pixel. The correlated and uncorrelated distributions used for the $\chi^{2}$ test in Section~\ref{ssec-dust-vs-pol} are shown as dotted and dashed lines, respectively.}
\label{fig-b-dust-ang}
\end{figure}
\subsection{Comparison of dust and magnetic field structures\label{ssec-dust-vs-pol}}
The magnetic field segments seem to roughly follow the direction of the ``spiral arm''. To examine this, we defined the dusty ``spiral arm'' axis as the line connecting the dusty sources (nodes) within the structure (shown in Fig.~\ref{fig-b-dust-ang} as a thick red line). Then, we obtained new maps with pixels of the size of the beam to ensure statistical independence. Each independent pixel with polarized emission was assigned a segment representing the local direction of the filament. This direction was defined as the line connecting the two nearest nodes (red vectors in Fig.~\ref{fig-b-dust-ang}). Finally, the difference between the magnetic field segment direction and the filament local direction was computed and binned. As shown in the histogram in Fig.~\ref{fig-b-dust-ang}, nine of the fourteen segments ($\sim$64\%) have differences of less than 20$^\circ$\ and none have an angular difference larger than 50$^\circ$. The number of independent measurement is relatively small, therefore, we performed a $\chi^{2}$ test to assess the statistical significance of the results. On the one hand, we compared the data to a flat distribution representing uncorrelated orientations. We found a $<1$\% probability to obtain larger $\chi^{2}$ values with random data, hence the null hypothesis of uncorrelated orientations was rejected. On the other hand, we compared the data to a simple distribution representing correlated orientations. We simplified this distribution to a linearly decreasing function that evolves from total correlation to none in half of the angular range covered (see Fig.~\ref{fig-b-dust-ang}). We found 98\% probability to obtain a larger $\chi^{2}$ value with random data, well above the standard 5\% rejection threshold of the $\chi^{2}$ test, hence the null hypothesis of correlated orientations cannot be rejected. Based on this analysis, we conclude that the orientations of filament and polarization segments are correlated.
\subsection{Energy state of the individual sources: the ``mass balance''\label{ssec-energy-sources}}
We analyze the main causes that are in interplay: gravity, magnetic field, thermal pressure and internal dynamics. On the one hand, gravity has the effect of bringing mass together. On the other hand, the rest of the causes exert the opposite effect either stopping the mass from accreting or dispersing it. We compute a series of meaningful parameters that relate these physical quantities. We also compare the relative strength of these causes in terms of the mass supported against gravitational collapse.
\subsubsection{Formulae}
We use the viral theorem to check whether the different cores are gravitationally bounded, and to estimate the maximum mass supported by the thermal and non-thermal motions. These take into account the different internal pressure components. In section~\ref{sec-disc}, we discuss the effect of the external pressure. The virial mass for the thermal component, $M_{\rm T}$, is
\begin{equation}
M_{\rm T}=\frac{k~c_{s}^{2}~R}{G},
\end{equation}
where $c_{s}$, $R$ and $G$ are the sound speed, the core radius and the gravitational constant, respectively. The parameter $k$ takes into account the specific density distribution of the core. We use $k=1$, which is the value for a density profile $\rho\propto r^{-2}$ \citep{maclaren88}. Similarly, for the non-thermal component the viral mass term, $M_{\rm NT}$, is:
\begin{equation}
M_{\rm NT}=\frac{k~\sigma_{\rm NT}^{2}~R}{G},
\end{equation}
where $\sigma_{\rm NT}$ is the full three-dimensional velocity dispersion of the gas due to non-thermal motions.
The support of the magnetic field can be included as an additional component in the virial mass. Thus, the mass for a critical mass-to-flux ratio is given by \citet{nakano78}:
\begin{equation}
M_{\rm mag}=\frac{\pi R^2 B}{\sqrt{4~\pi^{2}~G}}.
\label{eq-m-mag}
\end{equation}
where B is the field strength.
The non-thermal kinetic energy can be compared to thermal kinetic energy by the squared of the turbulent Mach number
\begin{equation}
\mathcal{M}_{s}^{2}=\left(\frac{\sigma_{\rm NT}}{c_{s}}\right)^{2}.
\label{eq-mach}
\end{equation}
$\mathcal{M}_{s}>1$ means that non-thermal motions are supersonic, and hence, more dynamically important than thermal motions. The thermal energy is compared to magnetic field energy by the plasma $\beta_{\rm T}$
\begin{equation}
\beta_{\rm T}=\frac{P_{\rm therm}}{P_{\rm mag}}=2\left(\frac{c_{s}}{v_{\rm A}}\right)^{2}.
\label{eq-beta}
\end{equation}
where $v_{\rm A}=B_{\rm 3D}/\sqrt{4~\pi~\rho}$ is the Alfv\'en speed. Similarly, magnetic fields compare to non-thermal motions by
\begin{equation}
\beta_{\rm NT}=\frac{P_{\rm NT}}{P_{\rm mag}}=2\left(\frac{\sigma_{\rm NT}^{\rm mol}}{v_{\rm A}}\right)^{2}.
\label{eq-beta-turb}
\end{equation}
$\beta_{\rm T}<1$ or $\beta_{\rm NT}<1$ imply that magnetic pressure overcomes thermal or kinetic pressure, respectively.
\begin{figure*}
\centering
\includegraphics[angle=270,width=\textwidth]{mass-4.pdf}
\caption{``Mass balance'' analysis. Comparison of the measured mass to the maximum supported mass by different forces. Cores are ordered according to their location in the central massive structure, in the filament, or isolated. {\it Black}: measured mass from continuum maps. {\it Gray}: Mass of the star embedded in the MM1 clump. {\it Red}: mass supported by magnetic fields assuming a uniform value across the source. {\it Light and dark blue}: mass supported by virialized gas motions due to internal dynamics and thermal dispersion, respectively.}
\label{fig-massBalance}
\end{figure*}
\begin{table*}
\caption{
Relative energy indicators and supported masses.
}
\begin{center}
\begin{tabular}{l l ccc ccc cccc}
\hline
\multicolumn{1}{c}{Structure} &
\multicolumn{1}{c}{Source} &
\multicolumn{1}{c}{$c_{s}$} &
\multicolumn{1}{c}{$\sigma_{\rm NT}$} &
\multicolumn{1}{c}{$v_{\rm A}$} &
\multicolumn{1}{c}{$\mathcal{M}_{s}^{2}$} &
\multicolumn{1}{c}{$\beta$} &
\multicolumn{1}{c}{$\beta_{\rm NT}$} &
\multicolumn{1}{c}{$M_{\rm T}^a$} &
\multicolumn{1}{c}{$M_{\rm NT}^b$} &
\multicolumn{1}{c}{$M_{\rm mag}$} &
\multicolumn{1}{c}{$M_{\rm obs}$}
\\
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{km~s$^{-1}$} &
\multicolumn{1}{c}{km~s$^{-1}$} &
\multicolumn{1}{c}{km~s$^{-1}$} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{$M_{\odot}$} &
\multicolumn{1}{c}{$M_{\odot}$} &
\multicolumn{1}{c}{$M_{\odot}$} &
\multicolumn{1}{c}{$M_{\odot}$}
\\
\hline
\multirow{3}{*}{Central}
&MM1
&1.20& 0.67& 0.96& 0.31& 3.14& 0.98& 12.05& 3.77& 3.10& 16.56 \\
&MM2
&0.49& 0.94& 1.69& 3.76& 0.16& 0.62& 3.75& 14.07& 11.24& 36.70 \\
&MM4
&0.49& 1.04& 1.71& 4.58& 0.16& 0.74& 3.34& 15.29& 8.92& 25.54 \\
\hline
\multirow{8}{*}{Filament}
&MM5
&0.49& 0.68& 3.47& 1.98& 0.04& 0.08& 2.51& 4.96& 5.78& 3.46 \\
&MM3b
&0.49& 0.99& 3.63& 4.18& 0.04& 0.15& 3.81& 15.91& 13.34& 11.05 \\
&MM3
&0.49& 0.83& 1.91& 2.95& 0.13& 0.38& 2.41& 7.10& 5.34& 10.11 \\
&MM6
&0.50& 0.95& 2.73& 3.52& 0.07& 0.24& 2.90& 10.21& 6.70& 6.97 \\
&MM10
&0.50& 0.93& 2.69& 3.42& 0.07& 0.24& 2.80& 9.58& 6.23& 6.46 \\
&MM12
&0.50& 1.23& 5.09& 5.98& 0.02& 0.12& 4.10& 24.50& 13.35& 5.64 \\
&MM7
&0.59& 1.27& 2.46& 4.70& 0.11& 0.53& 4.40& 20.65& 8.46& 12.15 \\
&MM7b
&0.59& 0.39& 3.33& 0.45& 0.06& 0.03& 3.91& 1.76& 6.70& 4.67 \\
\hline
\multirow{3}{*}{Isolated}
&MM9
&0.56& 0.48& 4.50& 0.72& 0.03& 0.02& 4.50& 3.26& 9.56& 4.22 \\
&MM11
&0.49& 0.23& 2.89& 0.22& 0.06& 0.01& 3.26& 0.70& 9.16& 9.62 \\
&MM13
&0.49& 0.62& 5.27& 1.64& 0.02& 0.03& 3.48& 5.71& 10.41& 3.50 \\
\hline
\end{tabular}
\end{center}
\smallskip
$^a$ To compute the sound speed, $c_{s}=\sqrt{\gamma~k_{\rm B}~T/\mu~m_{\rm H}}$,
we assume an idealized equation of state with adiabatic index $\gamma = 5/3$
\citep{tomida10}, a mean molecular weight of $\simeq2.33$, and the temperature
of the cores estimated by \citet{qiu11}.
\\
$^b$ The 3-D non thermal velocity dispersion is
$\sigma_{\rm NT}^2 = (\sigma_{obs}({\rm mol})^2 - \sigma_{\rm T}({\rm mol})^{2})$,
where $\sigma_{\rm obs}$ is the observed velocity dispersion
($\sigma_{\rm obs} = {\rm FWHM} / \sqrt{8~\ln{2}}$) and $\sigma_{\rm T}$ is
the thermal line broadening for a molecule of mass $m_{\rm mol}$
($\sigma_{\rm T}({\rm mol})=\sqrt{\gamma k_{\rm B}~T/m_{\rm mol}}$).
\label{tab-energy}
\end{table*}
\subsubsection{Energy ratios in NGC~7538~IRS~1\ sources}
The parameters described in the previous section that measure the energy balance among forces are listed in Table~\ref{tab-energy}. Sound speed ranges from 0.49~km~s$^{-1}$\ to 0.59~km~s$^{-1}$\ except for MM1 that hosts an O7.5 star and is significantly warmer. Non-thermal velocity dispersion $\sigma_{\rm NT}$ range from 0.23~km~s$^{-1}$\ to 1.27~kms. The mean value is $0.8\pm0.3$~km~s$^{-1}$, $\sim$60\% larger that the typical sound speed at 40~K. As a result, 70\% of the sources show supersonic gas motions $\mathcal{M}_{s}>1$.
We use for each source the magnetic field strength derived in Section~\ref{sssec-houde09-ngc7538}, 2.64~mG and 2.3~mG for the spiral arm and the central sources, respectively. For the isolated cores we used the average of 2.5~mG. This assumption implies that, within each region, the variation of the derived magnetic quantities depends on clump properties: $v_{\rm A}\propto n_{\rm H_{2}}^{-1}$ and $M_{\rm mag}\propto r^{2}$, where $r$ is the clump radius derived as half the diameter from Table~\ref{tab-cont} (see discussion in Section~\ref{ssec-disc-magfields}). The Alfv\'en speed ranges from 0.96~km~s$^{-1}$\ to 5.3~km~s$^{-1}$, with mean value of $3.0\pm1.1$~km~s$^{-1}$. All sources but MM1 have $v_{\rm A}>c_{s}$, and hence, magnetic pressure dominates locally over thermal pressure ($\beta<1$). Non-thermal kinetic energy is comparable to the magnetic energy in four sources: MM1, MM2, MM4, and MM7. For the rest of the sample, magnetic pressure dominates locally over non-thermal pressure.
\subsubsection{``Mass balance'' in NGC~7538~IRS~1\ sources}
A similar analysis can be performed in terms of the maximum mass supported by each force, listed in Table~\ref{tab-energy}. The ``mass balance'' accounts for all the available information at once. MM1 hosts an embedded O7.5 star whose mass is taken into account \citep[30~$M_{\odot}$:][]{pestalozzi04}. In Fig.~\ref{fig-massBalance}, we compare collapse forces versus support forces to derive the individual ``mass balance''. This analysis shows clearly two groups of sources in terms of stability, well correlated with their location in either ({\it i}) the central structure, or ({\it ii}) the spiral arm and isolated.
The sources in the central structure seem have more mass than that which can be supported: MM1 and MM2 have masses larger than the combined maximum supported mass, while in MM4 masses are comparable. In contrast, all sources located either at the filament or in isolation have virial masses larger than measured masses. In general, the main agent against gravity for sources at the filament is the internal dynamics, while for the isolated sources is the magnetic field. See Section~\ref{ssec-collapse} for a discussion on the implications.
\section{Discussion\label{sec-disc}}
NGC~7538~IRS~1--3 as a whole presents a complex and rich structure, velocity field, and magnetic field. In addition, a number of individual cores can be identified. In this section, we first discuss the global properties, and then proceed to analyze the state of the individual cores.
\subsection{Magnetic field properties\label{ssec-disc-magfields}}
NGC~7538~IRS~1--3 contains two different regions in terms of magnetic field properties as shown in Section~\ref{ssec-houde09}. These differences are important in two different but related aspects: the relative importance of non-ordered with respect to ordered magnetic fields, and the relative dynamical importance of magnetic fields in the overall picture.
For the central sources, the magnetic field segments are only correlated at distances slightly larger than the correlation distance of the beam. The turbulence of the magnetic field has a mild effect, but the data does not allow us to constrain accurately the contribution. In any case, the transitions to values for the angular dispersion function compatible to a random field happens at a very short distance ($\sim$3\arcsec, $\sim$39~mpc). The number of turbulent cells along the LOS seems to be large, which is a clue that the field is severely distorted. This suggests that non-ordered magnetic fields are more important than ordered magnetic fields. In fact, the ratio of energies $\langle B^2_{\rm t} \rangle / \langle B^2_0 \rangle$ is $\gsim$3. The size of the sources is larger than the correlation distance in all cases, and thus, this region needs to be analyzed source by source. The study of the energetic state of the individual sources suggests that non-thermal motions are dominant. In addition, the magnetic field tends to a random configuration at relatively short scales. These facts suggest that magnetic fields are not important in the overall picture. In summary, the relevance of the magnetic field is small in this region, and the field is mostly non-ordered.
The spiral arm shows different magnetic field properties with respect to the central structure. The correlation length due to the non-ordered magnetic field is accurately determined since it is $\sim$3 times larger than the beam correlation length. The distance between consecutive embedded sources within the filament of 2\arcsec\ (26~mpc) in average, smaller than the magnetic correlation length of 2\asec6 (33~mpc). This implies that the field properties among cores are not independent. Consequently, the filament analysis must take into account the whole complex. According to the statistical analysis, the spiral arm contains only one turbulent cell along the LOS, hence, it has relatively little turbulence. This is in agreement with the well aligned segments observed, and with the sources having non-independent magnetic fields. Finally, the field shows a $\propto l\,^2$ trend at large scales that suggests a smoothly varying field \citep{hildebrand09}. Moreover, the analysis in Section~\ref{ssec-dust-vs-pol} shows that the magnetic field direction is correlated to the dust morphology along the entire 11\arcsec-long filament. All these suggest that the magnetic field has a strong internal coherence within the filament, and that it is somehow tied to the dust structure.
\begin{figure*}
\centering
\includegraphics[angle=0,width=.49\textwidth]{n7538i1-selfcal3-co-simple.pdf}
\hfill
\includegraphics[angle=0,width=.45\textwidth]{ngc7538i1-cartoon-2.pdf}
\caption{{\it Left panel}: overlayed contours for dust continuum (gray, this work), and blue-- and red--shifted $^{13}$CO~2--1 outflow (blue and red, QZM11). Crosses mark the positions for dusty cores and red dots for IR sources. {\it Right panel}: schematic 3D cartoon of the proposed scenario for the NGC~7538~IRS~1--3 complex (Section~\ref{ssec-feedback}). Yellow sphere represents MM1, blue and red cones represent the blue and red outflow lobes, and the gray structure represents the spiral arm. Faded colors represent the structures behind MM1 in the LOS direction.}
\label{fig-dust-outflow}
\end{figure*}
\subsection{Energetics of the spiral arm}\label{ssec-dynamics}
The analysis of the kinematics in Section~\ref{ssec-kinetic} suggests that the spiral arm around NGC~7538~IRS~1\ is expanding, although we cannot discard a certain contribution from rotation. This result was already suggested based on CS observations by \citet{kawabe92}.
For the expansion to happen the filament must be gravitationally unbound to the total mass around IRS~1. The combined mass of MM1, MM2, MM4, plus the star embedded in MM1, is 110~$M_{\odot}$. The distance of the filament with respect to IRS~1 is $\sim$13\arcsec\ (0.17~pc at 2.65~kpc). For these quantities, the virialized rotation velocity of the filament is 1.7~km~s$^{-1}$, and the escape velocity is 2.4~km~s$^{-1}$. The velocity difference of the filament with respect to the central region is in the 1.5--4.5~km~s$^{-1}$\ range when projected in the POS, or 9~km~s$^{-1}$\ according to the best fitting kinematic model. Therefore, the spiral arm appears to be gravitationally unbound with respect to the massive MM1, MM2, and MM4 cores. We note that the measured mass is a lower limit due to the filtering effect of the SMA. However, it is required a mass of \tenpow{1.6}{3}~$M_{\odot}$\ within the central $\sim$10\arcsec\ to gravitationally bound the gas moving at 9~km~s$^{-1}$. The single-dish measured mass of the entire 1\arcmin\ clump is \tenpow{3.7}{3}~$M_{\odot}$\ \citep[corrected for the different $\kappa_{\rm dust}$ and distance used]{Momose01}. It seems unlikely that 40\% of the total mass is accumulated whitin the central 10\arcsec, and hence, it is unlikely that the spiral--arm is gravitationally bound.
Focusing on the filament, the total mass combining MM3, MM6, MM7, MM7b, MM10, and MM12 is 45.8~$M_{\odot}$. Hence, the total gravitational energy of the filament using a radius of 13\arcsec\ is $E_{\rm grav}$=\tenpow{2.6}{45}~erg. In addition, we can derive relevant dynamical parameters from the kinematic model. We assumed for the calculations that the expansion velocity is constant. Then, we considered two different scenarios: ({\it i}) a conservative one where the expansion velocity is the maximum velocity projected in the POS, 4.5~kms, and ({\it ii}) that drawn by the kinematic model that takes into account projection effects and results in a faster velocity, 9~km~s$^{-1}$. The conservative scenario delivers an age of $t_{\rm dyn}$$\simeq$\tenpow{3.6}{4}~yr and kinetic energy of $E_{\rm kin}$$\simeq$\tenpow{9.2}{45}~erg, while the deprojected scenario delivers $t_{\rm dyn}$$\simeq$\tenpow{1.8}{4}~yr and $E_{\rm kin}$$\simeq$\tenpow{3.7}{46}~erg. The gravitational well is a factor of 3.5--14 smaller, thus confirming the plausibility of an expanding filament.
To complete the picture, a crude estimate of the magnetic field energy can be done by multiplying the volume of mass permeated by the field times the overall magnetic pressure ($P_{\rm B}=B_{0}^{2}/(8\pi)$). An approximate area of 150\arcsec$^{2}$ and depth of 5\arcsec\ delivers a magnetic energy of $E_{\rm mag}$=\tenpow{1.3}{45}~erg, negligible for the overall filament dynamics when compared to kinetic energy.
The correlation of the field morphology to the dust morphology can now be tentatively explained: the 3--9 times more energetic expansion motions seem to push away both matter and magnetic field, shaping them in a similar morphology as shown in Fig.~\ref{fig-b-dust-ang}.
\subsection{Formation of the spiral arm through stellar feedback\label{ssec-feedback}}
The large scale expansion motions proposed request a powerful driving source. A plausible origin is the IRS~1 feedback, and in particular from the powerful NW-SW molecular outflow powered by IRS~1 (QZM11). We show in Fig.~\ref{fig-dust-outflow} a comparison of the $^{13}$CO outflow from MM1 to our continuum map. Projected on the POS, it seems that the outflow is perpendicular to the spiral arm, suggesting that it could be formed by swept material.
Quantitatively, QZM11 derive $E_{\rm outflow} = $\tenpow{4.9}{46}~erg, which is a factor of 1.2--3.7 larger than the combined kinetic, magnetic, and gravitational energy of the filament. In addition, energy losses are expected in the form of turbulence in this complex scenario. The average non-thermal velocity dispersion is $\Delta v^{\rm NT}$=1.8~km~s$^{-1}$, which implies an extra \tenpow{1.5}{45}~erg that the outflow can provide. Therefore, it is feasible that the molecular outflow is the energy source responsible for the expansion of the spiral arm. Furthermore, the outflow momentum for one lobe, \tenpow{2.3}{2}~$M_{\odot}$~km~s$^{-1}$, is comparable to the momentum of the spiral arm, $\simeq$\tenpow{2}{2}~$M_{\odot}$~km~s$^{-1}$, suggesting that the outflow is the unique cause to set the filament into motion. Finally, the outflow dynamical timescale of \tenpow{2}{4}~yr falls within the filament age range if expanding, suggesting that the two structures have contemporary births.
We can take into account the inclination effects for a better determination of the outflow effect. Based on our kinematic model, we found that the plane containing the logarithmic spiral lies almost parallel to the POS, tilted by $\sim$20$^\circ$\ (70$^\circ$ with respect to the LOS, see Table~\ref{tab-toymodel}). This information may help to understand the three-dimensional orientation of the system. One possibility is that outflow and spiral arm flow are parallel, and thus, the gas forming the spiral arm would be pushed by the outflow end. The opposite possibility is that the spiral arm is perpendicular to the outflow, implying that the material would be blown away from the cavity \citep[$\sim$80$^\circ$\ wide: ][]{Kraus06} and driven to the equatorial plane of the system. The aligned scenario implies an outflow lying almost on the POS, with an increase on de-projected outflow velocity by a factor of $\sim$3 and on $E_{\rm outflow}$ by a factor of $\sim$9, while the perpendicular scenario renders an outflow inclined by 20$^\circ$\ with respect to the LOS, and a mild increase of $\sim$6\% in outflow velocity and $\sim$13\% in $E_{\rm outflow}$. A possibly precessing outflow axis may increase the uncertainty in our analysis \citep{Kraus06}. Also, we might consider configurations between the parallel and perpendicular scenarios. The perpendicular case would imply that both structures have opposite velocity directions when projected on the POS. Therefore, the fact that the spiral arm is mostly red-shifted towards the same direction as the large scale outflow seems to favor the aligned scenario. There are other cases in the literature of sweeping up the ambient material as a snow-plow and accumulating it into a shell \citep{anglada95,girart05}.
Consequently, based on the morphology considerations in this section, and on the energy considerations in Section~\ref{ssec-dynamics}, we speculate that the dusty spiral arm is created by the accumulation of matter due to the IRS~1 outflow feedback.
\subsection{Gravitational collapse of the individual cores: a cluster in the making\label{ssec-collapse}}
Three active, bright IR sources indicate that star formation is ongoing in the NGC~7538~IRS~1--3 cluster. Moreover, through high-resolution IR interferometry, \citet{Kraus06} find 18 new faint stars and a NW-oriented, fan-shaped outflow arising from MM1. Interestingly, the positions of the stars are correlated well with the outflow, which they propose is precessing and triggering star formation. In this environment of high interaction, we target the cold dust to study the possible evolution of the mass reservoir in the cluster.
The three cores in the central dusty structure seem to be gravitationally dominated against the support forces (Section~\ref{ssec-energy-sources}, note that MM4 is only at the limit). The MM1 core has already formed a still accreting protostar, which has gathered about two thirds of the total mass locally available (star plus dust core system). If the same star-to-core mass ratio applies, MM2 and MM4 will form massive stars of $\sim$24~$M_{\odot}$\ and $\sim$17~$M_{\odot}$, respectively.
The situation is less clear for the dust cores located either in the spiral arm or isolated. The ``mass balance'' analysis shows that the total mass is insufficient to gravitationally bound the cores. Figure~\ref{fig-msup-mobs} shows an extension of the traditional virial parameter analysis with the inclusion of the magnetically supported mass. This parameter is usually fitted by a function of the form $M_{\rm obs}^{a}$ that delivers a typical $a=-0.68\pm0.06$, in agreement with previously reported trends \citep[see e.g.][and derived works]{bertoldi92}. The data can also be fitted by a function of the form $b\left[M_{\rm vir}/M_{\rm obs}\right]^{a}$, where the proportionality constant $b$ carries a physical meaning, varying from 2.06 for self-graviting clumps to 2.9 for pressure-confined clumps \citep{bertoldi92}. The best fit to the cores delivers $b=2.26\pm0.12$, in good agreement with the prediction of $b=2.12$ for magnetized critical cores. To examine the range of applicability of the model, we show as blue crosses in Fig.~\ref{fig-msup-mobs} the expected results using the masses of the clumps, the theoretical $b=2.12$, and the previously derived $a$. The prediction for magnetized critical cores is in good agreement with the magnetically dominated low-mass cores (see also Section~\ref{ssec-energy-sources} and Fig.~\ref{fig-massBalance}). On the contrary, the model prediction is less precise for the more massive cores, where the magnetic field is less important and are less likely to be magnetically critical.
The analyses performed assume cores in isolation and ignore the effects of the highly dynamical environment. In other words, the expansion powered by the outflow may help to pile up material as the filament is expanding through the ISM, and more importantly, it is creating a high external pressure along the filament. This pressure can help gravitation to overcome magnetic and turbulent energy. We can estimate if an external perturbation can have a significant impact on core evolution by using the typical crossing-time $t_{\rm cross}$=$R_{\rm core}/c_{s}$. In average, the filament cores have $R_{\rm core}\simeq$\tenpow{3.6}{3}~AU and $c_{s}\simeq$0.53~km~s$^{-1}$, resulting in a typical $t_{\rm cross}$$\simeq$\tenpow{3}{4}~yr. This value is comparable to the $t_{\rm dyn}$ estimates for the filament and outflow (Section~\ref{ssec-dynamics}). Therefore, a perturbation constantly acting for $t_{\rm dyn}$, such as stellar winds, can influence the evolution of the cores in the filament. Since all measures suggest that the outflow is behind the formation of the entire filament (Section~\ref{ssec-feedback}), comparable timescales at a core level make reasonable that the external pressure will also influence the cores.
We speculate that the external pressure from the winds acting for $t_{\rm dyn}$ could trigger the collapse of the cores in the filament, leading to the formation of a group of low-mass stars. This triggered star formation SE of MM1 is supported by the mirrored star formation towards NW, in a more evolved stage of evolution. These stars are older than the MM1 outflow and could have been formed through feedback from the older, more evolved IRS~2 star. This star is associated to a well studied H\,{\tiny II}\ region and powers a ``stellar wind bow shock'' \citep{bloomer98}. Consequently, triggered star formation in the NGC~7538~IRS~1--3 complex could be an episodic process following the evolution of the most massive stars. Such scenario would generate a small cluster with two stellar groups: ({\it i}) a few central high-mass stars, surrounded by ({\it ii}) a wealth of low-mass stars formed through feedback from the former group.
\begin{figure}
\includegraphics[width=\columnwidth]{n7538i1-msup-mobs.pdf}
\caption{Ratio of supported mass to observed mass as a function of the observed mass for the NGC~7538~IRS~1--3 dense cores. MM1 is represented by a red star and the other cores by black circles. The solid line shows the best fit to the starless cores while blue crosses show the expected values for critical, magnetized clumps (see Section~\ref{ssec-collapse}). The dotted line shows the limiting value for effective support against gravitational collapse.}\label{fig-msup-mobs}
\end{figure}
\section{Conclusions\label{sec-concl}}
We have carried out a molecular, dust, and polarimetric study of the NGC~7538~IRS~1--3 star-forming cluster. We used SMA high angular resolution observations at 880~$\mu$m with the compact configuration. Here, we summarize the main results.
\begin{enumerate}
\item We detect up to 14 dust cores in continuum emission, six of them newly discovered, spanning one order of magnitude in mass (from 3.5~$M_{\odot}$\ to 37~$M_{\odot}$). The brightest core is MM1, associated with IRS~1. IRS~2 and IRS~3 show no continuum counterpart. The dust cores are connected by diffuse gas, and are arranged in two larger scale structures: a central bar containing MM1, MM2, and MM4; and a filamentary spiral arm containing at least 6 cores. The total dust mass is $\simeq$160~$M_{\odot}$, almost equally split between the two large scale structures.
\item We detect C$^{17}$O~3--2 and H$^{13}$CO$^{+}$~4--3 large scale emission sharply tracing the two main large scale structures, unveiling a velocity gradient along the spiral arm. We developed a code to generate synthetic velocity cubes, \texttt{RATPACKS}, and reproduced the velocity gradient through a model of a spiral expanding at 9~km~s$^{-1}$\ with respect to the central MM1.
\item We broadly detect polarized emission in the compact cores and in the diffuse extended structures. Based on a statistical analysis, we derive a magnetic field strength of $\simeq$2.5~mG. The orientation of the magnetic field segments is significantly homogeneous along the spiral arm, and it is correlated at an 80\% confidence level to the direction of the dust main axis. This suggests that dust and magnetic field are tightly connected.
\item The spiral arm is gravitationally unbound with respect to the central bar. The gravitational and magnetic field energies combined are a factor of 2.3--9.5 smaller than the kinetic energy. Therefore, it is likely that the dominant expansion is shaping dust and magnetic field into a similar morphology.
\item The total energy, linear momentum, and dynamic age ($\simeq$\tenpow{4.2}{46}~erg, $\simeq$\tenpow{4}{2}~$M_{\odot}$~km~s$^{-1}$, and $\simeq$\tenpow{1.8}{4}~yr) of the spiral arm are compatible with the values of the MM1 outflow by QZM11 when de-projected. Both spiral arm and outflow are red-shifted, hence likely to flow in parallel. Consequently, it seems plausible that the dominant kinetic energy of the spiral arm has its origin in the MM1 outflow, which may be causing its formation in a snow-plow fashion in agreement to our expansion model.
\item We developed the ``mass balance'' analysis that compares collapse vs. support forces, accounting for all the available information on the energetics at core scales. On the one hand, the cores in the central bar seem to be gravitationally unstable, and prone to form massive stars. On the other hand, the combined support forces seem to dominate the cores located in the spiral arm or isolated, with non-thermal motions and magnetic fields being the main agents of support, respectively. However, the dynamically important external pressure from the outflow could trigger the gravitational collapse, and lead to the formation of low-mass stars as reported towards NW to MM1 \citep{Kraus06}.
\item We speculate that the NGC~7538~IRS~1 region is forming a small cluster with a few central high-mass stars, surrounded by a number of low-mass stars formed through proto-stellar feedback.
\end{enumerate}
\begin{acknowledgements}
We thank all members of the SMA staff that made these observations possible. This research has made use of NASA's Astrophysics Data System Bibliographic Services (\texttt{http://adsabs.harvard.edu/}), the SIMBAD database, operated at CDS, Strasbourg, France (\texttt{http://simbad.u-strasbg.fr/simbad/}), and the Splatalogue database for astronomical spectroscopy (\texttt{http://www.splatalogue.net}). We thank the anonymous referee for the useful comments. PF is supported by the Spanish CONSOLIDER project CSD2009-00038. PF and JMG are supported by the Spanish MINECO AYA2011-30228-C03-02, and Catalan AGAUR 2009SGR1172 grants.
\end{acknowledgements}
|
\section{Introduction}
\label{sec:introduction}
\subsection*{Computability Theory}
In computable analysis \cite{pourel,weihrauchd}, there has for a long time been an interest in how complicated the set of codes of some element in a suitable spaces may be.
\name{Pour-El} and \name{Richards} \cite{pourel} observed that any real number, and more generally, any point in a Euclidean space, has a Turing degree.
They subsequently raised the question whether the same holds true for any computable metric space.
\name{Miller} \cite{miller2} later proved that various infinite dimensional metric spaces such as the Hilbert cube and the space of continuous functions on the unit interval contain points which lack Turing degrees, i.e.~have no simplest code w.r.t.~Turing reducibility.
A similar phenomenon was also observed in algorithmic randomness theory \cite{nies}.
\name{Day} and \name{Miller} \cite{daymiller} showed that no neutral measure \cite{Levin76} has Turing degree by understanding each measure as a point in the infinite dimensional space consisting of probability measures on an underlying space.
These previous works convince us of the need for a reasonable theory of degrees of unsolvability of points in an arbitrary represented space.
To establish such a theory, we associate a substructure of the Medvedev degrees with a represented space, which we call its \emph{point degree spectrum}.
A wide variety of classical degree structures are realized in this way, e.g., Turing degrees \cite{SoareBook}, enumeration degrees \cite{friedberg}, continuous degrees \cite{miller2}, degrees of continuous functionals \cite{Hinman73}.
What is more noteworthy is that the concept of a point degree spectrum is closely linked to infinite dimensional topology.
For instance, all points in a Polish space have Turing degrees if and only if the small transfinite inductive dimension of that space exists.
In a broader context, there are various instances of smallness properties (i.e., $\sigma$-ideals) of spaces and sets that start making sense for points in an effective treatment; e.g., arithmetical (Cohen) genericity \cite{nies,OdiBook}, Martin-L\"of randomness \cite{nies}, and effective Hausdorff dimension \cite{lutz2}.
In all these cases, individual points can carry some amount of complexity -- e.g.~a Martin-L\"of random point is in some sense too complicated to be included in a computable $G_\delta$ set having effectively measure zero.
A recent important example \cite{PoZa12,Zap14} from forcing theory is genericity with respect to the $\sigma$-ideal generated by finite-dimensional compact metrizable spaces.
Our work provides an effective notion corresponding to topological invariants such as small inductive dimension or metrizability, and e.g.~allows us to say that certain points are too complicated to be (computably) a member of a (finite-dimensional) Polish space.
Additionally, the actual importance of point degree spectrum is not merely conceptual, but also applicative.
Indeed, unexpectedly, our notion of point degree spectrum turned out to be a powerful tool in descriptive set theory and infinite dimensional topology, in particular, in the study of Banach space theory and involved Borel isomorphism problems, as explained in more depth below.
\subsection*{Descriptive Set Theory}
A {\em Borel isomorphism problem} (see \cite{CenMau84,Maul76,Hrba78,Har78}) asks to find a nontrivial isomorphism type in a certain class of Borel spaces (i.e., topological spaces together with their Borel $\sigma$-algebras).
An {\em $\alpha$-th level Borel/Baire isomorphism} between $\repsp{X}$ and $\repsp{Y}$ is a bijection $f$ such that $E\subseteq \repsp{X}$ is of additive Borel/Baire class $\alpha$ if and only if $f[E]\subseteq \repsp{Y}$ is of additive Borel/Baire class $\alpha$.
These restricted Borel isomorphisms are introduced by \name{Jayne} \cite{Jayne74}, in Banach space theory, to obtain certain variants of the
\name{Banach-Stone Theorem} and the \name{Gelfand-Kolmogorov Theorem} for Banach algebras of the forms $\mathcal{B}_\alpha^*(\repsp{X})$ for realcompact spaces $\repsp{X}$.
Here, $\mathcal{B}_\alpha^*(\repsp{X})$ is the Banach algebra of bounded real valued Baire class $\alpha$ functions on a space $\repsp{X}$ with respect to the supremum norm and the pointwise operation \cite{Bade73,Dash74,Jayne74}.
The first and second level Borel/Baire isomorphic classifications have been studied by several authors (see \cite{JayRog79a,JayRog79b}).
However, it is not certain even whether there is an uncountable Polish space whose $G_{\delta\sigma}$-structure is neither isomorphic to the real line nor to the Hilbert cube:
\begin{problem}[The Second-Level Borel Isomorphism Problem]\label{mainproblem1}
Are all uncountable Polish spaces second-level Borel isomorphic either to $\mathbb{R}$ or to $\mathbb{R}^\mathbb{N}$?
\end{problem}
\name{Jayne}'s result \cite{Jayne74} shows that this is equivalent to asking the following problem on Banach algebras.
\begin{problem}[see also \name{Motto Ros} \cite{MRos13}]\label{mainproblem2}
If $\repsp{X}$ is an uncountable Polish space.
Then does there exist $n\in\mathbb{N}$ such that $\mathcal{B}_n^*(\repsp{X})$ is linearly isometric (or ring isomorphic) either to $\mathcal{B}_n^*([0,1])$ or to $\mathcal{B}_n^*([0,1]^\mathbb{N})$?
\end{problem}
The very recent successful attempts to generalize the Jayne-Rogers theorem and the Solecki dichotomy (see \cite{MRos13,PaSa12} and also \cite{KiNg} for a computability theoretic proof) revealed that two Polish spaces are second-level Borel isomorphic if and only if they are $\sigma$-homeomorphic.
Here, a topological space $\repsp{X}$ is $\sigma$-homeomorphic to $\repsp{Y}$ (written as $\repsp{X}\cong_\sigma^\mathfrak{T}\repsp{Y}$) if there are countable covers $\{\repsp{X}_i\}_{i\in\omega}$ and $\{\repsp{Y}_i\}_{i\in\omega}$ of $\repsp{X}$ and $\repsp{Y}$ such that $\repsp{X}_i$ is homeomorphic to $\repsp{Y}_i$ for every $i\in\omega$.
Therefore, the second-level Borel isomorphism problem can be reformulated as the following equivalent problem.
\begin{problem}[\name{Motto Ros} et al.~\cite{schlicht}]\label{prob:third}
Is any Polish space $\repsp{X}$ either $\sigma$-embedded into $\mathbb{R}$ or $\sigma$-homeomorphic to $\mathbb{R}^\mathbb{N}$?
\end{problem}
Unlike the classical Borel isomorphism problem, which was able to be reduced to the same problem on zero-dimensional Souslin spaces, the second-level Borel isomorphism problem is inescapably tied to {\em infinite dimensional topology} \cite{vMBook,vMBook2}, since all transfinite dimensional uncountable Polish spaces are mutually second-level Borel isomorphic.
The study of $\sigma$-homeomorphic maps in topological dimension theory dates back to a classical work by \name{Hurewicz-Wallman} \cite{hurewicz} characterizing transfinite dimensionality.
\name{Alexandrov} \cite{Alek51} asked whether there exists a weakly infinite dimensional compactum which is not $\sigma$-homeomorphic to the real line.
\name{Roman Pol} \cite{pol2} solved this problem by constructing such a compactum.
Roman Pol's compactum is known to satisfy a slightly stronger covering property, called property $C$ \cite{AdGr78,Anc85,Hav74}.
Our notion of {\em degree spectrum} on Polish spaces serves as an invariant under second-level Borel isomorphism.
Indeed, an invariant which we call {\em degree co-spectrum}, a collection of Turing ideals realized as lower Turing cones of points of a Polish space, plays a key role in solving the second-level Borel isomorphism problem.
We show that there is an embedding of an uncountable partial ordering into the $\sigma$-embeddability (the second-level Borel embeddability) ordering of metrizable $C$-compacta.
The key idea is measuring the quantity of all possible Scott ideals realized within the degree co-spectrum of a given space.
Our spaces are completely described in the terminology of computability theory (based on \name{Miller}'s work on the continuous degrees \cite{miller2}).
Nevertheless, the first of our examples turns out to be second-level Borel isomorphic to \name{Roman Pol}'s compactum.
Hence, our solution can also be viewed as a refinement of \name{Roman Pol}'s solution to \name{Alexandrov}'s problem.
\subsection*{Summary of Results}
This work is part of a general development to study the descriptive theory of represented spaces \cite{pauly-overview-arxiv}, together with approaches such as synthetic descriptive set theory proposed in \cite{paulydebrecht,pauly-descriptive}.
In Section \ref{sec:pointdegreespectra}, we introduce the notion of point degree spectrum, and clarify the relationship with countable-continuity.
In Section \ref{sec:intermediate}, we introduce the notion of an $\omega$-left-CEA operator in the Hilbert cube as an infinite dimensional analogue of an $\omega$-CEA operator (in the sense of classical computability theory), and show that the graph of a universal $\omega$-left-CEA operator is an individual counterexample to Problems \ref{mainproblem1}, \ref{mainproblem2}, and \ref{prob:third}.
In Section \ref{sec:intermediate-dimension}, we clarify the relationship between a universal $\omega$-left-CEA operator and Roman Pol's compactum.
In Section \ref{sec:piecewisehomeo}, we describe a general procedure to construct uncountably many mutually different compacta under $\sigma$-homeomorphism.
In Section \ref{sec:internalcharacterization}, we characterize represented spaces with effectively-fiber-compact representations (which are relevant for complexity approaches to complexity theory along the lines of \name{Weihrauch} 's \cite{weihrauchf}) as precisely the computable metric spaces.
In Section \ref{sec:quasi-Polish}, we also look at the degree structures of nonmetrizable spaces.
In Section \ref{sec:non-quasi-Polish}, we construct an admissibly represented space whose degree spectrum is strictly larger than that of any second-countable $T_0$ spaces up to an oracle.
The methods used in Sections \ref{sec:internalcharacterization}--\ref{sec:non-quasi-Polish} do not depend on those developed in Sections \ref{sec:intermediate}--\ref{sec:piecewisehomeo}.
\section{Preliminaries}
\subsection{Represented spaces}
We briefly present some fundamental concepts on represented spaces following \cite{pauly-synthetic-arxiv}. A \emph{represented space} is a pair $\repsp{X} = (X, \delta_X)$ of a set $X$ and a partial surjection $\delta_X : \subseteq \Baire \to X$. A function between represented spaces is a function between the underlying sets. For $f : \repsp{X} \to \repsp{Y}$ and $F : \subseteq \Baire \to \Baire$, we call $F$ a realizer of $f$, iff $\delta_Y(F(p)) = f(\delta_X(p))$ for all $p \in \dom(f\delta_X)$, i.e.~if the following diagram commutes:
$$\begin{CD}
\Baire @>F>> \Baire\\
@VV\delta_\repsp{X}V @VV\delta_\repsp{Y}V\\
\repsp{X} @>f>> \repsp{Y}
\end{CD}$$
A map between represented spaces is called {\em computable} ({\em continuous}), iff it has a computable (continuous) realizer. Similarly, we call a point $x \in \repsp{X}$ {\em computable}, iff there is some computable $p \in \Baire$ with $\delta_\repsp{X}(p) = x$.
Thus, a represented space is a kind of space equipped with the notion of computability.
Based on the \textrm{UTM}-theorem, we can introduce the space $\mathcal{C}(\repsp{X},\repsp{Y})$ of continuous functions between $\repsp{X}$ and $\repsp{Y}$ such that function evaluation and the other usual notions are computable. In the following, we will want to make use of two special represented spaces, the countable discrete space $\mathbb{N} = (\mathbb{N}, \delta_\mathbb{N})$ and the Sierpi\'nski space $\mathbb{S} = (\{\bot, \top\}, \delta_\mathbb{S})$. Their representations are given by $\delta_\mathbb{N}(0^n10^\mathbb{N}) = n$, $\delta_\mathbb{S}(0^\mathbb{N}) = \bot$ and $\delta_\mathbb{S}(p) = \top$ for $p \neq 0^\mathbb{N}$. It is straightforward to verify that the computability notion for the represented space $\mathbb{N}$ coincides with classical computability over the natural numbers.
We then have the space $\mathcal{O}(\repsp{X}) \cong \mathcal{C}(\repsp{X},\mathbb{S})$ of open subsets of a represented space $\repsp{X}$ by identifying a set with its characteristic function, and the usual set-theoretic operations on this space are computable, too. We write $\mathcal{A}(\repsp{X})$ for the space of closed subsets, where names are names of the open complement.
Traditionally in computability theory, a computable element of the hyperspace $\mathcal{O}(\repsp{X})$ is called a {\em $\Sigma^0_1$ set}, a {\em $\Sigma^0_1$ class} or a {\em c.e.~open set}, and a computable element of the hyperspace $\mathcal{A}(\repsp{X})$ is called a {\em $\Pi^0_1$ set}, a {\em $\Pi^0_1$ class} or a {\em co-c.e.~closed set}.
The canonic function $\kappa_\repsp{X} : \repsp{X} \to \mathcal{O}(\mathcal{O}(\repsp{X}))$ mapping $x$ to $\{U \in \mathcal{O} \mid x \in U\}$ is always computable. If it has a computable inverse, then we call $\repsp{X}$ {\em computably admissible}. Admissibility in this sense was introduced by \name{Schr\"oder} \cite{schroder,schroder5}. Intuitively, the computably admissible represented spaces are those that can be understood fully as topological spaces.
A particularly relevant subclass of represented spaces are the computable Polish spaces, which are derived from complete computable metric spaces by forgetting the details of the metric, and just retaining the representation (or rather, the equivalence class of representations under computable translations). Forgetting the metric is relevant when it comes to compatibility with definitions in effective descriptive set theory as shown in \cite{pauly-gregoriades-arxiv}.
\begin{example}\label{example:representation}
The following are examples of admissible representations.
\begin{enumerate}
\item A {\em computable metric space} is a tuple $\repsp{M} = (M, d, (a_n)_{n \in \mathbb{N}})$ such that $(M,d)$ is a metric space and $(a_n)_{n \in\mathbb{N}}$ is a dense sequence in $(M,d)$ such that the relation
\[ \{(t,u,v,w) \: |\: \nu_{\mathbb{Q}}(t) < d(a_u, a_v) <\nu_{\mathbb{Q}}(w) \}\]
is recursively enumerable.
The {\em Cauchy representation} $ \delta_{\repsp{M}} \: : \: \Baire \rightharpoonup M $ associated with the computable metric space $ \repsp{M} = (M, d, (a_n)_{n \in \mathbb{N}}) $ is defined by
\[ \delta_{\repsp{M}}(p) = x \: : \: \Longleftrightarrow \begin{cases}
d(a_{p(i)}, a_{p(k)}) \leq 2^{-i} \text{ for } i < k\\
\text{and } x = \lim\limits_{i\rightarrow \infty}a_{p(i)}
\end{cases} \]
\item Another, more general subclass are the quasi-Polish spaces introduced by \name{de~Brecht} \cite{debrecht6}. A represented space $\repsp{X} = (X, \delta_\repsp{X})$ is \emph{quasi-Polish}, if it is countably based, admissible and $\delta_\repsp{X} : \Baire \to \repsp{X}$ is total. These include the computable Polish spaces as well as $\omega$-continuous domains.
\item Generally, a topological $T_0$-space $\repsp{X}$ with a countable base $\mathcal{B}=\langle{B_n\rangle}_{n\in\mathbb{N}}$ is naturally represented by defining $\delta_{(\repsp{X},\mathcal{B})}(p)=x$ iff $p$ enumerates the code of a neighborhood basis for $x$, that is, ${\rm range}(p)=\{n\in\mathbb{N}:x\in B_n\}$.
\end{enumerate}
\end{example}
\subsection{Degree structures}
The Medvedev degrees $\mathfrak{M}$ \cite{medvedev} are a cornerstone of our framework. These are obtained by taking equivalence classes from Medvedev reducibility $\leq_M$, defined on subsets $A$, $B$ of Baire space $\Baire$ via $A \leq_M B$ iff there is a computable function $F : B \to A$. Important substructures of $\mathfrak{M}$ also relevant to us are the Turing degrees $\mathcal{D}_T$, the continuous degrees $\mathcal{D}_r$ and the enumeration degrees $\mathcal{D}_e$, these satisfy $\mathcal{D}_T \subsetneq \mathcal{D}_r \subsetneq \mathcal{D}_e \subsetneq \mathfrak{M}$.
Turing degrees are obtained from the usual Turing reducibility $\leq_T$ defined on points $p, q \in \Baire$ with $p \leq_T q$ iff there is a computable function $F : \subseteq \Baire \to \Baire$ with $F(q) = p$. We thus see $p \leq_T q \Leftrightarrow \{p\} \leq_M \{q\}$, and can indeed understand the Turing degrees to be a subset of the Medvedev degrees. The continuous degrees were introduced by \name{Miller} in \cite{miller2}. Enumeration degrees have received a lot of attention in computability theory, and were originally introduced by \name{Friedberg} and \name{Rogers} \cite{friedberg}. In both cases, we can provide a simple definition directly as a substructure of the Medvedev degrees later on.
We also use the standard notations from modern computability theory \cite{nies,SoareBook}.
For instance, $\Phi_e^x$ denotes the computation of the $e$-th Turing machine with oracle $x$, and $x^{(\alpha)}$ denotes the $\alpha$-th Turing jump of $x$.
\subsection{Isomorphism and Classification}\label{section:classification-problems}
We are now interested in isomorphisms of a particular kind, this always means a bijection in that function class, such that the inverse is also in that function class.
For instance, consider the following morphisms.
For a function $f:\repsp{X}\to\repsp{Y}$,
\begin{enumerate}
\item $f$ is {\em countably computable} (or {\em $\sigma$-computable}) if there are sets $(X_n)_{n \in \mathbb{N}}$ such that $\repsp{X} = \bigcup_{n \in \mathbb{N}} X_n$ and each $f|_{X_n}$ is computable.
\item $f$ is {\em countably continuous} (or {\em $\sigma$-continuous}) if there are sets $(X_n)_{n \in \mathbb{N}}$ such that $\repsp{X} = \bigcup_{n \in \mathbb{N}} X_n$ and each $f|_{X_n}$ is continuous.
\item $f$ is {\em $\mathbf{\Gamma}$-piecewise continuous} if there are $\mathbf{\Gamma}$-sets $(X_n)_{n \in \mathbb{N}}$ such that $\repsp{X} = \bigcup_{n \in \mathbb{N}} X_n$ and each $f|_{X_n}$ is continuous.
\item $f$ is {\em second-level Borel measurable} is $\mathbf{\Sigma}^0_3$ for every $\mathbf{\Sigma}^0_3$ set $A\subseteq\repsp{X}$.
\end{enumerate}
Note that if $\repsp{X}$ and $\repsp{Y}$ have uniformly proper representations (this includes all computable metric spaces), then the $\repsp{X}_{n}$ in the definition of countable continuity may be assumed to be $\Pi_2^0$-sets.
Moreover, by recent results from descriptive set theory (see \cite{KiNg,MRos13,PaSa12}), we have the following implication for functions on Polish spaces:
\[\mbox{$\mathbf{\Pi}^0_2$-piecewise continuous $\Rightarrow$ second-level Borel measurable $\Rightarrow$ countably continuous}\]
Consequently, the second-level Borel isomorphic classification and the countably-continuous isomorphic classification of Polish spaces are exactly the same.
More precisely, three classification problems, Problems \ref{mainproblem1}, \ref{mainproblem2} and \ref{prob:third} in Section \ref{sec:introduction} are equivalent.
Hereafter, for notation, let $\cong$ be computable isomorphism, $\cong^\mathfrak{T}$ continuous isomorphism (i.e., homeomorphism), $\cong_{\sigma}$ be isomorphism by countably computable functions and $\cong_{\sigma}^\mathfrak{T}$ is countably-continuous isomorphism.
For any of these notions, we write $\repsp{X} \leq \repsp{Y}$ with the same decorations on $\leq$ if $\repsp{X}$ is isomorphic to a subspace of $\repsp{Y}$ (i.e., $\repsp{X}$ is embedded into $\repsp{Y}$) in that way. If $\repsp{X} \leq \repsp{Y}$ and $\repsp{X}$ is not isomorphic to $\repsp{Y}$ in the designated way, then we also write $\repsp{X} < \repsp{Y}$, again with the suitable decorations on $<$. If neither $\repsp{X} \leq \repsp{Y}$ nor $\repsp{Y} \leq \repsp{X}$, we write $\repsp{X} \ | \ \repsp{Y}$ (again, with the same decorations).
The Cantor-Bernstein argument shows the following.
\begin{observation}\label{obs:Cantor-Bernstein}
Let $\repsp{X}$ and $\repsp{Y}$ be represented spaces.
Then, $\repsp{X}\cong_\sigma\repsp{Y}$ if and only if $\repsp{X}\leq_\sigma\repsp{Y}$ and $\repsp{Y}\leq_\sigma\repsp{X}$
\end{observation}
If underlying spaces are admissibly represented spaces, we also use the terminologies such as {\em $\sigma$-homeomorphism} and {\em $\sigma$-embedding} to denote countably-continuous isomorphism and countably-continuous embedding.
\subsection{Topological Dimension theory}
\label{subsec:dimension-theory-intro}
As general source for topological dimension theory, we point to \name{Engelking} \cite{EngBook}.
See also \name{van Mill} \cite{vMBook,vMBook2} for infinite dimensional topology.
A topological space $\repsp{X}$ is \emph{countable dimensional} if it is the union of countably many finite dimensional subspaces.
Recall that a Polish space is countable dimensional if and only if it is {\em transfinite dimensional}, that is, its transfinite small inductive dimension is less than $\omega_1$ (see \cite[pp.~50--51]{hurewicz}).
One can see that a Polish space $\repsp{X}$ is countable dimensional if and only if $\repsp{X}\leq_\sigma^\mathfrak{T}\Cantor$.
To investigate the structure of uncountable dimensional spaces, \name{Alexandrov} introduced the notion of weakly/strongly infinite dimensional space.
We say that $C$ is a {\em separator} of a pair $(A,B)$ in a space $\repsp{X}$ if there are two pairwise disjoint open sets $A'\supseteq A$ and $B'\supseteq B$ such that $A'\sqcup B'=\repsp{X}\setminus C$.
A family $\{(A_i,B_i)\}_{i\in\Lambda}$ of pairwise disjoint closed sets in $\repsp{X}$ is {\em essential} if whenever $C_i$ is a separator of $(A_i,B_i)$ in $\repsp{X}$ for every $i\in\mathbb{N}$, $\bigcap_{i\in\mathbb{N}}C_i$ is nonempty.
A space $X$ is said to be {\em strongly infinite dimensional} if it has an essential family of infinite length.
Otherwise, $X$ is said to be {\em weakly infinite dimensional}.
We also consider the following covering property for topological spaces.
Let $\mathcal{O}[{\repsp{X}}]$ be the collection of all open covers of a topological space $\repsp{X}$, and $\mathcal{O}_2[{\repsp{X}}]=\{\mathcal{U}\in\mathcal{O}^X:\#\mathcal{U}=2\}$.
Then, $\repsp{X}\in\mathcal{S}_c(\mathcal{A},\mathcal{B})$ if for any sequence $(\mathcal{U}_n)_{n\in\mathbb{N}}\in\mathcal{A}[\repsp{X}]^\mathbb{N}$, there is a sequence $(\mathcal{V}_n)_{n\in\mathbb{N}}$ of pairwise disjoint open sets such that $\mathcal{V}_n$ refines $\mathcal{U}_n$ for each $n\in\mathbb{N}$ and $\bigcup_{n\in\mathbb{N}}\mathcal{V}_n\in\mathcal{B}[\repsp{X}]$.
Note that a topological space $\repsp{X}$ is weakly infinite dimensional if and only if $\repsp{X}\in\mathcal{S}_c(\mathcal{O}_2,\mathcal{O})$.
We say that $\repsp{X}$ is a {\em $C$-space} \cite{AdGr78,Hav74} or {\em selectively screenable} \cite{Bab05} if $\repsp{X}\in\mathcal{S}_c(\mathcal{O},\mathcal{O})$.
We have the following implications:
\[\mbox{countable dimensional }\Rightarrow\mbox{ $C$-space }\Rightarrow\mbox{ weakly infinite dimensional}.\]
\name{Alexandrov}'s old problem was whether there exists a weakly infinite dimensional compactum $\repsp{X}>_\sigma^\mathfrak{T}\Cantor$.
This problem was solved by \name{R.~Pol} \cite{pol2} by constructing a metrizable compactum of the form $R\cup L$ for a strongly infinite dimensional totally disconnected subspace $R$ and a countable dimensional subspace $L$.
Such a compactum is called {\em Pol-type}.
Every Pol-type compactum is a $C$-space, but is not countable-dimensional.
Namely, \name{R.~Pol}'s theorem says that there are at least two $\sigma$-homeomorphism types of metrizable $C$-compacta.
There are previous studies on the structure of continuous isomorphism types (Fr\'echet dimension types) of various kinds of infinite dimensional compacta, e.g., strongly infinite dimensional Cantor manifolds (see \cite{Cha99,ChaEP99,EPol02}).
Concerning weakly infinite dimensional Cantor manifolds, \name{El{\.z}bieta Pol} \cite{EPol96} (see also \cite{Cha99}) constructed a $C$-compactum in which no separator of nonempty subspaces can be hereditarily weakly infinite dimensional.
We call such a space a {\em Pol-type Cantor manifold}.
\section{Point Degree Spectra}\label{sec:pointdegreespectra}
\subsection{Generalized Turing Reducibility}\label{subsection:generalized-Turing-reducibility}
Recall that the notion of a represented space involves the notion of computability.
Hence, we can associate analogies of Turing reducibility and Turing degrees with an arbitrary represented space.
\begin{definition}
Let $\repsp{X}$ and $\repsp{Y}$ be represented spaces.
We say that $y\in\repsp{Y}$ is {\em point-Turing reducible} to $x\in\repsp{X}$ (written as $y\leq_M^{\repsp{X},\repsp{Y}}x$, or simply, $y\leq_Mx$) if there is a partial computable function $f:\subseteq\repsp{X}\to\repsp{Y}$ such that $f(x)=y$, that is, $\delta^{-1}_\repsp{Y}(y)\leq_M\delta^{-1}_\repsp{X}(x)$.
\end{definition}
Based on this idea, we introduce the notion of point degree spectrum of a represented space.
\begin{definition}\label{def:point-degree-spectra}
For a represented space $\repsp{X}$ and a point $x\in\repsp{X}$, define
\begin{align*}
\textrm{Spec}(x) &= [\delta_{\repsp{X}}^{-1}(x)]_{\equiv_M} = \mbox{``the Medvedev degree of $\delta_{\repsp{X}}^{-1}(x)$''}\\
\textrm{Spec}(\repsp{X})& = \{\textrm{Spec}(x) \mid x \in \repsp{X}\} \subseteq \mathfrak{M}.
\end{align*}
In other words, ${\rm Spec}(x)$ is the point-Turing degree of $x$, and ${\rm Spec}(\repsp{X})$ is the point-Turing degrees of points in the space $\repsp{X}$.
We call $\textrm{Spec}(\repsp{X})$ the \emph{point degree spectrum} of $\repsp{X}$.
We also define the relativized point degree spectrum by ${\rm Spec}^p(x)=[\{p\}\times\delta^{-1}_\repsp{X}(x)]_{\equiv_M}$ and ${\rm Spec}^p(\repsp{X})=\{{\rm Spec}^p(x):x\in\repsp{X}\}$.
\end{definition}
Clearly, one can identify the Turing degrees $\mathcal{D}_T$, the continuous degrees $\mathcal{D}_r$ and the enumeration degrees $\mathcal{D}_e$ with degree spectra of some spaces as follows:
\begin{itemize}
\item $\spec(\Cantor) = \spec(\Baire) = \spec(\mathbb{R}) = \mathcal{D}_T$
\item (\name{Miller} \cite{miller2}) $\spec(\uint^\mathbb{N}) = \spec(\mathcal{C}(\uint,\uint)) = \mathcal{D}_r$
\item $\spec(\mathcal{O}(\mathbb{N})) = \mathcal{D}_e$
\end{itemize}
\begin{observation}\label{example:admissible-ON}
As any separable metric space embeds into the Hilbert cube $\uint^\mathbb{N}$, we find in particular that $\spec(\repsp{X}) \subseteq \mathcal{D}_r$ for any computable metric space $\repsp{X}$.
As any second-countable $T_0$ spaces embeds into the Scott domain $\mathcal{O}(\mathbb{N})$, we also have that $\spec(\repsp{X}) \subseteq \mathcal{D}_e$ for any second-countable $T_0$ space $\repsp{X}$.
In the latter case, the point degree of $x\in\repsp{X}$ corresponds to the enumeration degree of neighborhood basis as in Example \ref{example:representation}.
The Turing degrees will be characterized in Section \ref{subsec:spectra-dimensionth} in the context of topological dimension theory.
\end{observation}
The following lemma shows -- in \name{Miller}'s words -- that the continuous degrees are \emph{almost} Turing degrees.
To be more precise, any continuous degree is relativized into a Turing degree by all Turing degrees except the smaller ones.
\begin{lemma}[\name{Miller}]\label{lem:JM-almosttotal}
For any non-total continuous degree $q \in \mathcal{D}_r \setminus \mathcal{D}_T$ we find that for all $p \in \mathcal{D}_T$, $(p,q) \in \mathcal{D}_T$ iff $p \nleq_M q$.
\begin{proof}
Let $r=(r(n))_{n \in \mathbb{N}} \in \uint^\mathbb{N}$ be a representative of a non-total continuous degree $q \in \mathcal{D}_r \setminus \mathcal{D}_T$.
Let $I$ be the set of all $y\in\Cantor$ such that $y\leq_Mr$, which is a countable set.
Choose a real $x$ whose Turing degree is incomparable with $I$.
In particular, $x$ is algebraically transcendent with all reals in $I$.
So, there is an $x$-computable homeomorphism sending $r$ to a sequence of irrationals.
Hence, given any name of $(x, r)$, we first obtain $x$, and by using $x$, transform $r$ into irrationals, and then we get the least Turing degree name of $(x,r)$.
\end{proof}
\end{lemma}
\begin{remark}
One can also define the Muchnik degree spectrum of a point as the collection of all Turing degrees of names of the point.
Indeed, the degree spectrum of a countable structure $S$ in the sense of computable model theory (see \cite{HKSS02,Rich81}) is defined as the Muchnik degree spectrum of the corresponding point $[S]$ of a quotient space of countable structures by logic action, rather than the Medvedev degree spectrum of the point.
The notion of {\em degree spectrum on a cone} (i.e., degree spectrum up to an oracle) plays an important role in (computable) model theory (see \cite{Monta13,Mont14}).
The detailed investigation on the difference between Medvedev and Muchnik degrees can be found in \cite{kihara3,kihara3b,Stuk07}.
However, if an admissibly represented space $\repsp{X}$ is second countable and $T_0$, the point degree spectrum is equivalent to the Muchnik degree spectrum.
This is because $\repsp{X}$ is countably based admissible space, then ${\rm Spec}(\repsp{X})\subseteq\mathcal{D}_e$ by Observation \ref{example:admissible-ON}, and Medvedev and Muchnik reducibility coincide for $\textrm{Spec}(\mathcal{O}(\mathbb{N}))$ (see \cite{miller2,selman}).
\end{remark}
\subsection{Degree Spectra and Dimension Theory}\label{subsec:spectra-dimensionth}
One of the main tools in our work is the following characterization of the point degree spectra of represented spaces.
\begin{theorem}
\label{theo:spectrum-main}
The following are equivalent for admissibly represented spaces $\repsp{X}$ and $\repsp{Y}$:
\begin{enumerate}
\item ${\rm Spec}^r(\repsp{X})={\rm Spec}^r(\repsp{Y})$ for some oracle $r\in\Cantor$.
\item $\mathbb{N}\times\repsp{X}$ is countable-continuously isomorphic to $\mathbb{N}\times\repsp{Y}$, i.e., $\mathbb{N}\times\repsp{X}\cong_\sigma^\mathfrak{T}\mathbb{N}\times\repsp{Y}$.
\end{enumerate}
Moreover, if $\repsp{X}$ and $\repsp{Y}$ are Polish, then the following assertions (3) and (4) are also equivalent to the above assertions (1) and (2).
\begin{enumerate}
\item[(3)] $\mathbb{N}\times\repsp{X}$ is second-level Borel isomorphic to $\mathbb{N}\times\repsp{Y}$.
\item[(4)] The Banach algebra $\mathcal{B}_2^*(\mathbb{N}\times\repsp{X})$ is linearly isometric (ring isomorphic and so on) to $\mathcal{B}_2^*(\mathbb{N}\times\repsp{Y})$.
\end{enumerate}
\end{theorem}
One can also see that the following assertions are equivalent:
\begin{enumerate}
\item[(2$^\prime$)] $\mathbb{N}\times\repsp{X}$ is $G_\delta$-piecewise homeomorphic to $\mathbb{N}\times\repsp{Y}$.
\item[(3$^\prime$)] $\mathbb{N}\times\repsp{X}$ is $n$-th level Borel isomorphic to $\mathbb{N}\times\repsp{Y}$ for some $n\geq 2$.
\item[(4$^\prime$)] The Banach algebra $\mathcal{B}_n^*(\mathbb{N}\times\repsp{X})$ is linearly isometric (ring isomorphic and so on) to $\mathcal{B}_n^*(\mathbb{N}\times\repsp{Y})$ for some $n\geq 2$.
\end{enumerate}
By our argument in Section \ref{section:classification-problems}, the assertions (2$^\prime$) is equivalent to (2).
Obviously the assertions (3) and (4) imply (3$^\prime$) and (4$^\prime$), respectively.
The equivalence between (3) and (4) (and the equivalence between (3$^\prime$) and (4$^\prime$)) has already been shown by Jayne \cite{Jayne74} for second-countable (or more generally, realcompact) spaces $\repsp{X}$ and $\repsp{Y}$.
The implication from the assertion (3$^\prime$) to (2) is, as mentioned in Section \ref{section:classification-problems}, recently proved by \cite{MRos13,PaSa12}, and more recently, an alternative computability-theoretic proof is given by \cite{KiNg} using our framework of point degree spectra of Polish spaces.
Consequently, all assertions from (2) to (4$^\prime$) are equivalent.
To see the equivalence between (1) and (2), we characterize the point degree spectra of represented spaces in the context of countably-continuous isomorphism.
\begin{lemma}\label{lemma:spectrum-main}
The following are equivalent for represented spaces $\repsp{X}$ and $\repsp{Y}$:
\begin{enumerate}
\item $\textrm{Spec}(\repsp{X}) \subseteq \textrm{Spec}(\repsp{Y})$
\item $\repsp{X}\leq_\sigma\mathbb{N}\times\repsp{Y}$, i.e., $\repsp{X}$ is a countable union of subspaces that are computably isomorphic to subspaces of $\repsp{Y}$.
\end{enumerate}
\end{lemma}
\begin{proof}
We first show that the assertion (1) implies (2).
By assumption, for any $x \in \repsp{X}$ we find $\delta_{\repsp{X}}^{-1}(x) \equiv_M \delta_{\repsp{Y}}^{-1}(y_x)$ for some $y_x \in \repsp{Y}$. Let for $\repsp{Y}$ any $i, j \in \mathbb{N}$, let $\repsp{X}_{ij}$ be the set of all points where the reductions are witnessed by $\Phi_i$ and $\Phi_j$, and let $\repsp{Y}_{ij} = \{y_x \mid x \in \repsp{X}_{ij}\} \subseteq \Cantor$. Then $\Phi_i$, $\Phi_j$ also witness $\repsp{X}_{ij} \cong \repsp{Y}_{ij}$, and obviously $\repsp{X} = \bigcup_{\langle i, j\rangle \in \mathbb{N}} \repsp{X}_{ij}$.
Conversely, the point spectrum is preserved by computable isomorphism and $\textrm{Spec}\left (\bigcup_{n \in \mathbb{N}} \repsp{X}_n\right ) = \bigcup_{n \in \mathbb{N}} \textrm{Spec}(\repsp{X}_n)$, so the claim follows.
\end{proof}
\begin{proof}[Proof of Theorem \ref{theo:spectrum-main} (1) $\Leftrightarrow$ (2).]
It follows from relativizations of Lemma \ref{lemma:spectrum-main} and Observation \ref{obs:Cantor-Bernstein}.
Here, it is easy to see that the assertion (2) is equivalent to $\mathbb{N}\times\repsp{X}\leq_\sigma\mathbb{N}\times\repsp{Y}$.
\end{proof}
This simple argument completely solves a mystery about the occurrence of non-Turing degrees in proper infinite dimensional spaces.
Concretely speaking, by combining Lemma \ref{lemma:spectrum-main} and a dimension-theoretic fact (see Section \ref{subsec:dimension-theory-intro}), we can characterize the Turing degrees by transfinite dimensionality\footnote{The same observation was independently made by \name{Hoyrup}. \name{Brattka} and \name{Miller} had conjectured that dimension would be the crucial demarkation line for spaces with only Turing degrees (all personal communication).}.
\begin{corollary}
The following are equivalent for a Polish space $\repsp{X}$ endowed with an admissible representation:
\begin{enumerate}
\item $\textrm{Spec}^p(\repsp{X}) \subseteq \mathcal{D}_T$ for some oracle $p\in\Cantor$
\item $\repsp{X}$ is transfinite dimensional.
\end{enumerate}
\end{corollary}
Now, by Theorem \ref{theo:spectrum-main}, the countable-continuously isomorphic ($\sigma$-homeomorphic) classification can be viewed as a kind of degree theory dealing with the {\em degrees of degree structures} (on a cone).
For instance, one may ask whether Post's problem (the Friedberg-Muchnik theorem and so on) is true for degrees of degrees of uncountable Polish spaces.
More details of the structure of degree spectra of Polish space will be investigated in Sections \ref{sec:intermediate} and \ref{sec:piecewisehomeo}, and those of quasi-Polish space will be in Section \ref{sec:quasi-Polish}.
We also study the degree spectrum of a non-quasi-Polish space in Section \ref{sec:non-quasi-Polish}.
\section{Intermediate Point Degree Spectra in Computability Theory}\label{sec:intermediate}
\subsection{Intermediate Polish Spaces}
Let $\mathfrak{P}$ be the set of all uncountable Polish spaces.
In this section, we investigate the structure of $\mathfrak{P}/\cong_{\sigma}^\mathfrak{T}$, i.e.~either of the equivalence classes w.r.t.~$\sigma$-homeomorphisms, or equivalently, the structure of point degree spectra of uncountable Polish spaces up to relativization.
It is well-known that for every uncountable Polish space $X$:
\[\Cantor\leq_{c}^\mathfrak{T} X\leq_{c}^\mathfrak{T}[0,1]^\mathbb{N},\]
where, recall that $\leq_{c}^\mathfrak{T}$ is the topological embeddability relation (i.e., the ordering of Fr\'echet dimension types).
The structure of Fr\'echet types of uncountable dimensional Polish spaces has been developed by several authors (see \cite{Cha99,ChaEP99,EPol96,EPol02}).
In this section, we focus on Problem \ref{prob:third} asking whether there exists a Polish space $\repsp{X}$ satisfying the following:
\[\Cantor <_{\sigma}^\mathfrak{T}\repsp{X}<_{\sigma}^\mathfrak{T}[0,1]^\mathbb{N}.\]
One can see that there is no difference between the structures of $\sigma$-homeomorphism types of uncountable Polish spaces and uncountable metrizable compacta.
\begin{fact}\label{thm:equivalence}
Every Polish space is $\sigma$-homeomorphic to a compact metrizable space.
\end{fact}
\begin{proof}
All spaces of a given countable cardinality are clearly $\sigma$-homeomorphic, and there are compact metrizable spaces of all countable cardinalities.
So let $\repsp{X}$ be an uncountable Polish space.
\name{Lelek} \cite{lelek} showed that every Polish space $\repsp{X}$ has a compactification $\gamma \repsp{X}$ such that $\gamma \repsp{X}\setminus \repsp{X}$ is countable-dimensional.
Clearly $\repsp{X}\leq_{c}\gamma \repsp{X}$.
Then, we have $\gamma \repsp{X}\setminus \repsp{X}\leq_{\sigma}^\mathfrak{T}\Cantor\leq_{\sigma}^\mathfrak{T}\repsp{X}$, since $\repsp{X}$ is uncountable Polish and $\gamma \repsp{X}\setminus \repsp{X}$ is countable-dimensional.
Consequently, $\repsp{X},\gamma \repsp{X}\setminus \repsp{X}\leq_{\sigma}^\mathfrak{T}\repsp{X}$, and this implies $\gamma \repsp{X}=\repsp{X}\cup(\gamma \repsp{X}\setminus \repsp{X})\leq_{\sigma}^\mathfrak{T}\repsp{X}$.
\end{proof}
\subsection{The Graph Space of a Universal $\omega$-Left-CEA Operator}
Now, we provide a concrete example having an intermediate degree spectrum.
We say that a point $(r_n)_{n\in\mathbb{N}}\in[0,1]^\mathbb{N}$ is {\em $\omega$-left-CEA in} or an {\em $\omega$-left-pseudojump of} $x\in \Cantor$ if $r_{n+1}$ is left-c.e.~in $\langle x,r_0,r_1,\dots,r_n\rangle$ uniformly in $n\in\mathbb{N}$.
In other words, there is a computable function $\Psi:\Cantor\times[0,1]^{<\omega}\times\mathbb{N}^2\to\mathbb{Q}_{\geq 0}$ such that
\[r_{n}=\limsup_{s\to\infty}\Psi(x,r_0,\dots,r_{n-1},n,s)\]
for every $x,n,s$, where $\mathbb{Q}_{\geq 0}$ denotes the set of all nonnegative rationals.
Whenever $r_n\in[0,1]$ for all $n\in\mathbb{N}$, such a computable function $\Psi$ generates an operator $J_\Psi^\omega:\Cantor\to[0,1]^\mathbb{N}$ with $J_\Psi^\omega(x)=(r_0,r_1,\dots)$, which is called an {\em $\omega$-left-CEA operator}.
\begin{proposition}\label{prop:universaloperator}
There is an effective enumeration $(J^\omega_e)_{e\in\mathbb{N}}$ of all $\omega$-left-CEA operators.
\end{proposition}
\begin{proof}
It is not hard to see that $y\in[0,1]$ is left-c.e.~in $x\in\Cantor\times[0,1]^k$ if and only if there is a c.e.~set $W\subseteq\mathbb{N}\times \mathbb{Q}$ such that
\[y=J^{k}_W(x):=\sup\{\min\{|p|,1\}:x\in B^k_i\mbox{ for some }(i,p)\in W\},\]
where $B_i^k$ is the $i$-th rational open ball in $[0,1]^k$.
Thus, we have an effective enumeration of all left-c.e.~operators $J:\Cantor\times[0,1]^k\rightarrow[0,1]$ by putting $J^k_e=J^k_{W_e}$, where $W_e$ is the $e$-th c.e. subset of $\mathbb{N}\times \mathbb{Q}$.
Then, we define
\[J^\omega_e(x)=(x,J^0_{\langle e,0\rangle}(x),J^1_{\langle e,1\rangle}(x,J^0_{\langle e,0\rangle}(x)),\dots),\]
that is, $J^\omega_e$ is the $\omega$-left-CEA operator generated by the uniform sequence $(J^k_{\langle e,k\rangle})_{k\in\mathbb{N}}$ of left-c.e.~operators.
Clearly, $(J^\omega_e)_{e\in\mathbb{N}}$ is an effective enumeration of all $\omega$-left-CEA operators.
\end{proof}
Hence, we may define a {\em universal $\omega$-left-CEA operator} by $J^\omega(e,x)=J^\omega_e(x)$.
\begin{definition}
The {\em $\omega$-left-computably-enumerable-in-and-above space} $\omega\repspb{CEA}$ is a subspace of $\mathbb{N}\times \Cantor\times[0,1]^\mathbb{N}$ defined by
\begin{align*}
\repspb{\omega CEA}&=\{(e,x,r)\in\mathbb{N}\times \Cantor\times[0,1]^\mathbb{N}:r=J^\omega_e(x)\}\\
&=\mbox{``the graph of a universal $\omega$-left-CEA operator.''}
\end{align*}
\end{definition}
Note that in classical recursion theory, an operator $\Psi$ is called a {\em CEA-operator} (also known as an {\em REA-operator} or a {\em pseudojump}) if there is a c.e.~procedure $W$ such that $\Psi(A)=\langle{A,W(A)\rangle}$ for any $A\subseteq\mathbb{N}$ (see \name{Odifreddi} \cite[Sections XII and XIII]{OdiBook}).
An $\omega$-CEA operator is the $\omega$-th iteration of a uniform sequence of CEA-operators.
In general, computability theorists have studied $\alpha$-CEA operators for computable ordinals $\alpha$ in the theory of $\Pi^0_2$ singletons.
We say that a continuous degree is {\em $\omega$-left-CEA} if it contains a point $r\in[0,1]^\mathbb{N}$ which is $\omega$-left-CEA in a point $z\in \Cantor$ such that $z\leq_Mr$.
The point degree spectrum of the space $\repspb{\omega CEA}$ (as a subspace of $[0,1]^\mathbb{N}$) can be described as follows.
\[\spec(\repspb{\omega CEA})=\{\mathbf{a}\in\mathcal{D}_r:\mbox{$\mathbf{a}$ is $\omega$-left-CEA}\}.\]
Clearly,
\[\spec({\Cantor})\subseteq\spec({\repspb{\omega CEA}})\subseteq\spec({[0,1]^\mathbb{N}}).\]
\begin{lemma}
The $\omega$-left-CEA space $\repspb{\omega CEA}$ is Polish.
\end{lemma}
\begin{proof}
It suffices to show that $\repspb{\omega CEA}$ is $\Pi^0_2$.
The stage $s$ approximation to $J_{e}^k$ is denoted by $J_{e,s}^k$, that is, $J_{e,s}^k(z)=\max\{\min\{|p|,1\}:(\exists\langle i,p\rangle\in W_{e,s})\;x\in U_i\}$, where $W_{e,s}$ is the stage $s$ approximation to the $e$-th computably enumerable set $W_e$.
Note that the function $(e,s,k,z)\mapsto J_{e,s}^k(z)$ is computable.
We can easily see that $(e,x,r)\in\repspb{\omega CEA}$ if and only if
\[(\forall n,k\in\mathbb{N})(\exists s>n)\;d\left(\pi_{k}(r),J^k_{e,s}(x,\pi_0(r),\pi_1(r),\dots,\pi_{k-1}(r))\right)<2^{-n},\]
where $d$ is the Euclidean metric on $[0,1]$.
\end{proof}
We devote the rest of this section to a proof of the following theorem.
\begin{theorem}\label{thm:intermediatePolish}
The space $\repspb{\omega CEA}$ has an intermediate $\sigma$-homeomorphism type, that is,
\[\Cantor <_{\sigma}^\mathfrak{T} \repspb{\omega CEA} <_{\sigma}^\mathfrak{T} [0,1]^\mathbb{N}.\]
\end{theorem}
Consequently, the space $\repspb{\omega CEA}$ is a concrete counterexample to Problem \ref{prob:third}.
\subsection{Proof of $\repspb{\omega CEA}<_\sigma^\mathfrak{T}[0,1]^\mathbb{N}$}
The key idea is to measure {\em how similar the space $\repsp{X}$ is to a zero-dimensional space} by approximating each point in a space $\repsp{X}$ by a zero-dimensional space.
Recall from Section \ref{subsection:generalized-Turing-reducibility} that the point degree spectrum coincides with the Muchnik degree spectrum for any second-countable admissibly represented space.
Therefore, the spectrum ${\rm Spec}(x)$ of a point $x\in\repsp{X}$ can be identified with its Turing upper cone, that is,
\[{\rm Spec}(x)\simeq\{z\in \Cantor:x\leq_Mz\}.\]
We think of the spectrum ${\rm Spec}(x)$ as {\em the upper approximation of $x\in\repsp{X}$ by the zero-dimensional space $\Cantor$}.
Now, we need the notion of {\em the lower approximation of $x\in\repsp{X}$ by the zero-dimensional space $\Cantor$}.
We introduce the {\em co-spectrum} of a point $x\in\repsp{X}$ as its Turing lower cone
\[{\rm coSpec}(x)=\{z\in \Cantor:z\leq_Mx\},\]
and moreover, we define the {\em degree co-spectrum} of a represented space $\repsp{X}$ as follows:
\[{\rm coSpec}(\repsp{X})=\{{\rm coSpec}(x):x\in\repsp{X}\}.\]
Note that the degree spectrum of a represented space fully determines its co-spectrum, while the converse is not true. For every oracle $p\in \Cantor$, we may also introduce relativized co-spectra ${\rm coSpec}^p(x)=\{z\in \Cantor:z\leq_Mx\oplus p\}$, and the relativized degree co-spectra ${\rm coSpec}^p(\repsp{X})$ in the same manner.
\begin{observation}\label{obs:main-cospec-inv}
Let $\repsp{X}$ and $\repsp{Y}$ be admissibly represented spaces.
If ${\rm Spec}^p(\repsp{X})={\rm Spec}^p(\repsp{Y})$, then we also have ${\rm coSpec}^p(\repsp{X})={\rm coSpec}^p(\repsp{Y})$.
Therefore, by Theorem \ref{theo:spectrum-main}, the cospectrum of an admissibly represented space up to an oracle is invariant under $\sigma$-homeomorphism.
\end{observation}
A collection $\mathcal{I}$ of subsets of $\mathbb{N}$ is {\em realized as the co-spectrum of $x$} if ${\rm coSpec}(x)=\mathcal{I}$.
A countable set $\mathcal{I}\subseteq\mathcal{P}(\mathbb{N})$ is a {\em Scott ideal} if it is the standard system of a countable nonstandard model of Peano arithmetic, or equivalently, a countable $\omega$-model of ${\sf RCA}$+${\sf WKL}$.
\name{Miller} \cite{miller2} showed that every countable Scott ideal is realized as a co-spectrum in $[0,1]^\mathbb{N}$.
\begin{example}
The spectra and co-spectra of Cantor space $\Cantor$, the space $\omega{\rm CEA}$, and the Hilbert cube $[0,1]^\mathbb{N}$ are illustrated as follows (see also Figure \ref{fig:spec-cospec}):
\begin{enumerate}
\item The co-spectrum ${\rm coSpec}(x)$ of any point $x\in\Cantor$ is principal, and {\em meets with ${\rm Spec}(x)$} exactly at $\deg_T(x)$.
The same is true up to some oracle for an arbitrary Polish spaces $\repsp{X}$ such that $\repsp{X}\cong_\sigma^\mathfrak{T}\Cantor$.
\item For any point $z\in\omega{\rm CEA}$, the ``{\em distance}'' between ${\rm Spec}(z)$ and ${\rm coSpec}(z)$ has to be at most the $\omega$-th Turing jump (see Lemma \ref{lem:Scott-ideal}).
\item An arbitrary countable Scott ideal is realized as ${\rm coSpec}(y)$ of some point $y\in [0,1]^\mathbb{N}$.
Hence, ${\rm Spec}(y)$ and ${\rm coSpec}(y)$ can be separated by {\em an arbitrary distance}.
\end{enumerate}
\end{example}
\begin{figure}[t]
\begin{center}
\includegraphics[width=100mm]{spec.eps}
\end{center}
\caption{The upper and lower approximations of $\Cantor$, $\omega\repspb{CEA}$ and $[0,1]^\mathbb{N}$}
\label{fig:spec-cospec}
\end{figure}
This upper/lower approximation method clarifies the differences of $\sigma$-homeomorphism types of spaces because both relativized point-degree spectra and co-spectra are invariant under $\sigma$-homeomorphism by Theorem \ref{theo:spectrum-main} and Observation \ref{obs:main-cospec-inv}.
\begin{lemma}\label{lem:Scott-ideal}
For any oracle $p\in \Cantor$, there is a countable Scott ideal which cannot be realized as a $p$-co-spectrum of an $\omega$-left-CEA continuous degree.
\end{lemma}
\begin{proof}
Let $y=(e,x,r)\in\omega\repspb{CEA}$ be an arbitrary point.
Then, $x\in{\rm coSpec}(y)$, and $x^{(\omega)}\in{\rm Spec}(y)$ since $r$ is $\omega$-left-CEA in $x$.
Hence, ${\rm coSpec}(y)$ is not closed under the $\omega$-th Turing jump for any $y\in\omega\repspb{CEA}$.
Thus, for any oracle $p$, the Scott ideal $\mathcal{A}^p=\{x\in \Cantor:(\exists n\in\mathbb{N})\;x\leq_T p^{(\omega\cdot n)}\}$ cannot be realized as a co-spectrum in $\omega\repspb{CEA}$.
\end{proof}
Consequently, the $\omega$-left-CEA space is not $\sigma$-homeomorphic to the Hilbert cube.
Note that \name{Day} and \name{Miller} \cite{daymiller} showed that every countable Scott set $\mathcal{I}$ is realized by a neutral measure.
Hence, we can also conclude that there is a neutral measure whose continuous degree is not $\omega$-left-CEA.
\subsection{Proof of $\Cantor<_\sigma^\mathfrak{T}\repspb{\omega CEA}$}\label{subsec:proof-main-half}
Next, we have to show that the $\omega$-left-CEA space is not countable-dimensional.
For $\repsp{X}\subseteq[0,1]^\mathbb{N}$, we inductively define $\min \repsp{X}\in \repsp{X}$ as follows:
\[\pi_n(\min \repsp{X})=\min\pi_n[\{z\in \repsp{X}:(\forall i<n)\;\pi_i(z)=\pi_i(\min \repsp{X})\}],\]
where $\pi_n:[0,1]^\mathbb{N}\to[0,1]$ is the $n$-th projection.
\begin{lemma}\label{lem:basis}
If $\repsp{X}\subseteq[0,1]^\mathbb{N}$ is $\Pi^0_1(p)$ for some $p\in \Cantor$, then $\min \repsp{X}$ is $\omega$-left-CEA in $p$.
\end{lemma}
\begin{proof}
We first note that Hilbert cube $[0,1]^\mathbb{N}$ is computably compact in the sense that $A_{[0,1]^\mathbb{N}}:\mathcal{O}([0,1]^\mathbb{N})\to\mathbb{S}$ is computable, where $A_{[0,1]^\mathbb{N}}(U)=\top$ iff $U=[0,1]^\mathbb{N}$.
Equivalently, there is a computable enumeration of all finite collections $\mathcal{D}$ of basic open sets which covers the whole space, that is, $\bigcup\mathcal{D}=[0,1]^\mathbb{N}$.
It suffices to show that $\pi_{n+1}(\min X)$ is left-c.e.~in $\langle\pi_{i}(\min X)\rangle_{i\leq n}$ uniformly in $n$ relative to $p$.
Given a sequence $\mathbf{a}=(a_0,a_1,\dots,a_n)$ of reals and an real $q$, we denote by $C(\mathbf{a},q)$ the set of all points in $X$ of the form $(a_0,a_1,\dots,a_n,r,\dots)$ for some $r\leq q$, that is,
\[C(\mathbf{a},q)=X\cap \bigcap_{i\leq n}\pi_i^{-1}\{a_i\}\cap\pi_{n+1}^{-1}[0,q].\]
By computable compactness of Hilbert cube, one can see that $C^*(\mathbf{a})=\{q\in [0,1]:C(\mathbf{a},q)=\emptyset\}$ is $p$-c.e.~open uniformly in $\mathbf{a}$ since the complement of $C(\mathbf{a},q)$ is $p$-c.e.~open uniformly in $\mathbf{a}$ and $q$.
Therefore, if $C^*(\mathbf{a})$ is nonempty, then $\sup C^*(\mathbf{a})$ is $p$-left-c.e.~uniformly in $\mathbf{a}$.
Finally, we can easily see that $\pi_{n+1}(\min X)$ is exactly $\sup C^*(\langle\pi_{i}(\min X)\rangle_{i\leq n})$.
\end{proof}
We use the following relativized versions of \name{Miller}'s lemmas.
\begin{lemma}[\name{Miller} {\cite[Lemma 6.2]{miller2}}]\label{lemma:Miller-6-2}
For every $p\in \Cantor$, there is a multivalued function $\Psi^p:[0,1]^\mathbb{N}\to[0,1]^\mathbb{N}$ with a $\Pi^0_1(p)$ graph and nonempty, convex images such that, for all $e\in\mathbb{N}$, $\alpha\in[0,1]^\mathbb{N}$ and $\beta\in\Psi^p(\alpha)$, if for every representation $\lambda$ of $\alpha$, $\varphi_e^{\lambda\oplus p}$ is a representation of $x\in [0,1]$, then $\beta(e)=x$.
\end{lemma}
Note that \name{Kakutani}'s fixed point theorem ensures the existence of a fixed point of $\Psi$.
If $\alpha$ is a fixed point of $\Psi^p$, that is, $\alpha\in\Psi^p(\alpha)$, then ${\rm coSpec}^p(\alpha)=\{\alpha(n):n\in\mathbb{N}\}$.
Therefore, such an $\alpha$ has no Turing degree relative to $p$ (see \cite{miller2}).
\begin{lemma}[\name{Miller} {\cite[Lemma 9.2]{miller2}}]\label{lem:Miller2}
For every $p\in \Cantor$, there is an index $e\in\mathbb{N}$ such that for any $x\in[0,1]$, there is a fixed point $\alpha$ of $\Psi^p$ such that $\alpha(e)=x$.
\end{lemma}
\begin{lemma}\label{thm:nontotalcea}
For any oracle $p\in\Cantor$, there is an $\omega$-left-CEA continuous degree which is not contained in ${\rm Spec}^p(\Cantor)$.
\end{lemma}
\begin{proof}
Let ${\rm Fix}(\Psi^p)$ be the set of all fixed points of $\Psi^p$.
Then, ${\rm Fix}(\Psi^p)$ is $\Pi^0_1(p)$ since it is the intersection of the graph of a $\Pi^0_1(p)$ set and the diagonal set.
Let $e$ be an index in Lemma \ref{lem:Miller2}.
Clearly, $A=\{\alpha\in{\rm Fix}(\Psi^p):\alpha(e)=p\}$ is again a $\Pi^0_1(p)$ subset of $[0,1]^\mathbb{N}$.
By Lemma \ref{lem:basis}, $A$ contains an element $\alpha$ which is $\omega$-left-CEA in $p$.
By the property of $A$ discussed above, $\alpha$ has no Turing degree relative to $p$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:intermediatePolish}]
By Lemma \ref{lem:Scott-ideal}, ${\rm coSpec}^p(\omega\repspb{CEA})\subsetneq{\rm coSpec}^p([0,1]^\mathbb{N})$ for any oracle $p$.
Moreover, by Lemma \ref{thm:nontotalcea}, ${\rm Spec}^p(\Cantor)\subsetneq{\rm Spec}^p(\omega\repspb{CEA})$ for any oracle $p$.
Therefore, by Theorem \ref{theo:spectrum-main} and Observation \ref{obs:main-cospec-inv}, we conclude $\Cantor<_\sigma^\mathfrak{T}\omega\repspb{CEA}<_\sigma^\mathfrak{T}[0,1]^\mathbb{N}$.
\end{proof}
\section{Intermediate Point Degree Spectra in Dimension Theory}\label{sec:intermediate-dimension}
\subsection{Strongly Infinite Dimensional Totally Disconnected Polish Spaces}
In this section, we will shed light on dimension-theoretic perspectives of the $\omega$-left-CEA space.
Note that $\repspb{\omega CEA}$ is a totally disconnected infinite dimensional space.
We first compare our space $\repspb{\omega CEA}$ and a totally disconnected infinite dimensional space $\repspb{RSW}$ which is constructed by \name{Rubin}, \name{Schori}, and \name{Walsh} \cite{schori}.
A {\em continuum} is a connected compact metric space, and a continuum is {\em nondegenerated} if it contains at least two points.
It is known that the hyperspace ${\sf CK}(\repsp{X})$ of continua in a compact metrizable space $\repsp{X}$ equipped with the Vietoris topology forms a Polish space.
Hence, we may think of ${\sf CK}(\repsp{X})$ as a represented space, which corresponds to a positive-and-negative representation of the hyperspace in computable analysis.
We consider a closed subspace $S$ of ${\sf CK}(\repsp{X})$ consisting of all continua connecting opposite faces $\pi_0^{-1}\{0\}$ and $\pi_0^{-1}\{1\}$.
Then, fix a total Cantor representation of $S$, i.e., a continuous surjection $\delta_{\sf CK}$ from the Cantor set $C\subseteq[0,1]$ onto $S$.
We define the {\em Rubin-Schori-Walsh space} $\repspb{RSW}$ \cite{schori,vMBook} as follows:
\begin{align*}
\repspb{RSW}&=\{\min (\delta_{\sf CK}(p)^{[p]}):p\in C\},\\
&=\{\min A^{[p]}:\mbox{$A$ is the $p$-th continuum of $[0,1]^\mathbb{N}$ with $[0,1]\subseteq\pi_0[A]$}\},
\end{align*}
where $A^{[p]}=A\cap\pi_0^{-1}\{p\}=\{z\in A:\pi_0(z)=p\}$.
A compactification of $\repspb{RSW}$ is well-known in the context of \name{Alexandrov}'s old problem in dimension theory.
\name{Pol}'s compactum $\repspb{RP}$ is given as a compactification in the sense of \name{Lelek} of the space $\repspb{RSW}$.
Hence, we can see that $\repspb{RP}$ and $\repspb{RSW}$ have the same point degree spectra (modulo an oracle) as in the proof of Fact \ref{thm:equivalence}.
Surprisingly, these spaces have the same degree spectra as the space $\omega\repspb{CEA}$ up to an oracle.
\begin{theorem}\label{thm:degspec-dim}
The following spaces are all $\sigma$-homeomorphic to each other.
\begin{enumerate}
\item The $\omega$-left-CEA space $\repspb{\omega CEA}$.
\item Rubin-Schori-Walsh's totally disconnected strongly infinite dimensional space $\repspb{RSW}$.
\item Roman Pol's counterexample $\repspb{RP}$ to Alexandrov's problem.
\end{enumerate}
\end{theorem}
\begin{lemma}\label{lem:degspec-dim1}
Every point of $\repspb{RSW}$ is $\omega$-left-CEA.
\end{lemma}
\begin{proof}
By Lemma \ref{lem:basis}, $\min A^{[p]}$ is $\omega$-left-CEA in $p$, since $A^{[p]}$ is $\Pi^0_1(p)$.
Moreover, clearly, $p\leq_M\min A^{[p]}$.
Thus, $\min A^{[p]}$ is $\omega$-left-CEA.
\end{proof}
For notational convenience, without loss of generality, we may assume that the $e$-th $z$-computable continuum is equal to the $\langle e,z\rangle$-th continuum.
\begin{lemma}\label{lem:degspec-dim2}
Suppose that $x\in[0,1]^\mathbb{N}$ is $\omega$-left-CEA in a point $z\in \Cantor$.
Then, there is a nondegenerated $z$-computable continuum $A\subseteq[0,1]^\mathbb{N}$ such that $[0,1]\subseteq\pi_0[A]$ and $\min A^{[p]}=(p,x)$ for a name $p$ of $A$.
\end{lemma}
\begin{proof}
Given $p$, we will effectively construct a name $\Psi(p)$ of a continuum $A$.
By Kleene's recursion theorem (see \cite{SoareBook}), we may fix $p$ such that the $p$-th continuum is equal to the $\Psi(p)$-th continuum.
Fix an $\omega$-left-CEA operator $J$ generated by $\langle W_n\rangle_{n\in\mathbb{N}}$ such that $J(z)=x$.
Here, as in the proof of Proposition \ref{prop:universaloperator}, each $W_n$ is a c.e.~list of pairs $(i,p)$ indicating $B_i^n\subseteq(J^n_{W_n})^{-1}[p,1]$.
Since $p=\langle e,z\rangle$ for some $e\in\mathbb{N}$, we have a computable function $\pi$ with $\pi(p)=z$, and then, redefine $W_0$ to be $W_0\circ\pi$.
In this way, we may assume that $J(p)=x$.
At stage $0$, $\Psi$ constructs $A_0=[0,1]\times[0,1]^\mathbb{N}$.
At stage $s+1$, if we find some rational open ball $B_i^n\subseteq[0,1]^{n}$ and a rational $q\in\mathbb{Q}$ such that $W_{n,s}$ declares $B_i^n\subseteq (J^n_{W_n})^{-1}[q,1]$ by enumerating $(i,q)$, then $\Psi$ removes $\pi_0^{-1}[B(p;2^{-s})]\cap(B_i^n\times[0,q)\times[0,1]^\mathbb{N})$ from the previous continuum $A_{s-1}$, where $B(p;2^{-s})$ is the rational open ball with center $p$ and radius $2^{-s}$.
Now, we show $\min A^{[p]}=x:=(x_0,x_1,\dots)$.
Assume that $x_0,\dots,x_{n-1}$ is an initial segment of $\min A^{[p]}$.
We will show that $x_n=\pi_n(\min A^{[p]})=\min\pi_n[\{z\in A^{[p]}:(\forall i<n)\;\pi_i(z)=x_i\}]$.
Since $J^n_{W_n}(p,x_0,\dots,x_{n-1})=x_n$, for any rational $q<x_n$, there is $i$ such that $(i,q)\in W_n$ and $(p,x_0,\dots,x_{n-1})\in B^n_i$.
Therefore, $A\cap(\pi_0^{-1}[B(p;2^{-s})]\cap(B^n_i\times[0,q)\times[0,1]^\mathbb{N}))=\emptyset$.
Hence, if $y<x_n$, then no extension of $(p,x_0,\dots,x_{n-1},y)$ is contained in $A$.
Moreover, if $(p,x_0,\dots,x_{n-1})\in B^n_i$ and $q<x_n$, then $(i,q)\not\in W_n$.
Hence, $x_n=\pi_n(\min A^{[p]})$ as desired.
Now, clearly $\min A^{[p]}=(p,x)$.
Note that $\Psi$ defines a $z$-computable continuum $A$ in a uniform manner.
The computability is ensured because we only remove a subset of $\pi_0^{-1}[B(p;2^{-s})]$ after stage $s$.
For the connectivity, if $L$ is any closed subset of $[0,1]^\mathbb{N}\setminus A$, then by compactness of $L$, it is covered by a finite collection of open sets of the form $B^n_i\times[0,q)\times[0,1]^\mathbb{N}$.
Consider $L^\complement=[0,1]^\mathbb{N}\setminus L$
If $n_0$ is a number greater than all such $n$'s, then any $y=(y_n)_{n\in\mathbb{N}}\in L^\complement$ is connected to $(y_0,y_1,\dots,y_{n_0-1},1,1,1,\dots)\in L^\complement$ by a line segment inside $L^\complement$.
Moreover, any $(y_0,\dots,y_k,\vec{1})\in L^\complement$ is connected to $(y_0,\dots,y_{k-1},1,\vec{1})\in L^\complement$ by a line segment inside $L^\complement$.
Therefore, any point $y\in L^\complement$ is connected to $\vec{1}\in L^\complement$ by a polygonal line inside $L^\complement$.
Hence, $L$ cannot separate $A$.
Consequently, $A$ is connected.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:degspec-dim}]
By Theorem \ref{theo:spectrum-main}, Lemmata \ref{lem:degspec-dim1} and \ref{lem:degspec-dim2}.
\end{proof}
The properness of $\repspb{RSW} <_{\sigma}^\mathfrak{T} [0,1]^\mathbb{N}$ can also be obtained by some relatively recent work on infinite dimensional topology: the Hilbert cube (indeed, any strongly infinite dimensional compactum) is not $\sigma$-hereditary-disconnected (see \cite{Radul}).
However, our alternative proof is naturally extended to a new construction of infinite dimensional spaces, which will be discussed in Section \ref{sec:piecewisehomeo}.
Now, one can also define the graph $n\repspb{CEA}\subseteq\mathbb{N}\times\Cantor\times[0,1]^n$ of a universal $n$-left-CEA operator (as an analogy of an $n$-REA operator) in a straightforward manner.
Then, the space $n\repspb{CEA}$ has the following properties.
\begin{theorem}
The space $n\repspb{CEA}$ is a totally disconnected $n$-dimensional Polish space.
Moreover, the countable product $n\repspb{CEA}^\mathbb{N}$ is again $n$-dimensional.
\end{theorem}
\begin{proof}
Clearly, $n\repspb{CEA}$ is totally disconnected and Polish.
To check the $n$-dimensionality, we think of $n\repspb{CEA}$ as a subspace of $[0,1]^{n+1}$ by identifying $(e,x)\in\mathbb{N}\times\Cantor$ with $\iota(0^e1x)\in[0,1]$, where $\iota$ is a computable embedding of $\Cantor$ into $[0,1]$.
We claim that $n\repspb{CEA}$ intersects with all continua $A\subseteq [0,1]^{n+1}$ such that $[0,1]\subseteq\pi_0[A]$.
We have a computable function $d$ such that the $d(e)$-th $n$-left-CEA procedure $J^n_{d(e)}(x)$ for a given input $x\in\Cantor$ outputs the value $y\in[0,1]^n$ such that $(\iota(0^e1x),y)=\min A_{e,x}^{[\iota(0^e1x)]}$, where $A_{e,x}$ is the $e$-th $x$-computable continuum in $[0,1]^{n+1}$ such that $[0,1]\subseteq\pi_0[A_{e,x}]$.
By Kleene's recursion theorem (see \cite{SoareBook}), there is $r$ such that $J^n_{d(r)}=J^n_r$.
Hence, $(\iota(0^r1x),J^n_r(x))\in n\repspb{CEA}\cap A_{e,x}$, which verifies the claim.
The claim implies that $n\repspb{CEA}$ is $n$-dimensional (see \name{van Mill} \cite{vMBook}).
To verify the second assertion, consider the (computably) continuous map $g$ from the square $n\repspb{CEA}^2$ into $n\repspb{CEA}$ such that for two points $\mathbf{x}=(e,r,x_0,\dots,x_{n-1})$ and $\mathbf{y}=(d,s,y_0,\dots,y_{n-1})$ in $n\repspb{CEA}$,
\[g(\mathbf{x},\mathbf{y})=(\langle e,d\rangle,r\oplus s,(x_0+y_0)/2,\dots,(x_{n-1}+y_{n-1})/2).\]
It is not hard to verify that $g^{-1}$ is also (computably) continuous.
Hence, $n\repspb{CEA}^2$ computably embedded into $n\repspb{CEA}$.
In particular, it is $n$-dimensional.
Then, we can conclude that $n\repspb{CEA}^\mathbb{N}$ is also $n$-dimensional (by the same argument as in \name{van Mill} \cite[Theorem 3.9.5]{vMBook2}).
\end{proof}
\subsection{Nondegenerated Continua and $\omega\repspb{CEA}$ Degrees}
We may extract computability-theoretic contents from the construction of Rubin-Schori-Walsh's strongly infinite-dimensional totally disconnected space $\repspb{RSW}$.
The standard proof of non-countable-dimensionality of $\repspb{RSW}$ (hence, the existence of a non-Turing degree in $\repspb{RSW}$) indeed implies the following computability theoretic result.
\begin{theorem}\label{thm:avoid-Turing}
There exists a nondegenerated continuum $A\subseteq[0,1]^\mathbb{N}$ in which no point has Turing degree.
\end{theorem}
\begin{proof}
Define $\repsp{H}_{\langle{i,j\rangle}}\subseteq[0,1]^\mathbb{N}$ to be the set of all points which can be identified with an element in $\Cantor$ via the witnesses $\Phi_i$ and $\Phi_j$ (as in the proof of Lemma \ref{lemma:spectrum-main}).
Then, $\bigcup_{n}\repsp{H}_{n}$ is the set of all points in $[0,1]^\mathbb{N}$ having Turing degrees.
Note that each $\repsp{H}_{n}$ is zero-dimensional since it is homeomorphic to a subspace of $\Cantor$.
Consider the hyperplane $P_n^i=[0,1]^n\times\{i\}\times[0,1]^\mathbb{N}$ for each $n\in\mathbb{N}$ and $i\in\{0,1\}$.
It is well known that $\{(P_n^0,P_n^1)\}_{n\in\mathbb{N}}$ is essential in $[0,1]^\mathbb{N}$.
Then, by using the dimension-theoretic fact (see \name{van Mill} \cite[Theorem 4.2.2 (5)]{vMBook}), we can find a separator $L_{n}$ of $(P_{n+1}^0,P_{n+1}^1)$ in $[0,1]^\mathbb{N}$ such that $L_{n}\cap\repsp{H}_{n}=\emptyset$ since $\repsp{H}_{n}$ is zero-dimensional.
Put $L=\bigcap_nL_{n}$.
Then, $L$ contains no point having Turing degree, since $L\cap\repsp{H}_n=\emptyset$ for every $n\in\mathbb{N}$.
Moreover, $L$ contains a continuum $A$ from $P_0^0$ to $P_0^1$ (see \name{van Mill} \cite[Proposition 4.7.8]{vMBook}).
\end{proof}
\begin{theorem}\label{thm:Pol-deg}
Every nondegenerated continuum $A\subseteq[0,1]^\mathbb{N}$ contains a point of an $\omega$-left-CEA continuous degree.
\end{theorem}
\begin{proof}
Note that there is $n\in\omega$ such that $P_n^{[0,p]}$ and $P_n^{[q,1]}$ with some rationals $p<q\in\mathbb{Q}$ intersect with $A$, since $A$ is nondegenerated, where $P_n^{[a,b]}=[0,1]^n\times[a,b]\times[0,1]^\mathbb{N}$.
Clearly, there is no separator $C$ of $P_n^{[0,p]}$ and $P_n^{[q,1]}$ with $C\cap A=\emptyset$ (i.e., the pair $(P_n^{[0,p]},P_n^{[q,1]})$ is essential in $A$), since $A$ is not zero-dimensional.
Therefore, the pair $(P_n^{p},P_n^{q})$ is essential in the compact subspace $A^*=A\cap P_n^{[p,q]}$.
Hence, $A^*\subseteq P_n^{[p,q]}$ contains a continuum intersecting with $P_n^{p}$ and $P_n^{q}$ (see van Mill \cite[Proposition 4.7.8]{vMBook}).
Consider a computable homeomorphism $h:P_n^{[p,q]}\cong[0,1]^\mathbb{N}$ mapping $P_n^{p}$ and $P_n^{q}$ to $P_0^{0}$ and $P_0^{1}$, respectively.
Then, $h[A^*]$ is a continuum intersecting with Rubin-Schori-Walsh's space $\repspb{RSW}$.
Hence, it has an $\omega$-left-CEA continuous degree by Theorem \ref{thm:degspec-dim}.
\end{proof}
As a corollary, we can see that every compactum $A\subseteq[0,1]^\mathbb{N}$ of positive dimension contains a point of an $\omega$-left-CEA continuous degree.
Our proof of Theorem \ref{thm:avoid-Turing} is essentially based on the fact that for any sequence of zero-dimensional spaces $\{X_i\}_{i\in\mathbb{N}}$, there exists a continuum avoiding all $X_i$'s.
Contrary to this fact, Theorem \ref{thm:Pol-deg} says that $\{X_i\}_{i\in\mathbb{N}}$ cannot be replaced with a sequence of totally disconnected spaces.
\begin{corollary}
There exists a sequence $\{X_i\}_{i\in\mathbb{N}}$ of totally disconnected subspaces of $[0,1]^\mathbb{N}$ such that every compact subspace of $Y=[0,1]^\mathbb{N}\setminus\bigcup_{i\in\mathbb{N}}X_i$ is zero-dimensional, while $Y$ is infinite dimensional.
\end{corollary}
\begin{proof}
Define $X_{\langle i,j\rangle}$ to be the set of all points which can be identified with an element in $\omega\repspb{CEA}$ via the witnesses $\Phi_i$ and $\Phi_j$.
Then, $X_{\langle i,j\rangle}$ is totally disconnected since it is homeomorphic to a subspace of $\omega\repspb{CEA}$.
Clearly, no point $Y=[0,1]^\mathbb{N}\setminus\bigcup_{i,j\in\mathbb{N}}X_{\langle i,j\rangle}$ has an $\omega$-left-CEA continuous degree.
Assume that $Z$ is a compact subspace of $Y$ of positive dimension.
Then $Z$ has a nondegenerated subcontinuum $A$.
However, by Theorem \ref{thm:Pol-deg}, $A$ contains a point of an $\omega$-left-CEA continuous degree.
\end{proof}
\section{Structure of $\sigma$-Homeomorphism Types}\label{sec:piecewisehomeo}
\subsection{Almost Arithmetical Degrees}
In this section, we generalize our proof idea in Section \ref{sec:intermediate} to construct a compact metrizable space whose points realize a given well-behaved family of {\em ``almost'' arithmetical degrees} as cospectra.
As a consequence, we obtain the following embeddability result on the structure of $\sigma$-homeomorphism types of metrizable compacta.
\begin{theorem}\label{thm:maintheorem_emb}
There is an embedding of the inclusion ordering $([\omega_1]^{\leq\omega},\subseteq)$ of countable subsets of the smallest uncountable ordinal $\omega_1$ into the $\sigma$-embeddability ordering of Pol-type Cantor manifolds.
\end{theorem}
As a corollary, there are a continuum chain and a continuum antichain of $\sigma$-homeomorphism types of Polish spaces.
As seen in the previous sections, the notion of a co-spectrum plays a role of a $\sigma$-topological invariant.
Roughly speaking, closure properties of co-spectra reflect $\sigma$-homeomorphism types of Polish spaces.
The following notion estimates the strength of closure properties of functions up to the arithmetical equivalence.
\begin{definition}\label{def:almostarithmetical}
Let $g$ and $h$ be total Borel measurable functions from $\Cantor$ into $\Cantor$.
\begin{enumerate}
\item We inductively define $g^0(x)=x$ and $g^{n+1}(x)=g^{n}(x)\oplus g(g^{n}(x))$.
\item For every oracle $r\in \Cantor$, consider the following jump ideal defined as
\[\mathcal{J}_a(g,r)=\{z\in \Cantor:(\exists n\in\mathbb{N})\;x\leq_ag^n(r)\},\]
where $\leq_a$ denotes the arithmetical reducibility (see \cite{OdiBook1}), that is, $p\leq_aq$ is defined by $p\leq_Tq^{(m)}$ for some $m\in\mathbb{N}$.
\item A function $g$ is {\em almost arithmetical reducible to} a function $h$ (written as $g\leq_{aa}h$) if for any $r\in\Cantor$ there is $x\in\Cantor$ with $x\geq_Tr$ such that
\[\mathcal{J}_a(g,x)\subseteq\mathcal{J}_a(h,x).\]
\item Let $\mathcal{G}$ and $\mathcal{H}$ be countable sets of total functions.
We say that $\mathcal{G}$ is {\em $aa$-included in} $\mathcal{H}$ (written as $\mathcal{G}\subseteq_{aa}\mathcal{H}$) if for all $g\in\mathcal{G}$, there is $h\in\mathcal{H}$ such that $g\equiv_{aa}h$ (i.e., $g\leq_{aa}h$ and $h\leq_{aa}g$).
\end{enumerate}
\end{definition}
A function $g:\Cantor\to \Cantor$ is said to be {\em monotone} if $x\leq_Ty$ implies $g(x)\leq_Tg(y)$.
An {\em oracle $\mathbf{\Pi}^0_2$-singleton} is a total function $g:\Cantor\to \Cantor$ whose graph is $G_\delta$.
Clearly, every oracle $\mathbf{\Pi}^0_2$-singleton is Borel measurable, whereas there is no upper bound of Borel ranks of oracle $\mathbf{\Pi}^0_2$-singletons.
For instance, the $\alpha$-th Turing jump $j_\alpha(x)=x^{(\alpha)}$ is a monotone oracle $\mathbf{\Pi}^0_2$-singleton for every computable ordinal $\alpha$ (see \cite{OdiBook1}).
The following is the key lemma in our proof.
\begin{lemma}[Realization Lemma]\label{thm:refle}
There is a map $\repspb{Rea}$ transforming each countable set of monotone oracle $\mathbf{\Pi}^0_2$-singletons into a Polish space such that
\[\repspb{Rea}(\mathcal{G})\leq^\mathfrak{T}_\sigma\repspb{Rea}(\mathcal{H})\;\Longrightarrow\;\mathcal{G}\subseteq_{aa}\mathcal{H}.\]
\end{lemma}
\subsection{Construction}
We construct a Polish space whose co-spectrum codes almost arithmetical degrees contained in a given countable set $\mathcal{G}$ of oracle $\Pi^0_2$ singletons.
For notational simplicity, given $x\in[0,1]^\mathbb{N}$, we write $x_n$ for the $n$-th coordinate of $x$, and moreover, $x_{<n}$ and $x_{\leq n}$ for $(x_i)_{i<n}$ and $(x_i)_{i\leq n}$ respectively.
\begin{definition}\label{def:reaspace}
Let $\mathcal{G}=(g_n)_{n\in\mathbb{N}}$ be a countable collection of oracle $\mathbf{\Pi}^0_2$-singletons.
The space $\repspb{\omega CEA}(\mathcal{G})$ consists of $(n,d,e,r,x)\in\mathbb{N}^3\times \Cantor\times[0,1]^\mathbb{N}$ such that for every $i$,
\begin{enumerate}
\item either $x_i=g_n^i(r)$, or
\item there are $u\leq v\leq i$ such that $x_i\in [0,1]$ is the $e$-th left-c.e.~real in $\langle r,x_{<i},x_{l(u)}\rangle$ and $x_{l(u)}=g_n^{l(u)}(r)$, where $l(u)=\Phi_d(u,r,x_{<v})\geq i$.
\end{enumerate}
Here, $x_{<i}$ is an abbreviation for the sequence $(x_0,\dots,x_{i-1})$.
Moreover, for a set $P\subseteq[0,1]^\mathbb{N}$, define $\repspb{\omega CEA}(\mathcal{G},P)$ to be the set of all points $(d,e,r,x)\in\repspb{\omega CEA}(\mathcal{G})$ with $(r,x)\in P$.
\end{definition}
\begin{lemma}
Suppose that $\mathcal{G}$ is an oracle $\mathbf{\Pi}^0_2$-singleton, and $P$ is a $\mathbf{\Pi}^0_2$ subset of $[0,1]^\mathbb{N}$.
Then, $\repspb{\omega CEA}(\mathcal{G},P)$ is Polish.
\end{lemma}
\begin{proof}
It suffices to show that $\repspb{\omega CEA}(\mathcal{G})$ is $\mathbf{\Pi}^0_2$.
The condition (1) in Definition \ref{def:reaspace} is clearly $\mathbf{\Pi}^0_2$.
Let $\forall a\exists b>a\;G(a,b,n,l,r,x)$ be a $\mathbf{\Pi}^0_2$ condition representing $x=g_n^l(r)$, and $l(u)[s]$ be the stage $s$ approximation of $l(u)$.
The condition (2) is equivalent to the statement that there are $u\leq v\leq i$ such that
\begin{align*}
(\forall t\in\mathbb{N})(\exists s>t)\;&l(u)[s]\downarrow\geq i,\;d(x_i,J_{e,s}^{i+1}(r,x_{<i},x_{l(u)[s]}))<2^{-t},\\
&\mbox{ and }G(t,s,n,l(u)[s],r,x_{l(u)[s]}).
\end{align*}
Clearly, this condition is $\mathbf{\Pi}^0_2$.
\end{proof}
\begin{remark}
The space $\repspb{\omega CEA}(\mathcal{G})$ is totally disconnected for any countable set $\mathcal{G}$ of oracle $\mathbf{\Pi}^0_2$ singletons, since for any fixed $(n,d,e,r)\in\mathbb{N}^3\times \Cantor$, its extensions form a finite-branching infinite tree $T\subseteq[0,1]^{<\omega}$.
\end{remark}
Recall from Section \ref{subsec:proof-main-half} that \name{Miller} \cite[Lemma 6.2]{miller2} constructed a $\Pi^0_1$ set ${\rm Fix}(\Psi)$ such that ${\rm coSpec}(x)=\{x_i:i\in\mathbb{N}\}$ for every $x=(x_i)_{i\in\mathbb{N}}\in{\rm Fix}(\Psi)$.
By Lemma \ref{lemma:Miller-6-2}, without loss of generality, we may assume that ${\rm Fix}(\Psi)\cap\pi_0^{-1}\{r\}\not=\emptyset$ for every $r\in [0,1]$.
Now, consider the space $\repspb{Rea}(\mathcal{G})=\repspb{\omega CEA}(\mathcal{G},{\rm Fix}(\Psi))$.
To state properties of $\repspb{Rea}(\mathcal{G})$, for an oracle $\mathbf{\Pi}^0_2$-singleton $g$ and an oracle $r\in \Cantor$, we use the following Turing ideal:
\[\mathcal{J}_T(g,r)=\{z\in \Cantor:(\exists n\in\mathbb{N})\;x\leq_Tg^n(r)\}.\]
The following is the key lemma, which states that any collection of jump ideals generated by countably many oracle $\mathbf{\Pi}^0_2$-singletons has to be the degree co-spectrum of a Polish space up to the almost arithmetical equivalence!
\begin{lemma}\label{lem:keyref}Suppose that $\mathcal{G}=(g_n)_{n\in\mathbb{N}}$ is a countable set of oracle $\mathbf{\Pi}^0_2$-singletons.
\begin{enumerate}
\item For every $x\in\repspb{Rea}(\mathcal{G})$, there are $r\in \Cantor$ and $n\in\mathbb{N}$ such that
\[\mathcal{J}_T(g_n,r)\subseteq{\rm coSpec}(x)\subseteq\mathcal{J}_a(g_n,r).\]
\item For every $r\in \Cantor$ and $n\in\mathbb{N}$, there is $x\in\repspb{Rea}(\mathcal{G})$ such that
\[\mathcal{J}_T(g_n,r)\subseteq{\rm coSpec}(x).\]
\end{enumerate}
\end{lemma}
\begin{proof}[Proof of Lemma \ref{lem:keyref} (1)]
We have $(r,x)\in{\rm Fix}(\Psi)$ for every $(n,d,e,r,x)\in\repspb{Rea}(\mathcal{G})$.
For every $i\in\mathbb{N}$, we inductively assume that for every $j<i$, $x_j$ is arithmetical in $g_n^k(r)$ for some $k\in\mathbb{N}$.
Now, either $x_i=g_n^i(r)$ or $x_i$ is left-c.e.~in $(r,x_{<i},g_n^l(r))$ for some $l$.
In both cases, $x_i$ is arithmetical in $g_n^k(r)$ for some $k$.
Moreover, $x_i=g_n^i(r)$ for infinitely many $i\in\mathbb{N}$, since either $x_i=g^n_i(r)$ holds or there is $l\geq i$ such that $x_l=g^l_n(r)$ by the condition (2) in Definition \ref{def:reaspace}.
Therefore, $g_n^k(r)\leq_Tx$ for all $k\in\mathbb{N}$.
\end{proof}
To verify the assertion (2) in Lemma \ref{lem:keyref}, indeed, for any $n\in\mathbb{N}$, we will construct indices $d$ and $e$ such that for every $r\in \Cantor$, there is $x$ with $(n,d,e,r,x)\in\repspb{Rea}(\mathcal{G})$, where $x_i=g_n^i(r)$ for infinitely many $i\in\mathbb{N}$.
The $e$-th left-c.e.~procedure $J^{i+1}_e(r,x_{<i},x_{l(u)})$ is a simple procedure extending $r,x_{<i},x_{l(u)}$ to a fixed point of $\Psi$.
The function $\Phi_d$ searches a {\em safe coding location} $c(n)$ from a given {\em name} of $x_{\leq c(n-1)}$, where $c(n-1)$ is the previous coding location.
To make sure the search of the next coding location is bounded, as in Definition \ref{def:reaspace}, we have to restrict the set of names of a $v$-tuple $x_{<v}$ to at most $v+1$ candidates.
It is known that a separable metrizable space is at most $n$-dimensional if and only if it is the union of $n+1$ many zero-dimensional subspaces (see \cite{EngBook}).
We say that an admissibly represented Polish space is {\em computably at most $n$-dimensional} if it is the union of $n+1$ many subspaces that are computably homeomorphic to subspaces of $\mathbb{N}^\mathbb{N}$.
\begin{lemma}\label{lem:compu-n-dim}
Suppose that $(\repsp{X},\rho_X)$ is a computably at most $n$-dimensional admissibly represented space.
Then, there is a partial computable injection $\nu_X:\subseteq (n+1)\times\repsp{X}\to\mathbb{N}^\mathbb{N}$ such that for every $x\in\repsp{X}$, there is $k\leq n$ such that $(k,x)\in{\rm dom}(\nu_X)$ and $\rho_X\circ\nu_X(k,x)=x$.
\end{lemma}
\begin{proof}
By definition, $\repsp{X}$ is divided into $n+1$ many subspaces $S_0,\dots,S_n$ such that $S_k$ is homeomorphic to $N_k\subseteq\mathbb{N}^\mathbb{N}$ via computable maps $\tau_k$ and $\tau_k^{-1}$.
Then, the partial computable injection $\tau_k^{-1}:\subseteq\mathbb{N}^\mathbb{N}\to\repsp{X}$ has a computable realizer $\tau^*_k$, i.e., $\tau_k^{-1}=\rho_X\circ\tau^*_k$.
Define $\nu_X(k,x)=\tau^*_k\circ\tau_k(x)$ for $x\in S_k$.
Then, we have $\rho_X\circ\nu_X(k,x)=\tau_k^{-1}\circ\tau_k(x)=x$ for $x\in S_k$.
\end{proof}
The Euclidean $n$-space $\mathbb{R}^n$ is clearly computably $n$-dimensional, e.g., let $S_k$ be the set of all points $x\in\mathbb{R}^n$ such that exactly $k$ many coordinates are irrational.
Furthermore, one can effectively find an index of $\nu_n:=\nu_{\mathbb{R}^n}$ in Lemma \ref{lem:compu-n-dim} uniformly in $n$.
Hereafter, let $\rho_i$ be the usual Euclidean admissible representation of $\mathbb{R}^i$.
Now, a coding location $c(n)$ will be obtained as a fixed point in the sense of Kleene's recursion theorem (see \cite{SoareBook}).
Hence, one can effectively find such a location in the following sense.
\begin{lemma}[\name{Miller} {\cite[Lemma 9.2]{miller2}}]
Suppose that $(r,x_{<i})$ can be extended to a fixed point of $\Psi$, and fix a partial computable function $\nu$ which sends $x_{<i}$ to its name, i.e., $\rho_i\circ\nu(x_{<i})=(x_{<i})$.
From an index $t$ of $\nu$ and the sequence $x_{<i}$, one can effectively find a location $p=\Gamma(t,r,x_{<i})$ such that for every real $y$, the sequence $(r,x_{<i})$ can be extended to a fixed point $(r,x)$ of $\Psi$ such that $x_p=y$.
\end{lemma}
Let $t(n,k)$ be an index of the partial computable function $x\mapsto\nu_n(k,x)$.
We define $\Phi_d(u,r,x_{<v})$ to be $\Gamma(t(v,u),r,x_{<v})$ for every $u\leq v$.
Note that indices $d$ and $e$ do not depend on $g_n$.
\begin{proof}[Proof of Lemma \ref{lem:keyref} (2)]
Now, we claim that for every $r\in \Cantor$ and $n\in\mathbb{N}$, there is $x$ with $(n,d,e,r,x)\in\repspb{Rea}(\mathcal{G})$, where $x_i=g_n^i(r)$ for infinitely many $i\in\mathbb{N}$.
We follow the argument by \name{Miller} \cite[Lemma 9.2]{miller2}.
Suppose that $i$ is a coding location of $g_n^i(r)$, and $(r,x_{\leq i})$ is extendible to a fixed point of $\Psi$.
Then, there is $k\leq i+1$ such that $p=\Phi_d(k,r,x_{\leq i})$ is defined, and then we set $x_p=g_n^p(r)$.
By the property of $\Phi_d$, $(r,x_{\leq i},x_p)$ can be extended to a fixed point of $\Psi$.
Then, the $e$-th left-c.e.~procedure automatically produces $x_{\leq p}$ which is extendible to a fixed point of $\Psi$.
Note that the condition (2) in Definition \ref{def:reaspace} is ensured via $u=k$, $v=i+1$, and $l(u)=p$.
Eventually, we obtain $(r,x)\in{\rm Fix}(\Psi)$ such that $z=(n,d,e,r,x)\in\repspb{Rea}(\mathcal{G})$.
Clearly, $g_n^k(r)\in{\rm coSpec}(z)$ for every $k\in\mathbb{N}$, since ${\rm coSpec}(z)$ is a Turing ideal, and $g_n^k(r)\leq_Tg_n^{k+1}(r)$.
Consequently, $\mathcal{J}_T(g_n,r)\subseteq{\rm coSpec}(z)$.
\end{proof}
\begin{proof}[Proof of Lemma \ref{thm:refle}]
Suppose that $\repspb{Rea}(\mathcal{G})\leq^\mathfrak{T}_\sigma\repspb{Rea}(\mathcal{H})$.
Then, $\mathbb{N}\times\repspb{Rea}(\mathcal{G})\leq^\mathfrak{T}_\sigma\mathbb{N}\times\repspb{Rea}(\mathcal{H})$, and by Theorem \ref{lemma:spectrum-main} and Observation \ref{obs:main-cospec-inv}, the degree cospectrum of $\repspb{Rea}(\mathcal{G})$ is a sub-cospectrum of that of $\repspb{Rea}(\mathcal{H})$ up to an oracle $p$.
Fix enumerations $\mathcal{G}=(g_n)_{n\in\mathbb{N}}$ and $\mathcal{H}=(h_n)_{n\in\mathbb{N}}$.
\begin{claim}
For any $n$ and $u$, there are $m$ and $v$ such that $\mathcal{J}_a(g_n,u)=\mathcal{J}_a(h_m,v)$.
\end{claim}
By Lemma \ref{lem:keyref} (2), for any $n$ and $u\geq_Tp$, there is $x\in\repspb{Rea}(\mathcal{G})$ such that $\mathcal{J}_T(g_n,u)\subseteq{\rm coSpec}(x)\subseteq\mathcal{J}_a(g_n,u)$.
Then, there is $y\in\repspb{Rea}(\mathcal{H})$ such that ${\rm coSpec}^p(x)={\rm coSpec}^p(y)$.
We may assume that $p\leq_My$, otherwise $(y,p)$ has Turing degree by Lemma \ref{lem:JM-almosttotal}.
By Lemma \ref{lem:keyref} (1), there exist $m$ and $v$ such that $\mathcal{J}_T(h_m,v)\subseteq{\rm coSpec}(y)\subseteq\mathcal{J}_a(h_m,v)$.
Now, ${\rm coSpec}(x)={\rm coSpec}(y)$ holds, and note that $\mathcal{J}_T(h_m,v)\subseteq\mathcal{J}_a(g_n,u)$ implies $\mathcal{J}_a(h_m,v)\subseteq\mathcal{J}_a(g_n,u)$.
This verifies the claim.
For a fixed $n$, $\beta_n(u)$ chooses $m$ fulfilling the above claim for some $v$.
It is not hard to see that there is $m(n)$ such that $\beta_n(u)=m(n)$ for cofinally many $u$.
Then, for cofinally many $v$, there is $u$ such that $\mathcal{J}_a(g_n,u\oplus v)=\mathcal{J}_a(h_{m(n)},u\oplus v)$ by monotonicity.
Therefore, $g_n\equiv_{aa}h_{m(n)}$.
Consequently, $\mathcal{G}\subseteq_{aa}\mathcal{H}$.
\end{proof}
\begin{lemma}
For any $\mathcal{G}$, there exists a Pol-type Cantor manifold $\repsp{Z}(\mathcal{G})$ such that $\omega\repspb{CEA}\oplus\repspb{Rea}(\mathcal{G})\equiv_\sigma^\mathfrak{T}\repsp{Z}(\mathcal{G})$.
\end{lemma}
\begin{proof}
Recall from Theorem \ref{thm:degspec-dim} that $\omega\repspb{REA}$ is $\sigma$-homeomorphic to a strongly infinite dimensional space $\repspb{RSW}$.
Let $\repsp{R}_0$ and $\repsp{R}_1$ be homeomorphic copies of $\repspb{RSW}$, and let $\repsp{X}$ be a compactification of $R_0\oplus R_1\oplus\repspb{Rea}(\mathcal{G})$ in the sense of \name{Lelek}.
Then, $\repsp{X}$ is $\sigma$-homeomorphic to $\omega\repspb{CEA}\oplus\repspb{Rea}(\mathcal{G})$.
We follow the construction of \name{El\.{z}abieta Pol} \cite[Example 4.1]{EPol96}.
Now, $\repsp{R}_0$ has a hereditarily strongly infinite dimensional subspace $\repsp{Y}$ \cite{Rub80}.
Choose a point $p\in\repsp{Y}$ and a closed set $F\subseteq\repsp{Y}$ containing $p$ such that every separator between $p$ and ${\rm cl}_\repsp{X}F$ is strongly infinite dimensional as in \cite[Example 4.1 (A)]{EPol96}.
Define $\repsp{K}=\repsp{X}/{\rm cl}_\repsp{X}F$ as in \cite[Example 4.1 (A)]{EPol96}.
To see that $\repsp{K}$ is $\sigma$-homeomorphic to $\repsp{X}$, we note that ${\rm cl}_{\repsp{X}}F\cap(\repsp{R}_1\cup\repspb{Rea}(\mathcal{G}))=\emptyset$ since $\repsp{R}_0$, $\repsp{R}_1$ and $\repspb{Rea}(\mathcal{G})$ are separated in $\repsp{X}$.
Therefore, ${\rm cl}_{\repsp{X}}F$ is covered by the union of $\repsp{R}_0$ (which is homeomorphic to $\repsp{R}_1$) and a countable dimensional space.
Define $\repsp{Z}$ as a Pol-type Cantor manifold in \cite[Example 4.1 (C)]{EPol96}.
Then, $\repsp{Z}(\mathcal{G}):=\repsp{Z}$ is the union of a finite dimensional space and countably many copies of $\repsp{K}$.
Consequently, $\repsp{Z}(\mathcal{G})$ is $\sigma$-homeomorphic to $\repspb{Rea}(\mathcal{G})$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:maintheorem_emb}]
Let $S$ be a countable subset of $\omega_1$.
Note that $\sup S$ is countable by regularity of $\omega_1$.
Then, there is an oracle $p$ such that $\sup S<\omega_1^{{\rm CK},p}$, where $\omega_1^{{\rm CK},p}$ is the smallest noncomputable ordinal relative to $p$.
Now, the $\alpha$-th Turing jump operator $j^p_\alpha$ for $\alpha<\omega_1^{{\rm CK},p}$ is defined via a $p$-computable coding of $\alpha$.
By Spector's uniqueness theorem, the Turing degree of $j^p_\alpha(x)$ for $x\geq_Tp$ is independent of the choice of coding of $\alpha$, so is $\mathcal{J}_a(j^p_\alpha,x)$.
Therefore, we simply write $j_\alpha$ for $j^p_\alpha$.
Define $\mathcal{G}_S=\{j_{\omega^{1+\alpha}}:\alpha\in S\}$.
We show that $S\subseteq T$ if and only if $\mathcal{G}_S\subseteq_{aa}\mathcal{G}_T$.
Suppose $\alpha\not=\beta$, say $\alpha<\beta$.
Clearly, $j_{\omega^{\alpha}}\leq_{aa}j_{\omega^\beta}$.
Suppose for the sake of contradiction that $j_{\omega^{\beta}}\leq_{aa}j_{\omega^\alpha}$.
Then, in particular, for every $x\leq_a\emptyset^{(\omega^\beta\cdot t)}$ with $t\in\mathbb{N}$, we must have $\emptyset^{(\omega^\beta\cdot(t+1))}\leq_ax^{(\omega^{\alpha}\cdot m)}$ for some $m\in\mathbb{N}$.
Thus, there is $n$ such that $\emptyset^{(\omega^\beta\cdot t+\omega^\beta)}\leq_T\emptyset^{(\omega^\beta\cdot t+\omega^{\alpha}\cdot m+n)}<_T\emptyset^{(\omega^\beta\cdot t+\omega^{\alpha+1})}$.
This is a contradiction.
\end{proof}
\begin{corollary}
There exists a collection $(\repsp{X}_\alpha)_{\alpha<2^{\aleph_0}}$ of continuum many Pol-type Cantor manifolds satisfying the following conditions:
\begin{enumerate}
\item If $\alpha\not=\beta$, then $\repsp{X}_\alpha$ is not $n$-th level Borel isomorphic to $\repsp{X}_\beta$ for all $n\in\mathbb{N}$.
\item If $\alpha\not=\beta$, then the Banach algebra $\mathcal{B}^*_n(\repsp{X}_\alpha)$ is not linearly isometric (not ring isomorphic etc.) to $\mathcal{B}^*_n(\repsp{X}_\beta)$ for all $n\in\mathbb{N}$.
\end{enumerate}
\end{corollary}
\begin{proof}
By Theorems \ref{theo:spectrum-main} and \ref{thm:maintheorem_emb}.
Here, we note that if $\repsp{X}$ is $n$-th level Borel isomorphic to $\repsp{Y}$, then $\mathbb{N}\times\repsp{X}$ is again $n$-th level Borel isomorphic to $\mathbb{N}\times\repsp{Y}$.
\end{proof}
\subsection{An Ordinal Valued $\sigma$-Topological Invariant}
Although we constructed continuum many mutually different spaces, it is difficult to discern dimension-theoretic differences among these spaces.
For instance, all of our spaces have the same transfinite Steinke dimensions \cite{ArChPu00,Radul}, game dimensions \cite{FedOsi08}, and so on (see \name{Chatyrko} and \name{Hattori} \cite{ChaHat13} for the thorough treatment of the notion of various kinds of transfinite dimensions).
We now focus on an $\aleph_1$ chain of $\sigma$-homeomorphism types of Polish spaces:
\[\mathbb{R}^n<^\mathfrak{T}_\sigma\repspb{Rea}(\{j_1\})<^\mathfrak{T}_\sigma\repspb{Rea}(\{j_\omega\})<^\mathfrak{T}_\sigma\repspb{Rea}(\{j_{\omega^2}\})<^\mathfrak{T}_\sigma\repspb{Rea}(\{j_{\omega^3}\})<^\mathfrak{T}_\sigma\dots\]
Our key observation was that closure properties of Scott ideals reflects piecewise homeomorphism types of Polish spaces.
The first purpose here is to provide a topological understanding of our method.
\begin{definition}
Let $\repsp{X}$ be a topological space.
Let $\mathcal{C}(_\subseteq\repsp{X},\mathbb{R})$ denote the collection of all continuous functions from subspaces of $\repsp{X}$ into $\mathbb{R}$.
Suppose that a collection $\mathcal{H}(_\subseteq\mathbb{R},\mathbb{R})$ of functions from subspaces of $\mathbb{R}$ into $\mathbb{R}$ is given.
A countable set $F\subseteq\mathcal{C}(_\subseteq\repsp{X},\mathbb{R})$ {\em avoids $\mathcal{H}$ on $\repsp{X}$} if for any countable set $G\subseteq\mathcal{C}(_\subseteq\repsp{X},\mathbb{R})$ and countable set $H\subseteq\mathcal{H}(_\subseteq\mathbb{R},\mathbb{R})$, there exists a point $x\in\repsp{X}$ such that
\[
(\forall g\in G|_x)(\exists f\in F|_x)(\forall h\in H|_{g(x)})\;f(x)\not=h\circ g(x),
\]
where $E|_y:=\{f\in E:y\in{\rm dom}(f)\}$.
We then say that $\repsp{X}$ is {\em $\alpha$-avoiding} if there is a countable set that avoids $\mathcal{B}_\alpha$ on $\repsp{X}$, where $\mathcal{B}_\alpha$ is the class of all Baire $\alpha$ functions, and $\mathcal{B}_0=\mathcal{C}$
\end{definition}
The {\em jump dimension} ${\rm jdim}(\repsp{X})$ of $\repsp{X}$ is the supremum of countable ordinals $\alpha<\omega_1$ such that $\repsp{X}$ is $\beta$-avoiding for all $\beta<\omega^\alpha$, where $\omega^0=1$.
If such $\alpha$ does not exist, then ${\rm jdim}(\repsp{X})=-1$.
Hence, ${\rm jdim}(\repsp{X})=-1$ if and only if $\repsp{X}$ is not $0$-avoiding.
This notion provides a new characterization of countable-dimensionality for Polish spaces, and we also see that the jump dimension is invariant under $\sigma$-homeomorphism.
\begin{theorem}\label{thm:jdim-countable-dimension}
Let $\repsp{X}$ and $\repsp{Y}$ be separable metrizable spaces.
\begin{enumerate}
\item $\repsp{X}$ is countable dimensional if and only if ${\rm jdim}(\repsp{X})=-1$ (i.e., $\repsp{X}$ is not $0$-avoiding).
\item If $\repsp{X}\leq_\sigma^\mathfrak{T}\repsp{Y}$, then ${\rm jdim}(\repsp{X})\leq{\rm jdim}(\repsp{Y})$.
\item For every countable ordinal $\alpha$, there is a Pol-type Cantor manifold $\repsp{X}$ such that ${\rm jdim}(\repsp{X})=\alpha$.
\end{enumerate}
\end{theorem}
To show Theorem \ref{thm:jdim-countable-dimension}, we need the following effective interpretation of jump-dimension.
We say that $\mathcal{I}\subseteq\Cantor$ is {\em $\alpha$-principal} if there is $p\in\mathcal{I}$ such that $q\leq_Tp^{(\alpha)}$ for all $q\in\mathcal{I}$.
\begin{lemma}\label{lem:jdimequiv}
An admissibly represented separable metrizable space $\repsp{X}$ is $\alpha$-avoiding if and only if relative to some oracle $r$, for all $z\in\Cantor$ there is a point $x\in\repsp{X}$ such that $z\in{\rm coSpec}^r(x)$ and ${\rm coSpec}^r(x)$ is not $\alpha$-principal.
\end{lemma}
\begin{proof}
Suppose that $F=\{f_n\}_{n\in\omega}$ is a countable set avoiding $\mathcal{B}_{\alpha}$ on $\repsp{X}$.
Then, almost all oracles $r$ satisfy that $\repsp{X}$ is $r$-computably embedded into Hilbert cube, $\alpha<\omega_1^{{\rm CK},r}$ and every $f\in F$ is computable relative to $r$.
Then, clearly $\{f(x):{f\in F},\;x\in{\rm dom}(f)\}\subseteq{\rm coSpec}^r(x)$ holds for all $x\in\repsp{X}$.
Fix $z\geq_Tr$.
Let $G$ be the set of all partial $z$-computable functions from $\repsp{X}$ into $\mathbb{R}$, and $H$ be the set of all $r$-effective Baire $\alpha$ functions (i.e., functions of the form $p\mapsto\Phi_e^{p^{(\alpha)}}$).
Let $x\in X$ be a point witnessing the avoiding property of $F$ for given $G$ and $H$.
Now, every $p\in{\rm coSpec}^r(x)$ is of the form $g(x)$ for some $g\in G$.
By the avoiding property of $F$, there is $f\in F$ such that $f(x)\not=h\circ g(x)$ for all $h\in H$.
In other words, for all $p\in{\rm coSpec}^r(x)$, we have $f(x)\not=\Phi_e^{p^{(\alpha)}}$ for all $e\in\mathbb{N}$, i.e., $f(x)\not\leq_Tp^{(\alpha)}$.
Then, put $q:=f(x)\in{\rm coSpec}^r(x)$.
We claim that $z\in{\rm coSpec}^r(x)$, i.e., $z\not\leq_M(x,r)$.
Otherwise, by Lemma \ref{lem:JM-almosttotal}, $(x,r,z)$ has a Turing degree since $(x,r)$ has a continuous degree.
Therefore, there is $g\in G$ such that $g(x)\equiv_M(x,r,z)$.
Thus, $f(x)$ is of the form $h\circ g(x)$ for some $r$-computable function $h\in H$.
Conversely, suppose that the condition in Lemma \ref{lem:jdimequiv} holds for $r$.
We show that the set $F$ of all partial $r$-computable functions from $\repsp{X}$ into $\mathbb{R}$ avoids $\mathcal{B}_{\alpha}$ on $\repsp{X}$.
Given $G$ and $H$, one can find an oracle $z\geq_Tr$ such that every $g\in G$ is $z$-computable and every $h\in H$ is $z$-effectively Baire $\alpha$.
Then we have a point $x\in\repsp{X}$ with $z\in{\rm coSpec}^r(x)$ such that for all $p\in {\rm coSpec}^r(x)$, there exists $q\in{\rm coSpec}^r(x)$ such that $q\not\leq_T(z\oplus p)^{(\alpha)}$.
Since ${\rm coSpec}^r(x)={\rm coSpec}^z(x)$, ${\rm coSpec}^r(x)$ contains $g(x)$ for all $g\in G$, and hence, $z\oplus g(x)\in{\rm coSpec}^r(x)$.
Therefore, $q\not=h\circ g(x)$ since $h$ is of the form $p\mapsto\Phi_e^{(z\oplus p)^{(\alpha)}}$ for some index $e$.
Therefore, $F$ avoids $\mathcal{B}_\alpha$ since $\{f(x):{f\in F},\;x\in{\rm dom}(f)\}={\rm coSpec}^r(x)$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:jdim-countable-dimension}]
By Miller's result \cite{miller2}, we can deduce that a point $x\in[0,1]^\mathbb{N}$ has a Turing degree relative to $r$ if and only if ${\rm coSpec}^r(x)$ is principal (i.e., $0$-principal).
Hence, if $\repsp{X}$ is countable dimensional, all cospectra are $0$-principal up to some oracle.
Therefore, $\repsp{X}$ is not $0$-avoiding.
Conversely, suppose that $\repsp{X}$ is not countable dimensional.
We claim that for all $z\in\Cantor$, there is $x\in\repsp{X}$ such that $z\leq_Mx$ and $x$ has no Turing degree.
Otherwise, $(x,z)$ has a Turing degree by Lemma \ref{lem:JM-almosttotal}.
In this case, ${\rm Spec}^z(\repsp{X})\subseteq\mathcal{D}_T$.
This implies that $\repsp{X}$ is countable dimensional.
Now, our claim clearly implies the desired condition by Lemma \ref{lem:jdimequiv}.
The second assertion follows from Lemma \ref{lem:jdimequiv} since the cospectrum is invariant under $\sigma$-homeomorphism by Observation \ref{obs:main-cospec-inv}.
Now, we show the third assertion.
We first see that the jump-dimension of $\repspb{Rea}(j_{\omega^{\alpha}})$ is $\alpha+1$.
We have ${\rm jdim}(\repspb{Rea}(j_{\omega^{\alpha}}))\geq\alpha+1$ because for any $z$, there is $x\in\repspb{Rea}(j_{\omega^{\alpha}})$ such that $\mathcal{J}_T(j_{\omega^\alpha},z)\subseteq{\rm coSpec}(x)\subseteq\mathcal{J}_a(j_{\omega^\alpha},z)$ by Lemma \ref{lem:keyref}.
If $y\in{\rm coSpec}(x)$, then $y\in \mathcal{J}_a(j_{\omega^\alpha},z)$.
Therefore, $y^{(\omega^\alpha+n)}\in \mathcal{J}_T(j_{\omega^\alpha},z)\subseteq{\rm coSpec}(x)$ for all $n\in\mathbb{N}$.
Hence, ${\rm coSpec}^z(x)={\rm coSpec}(x)$ is closed under the $\beta$-th Turing jump for all $\beta<\omega^{\alpha+1}$.
To see ${\rm jdim}(\repspb{Rea}(j_{\omega^{\alpha}}))<\alpha+2$, we note for any $x\in\repspb{Rea}(j_{\omega^{\alpha}})$ that $\mathcal{J}_T(j_{\omega^\alpha},z)\subseteq{\rm coSpec}(x)\subseteq\mathcal{J}_a(j_{\omega^\alpha},z)$ for some $z$ by Lemma \ref{lem:keyref}.
Then, $z\in{\rm coSpec}(x)$, but ${\rm coSpec}(x)$ is covered by the Turing ideal generated by $z^{(\omega^{\alpha+1})}$.
If $\alpha$ is a limit ordinal, then consider $\repsp{X}=\repspb{Rea}(\{j_{\omega^{\beta}}\}_{\beta<\alpha})$.
\end{proof}
\begin{example}
\begin{enumerate}
\item The jump-dimension of Hilbert cube $[0,1]^\mathbb{N}$ is $\omega_1$.
This is because every countable Scott ideal is realized as a cospectrum in the Hilbert cube \cite{miller2} and by Lemma \ref{lem:jdimequiv}.
\item The jump-dimension of $\repspb{Rea}(\mathcal{G})$ cannot be $\omega_1$ for every countable set of $\mathcal{G}$ of oracle $\mathbf{\Pi}^0_2$ singletons.
This is because every oracle $\mathbf{\Pi}^0_2$ singleton is Borel measurable.
Therefore, there is a countable ordinal $\alpha$ which bounds all Borel ranks of functions contained in $\mathcal{G}$ since $\aleph_1$ is regular.
Thus, for any $g\in\mathcal{G}$, we have $g(r)\leq_T(r\oplus z)^{(\alpha)}$ for some oracle $z$.
One can see that the cospectrum of a point in $\repspb{Rea}(\mathcal{G})$ is $(\alpha\cdot\omega)$-principal.
\item We have $0\leq{\rm jdim}({\rm \omega CEA})\leq 1$.
This is because the cospectrum of a point in $\omega{\rm CEA}$ is $\omega$-principal by the proof of Lemma \ref{lem:Scott-ideal}.
\end{enumerate}
\end{example}
\section{Internal Characterization of Degree Structures}\label{sec:internalcharacterization}
\subsection{Characterizing Continuous Degrees Through a Metrization Theorem}
In this section, we will provide a rather strong metrization theorem, namely that any computably admissible space with an effectively fiber-compact representation can be computably embedded in a computable metric space. Our result is a slightly stronger version of a result by \name{Schr\"oder} that an admissible space with a proper representation is metrizable \cite{schroder6}. This also gives us a characterization of the continuous degrees inside the Medvedev degrees that does not refer to represented spaces at all.
For some closed set $A \subseteq \Cantor$, let $T(A) \subseteq \Cantor$ be the set of trees for $A$, where each infinite binary tree is identified with an element of Cantor space. Now let $\delta : \subseteq \Cantor \to \repsp{X}$ be an {\em effectively fiber-compact representation}, i.e.~let $x \mapsto \delta^{-1}(\{x\}) : \repsp{X} \to \mathcal{A}(\Cantor)$ be computable. Then $T(\delta^{-1}(\{x\})) \leq_M \delta^{-1}(\{x\})$. If $\delta$ is computably admissible, we also have $\delta^{-1}(\{x\}) \leq_M T(\delta^{-1}(\{x\}))$. Note that being effectively fiber-compact is equivalent to being effectively proper, as the union of compactly many compact sets is compact. It is known that any computable metric space has a computably admissible effectively fiber-compact representation (e.g.~\cite{weihrauchf}).
We shall prove that the converse holds, too.
\begin{theorem}\label{thm:internal_characterization}
A represented space $\repsp{X}$ admits a computably admissible effectively fiber-compact representation iff $\repsp{X}$ embeds computably into a computable metric space.
\end{theorem}
\begin{corollary}
\label{corr:contdegreecharac}
$A \subseteq \Cantor$ has continuous degree iff there is $B \in \mathcal{A}(\Cantor)$ such that $A \equiv_M B \equiv_M T(B)$.
\end{corollary}
To prove Theorem \ref{thm:internal_characterization}, we need the following two lemmata and a result by \name{Weihrauch}.
\begin{lemma}
\label{lemma:applysequence}
Let $\repsp{X}$ admit an effectively fiber-compact representation. Then there is a space $\repsp{Y}$ such that:
\begin{enumerate}
\item $\repsp{X} \hookrightarrow \repsp{Y}$ (as a closed subspace),
\item $\repsp{Y}$ has an effectively fiber-compact representation,
\item $\repsp{Y}$ has a computable dense sequence,
\item if $\repsp{X}$ is computably admissible, so is $\repsp{Y}$.
\end{enumerate}
\begin{proof}
{\bf Construction of $\repsp{Y}$:} We start with some preliminary technical notation. Let $\operatorname{Wrap} : \Cantor \to \Cantor$ be defined by $\operatorname{Wrap}(p)(2i) = p(i)$ and $\operatorname{Wrap}(p)(2i+1) = 0$. Let $\operatorname{Prefix} :\subseteq \Cantor \to \{0,1\}^*$ be defined by $\operatorname{Prefix}(p) = w$ iff $p = 0w(1)0w(2)0\ldots011q$ for some $q \in \Cantor$. Note that $\dom(\operatorname{Prefix}) \cap \dom(\operatorname{Wrap}^{-1}) = \emptyset$ and $\dom(\operatorname{Prefix}) \cup \dom(\operatorname{Wrap}^{-1}) = \Cantor$.
Let the presumed representation of $\repsp{X}$ be $\delta_\repsp{X} : \subseteq \Cantor \to X$. Our construction of $\repsp{Y}$ will utilize a notation $\nu_\repsp{Y} : \{0,1\}^* \to Y'$ as auxiliary part, this notation (or alternatively, equivalence relation on $\{0,1\}^*$) will be dealt with later. We set $Y = X \cup Y'$ (in particular, we add only countably many elements to $\repsp{X}$) and then define $\delta_\repsp{Y}$ via $\delta_\repsp{Y}(p) := \delta_\repsp{X}(\operatorname{Wrap}^{-1}(p))$ if $p \in \dom(\delta_\repsp{X}\circ\operatorname{Wrap}^{-1})$ and $\delta_\repsp{Y}(p) = \nu_\repsp{Y}(\operatorname{Prefix}(p))$ if $p \in \dom(\operatorname{Prefix})$.
In order to define $\nu_\repsp{Y}$, we do need to refer to the effective fiber-compactness of $\delta_\repsp{X}$. From the function realizing $x \mapsto \delta_\repsp{X}^{-1}(\{x\}) : \repsp{X} \to \mathcal{A}(\Cantor)$ we can obtain an indexed family of finite trees $(T_w)_{w \in \{0,1\}^*}$ with the following properties:
\begin{enumerate}
\item Each $T_w$ has height $|w|$.
\item If $w \prec u$, then $T_u \cap \{0,1\}^{\leq |w|} = T_w$.
\item $w \in T_w$.
\item For any $p \in \dom(\delta_\repsp{X})$, some $q \in \Cantor$ is an infinite path through $\bigcup_{n \in \mathbb{N}} T_{p_{\leq n}}$ iff $\delta_\repsp{X}(q) = \delta_\repsp{X}(p)$.
\end{enumerate}
Now we set $\nu_\repsp{Y}(w) = \nu_\repsp{Y}(u)$ iff $T_w = T_u$. Note in particular that $T_w = T_u$ is a decidable property.
{\bf Proof of the properties:} To see that $\repsp{X} \hookrightarrow \repsp{Y}$ it suffices to note that both $\operatorname{Wrap}$ and $\operatorname{Wrap}^{-1}$ are computable. That $\repsp{X}$ embeds as a closed subspace follows from $\dom(\operatorname{Wrap}^{-1})$ being closed in $\Cantor$.
Next we shall see that $\delta_\repsp{Y}$ is effectively fiber-compact by reversing the step from the function $x \mapsto \delta_\repsp{X}^{-1}(\{x\}) : \repsp{X} \to \mathcal{A}(\Cantor)$ to the family $(T_w)_{w \in \{0,1\}^*}$. First, we define a version of $\operatorname{Wrap}$ for finite trees via $\operatorname{T-Wrap}(T) = \{0w(1)0w(2)\ldots w(|w|)\mid w \in T\} \cup \{0w(1)0w(2)\ldots w(|w|)0\mid w \in T\}$. Given some set $W \subseteq \{0,1\}^*$, let the induced tree of height be defined via $T(W,n) = \{u \exists w \in W \ u \prec w\} \cup \{u \in \{0,1\}^n \mid \exists w \in W \ \wedge w \prec u\}$. Then we define a derived family $(T'_w)_{w \in \{0,1\}^*}$ by $T'_{0w(1)0w(2)\ldots0w(|w|)} = T'_{0w(1)0w(2)\ldots0w(|w|)0} = \operatorname{T-Wrap}(T_w)$ and $T'_{0w(1)\ldots w(|w|)1v} = T(\{u \mid T_u = T_w\},|w|+|v|)$. This construction too satisfies that if $w \prec u$, then $T'_u \cap \{0,1\}^{\leq \operatorname{height}(T'(w))} = T'_w$. Thus, the function that maps $p$ to the set of all infinite pathes through $\bigcup_{n \in \mathbb{N}} T'_{p_{\leq n}}$ does define some function $t : \Cantor \to \mathcal{A}(\Cantor)$, and one can verify readily that $t(p) = \delta_\repsp{Y}^{-1}(\delta_\repsp{Y}(p))$ whenever $p \in \dom(\delta_\repsp{Y}(p))$.
It is clear that $\repsp{Y}$ has a computable dense sequence: Fix some standard enumeration $\nu : \mathbb{N} \to \{0,1\}^*$, and consider $(y_n)_{n \in \mathbb{N}}$ with $y_n = \delta_\repsp{Y}(0\nu(n)(1)\ldots0\nu(n)(|\nu(n)|)1^\omega)$.
It remains to show that if $\delta_\repsp{X}$ is admissible, so is $\delta_\repsp{Y}$. It is this step which requires the identification of some points via $\nu_\repsp{Y}$, and through this, also depends on $\delta_\repsp{X}$ being effectively-fiber-compact. Given some tree encoding some $\delta_\repsp{Y}^{-1}(\{x\})$, we need to be able to compute a path through it. As long as the tree seems to have a path without repeating $1$'s, we lift the corresponding map for $\delta_\repsp{X}$. If $x = \nu_\repsp{Y}(w)$ for some $w \in \{0,1\}^*$, we notice eventually, and can extend the current path in a computable way by virtue of the identifications.
\end{proof}
\end{lemma}
The preceding lemma produces spaces with a somewhat peculiar property: The designated dense sequence is an open subset of the space, unlike the usual examples. In \cite{gregoriades3}, \name{Gregoriades} has explored a general construction yielding Polish spaces with such properties (cf.~\cite[Theorem 2.5]{gregoriades}), which in particular serves to prevent effective Borel isomorphisms between spaces.
\begin{lemma}
\label{lemma:regular}
Let $\repsp{X}$ admit a computably admissible effectively fiber-compact representation. Then $\repsp{X}$ is computably regular.
\begin{proof}
The properties of the representations mean that we can consider $\repsp{X}$ as a subspace of $\mathcal{A}(\Cantor)$ containing only pair-wise disjoint sets. Let $A \in \mathcal{A}(\mathcal{A}(\Cantor))$ be a closed subset in $\repsp{X}$. Note that we can compute $\bigcup A \in \mathcal{A}(\Cantor)$, as every infinite path computes the relevant tree. Furthermore, given $x \in \repsp{X} \subseteq \mathcal{A}(\Cantor)$ and $A$, we can compute $x \cap \bigcup A \in \mathcal{A}(\Cantor)$. As $\repsp{X}$ only contains pair-wise disjoint points, this set is empty if and only if $x$ and $A$ are disjoint. As $\Cantor$ is compact, the corresponding tree will have to die out at some finite level, which means that the trees for $x$ and $A$ are disjoint below this level. Let $I$ be the vertices at this level belonging to $x$. We may now define two open sets $U_I, U_{I^C} \in \mathcal{O}(\mathcal{A}(\Cantor))$ by letting $U_{X}$ for $X \in \{I, I^C\}$ accept its input sets $A$ as soon as $A \cap X\Cantor = \emptyset$ is verified. Then $U_{I} \cap U_{I^C} = \{\emptyset\}$, thus $\repsp{X} \cap U_I$ and $\repsp{X} \cap U_{I^C}$ are disjoint open sets. Moreover, we find $A \subseteq U_{I}$ and $x \in U_{I^C}$, so the two open sets are those we needed to construct for computable regularity.
\end{proof}
\end{lemma}
\begin{proof}[Proof of Theorem \ref{thm:internal_characterization}]
The $\Leftarrow$-direction is present e.g.~in \cite{weihrauchf}. We can use Lemma \ref{lemma:applysequence} to make sure w.l.o.g.~that $\repsp{X}$ has a computable dense sequence. By Lemma \ref{lemma:regular}, the space is computably regular. As shown in \cite{grubba3,weihrauchm}, a computably regular space with a computable dense sequence admits a compatible metric.
\end{proof}
\name{Miller} showed that the Turing degrees below any non-total continuous degree form a Scott ideal \cite{miller2}, heavily drawing on topological arguments. However, based on Corollary \ref{corr:contdegreecharac} we see that the statement itself can be phrased entirely in the language of trees, points and Medvedev reducibility. So far, we do not know of a direct proof involving only these concepts:
\begin{proposition}
Let $A \subseteq \Cantor$ be such that $A \equiv_M T(A)$ and that there is no $r \in \Cantor$ with $A \equiv_M \{r\}$. Then $T(B) \leq_M \{p\} <_M A$ for $p \in \Cantor$, $B \subseteq \Cantor$ implies $B \leq_M A$.
\end{proposition}
\subsection{Enumeration Degrees and Overtness}
\label{subsec:fiberovert}
The often overlooked dual notion to compactness is \emph{overtness} (see \cite{taylor,taylor2}). Intuitively, overtness makes existential quantification well-behaved: a space $\repsp{X}$ is overt if $E_\repsp{X}:\mathcal{O}(\repsp{X})\to\mathcal{S}$ is continuous, where $E_\repsp{X}(U)=\top$ iff $U$ is nonempty. Therefore, if $\repsp{X}$ is overt and $P \subseteq \repsp{X} \times \repsp{Y}$ is open, then $\{y \in \repsp{Y} \mid \exists x \in \repsp{X} \ (x,y) \in P\}$ is open, too. Classically, this is a trivial notion, however, the situation is different from an effective point of view.
One may identify an overt subspace $A$ of $\repsp{X}$ with $E_A$, or equivalently, its overtness witness $\{U\in\mathcal{O}(\repsp{X}):A\cap U\not=\emptyset\}$ as a point in the represented space $\mathcal{O}(\mathcal{O}(\repsp{X}))$.
Via this identification, we obtain the hyperspace $\mathcal{V}(\repsp{X})$ of representatives $\overline{A}$ of all overt subspaces $A$ of $\repsp{X}$ (see also \cite{pauly-synthetic-arxiv}).
Note that this corresponds to the lower Vietoris topology on the hyperspace of closed sets.
A computable point in $\mathcal{V}(\repsp{X})$ is also called a {\em c.e.~closed} set in computable analysis.
Now we call a representation $\delta : \subseteq \Baire \to \repsp{X}$ \emph{effectively fiber-overt}, iff $\overline{\delta^{-1}} : \repsp{X} \to \mathcal{V}(\Baire)$ is computable. A straightforward argument shows that this is equivalent to $\delta$ being effectively open, i.e.~$U \mapsto \delta[U] : \mathcal{O}(\Baire) \to \mathcal{O}(\repsp{X})$ being computable. Now we see that every space with an effectively fiber-overt representation inherits an effective countable basis from $\Baire$, while on the other hand, the standard representations of countably based spaces introduced in Example \ref{example:representation} are all effectively fiber-overt. Thus we see that while effectively fiber-compact representations characterize metrizability, effectively fiber-overt representations characterize second-countability.
\section{Point Degree Spectra of Quasi-Polish Spaces}\label{sec:quasi-Polish}
\subsection{Lower Reals and Semirecursive Enumeration Degrees}
Let us move on to the $\sigma$-isomorphic classification of quasi-Polish spaces \cite{debrecht6}.
We now focus on the following chain of quasi-Polish spaces:
\[\mathcal{D}_T\subsetneq\mathcal{D}_r\subsetneq\mathcal{D}_e\mbox{, and }\Cantor<_\sigma^\mathfrak{T}[0,1]^\mathbb{N}<_\sigma^\mathfrak{T}\mathcal{O}(\mathbb{N}).\]
Here, the proper inclusion $[0,1]^\mathbb{N}<_\sigma^\mathfrak{T}\mathcal{O}(\mathbb{N})$ follows from relativizing \name{Miller}'s observation in \cite{miller2} that no quasi-minimal degree has continuous degree.
In quasi-Polish case, the notion of the specialization order is quite useful.
Indeed, \name{Motto Ros} has already used the specialization order to give an alternative way to show the properness of $[0,1]^\mathbb{N}<_\sigma^\mathfrak{T}\mathcal{O}(\mathbb{N})$.
Recall that the {\em specialization order} $\prec$ on a topological space $\repsp{X}$ is defined via $x \prec y :\Leftrightarrow x \in \overline{\{y\}}$. In particular, the specialization order on $\mathcal{O}(\mathbb{N})$ coincides with subset-inclusion. The $T_1$ separation property asserts that no two elements are comparable w.r.t.~$\prec$, i.e.~that the specialization order is a single antichain.
\begin{theorem}\label{thm:quasi-polish-main}
There is a map $\repsp{Q}$ transforming each countable set $S\subseteq\omega_1$ into a nonmetrizable quasi-Polish space $\repsp{Q}(S)\not\leq_\sigma^\mathfrak{T}[0,1]^\mathbb{N}$ such that
\[S\not\subseteq T\;\Longrightarrow\;\repsp{Q}(S)\not\leq_\sigma^\mathfrak{T}\repsp{Q}(T).\]
\end{theorem}
Let $\mathbb{R}_<$ be the real line endowed with the lower topology, that is, its topology is generated by open intervals of the form $(p,\infty)$.
One can easily see that $\mathbb{R} \ |_{\sigma}^\mathfrak{T} \ \mathbb{R}_<$ by comparing their specialization orders.
From the computability theoretic viewpoint, the property $\mathbb{R}\not\leq^\mathfrak{T}_\sigma\mathbb{R}_<$ can be strengthened as follows.
\begin{lemma}[Co-spectrum Preservation]\label{thm:cospectrum-preservation}
Let $\repsp{X}$ be an admissibly represented Polish space.
Then,
\[{\rm coSpec}(\repsp{X}\times\mathbb{R}_{<})\subseteq{\rm coSpec}(\repsp{X})\cup{\rm coSpec}(\Cantor).\]
In particular, if such an $\repsp{X}$ is uncountable, then there is an oracle $r\in\Cantor$ such that
\[{\rm coSpec}^r(\repsp{X}\times\mathbb{R}_{<})={\rm coSpec}^r(\repsp{X}).\]
\end{lemma}
\begin{lemma}\label{lem:cospectrum-preservation}
Let $\repsp{X}$ admit an effectively fiber-overt representation $\delta_\repsp{X}$ (cf.~Subsection \ref{subsec:fiberovert}), $x\in \repsp{X}$, $y\in\mathbb{R}_<$, and $z\in\Cantor$.
If $z\leq_M(x,y)$, then either $z\leq_M x$ or $-y\leq_M x$ holds.
\end{lemma}
\begin{proof}
Let computable $f :\subseteq \repsp{X} \times \mathbb{R}_< \to \Cantor$ witness the reduction $z\leq_M(x,y)$.
By extending the domain of $f$ if necessary, it can be identified with a c.e.~open set $U\subseteq\Cantor\times\mathbb{Q}\times\mathbb{N}\times\{0,1\}$ satisfying that $f(x,y)(n)=i$ if and only if the following two condition holds:
\begin{enumerate}
\item For any $p \in \delta_\repsp{X}^{-1}(x)$ there is some rational $s < y$ such that $(p, s, n, i) \in U$.
\item For any $p \in \delta_\repsp{X}^{-1}(x)$ and any rational $s<y$, $(p, s, n, 1-i) \not\in U$.
\end{enumerate}
As $\delta_\repsp{X}$ is effectively fiber-overt, the set $U' := \{(x',t,n,i) \mid \exists p \in \delta_\repsp{X}^{-1}(x') \ (p,t,n,i) \in U\}$ is also computable as an open subset of $\repsp{X} \times \mathbb{Q} \times \mathbb{N} \times \{0,1\}$. Now we can distinguish two cases:
\begin{enumerate}
\item For any $\varepsilon>0$, there exist rationals $t<s<y+\varepsilon$ such that $(x,t,n,i) \in U'$ and $(x,s,n,1-i) \in U'$ for some $n\in\mathbb{N}$ and $i\in\{0,1\}$.
\item Otherwise, there exists $\varepsilon>0$ such that for all $t<y+\varepsilon$, if $(x,t,n,i) \in U'$ for some $n\in\mathbb{N}$ and $i\in\{0,1\}$, then we must have $i = f(x,y)(n)$.
\end{enumerate}
Note that if $(x, t, n, i) \in U'$ and $(x,s,n,1-i) \in U'$ for $t < s$, then we automatically have $y \leq s$.
Therefore, in the first case we can compute $-y \in \mathbb{R}_<$ as the supremum of $-s$ over all witnesses $s$, thus find $-y \leq_M x$. In the second case, there will be some rational number $y_0$ with $y \leq y_0 < y + \varepsilon$. Using $y_0$ in place of $y$ leaves the value $f$ is producing unchanged, thus we have that $z \leq_M x$.
\end{proof}
\begin{proof}[Proof of Lemma \ref{thm:cospectrum-preservation}]
Suppose that $y\in\mathbb{R}_<$ and $x\in\repsp{X}$.
If $-y\not\leq_Mx$, then ${\rm coSpec}(x,y)={\rm coSpec}(x)$ by Lemma \ref{lem:cospectrum-preservation}.
Otherwise, $(x,y,-y)\equiv_M(x,y)$.
If $y\leq_Mx$, then clearly, ${\rm coSpec}(x,y)={\rm coSpec}(x)$.
Otherwise, $(y,-y)\not\leq_Mx$.
Obviously, $(y,-y)$ has Turing degree.
By Lemma \ref{lem:JM-almosttotal}, we have $(x,y,-y)\in\mathcal{D}_T$.
Hence, ${\rm coSpec}(x,y)\in{\rm coSpec}(\Cantor)$.
For the latter half of Lemma \ref{thm:cospectrum-preservation}, if $\repsp{X}$ is uncountable, then there is an $r$-computable embedding of $\Cantor$ into $\repsp{X}$ for some oracle $r$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:quasi-polish-main}]
Let $\mathcal{G}_S$ be the countable set of monotone oracle $\mathbf{\Pi}^0_2$ singletons constructed in the proof of Theorem \ref{thm:maintheorem_emb}.
By Lemma \ref{thm:cospectrum-preservation}, the quasi-Polish space $\repsp{Q}(S):=\repspb{Rea}(\mathcal{G}_S)\times\overline{\mathbb{R}}_<$ has the same cospectrum as $\repspb{Rea}(\mathcal{G}_S)$, where $\overline{\mathbb{R}}_<:=\mathbb{R}_<\cup\{\infty\}$ is a quasi-completion of $\mathbb{R}_<$.
By the proofs of Theorem \ref{thm:maintheorem_emb} and Lemma \ref{thm:refle}, if $S\not\subseteq T$, then the cospectrum of $\repspb{Rea}(\mathcal{G}_S)$ is not a sub-cospectrum of $\repspb{Rea}(\mathcal{G}_T)$ relative to all oracles.
Therefore, by Observation \ref{obs:main-cospec-inv}, we have $\repsp{Q}(S)\not\leq_\sigma^\mathfrak{T}\repsp{Q}(T)$.
\end{proof}
As a consequence of Lemma \ref{thm:cospectrum-preservation}, any lower real can compute only a $\Delta^0_2$ real:
\[{\rm coSpec}(\mathbb{R}_{<})=\{\{x\in \Cantor:x\leq_Ty\}:y\mbox{ is right-c.e.}\}\]
Indeed, Lemma \ref{lem:cospectrum-preservation} provides a very simple and natural construction of a quasi-minimal enumeration degree.
\begin{corollary}[see also {\name{Arslanov}, \name{Kalimullin} \& \name{Cooper} \cite[Theorem 4]{kalimullin}}]
Suppose that $z\in\mathbb{R}$ is neither left-c.e nor right-c.e.
Then, the enumeration degree of the cut $\{q\in\mathbb{Q}:q<z\}$ is quasi-minimal.
\end{corollary}
On the one hand, we deduced the property $[0,1]^\mathbb{N}<^\mathfrak{T}_\sigma\mathcal{O}(\mathbb{N})$ from the topological argument concerning the specialization order on the lower real $\mathbb{R}_<$.
On the other hand, \name{Miller}'s original proof used the existence of a quasi-minimal enumeration degree to show ${\rm Spec}([0,1]^\mathbb{N})\subsetneq{\rm Spec}(\mathcal{O}(\mathbb{N}))$.
Surprisingly, however, the previous argument clarifies that these two seemingly unrelated approaches are essentially equivalent.
Note that the point degree spectrum of the lower real $\mathbb{R}_<$ is indeed strongly connected with the notion of a semirecursive set in the context of the enumeration degrees.
Recall from \cite{jockusch2} that a set $A \subseteq \mathbb{N}$ is called {\em semirecursive}, if there is a computable function $f : \mathbb{N} \times \mathbb{N} \to \mathbb{N}$ such that for all $n, m \in \mathbb{N}$ we find $f(n,m) \in \{n,m\}$, and if $n \in A$ or $m \in A$, then $f(n,m) \in A$. We call an enumeration degree $q \in \mathcal{D}_e$ semirecursive, if it is the degree of a semirecursive point in $\mathcal{O}(\mathbb{N})$.
\name{Jockusch} \cite{jockusch2} pointed out that every left-cut (i.e., every lower real $x\in\mathbb{R}_<$) is semirecursive, and conversely, \name{Ganchev} and \name{Soskova} \cite{GanSos15} showed that every semirecursive enumeration degree contains a left-cut.
Consequently, the point degree spectra of the lower real $\mathbb{R}_<$ can be characterized as follows:
\[{\rm Spec}(\mathbb{R}_<)=\{\mathbf{d}\in\mathcal{D}_e:\mathbf{d}\mbox{ is a semirecursive enumeration degree}\}.\]
\subsection{Higher Dimensional Lower Cubes}
We can also consider the higher dimensional lower real cubes $\mathbb{R}_<^n$.
Surprisingly, the spectra of $\mathbb{R}_<^n$ form a proper hierarchy as follows.
\begin{theorem}\label{thm:lowerrealhierarchy}
If $\repsp{X}$ is a second-countable $T_1$ space, then $\mathbb{R}_<^{n+1}\ |_\sigma^\mathfrak{T} \ \repsp{X}\times\mathbb{R}_<^{n}$ for every $n$.
\end{theorem}
To show the above theorem, we use the following order theoretic lemma.
Let $\Lambda^n=(\{0,1\}^n,\leq)$ be a partial order on $\{0,1\}^n$ obtained as the $n$-th product of the ordering $0<1$.
\begin{lemma}\label{lem:lowerrealhierarchy}
For every countable partition $(P_i)_{i\in\omega}$ of the $n$-dimensional hypercube $[0,1]^n$ (endowed with the standard product order), there is $i\in\omega$ such that $P_i$ has a subset which is order isomorphic to the product order $\Lambda^n$.
\end{lemma}
\begin{proof}
We use Vaught's ``non-meager'' quantifier $\exists^*x\varphi(x)$, which states that the set $\{x:\varphi(x)\}$ is not meager in $[0,1]$ (with respect to the standard Euclidean topology).
We claim that for every countable partition $(P_i)_{i\in\omega}$ of $[0,1]^n$, there is $i\in\omega$ such that
\[\exists^*x_1\exists^*x_2\dots\exists^*x_n\;(x_1,x_2,\dots,x_n)\in P_i\]
Inductively assume that the above claim is true for $n-1$.
If the above claim does not hold for $n$, then by the Baire category theorem, there are comeager many $x_1$ such that
\[\neg\exists^*x_2\dots\exists^*x_n\;(x_1,x_2,\dots,x_n)\in\bigcup_iP_i.\]
However, for any such $x_1$, by the induction hypothesis, the $x_1$-sections of $P_i$'s do not cover the $x_1$-section of $[0,1]^n$.
In particular, $\bigcup_iP_i$ cannot cover the $n$-hypercube $[0,1]^n$, which verifies the claim.
Now, let $S$ be a nonmeager set consisting of all $x_1$'s in the above claim.
Note that since there are non-meager many $x_1\in S$, there is a nonempty open set $U$ such that for any nonempty open set $V\subseteq U$, one can find uncountably many such $x_1\in V\cap S$.
Otherwise, $S$ is covered by the closure of the union of the collection $\mathcal{B}$ of all rational open balls $B$ such that $B\cap S$ is countable.
Therefore, $S$ is divided into the union of the nowhere dense set $\partial\bigcup\mathcal{B}$ and the countable set $\bigcup_{B\in\mathcal{B}}B\cap S$, which contradicts the fact that $S$ is nonmeager.
We fix such a nonempty open set $U$.
Now, for any $x_1\in S$, we may inductively assume that the $x_1$-th section of $P_i$ has a subset $L(x_1)$ which is order isomorphic to $\Lambda^{n-1}$.
Let $\hat{L}(x_1)$ be the region bounded by $L(x_1)$, which is homeomorphic to $[0,1]^{n-1}$.
We may also inductively assume that $P_i$ is dense in $\hat{L}(x_1)$.
Therefore, since $\hat{L}(x_1)$ for any $x_1\in S$ has positive $(n-1)$-dimensional Lebesgue measure, for any nonempty open set $V\subseteq U$ one can find $x_1^0<x_1^1$ in $V\cap S$ such that the intersection $\pi\circ\hat{L}(x_1^0)\cap\pi\circ\hat{L}(x_1^1)$ also has positive $(n-1)$-dimensional Lebesgue measure, where $\pi:[0,1]^n\to[0,1]^{n-1}$ is the projection defined by $\pi(x_1,x_2,\dots,x_n)=(x_2,\dots,x_n)$.
By density of $P_i$, one can find a smaller $(n-1)$-cubes $L^*(x^0_1),L^*(x^1_1)\subseteq\pi\circ\hat{L}(x_1^0)\cap\pi\circ\hat{L}(x_1^1)$ such that $(\{x^0_1\}\times L^*(x^0_1))\cup(\{x^1_1\}\times L^*(x^1_1))\subseteq P_i$ is order isomorphic to $\Lambda^{n}$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:lowerrealhierarchy}]
Note that the specialization order on the space $\mathbb{R}_<^{n+1}$ is exactly the same as the standard product order on $\mathbb{R}^{n+1}$.
By Lemma \ref{thm:lowerrealhierarchy}, for every countable partition $(P_i)_{i\in\omega}$ of $\mathbb{R}_<^{n+1}$, there is $i\in\omega$ such that the specialization order on $P_i$ has a subset which is order isomorphic to the product order $\Lambda^{n+1}$ whose order dimension is $n+1$.
If $P_i$ is embedded into the specialization order on $\repsp{X}\times\mathbb{R}_<^n$, then the embedded image of an isomorphic copy of $\Lambda^{n+1}$ has to be contained in a connected component of the order of $\repsp{X}\times\mathbb{R}_<^n$.
However, the specialization order on $\repsp{X}\times\mathbb{R}_<^n$ is now ${\rm card}(\repsp{X})$ many copies of that on $\mathbb{R}_<^n$ since $\repsp{X}$ is $T_1$.
Therefore, every connected component of the specialization order on $\repsp{X}\times\mathbb{R}_<^n$ is isomorphic to the product order on $\mathbb{R}^n$ whose order dimension is $n$.
Hence, $\mathbb{R}_<^{n+1}$ cannot be $\sigma$-embedded into $\repsp{X}\times\mathbb{R}_<^n$.
Conversely, suppose that $\repsp{X}\times\mathbb{R}_<^n$ is $\sigma$-embedded into $\mathbb{R}_<^{n+1}$.
By Lemma \ref{thm:lowerrealhierarchy}, for every countable partition $(P_i)_{i\in\omega}$ of $\repsp{X}\times\mathbb{R}_<^{n}$, there must exist $i\in\omega$ such that $P_i$ contains a uncountable family $(\Lambda^n_\alpha)_{\alpha\in\aleph_1}$ of pairwise incomparable suborders of $P_i$ which are order isomorphic to $\Lambda^n$.
Let $L_\alpha$ be the embedded image of $\Lambda^n_\alpha$ in $\mathbb{R}_<^{n+1}$, and $\hat{L}_\alpha$ be the region bounded by $L_\alpha$, which is homeomorphic to $[0,1]^n$.
As in the proof of Lemma \ref{thm:lowerrealhierarchy}, we may also assume that the embedded image $P_i^*\subseteq\mathbb{R}_<^{n+1}$ of $P_i$ is dense in $\hat{L}_\alpha$ for any $\alpha<\aleph_1$.
For any $\alpha<\aleph_1$, the projection $\pi_k[\hat{L}_\alpha]=\{(x_0,\dots,x_{k-1},x_{k+1},\dots,x_{n}):(x_0,\dots,x_n)\in\hat{L}_\alpha\}$ of $\hat{L}_\alpha$ for some $k\leq n$ has positive $n$-dimensional Lebesgue measure.
Fix $k<n+1$ such that $\pi_k[\hat{L}_\alpha]$ has positive $n$-dimensional Lebesgue measure for uncountably many $\alpha$.
Then, there are $\alpha\not=\beta$ such that $\pi_k[\hat{L}_\alpha]\cap\pi_k[\hat{L}_\beta]$ also has positive $(n+1)$-dimensional Lebesgue measure.
It is not hard to see that it contradicts our assumption that $L_\alpha$ and $L_\beta$ are incomparable.
\end{proof}
Note that Theorem \ref{thm:lowerrealhierarchy} has immediate computability-theoretic corollaries:
\begin{corollary}
For every $n \in \mathbb{N}$ there are enumeration degrees $p_n$, $q_n$ such that
\begin{itemize}
\item $p_n$ is the product of $n + 1$ semirecursive degrees, but not of $n$ semirecursive degrees and a Turing degree.
\item $q_n$ is the product of $n$ semirecursive degrees and a Turing degree, but not of $n-1$ semirecursive degrees and a Turing degree, or of $n + 1$ semirecursive degrees.
\end{itemize}
\end{corollary}
\subsection{The co-spectrum of a Universal Quasi-Polish Space}
Recall that the co-spectrum of the universal Polish space $[0,1]^\mathbb{N}$ consists of all principal countable Turing ideals and all countable Scott ideals.
However, there are many non-principal countable Turing ideals that are not Scott ideals, e.g., countable $\omega$-models of ${\sf WWKL}+\neg{\sf WKL}$, ${\sf RT}^2_2+\neg{\sf WKL}$ and so on.
We now see that every countable Turing ideal is realized as a co-spectrum of the universal quasi-Polish space $\mathcal{O}(\mathbb{N})$ by modifying the standard forcing construction of quasi-minimal enumeration degrees.
\begin{theorem}\label{thm:cospec-univscottdom}
Every countable Turing ideal is realized as a co-spectrum in the universal quasi-Polish space $\mathcal{O}(\mathbb{N})$.
In particular, ${\rm coSpec}^r([0,1]^\mathbb{N})\subsetneq{\rm coSpec}^r(\mathcal{O}(\mathbb{N}))$ for every $r\in\mathbb{N}$.
\end{theorem}
\begin{proof}
It suffices to show that, for any sequence $(x_i)_{i\in\mathbb{N}}$ of reals and oracle $r$, there is $A\in\mathcal{O}(\mathbb{N})$ whose $r$-co-spectrum ${\rm coSpec}^r(A)$ is equal to all $y\in \Cantor$ such that $y\leq_Tr\oplus\bigoplus_{m\leq n}x_m$ for some $n\in\mathbb{N}$.
Without loss of generality, we may assume that $x_0=r$.
Suppose $\bot\not\in\mathbb{N}$, and let $\mathbb{N}_\bot=\mathbb{N}\cup\{\bot\}$.
We say that a sequence $\sigma\in\mathbb{N}_\bot$ strongly extends $\tau\in\mathbb{N}_\bot$ if $\tau$ is an initial segment of $\sigma$ as a $\mathbb{N}_\bot$-valued sequence.
A sequence $\sigma\in\mathbb{N}_\bot$ extends $\tau\in\mathbb{N}_\bot$ if $\sigma$ extends $\tau$ as a partial function on $\mathbb{N}$, where the equality $\sigma(n)=\bot$ is interpreted as meaning that $\sigma(n)$ is undefined, that is, $n\not\in{\rm dom}(\sigma)$.
Every partial function $\varphi:\subseteq\mathbb{N}\to\mathbb{N}$ generates a tree $T_\varphi\subseteq\mathbb{N}_\bot^{<\omega}$ by
\[T_\varphi=\{\sigma\in\mathbb{N}_\bot^{<\omega}:(\forall n<|\sigma|)\;\varphi(n)\downarrow\;\rightarrow\;\sigma(n)=\varphi(n)\}.\]
Let $\mathbb{P}$ be the collection of pairs $(\sigma,\varphi)$ of a string $\sigma\in \mathbb{N}_\bot^{<\omega}$ and a partial function $\varphi$ such that $\sigma\in T_\varphi$ and ${\rm dom}(\varphi)$ is of the form $D(A)=\{(m,n):n\in\mathbb{N}\;\&\;m\in A\}$ for some finite set $A\subseteq\mathbb{N}$.
We write $(\tau,\psi)\leq(\sigma,\varphi)$ if $\tau$ strongly extends $\sigma$, $\psi$ extends $\varphi$, and $\psi\upharpoonright|\sigma|=\varphi\upharpoonright|\sigma|$.
By induction, we assume that $(\sigma_0,\varphi_0)$ is the pair of an empty string and an empty function, and $(\sigma_s,\varphi_s)\in\mathbb{P}$ has already been defined.
Moreover, we inductively assume that the tree $T_{\varphi_s}$ is computable in $\bigoplus_{2t<s}x_t$.
We now have ${\rm dom}(\varphi_s)=D(A_s)$ for some $s\in\mathbb{N}$ by the definition of $\mathbb{P}$.
If $s=2e$ for some $e\in\mathbb{N}$, then choose sufficiently large $m_{s+1}\not\in A_s$ with $m_{s+1}>|\sigma_{s}|$.
Then, put $\sigma_{s+1}=\sigma_s$, and define $\varphi_{s+1}(m_{s+1},n)=x_e(n)$ for every $n\in\mathbb{N}$.
Clearly, the tree $T_{\varphi_{s+1}}$ is computable in $\bigoplus_{2t\leq s}x_t$.
If $s=2e+1$ for some $e\in\mathbb{N}$, we look for a string $\tau\in T_{\varphi_s}$ strongly extending $\sigma_s$ which forces the $e$-th computation $\Psi_e$ to be inconsistent, that is, two different values $\Psi_e(\tau)(n)=i$ and $\Psi_e(\tau)(n)=j$ for some $n$ and $i\not=j$ are enumerated.
If there is such a $\tau$, define $\sigma_{s+1}=\tau$ and $\varphi_{s+1}=\varphi_s$.
If there is no such a $\tau$, we look for strings $\eta,\theta\in T_{\varphi_s}$ strongly extending $\sigma_{s}$ such that the $e$-th computations $\Psi_e$ on $\eta$ and $\theta$ split and are consistent, that is, the consistent computations $\Psi_e(\eta)(n)=i$ and $\Psi_e(\theta)(n)=j$ for some $n$ and $i\not=j$ are enumerated.
In this case, for a sufficiently large $k>\max{|\eta|,|\theta|}$, define $\sigma_{s+1}$ to be the rightmost node of $T_{\varphi_s}$ strongly extending $\sigma_s$, where we declare that $\bot$ is the rightmost element in $\mathbb{N}_\bot$ in the sense that $n<\bot$ for every $n\in\mathbb{N}$.
Note that $\eta$ (resp.~$\theta$) (non-strongly) extends $\sigma_{s+1}\upharpoonright|\eta|$ (resp.~$\sigma_{s+1}\upharpoonright|\theta|$) since $\sigma_{s+1}$ chooses as many $\bot$'s as possible.
Then, define $\varphi_{s+1}=\varphi_s$.
Otherwise, define $\sigma_{s+1}=\sigma_s$ and $\varphi_{s+1}=\varphi_s$.
Finally, we obtain a partial function $\Phi$ on $\mathbb{N}$ by combining $\{\varphi_s\}_{s\in\mathbb{N}}$.
As in the usual argument, we will show that $\Phi$ is quasi-minimal above the collection $\{\bigoplus_{m\leq n}x_m:n\in\mathbb{N}\}$.
Clearly, $\bigoplus_{m\leq e}x_m$ is computable in $\Phi$ by our strategy at stage $2e$.
To show quasi-minimality of $\Phi$, consider the $e$-th computation $\Psi_e$.
If we find an inconsistent computation on some $\tau$ at stage $s=2e+1$, then clearly, $\Psi_e(\Phi)$ does not define an element of $\Cantor$.
If we find a consistent $e$-splitting $\eta$ and $\theta$ on an input $n$ at stage $s=2e+1$, $\Psi_e(\Phi)(n)$ is undefined, since otherwise $\Psi_e(\Phi)(n)=k$ implies $\Psi_e(\eta)=\Psi_e(\theta)=k$.
Otherwise, for every $n\in\mathbb{N}$, if $\Psi_e(\Phi)(n)$ is defined, then it is consistent, and uniquely determined inside $T_{\varphi_{s}}$.
Therefore, $\Psi_e(\Phi)(n)=k$ if and only if there is $\tau\in T_{\varphi_s}$ strongly extending $\sigma_s$ such that $\Psi_e(\tau)(n)=k$.
Consequently, $\Psi_e(\Phi)$ is computable in $\bigoplus_{m\leq e}x_m$, since $T_{\varphi_s}$ is a pruned $\bigoplus_{m\leq e}x_m$-computable tree by induction.
\end{proof}
\begin{corollary}
For any separable metrizable space $\repsp{X}$, we have $\repsp{X}\times\mathbb{R}_<<_\sigma^\mathfrak{T}\mathcal{O}(\mathbb{N})$.
\end{corollary}
\begin{proof}
By Observation \ref{obs:main-cospec-inv}, Lemma \ref{lem:cospectrum-preservation} and Theorem \ref{thm:cospec-univscottdom}.
\end{proof}
\section{Admissibly Represented Spaces which are not Quasi-Polish}\label{sec:non-quasi-Polish}
Recall that the class of admissibly represented space coincides with that of $T_0$ spaces that are quotients of second-countable $T_0$ spaces \cite{schroder}.
In particular, there is an admissibly represented space which is not second-countable. \name{Schr\"oder} and \name{Selivanov} have studied hierarchies of such spaces in \cite{selivanov4,selivanov4b}.
In this section, we construct an admissibly represented space which is not $\sigma$-embedded into any second-countable $T_0$ space.
\begin{theorem}\label{thm:non-quasi-Polish}
There is an admissibly represented space $\repsp{X}$ such that $\mathcal{O}(\mathbb{N})<_\sigma^\mathfrak{T}\repsp{X}$.
\end{theorem}
Let $\mathcal{O}_\infty(\mathbb{N})$ be a subspace of $\mathcal{O}(\mathbb{N})$ consisting of infinite subsets of $\mathbb{N}$.
The space $\mathbb{Z}_<$ represents the set of integers equipped with lower topology.
In this section, we develop the degree spectrum of the function space $C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$.
We represent each continuous functional $H\in C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$ by an enumeration $\tilde{H}$ of pairs $(C,n)$ of a finite set $C\subseteq\mathbb{N}$ and an integer $n\in\mathbb{Z}$ indicating that $H(Y)\geq n$ for any infinite set $Y\supseteq C$.
This representation is automatically given as the usual category-theoretic exponential in the Cartesian closed category of admissibly represented spaces.
\begin{lemma}\label{lem:non-quasi-Polish1}
The Scott domain $\mathcal{O}(\mathbb{N})$ is computably embedded into the function space $C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$.
In particular, the degree spectrum of $C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$ contains all enumeration degrees.
\end{lemma}
\begin{proof}
For any sets $X,A,B\subseteq\mathbb{N}$, define $e_A(X\oplus B)=1$ if $A\cap B\not=\emptyset$, and $e_A(X\oplus B)=0$ otherwise.
Then, the function $e:\mathcal{O}(\mathbb{N})\to C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$ defined by $e(A)=e_A$ is a computable embedding.
To see $A\leq_Me_A$, for any $n\in\mathbb{N}$, we have $n\in A$ if and only if $e_A(\mathbb{N}\oplus\{n\})=1$.
To see $e_A\leq_MA$, for any $Y=Y_0\oplus Y_1\subseteq\mathbb{N}$, $e_A(Y)=1$ if and only if $A\cap Y_1\not=\emptyset$.
Thus, to compute $e_A$ from $A$, given a finite set $D=D_0\oplus D_1\subseteq\mathbb{N}$, we enumerate $(D,1)$ into $\tilde{e}_A$ when $A\cap D_1\not=\emptyset$ is witnessed.
\end{proof}
Indeed, $C(\mathcal{O}(\mathbb{N}),\mathbb{Z}_<)$ is computably embedded into $C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$ by transforming each $f$ into $\hat{f}$ defined by $\hat{f}(X\oplus Y)=f(Y)$ for any $X,Y\subseteq\mathbb{N}$.
\begin{lemma}\label{lem:non-quasi-Polish2}
There is a point $F$ in the function space $C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$ such that $F$ has no enumeration degree.
\end{lemma}
\begin{proof}
We follow the argument by Hinman \cite{Hinman73}.
We construct a continuous functional $F:\mathcal{O}_\infty(\mathbb{N})\to\mathbb{Z}_<$ as the limit of an increasing sequence of partial continuous functionals $F_n$ with domain $[\mathcal{A}_n]\subseteq\mathcal{O}_\infty(\mathbb{N})$, where $\mathcal{A}_n$ is a collection of finite sets such that $\bigcup\mathcal{A}_n$ is coinfinite, and $[\mathcal{A}_n]$ denote the set of all infinite sets $X\supseteq L$ for $L\in\mathcal{A}_n$.
At stage $n$, we look for $G\in C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$ not exceeding $F_{n-1}$ on $\mathcal{A}_{n-1}$ and $A\in\mathcal{O}(\mathbb{N})$ such that $G$
has the same degree with $A$ via indices $n=\langle a,b\rangle$, i.e., $G$ is computable in $A$ via $\Phi_b$ and $A$ is computable in $G$ via $\Phi_a$.
If such $G$ and $A$ exist, we choose any $B\subseteq\mathbb{N}\setminus\bigcup\mathcal{A}_{n-1}$.
Since $\Phi_b(A)(B)=G(B)$, we may find a finite set $D\subseteq A$ such that $(E,G(B))$ for some finite set $E\subseteq B$ is enumerated into $\Phi_b(D)$, that is, there are finite sets $D\subseteq A$ and $E\subseteq B$ such that for any $X,Y\subseteq\mathbb{N}$,
\[X\supseteq D\mbox{ and }Y\supseteq E\;\longrightarrow\;\Phi_b(X)(Y)\geq \Phi_b(A)(B)=G(B).\]
Here, we may assume that $G(E)=G(B)$ since the value $G(B)$ is determined by a sufficiently large finite subset of $B$.
Conversely, since $\Phi_a(G)=A\supseteq D$, there is a finite sublist $\tilde{L}\subseteq\tilde{G}$ such that for any $H\in C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$, if $H(C)\geq n$ for every $(C,n)\in\tilde{L}$, then $\Phi_a(H)\supseteq D$.
By choosing a slow enumeration of $G$ as a name, we may assume that $|C|>|E|$ for any $(C,n)\in\tilde{L}$.
Since $\tilde{L}_0=\{C\subseteq\mathbb{N}:(C,n)\in\tilde{L}\}$ is a finite collection of finite sets, we can find an infinite set $I_n\subseteq\omega\setminus(\bigcup\mathcal{A}_{n-1}\cup\bigcup\tilde{F}_0)$ such that $I_n\cup\bigcup\mathcal{A}_{n-1}\cup\bigcup\tilde{L}_0$ is coinfinite.
Define $K_n=I_n\cup E$.
Then, $\tilde{F}_n$ is defined as follows:
\[\tilde{F}_n=\tilde{F}_{n-1}\cup\tilde{L}\cup\{(\{t\},G(E)-1):t\in I_n\}\cup J_n,\]
where $J_n=\{(n,z)\}$ for a sufficiently small value $z\in\mathbb{Z}$ if $n\not\in\bigcup_{m\leq n}K_m$, and $J_n=\emptyset$ otherwise.
Eventually, we get a function $F\in C(\mathcal{O}_\infty(\mathbb{N}),\mathbb{Z}_<)$.
Note that $F_n\upharpoonright\mathcal{A}_{n-1}=F_{n-1}\upharpoonright\mathcal{A}_{n-1}$ since $G$ does not exceed $F_{n-1}$ on $\mathcal{A}_{n-1}$.
Now, suppose $F\equiv_MS$ for some $S\subseteq\mathbb{N}$ via an index $n=(a,b)$.
So at stage $n$, the strategy acts via $G$ and $A$.
First note that $\Phi_a(F)=S\supseteq D$ since $\tilde{L}\subset\tilde{F}$.
Therefore, by our choice of $E$,
\[F(E\cup K_n)=\Phi_b(S)(E\cup K_n)\geq \Phi_b(A)(E\cup K_n)=G(E\cup K_n).\]
However, $F(E\cup K_n)\leq G(E)-1<G(E\cup K_n)$, a contradiction.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:non-quasi-Polish}]
By Lemmata \ref{lem:non-quasi-Polish1} and \ref{lem:non-quasi-Polish2}.
Here, the relativization of Lemma \ref{lem:non-quasi-Polish2} is obviously true.
\end{proof}
\section*{Acknowledgements}
The work has benefited from the Marie Curie International Research Staff Exchange Scheme \emph{Computable
Analysis}, PIRSES-GA-2011- 294962.
The first author is partially supported by a Grant-in-Aid for JSPS fellows.
The authors are also grateful to Masahiro Kumabe, Joseph Miller, Luca Motto Ros, Philipp Schlicht, and Takamitsu Yamauchi for their insightful comments and discussions.
\bibliographystyle{eptcs}
|
\section{Introduction}
\label{sec1}
Over the past few decades there has been significant progress in the understanding of dissipative phenomena, both in classical and quantum mechanics.
This has been particularly achieved through the realization that dissipative parameters such as the lifetime of an unstable particle or the diffusion coefficient of Brownian motion are ultimately connected, at the most basic level, to complex eigenvalues of the Hamiltonian or the Liouville operators \cite{Nakanishi58,Prigogine73,Sudarshan78,Bohm89,Tasaki91,Petrosky96,Petrosky97,Hatano10,Klaiman11}.
These eigenvalues correspond to generalized eigenfunctions. Alternatively, the complex eigenvalues can be associated with eigenfunctions of an effective Hamiltonian or effective Liouvillian that are non-Hermitian operators \cite{Livshits56,Feshbach58,Feshbach62,Rotter91,Petrosky96,Petrosky97,Rotter09,Nakano11,Hatano13}.
(In some references\cite{Petrosky96,Petrosky97}, these non-Hermitian operators were called collision operators.)
Either way, the conclusion is that dissipative phenomena need \textit{not} be formulated as approximations or coarse graining of basic dynamics;
note that perturbation approximation often breaks the unitarity of the time-evolution operator and thereby breaks time-reversal symmetry.
Instead, they can be formulated in terms of complex eigenvalues of the basic dynamical operators, without resorting to approximations.
A deeper question is the origin of time-reversal symmetry breaking or irreversibility and how it is connected to time-reversible dynamics.
Previous work has addressed this problem by introducing time-reversal symmetry breaking in the complete set of eigenfunctions of the dynamical operator (for example the Hamiltonian in quantum mechanics).
This was achieved by starting with a set of eigenfunctions with continuous real eigenvalues, and then deforming the integration contour of the complete set to include complex eigenvalues on either the lower half plane (resonant states corresponding to future-oriented evolution) or on the upper half plane (anti-resonant states, corresponding to past-oriented evolution) \cite{Nakanishi58,Sudarshan78,Tasaki91,Ordonez01,Petrosky01}.
The contour deformation, however, is not unique.
One can select certain complex eigenvalues by encircling them, while ignoring other complex eigenvalues.
The remainder contour becomes a ``background integral''.
The resulting spectrum of eigenvalues thus includes both discrete complex eigenvalues and a continuum coming from the background integral. Since this separation into discrete eigenvalues and background integral is not unique, the physical interpretation of this construction is not very clear.
Another common view on irreversibility is that it is connected to the initial conditions of the system in question, or even the whole universe.
It is assumed that the initial state is a state of very low entropy;
the second law of thermodynamics then explains irreversibility.
This view, however, does not make a precise connection with dynamics and again it relies on approximations or coarse graining in order to derive the second law.
In this paper we synthesize the views described above by formulating a complete set of eigenfunctions of the Hamiltonian for a class of quantum mechanical systems that include both resonant and anti-resonant states.
Our complete set is explicitly time-reversal symmetric.
We avoid the introduction of any arbitrary background integral, as we obtain a complete set of eigenfunctions corresponding to {\it all} the real and complex discrete eigenvalues (\textit{i.e.}\ all point spectra) of the Hamiltonian.
We find that for a time-reversal symmetric condition at time $t=0$, time-reversal symmetry is broken for $t\ne 0$ depending on the sign of $t$.
Mathematically, this corresponds to choosing whether we solve an initial-condition problem or a terminal-condition problem, which may favor resonant or anti-resonant states, respectively.
Moreover, we find that for certain specifically prepared conditions which are \textit{not} time-reversal symmetric, anti-resonant states dominate during a part of the time evolution and resonant states dominate during another part.
For example, by performing a momentum inversion on a wave function emitted from a quantum dot, the wave function will collect itself back into the quantum dot (like a movie played backwards, showing water waves collecting themselves toward a point where a rock was dropped).
During this period anti-resonant states dominate.
Subsequently the wave function is re-emitted, a process during which resonant states dominate.
By maintaining time-reversal symmetry in the set of eigenfunctions, the selection of future-oriented resonant states or past-oriented anti-resonant states is uniquely determined by the overlap between either the initial or terminal conditions and the discrete eigenfunctions of the Hamiltonian.
Our set of time-reversal symmetric eigenstates is limited to a class of quantum systems with a tight-binding Hamiltonian.
In different contexts, sets of time-reversal symmetric eigenstates have been presented for scattering problems\cite{GarciaCalderon10} and for the Friedrichs-Lee model\cite{Kim14}.
However, our formulation, based on the solution of a quadratic eigenvalue problem, is general enough that it can be extended to other systems, including systems considered in non-equilibrium statistical mechanics.
Such extensions are left for future work.
In the present paper we will focus on the tight-binding systems because we can then present the main ideas in enough detail that they can be subsequently generalized.
The outline of the present paper is as follows.
Sections~\ref{sec2} and~\ref{sec3} introduce basic concepts and models which we use throughout the paper.
Section~\ref{sec4} presents an overview of the results of the paper.
Sections~\ref{sec5}--\ref{sec10} give the derivation of the results step by step in details.
Using the results, we calculate the survival probability in Sec.~\ref{sec11} and the escaping probability in Sec.~\ref{sec12}.
Finally in Secs.~\ref{sec13} and~\ref{sec14}, we analyze dynamics of wave packets, breaking them down into resonant and anti-resonant states.
Section~\ref{sec15} summarizes the paper.
\section{Resonant state as an eigenstate that breaks the time reversal symmetry: a short review}
\label{sec2}
In this section, we present a concise review of the resonant state as an eigenstate of the Schr\"{o}dinger equation.
Let us consider for the moment the standard Schr\"{o}dinger equation in a one-dimensional space with a real potential on a compact support around the origin:
\begin{align}\label{eq2-10}
\left(-\frac{d^2}{dx^2}+V(x)\right)\psi(x)=E\psi(x),
\end{align}
where $V(x)\in\mathbb{R}$ and $V(x)=0$ for $|x|>L$.
(We will switch to the tight-binding model on a discretized space in the next section.)
Note that Eq.~\eqref{eq2-10} observes the time-reversal symmetry;
external magnetic fields are absent.
The wave function $\psi(x)$ for a real eigenvalue therefore can be made a real function.
The Schr\"{o}dinger equation~\eqref{eq2-10} has eigenvalues with point spectra and those with a continuous spectrum; see Fig.~\ref{fig1}.
\begin{figure}
\centering
\includegraphics[width=0.3\textwidth]{fig1.eps}
\caption{The positions of the discrete and continuous eigenvalues in (a) the first and second Riemann sheets of the complex $E$ plane and (b) the complex $k$ plane.}
\label{fig1}
\end{figure}
Let us refer to the former as the discrete eigenstates and the latter as the continuous eigenstates.
The discrete eigenstates have four types, namely the bound states, the anti-bound states, the resonant states and the anti-resonant states.
All of them are given by the wave function under the Siegert boundary condition\cite{Gamow28,Siegert39,Peierls59,leCouteur60,Zeldovich60,Hokkyo65,Romo68,Berggren70,Gyarmati71,Landau77,Romo80,Berggren82,Berggren96,Madrid05,Hatano08,GarciaCalderon10}
\begin{align}\label{eq2-30}
\psi(x)\propto e^{ik|x|}
\quad\mbox{for $|x|>L$}
\end{align}
with the eigenvalue $E=k^2$.
This is indeed equivalent to seeking the poles of the $S$-matrix.\cite{Landau77,Hatano08}
The Schr\"{o}dinger equation~\eqref{eq2-10}, although its Hamiltonian appears to be Hermitian, nonetheless can harbor complex eigenenergies because the system is open.
The Hamiltonian is Hermitian inside the Hilbert space but not outside it.\cite{Hatano08,Hatano13}
The eigenfunctions for the complex energy eigenvalues indeed diverge spatially and hence reside outside the Hilbert space.
Out of the four types of the discrete states, the bound states are located on the positive imaginary axis of the complex $k$ plane, \textit{i.e.}\ $k=i\kappa$ with $\kappa>0$, and hence are on the negative real axis of the first Riemann sheet of the complex $E$ plane.
Their eigenfunction decay exponentially in the form $\exp(-\kappa|x|)$, which can be confirmed by inserting $k=i\kappa$ into Eq.~\eqref{eq2-30}.
The bound states are the only discrete eigenstates that are inside the Hilbert space.
The anti-bound states are on the negative imaginary axis of the complex $k$ plane, \textit{i.e.}\ $k=i\kappa$ with $\kappa<0$, and hence on the negative real axis of the second Riemann sheet of the complex $E$ plane.
Their eigenfunctions diverge spatially in the form $\exp(|\kappa||x|)$.
We may understand the origin of these states in the following way.
Consider a closed system where we have an attractive potential around the origin with infinitely high walls on the right and left boundaries (Fig.~\ref{fig2}).
\begin{figure}
\includegraphics[width=0.4\textwidth]{fig2.eps}
\caption{(a) A closed system with a bound state~(i) inside the potential and a bound state~(ii) outside it.
(b) When the boundaries are taken away so that the system may be open, the bound state~(i) remains a bound state, but the bound state~(ii) diverges spatially, turning into an anti-bound state.}
\label{fig2}
\end{figure}
We would have only bound states, some of which would be inside the range of the attractive potential but some outside, analogously to the bonding and anti-bonding orbitals of a chemical bonding.
If we move away the boundary walls to make the system open, the states outside the potential range would turn into the anti-bound states.
This is why the wave functions of the anti-bound states spatially diverge.
The resonant states are located in the fourth quadrant of the complex $k$ plane and hence in the lower half of the second Riemann sheet of the complex $E$ plane.
Note that the term ``resonant states" here refers to eigenstates of the time-independent Schr\"{o}dinger equation~\eqref{eq2-10};
it does not refer to resonant phenomena found in the time evolution of an incoming wave packet scattered by a trapping potential as a solution of the time-dependent Schr\"{o}dinger equation.
(We will analyze the time evolution of wave packets in Secs.~\ref{sec13} and~\ref{sec14}.)
We can visualize the resonant states as eigenstates of the static Schr\"{o}dinger equation as shown in Fig.~\ref{fig3}(a).
\begin{figure}
\includegraphics[width=0.35\textwidth]{fig3.eps}
\caption{(a) Visualization of a resonant state as an eigenstate of the static Schr\"{o}dinger equation.
The number of particles decreases exponentially in the trapping potential according to the negative imaginary part of the eigenenergy.
The corresponding amount of the particles leak from the potential according to the positive real part of the eigen-wave number.
(b) Visualization of an anti-resonant state, which is the time reversal of a resonant state.
Particles gather into the trapping potential according to the negative real part of the eigen-wave number.
The number of the particles thereby increases exponentially according to the positive imaginary part of the eigenenergy.}
\label{fig3}
\end{figure}
On one hand, the state decays exponentially in time because the imaginary part of its eigenenergy $E$ is negative in the exponent of the time-evolution factor $e^{-iEt}$.
On the other hand, a resonant state has only out-going waves according to Eq.~\eqref{eq2-30} because the real part of its eigen-wave number $k$ is positive.
These two facts combined yield the view that the particles escape away from the trapping potential.
We can indeed prove the particle-number conservation on the basis of this view by noting the proportionality between the imaginary part of the eigenenergy and the real part of the eigen-wave number.\cite{Hatano08,Hatano09,Goldzak10}
Because the imaginary part of the eigen-wave number is negative, the eigenfunction spatially diverges.
This makes possible for the seemingly Hermitian Hamiltonian to have complex eigenvalues, as we explained above.
From a physical point of view, the divergence in space means that the particles eventually escape away from the potential.
The anti-resonant states are the time reversal of the resonant states as visualized in Fig.~\ref{fig3}(b).
An anti-resonant state has only in-coming waves because the real part of its eigen-wave number $k$ is negative.
The state grows exponentially in time because the imaginary part of its eigenenergy $E$ is positive in the exponent of $e^{-iEt}$.
We thereby have the view that the particles are injected into the trapping potential.
We stress here that each of resonant and anti-resonant states has an arrow of time, breaking the time-reversal symmetry.
A resonant state and an anti-resonant state always appear as a complex conjugate pair, together recovering the time-reversal symmetry that the original Schr\"{o}dinger equation observes.
We can therefore regard the appearance of the pair of time-reversal asymmetric states out of a time-reversal symmetric equation as the seed of spontaneous breaking of time reversal symmetry.
A key feature of our expansion of the Green's function and the time-evolution operator, which we will present below, is that it contains the resonant and anti-resonant states parallelly, thereby retaining the time-reversal symmetry until the last moment.
Mathematically, the time-reversal symmetry is finally broken when we choose to observe the time evolution from an initial condition or the one towards a terminal condition.
\section{System in question}
\label{sec3}
In this section, we introduce the system in question.
We consider a class of open quantum systems consisting of a dot of $N$ sites with semi-infinite leads attached to some of the dot sites (Fig~\ref{fig4}(a)).
Note that the leads must be semi-infinite in order for us to have resonant states;
otherwise the leak shown in Fig.~\ref{fig3}(a) would come back to the trapping potential and destroy the resonant states.
Any number of leads can be attached to any sites of the dot.
Let the label $i\alpha$ denote the $\alpha$th lead attached to the $i$th site of the dot, $d_i$.
The Hamiltonian which we consider here consists of three parts; namely, the dot Hamiltonian, the lead Hamiltonian and the contact Hamiltonian:
\begin{align}\label{eq10}
H\defineH_\mathrm{d}+H_\mathrm{leads}+H_\mathrm{contacts}.
\end{align}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{fig4a.eps}
\\
\vspace{\baselineskip}
\includegraphics[width=0.6\textwidth]{fig4b.eps}
\caption{(a) A schematic view of the general system which we consider in the present paper.
(b) An $(N-2)$-level quantum dot with two quantum wires. The model (a) can be a prototype of the system (b).}
\label{fig4}
\end{figure}
The first term is the tight-binding Hamiltonian inside the dot:
\begin{align}\label{eq40}
H_\mathrm{d}:=
-\sum_{\substack{i,j=1 \\ i\neq j}}^Nt_{ij}|d_i\rangle\langle d_j| + \sum_{i=1}^N \varepsilon_i |d_i\rangle\langle d_i|,
\end{align}
where $|d_i\rangle$ denotes the site basis of the $i$th site of the dot and $t_{ij}=t_{ji}\in\mathbb{R}$, so that we have no magnetic fields and the time reversal symmetry is not explicitly broken.
The chemical potentials $\varepsilon_i$ at the dot sites are all real, $\varepsilon_i\in\mathbb{R}$, and hence the dot Hamiltonian $H_\mathrm{d}$ is real symmetric under the basis set of $\{|d_i\rangle|i=1,2,\ldots,N\}$.
The second term of~\eqref{eq10} is the tight-binding Hamiltonian for the semi-infinite leads:
\begin{align}
H_\mathrm{leads}&:=\sum_{i=1}^N\sum_{{\alpha}=1}^{n_i}H_{i\alpha}^{\mathrm{lead}},
\\
H_{i\alpha}^{\mathrm{lead}}&:=
-t_\mathrm{lead}\sum_{x_{i\alpha}=1}^\infty \left(
|x_{i\alpha}+1\rangle\langle x_{i\alpha}|+|x_{i\alpha}\rangle\langle x_{i\alpha}+1|\right)
\end{align}
where $H_{i\alpha}^\mathrm{lead}$ is the tight-binding Hamiltonian of the $\alpha$th of all $n_i$ leads that are attached to the $i$th site of the dot, $|x_{i\alpha}\rangle$ with $1\leq x_{i\alpha}<\infty$ denotes the basis of a site on the lead and the hopping amplitude $t_\mathrm{lead}\in\mathbb{R}$ is common to all leads.
We hereafter put
\begin{align}
t_\mathrm{lead}=1,
\end{align}
making it the unit of the energy.
We have also put the lattice constant to unity.
The dispersion relation on a lead is therefore given by
\begin{align}\label{eq60}
E=-(e^{ik}+e^{-ik})=-2\cos k,
\end{align}
where $k$ is the wave number limited to the first Brillouin zone
\begin{align}
-\pi\leq k\leq \pi
\end{align}
because the leads consist of regular lattices.
The last term of~\eqref{eq10} is the tight-binding coupling Hamiltonian between the dot and the leads:
\begin{align}
H_\mathrm{contacts}&:=\sum_{i=1}^N\sum_{\alpha=1}^{n_i}H_{i\alpha}^{\mathrm{contact}},
\\
H_{i\alpha}^{\mathrm{contact}}&:=
-t_{i\alpha}\left(|1_{i\alpha}\rangle\langle d_i|+|d_i\rangle\langle 1_{i\alpha}|\right),
\end{align}
where $1_{i\alpha}$ denotes the end site of the lead which directly couples to the dot site $d_i$ and $t_{i\alpha}\in\mathbb{R}$.
This Hamiltonian can be a prototype model of various open quantum systems.
For example, the model in Fig.~\ref{fig4}(a) can describe an $(N-2)$-level quantum dot which is connected to two quantum wires;\cite{Sasada11} see Fig.~\ref{fig4}(b).
We next define projection operators which separate the dot space and the lead space:
\begin{align}\label{eq80}
P&:= \sum_{i=1}^N|d_i\rangle\langle d_i|,
\\
Q&:= I_\infty-P=\sum_{i=1}^N\sum_{\alpha=1}^{n_i}\sum_{x_{i\alpha}=1}^\infty|x_{i\alpha}\rangle\langle x_{i\alpha}|,
\end{align}
where $I_\infty$ is the identity operator in the whole space spanned by all site bases.
Note the relations
\begin{align}
PHP&=H_\mathrm{d},
\\
PHQ+QHP&=H_\mathrm{contacts},
\\
QHQ&=H_\mathrm{leads}.
\end{align}
As long as we restrict ourselves to the $P$ subspace, which is spanned by the dot sites $\{|d_i\rangle\}$, the $P$ operator~\eqref{eq80} is the $N\times N$ identity matrix, and therefore we will refer to it as $I_N$ where appropriate.
As we will show below in Secs.~\ref{sec6} and~\ref{sec7}, this open system generally has $2N$ pieces of discrete states with point spectra, which contain bound, anti-bound, resonant and anti-resonant states, as was reviewed in Sec.~\ref{sec2}.
The locations of the discrete states are modified from the ones indicated in Fig.~\ref{fig1} for the spatially continuous models to the ones indicated in Table~\ref{tab1} for the tight-binding models.
\begin{table}
\caption{The positions of the discrete states (the states with point spectra) in the complex $E$ plane, the complex $k$ plane and the complex $\lambda$ plane for tight-binding systems.
See Fig.~\ref{fig5} for the symbols.}
\label{tab1}
\vspace{\baselineskip}
\begin{tabular}{lc}
\hline
state & $E_n$, $k_n$ and $\lambda_n$ \\
\hline\hline
bound & on the lines AC or BD \\
\hline
anti-bound & on the lines AE or BF \\
\hline
resonant & in the plane $\delta$ \\
\hline
anti-resonant & in the plane $\gamma$ \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\centering
\includegraphics[width=0.3\textwidth]{fig5.eps}
\caption{The correspondence among the three complex planes: (a) the complex $E$ plane with the first and second Riemann sheets; (b) the complex $k$ plane; (c) the complex $\lambda$ plane.
Parts of the planes with the same index $\alpha$, $\beta$, $\gamma$, or $\delta$ correspond to each other.
The points with the same index A, B, C, D, E, or F correspond to each other.
The segment AB is the cosine energy band, which connects the two Riemann sheets as a branch cut.
The band edges A and B are the branch points.}
\label{fig5}
\end{figure}
Because the present system has the time-reversal symmetry, every term in the Hamiltonian~\eqref{eq10} can be expressed as a real symmetric matrix.
We can thereby obtain the following relations for the discrete eigenstates.
We can express the states with real eigenvalues, namely the bound and anti-bound states, as real vectors.
In other words, these states do not break the time-reversal symmetry, which the Schr\"{o}dinger equation observes.
We can therefore transpose the Schr\"{o}dinger equation $H|\psi_n\rangle=E_n|\psi_n\rangle$ to have
\begin{align}\label{eq171}
|\psi_n\rangle^T H=E_n|\psi_n\rangle^T.
\end{align}
We thereby find that the left-eigenvector is not only the Hermitian conjugate but the real transpose of the right-eigenvector:
\begin{align}
\langle\tilde{\psi}_n|:=|\psi_n\rangle^T=|\psi_n\rangle^\dag=:\langle\psi_n|
\quad\mbox{for $n\in$ bound or anti-bound states.}
\end{align}
We here unnecessarily added the tilde symbol to the left-eigenvector on the left in order to unify the symbol for the resonant and anti-resonant states, for which the left-eigenvectors are not the Hermitian conjugate of the right-eigenvector, as we will show now.
Wherever appropriate, we also use the symbol without the tilde for the left-eigenvectors, which we let denote the Hermitian conjugate of the right-eigenvectors.
Indeed, the right-eigenvectors for the resonant and anti-resonant states are not generally real vectors because their eigenvalues are complex.
In other words, these states break the time-reversal symmetry.
Since the anti-resonant state is the time reversal of the corresponding resonant state, we have
\begin{align}\label{eq:n-nbar}
E_{\bar{n}}={E_n}^\ast
\quad\mbox{and}\quad
|\psi_{\bar{n}}\rangle=|\psi_n\rangle^\ast
\quad\mbox{for $n\in$ resonant states and $\bar{n}\in$ anti-resonant states,}
\end{align}
where we let the subscript $\bar{n}$ denote the anti-resonant state that corresponds to the resonant state labeled by $n$.
The pair of a resonant state and its complex conjugate anti-resonant state $(n,\bar{n})$ recovers the time-reversal symmetry of the whole system of solutions.
Because of Eq.~\eqref{eq171}, the left-eigenvector of each resonant or anti-resonant state is real transpose but \textit{not} the Hermitian conjugate of the right-eigenvector:\cite{Zeldovich60,Hokkyo65,Romo68,Moiseyev11}
\begin{align}\label{eq201}
\langle\tilde{\psi}_n|=|\psi_n\rangle^T\neq|\psi_n\rangle^\dag
\quad\mbox{for $n\in$ resonant or anti-resonant.}
\end{align}
We thereby arrive at the relations
\begin{align}\label{eq:n-nbar1}
\langle\tilde{\psi}_n|&=|\psi_{\bar{n}}\rangle^\dag,
\\\label{eq:n-nbar2}
\langle\tilde{\psi}_{\bar{n}}|&=|\psi_n\rangle^\dag
\quad\mbox{for $n\in$ resonant states and $\bar{n}\in$ anti-resonant states;}
\end{align}
see Appendix~A of Ref.~\cite{Sasada11} for the relations with and without magnetic fields.
\section{Overview of the results}
\label{sec4}
Let us here present the results that we will prove below for the class of open quantum systems defined in Sec.~\ref{sec3}.
The main result in the present paper is the new resolution of unity in the form
\begin{align}\label{eq179}
I_N&=\sum_{n=1}^{2N}P|\psi_n\rangle\langle\tilde{\psi}_n|P.
\end{align}
The subscript $n$ in Eq.~\eqref{eq179} denotes each of all $2N$ discrete eigenstates including resonant, anti-resonant, bound and anti-bound states.
The bra and ket vectors $\langle\tilde{\psi}_n|$ and $|\psi_n\rangle$ are their left- and right-eigenvectors, respectively.
We can use this new complete set of $2N$ states to expand the Green's function $G(E)$ and the time-evolution operator $e^{-iHt}$ in the forms
\begin{align}\label{eq140}
PG(E)P&=P\frac{1}{EI_\infty-H}P=\sum_{n=1}^{2N}
P|\psi_n\rangle\frac{\lambda\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P,
\\\label{eq141}
Pe^{-iHt}P&=
\frac{1}{2\pi i}\sum_{n=1}^{2N}\int_{C_2}
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
P|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P
\left(-\lambda+\frac{1}{\lambda}\right)d\lambda,
\end{align}
where
\begin{align}\label{eq145}
\lambda=e^{ik}
\end{align}
converts the dispersion relation~\eqref{eq60} into the form
\begin{align}\label{eq146}
E(\lambda)=-\lambda-\frac{1}{\lambda}.
\end{align}
The $2N$ pieces of discrete eigenvalues $E_n$ are here represented by $\lambda_n$ defined in
\begin{align}\label{eq170}
E_n=-2\cos k_n=E(\lambda_n)=-\lambda_n-\frac{1}{\lambda_n};
\end{align}
we choose an appropriate one of $k_n$ and $\lambda_n$ from the two solutions of Eq.~\eqref{eq170} on the basis of Table~\ref{tab1}.
The integration contour $C_2$ in Eq.~\eqref{eq141} is to be specified below in Fig.~\ref{fig7}(c).
The expansion of the Green's function in the $P$ subspace, Eq.~\eqref{eq140}, as well as the basic idea of the following proofs first appeared in Ref.~\cite{Klaiman11} for a one-dimensional open quantum system.
The Green's function in the $P$ subspace is particularly important because it gives the transmission coefficient from a lead $i\alpha$ to another lead $j\beta$ in the form\cite{Fisher81,Datta95}
\begin{align}
T_{i\alpha,j\beta}(E)=(t_{i\alpha}t_{j\beta})^2\sin^2k\left|\langle d_i|G(E)|d_j\rangle\right|^2,
\end{align}
and hence the Landauer formula~\cite{Landauer57,Datta95} can convert it to the electric conductance as
\begin{align}
\mathcal{G}_{i\alpha,j\beta}=\frac{2e^2}{h}T_{i\alpha,j\beta},
\end{align}
where $e$ is the charge of an electron and $h$ is the Planck constant.
The time evolution in the $P$ subspace, Eq.~\eqref{eq141}, is useful in computing the survival probability of a particle in an excited state, which we will do in Sec.~\ref{sec11}.
A remarkable point of the expansions~\eqref{eq179}--\eqref{eq141} is the absence of the background integral.
In the conventional analysis, the resonant states were taken into account by modifying the contour of the background integral as follows.
R.~Newton~\cite{Newton60,Newton82} proved that the bound states and the continuum scattering eigenstates form a complete set of the open quantum system.
We can straightforwardly convert the proof to the tight-binding system to have
\begin{align}\label{eq220}
I_\infty=&\sum_{n\in\mbox{\scriptsize bound}}|\phi_n\rangle\langle\phi_n|
+\int_{-\pi}^\pi \frac{dk}{2\pi}|\phi_k\rangle\langle\phi_k|,
\\\label{eq230}
\frac{1}{EI_\infty-H}=&\sum_{n\in\mbox{\scriptsize bound}}|\phi_n\rangle\frac{1}{E-E_n}\langle\phi_n|
+\int_{-\pi}^\pi \frac{dk}{2\pi}|\phi_k\rangle\frac{1}{E-E_k}\langle\phi_k|,
\end{align}
where the first summation runs over all bound states and $|\phi_k\rangle$ denotes the continuum scattering eigenstate with the dispersion relation~\eqref{eq60}.
We here denoted the eigenstates by $|\phi_n\rangle$ intentionally because its normalization is different from the one of $|\psi_n\rangle$ in Eqs.~ \eqref{eq179}--\eqref{eq141} as will be given below in Eqs.~\eqref{eq841} and~\eqref{eq842}.
Note that for the bound and scattering eigenstates, the left-eigenvectors are the Hermitian conjugate of the right-eigenvectors and hence we omitted the tilde symbols from the left-eigenvectors.
We could extract the contributions of \textit{some} of the resonant states by modifying the contour on the real axis into the fourth quadrant of the complex $k$ plane:~\cite{Berggren68,Berggren70,Berggren96}
\begin{align}\label{eq240}
\int_{-\pi}^\pi \frac{dk}{2\pi}|\phi_k\rangle\langle\phi_k|
=\sum_{n\in\mbox{\scriptsize some res.}}|\phi_n\rangle\langle\tilde{\phi}_n|
+\int_C \frac{dk}{2\pi}|\phi_k\rangle\langle\phi_k|;
\end{align}
see Fig.~\ref{fig6}.
\begin{figure}
\centering
\includegraphics[width=0.4\textwidth]{fig6.eps}
\caption{Modification of the integration contour in Eqs.~\eqref{eq220} and~\eqref{eq230} over the range $-\pi\leq k \leq \pi$ on the real axis.
The portion $0\leq k\leq \pi$ is lowered to include some of the resonant states in the fourth quadrant whereas the other resonant states are excluded.
The portion $-\pi \leq k\leq 0$ is raised to keep the symmetry $-k\leftrightarrow k$.
All anti-resonant states in the third quadrant are therefore excluded.}
\label{fig6}
\end{figure}
We refer to the second term on the right-hand side as the background integral.
The approach has two drawbacks.
First, it is arbitrary to choose which resonant states we include and which we exclude.
In other words, we are splitting the left-hand side of Eq.~\eqref{eq240} arbitrarily into the resonant contributions and the background integral.
Therefore, in general, the background integral has no clear physical meaning.
Second, the formulation explicitly breaks the time-reversal symmetry.
It extracts the contributions of resonant states but not those of the time-reversal anti-resonant states.
In contrast, our approach produces expansions which perfectly maintain the time-reversal symmetry.
Furthermore, the expansions do not have any background integrals as long as they are considered in the $P$ subspace.
An integral indeed appears when we include the $Q$ subspace but with a critical difference.
For example, the resolution of unity takes the form
\begin{align}\label{eq245}
I_\infty
= \sum_{n=1}^{2N}P|\psi_n\rangle\langle\tilde{\psi}_n|P
+ \sum_{i\alpha}\int \frac{dk_{i\alpha}}{2\pi} \, Q|k_{i\alpha}\rangle\langle k_{i\alpha} |Q
\end{align}
as we will show below in Eq.~\eqref{eqIinfty}, where $|k_{i\alpha}\rangle$ denotes the scattering eigenstate~\eqref{plainwave} of the \textit{un}perturbed lead Hamiltonian $H_\mathrm{leads}=QHQ$.
Since the integral in the second term on the right-hand side is with respect to the unperturbed states, we can generally carry it out rigorously for specific matrix elements and thereby eliminate the background integral.
It is remarkable that the perturbations $H_\mathrm{contacts}=PHQ+QHP$ affect only the first term in the $P$ subspace in our formulation.
This is because we factor out the $Q$ subspace as we will show in the next section.
\section{Effective Hamiltonian}
\label{sec5}
Throughout the three sections~\ref{sec5}--\ref{sec7}, we will map the eigenvalue problem of the open quantum systems in infinite dimensions, first into a nonlinear eigenvalue problem in $N$ dimensions, and then into a generalized linear eigenvalue problem in $2N$ dimensions.
When we compute physical quantities in Secs.~\ref{sec8}--\ref{sec12}, we start from the $2N$-dimensional space, trace back first into the $N$-dimensional space and then into the infinite-dimensional space.
In the present section, we show how we factor out the $Q$ subspace and focus on the $N$-dimensional $P$ subspace.
We utilize the effective Hamiltonian for an open quantum system defined in the $P$ subspace:\cite{Livshits56,Feshbach58,Feshbach62,Rotter91,Albeverio96,Petrosky96,Petrosky97,Fyodorov97,Dittes00,Pichugin01,Sadreev03,Okolowicz03,Kunz06,Kunz08,Rotter09,Sasada11}
\begin{align}\label{eq250}
\Ham_\mathrm{eff}(E):= PHP+PHQ\frac{1}{E-QHQ}QHP.
\end{align}
It has the same discrete eigenvalues as those of the full Hamiltonian $H$; see Appendix~\ref{appA} for derivation.
Equivalently, the Green's function of the full Hamiltonian
\begin{align}
G(E):=\frac{1}{EI_\infty-H}
\end{align}
is equal to that of the effective Hamiltonian in the $P$ subspace:\cite{Sasada11}
\begin{align}\label{eq260}
PG(E)P=
P\frac{1}{EI_\infty-H}P=\frac{1}{EI_N-\Ham_\mathrm{eff}(E)}=:G_\mathrm{eff}(E);
\end{align}
see Appendix~\ref{appB} for a proof.
For the specific open quantum system defined in Fig.~\ref{fig4}(a), we can easily write down the effective Hamiltonian as follows:
\begin{align}\label{eq70}
H_\mathrm{eff}(E)&=H_\mathrm{d}+\Sigma(E),
\\\label{eq71}
\Sigma(E)&:= -e^{ik}PHQHP=\sum_{i=1}^N\sum_{\alpha=1}^{n_i}\Sigma_{i\alpha}(E),
\end{align}
where Eq.~\eqref{eqb280} gives the self-energy for the ($i\alpha$)th lead in the form~\cite{Sasada08,Sasada11}
\begin{align}\label{eq72}
\Sigma_{i\alpha}(E):=
-\left(t_{i\alpha}\right)^2e^{ik}|d_i\rangle\langle d_i|
=-\left(t_{i\alpha}\right)^2\lambda|d_i\rangle\langle d_i|
\end{align}
with the wave number $k$ being related to the particle energy $E$ as in the dispersion relation~\eqref{eq60} and $\lambda$ being defined in Eq.~\eqref{eq145}.
Note that the effective Hamiltonian~\eqref{eq70} is an $N\times N$ matrix in the basis set $\{|d_i\rangle | i=1,2,\ldots,N\}$ of the $P$ subspace.
The eigenvalue problem therefore reads as follows:
\begin{align}\label{eq100}
\left(E_nI_N-H_\mathrm{eff}(E)\right)P|\psi_n\rangle=0,
\end{align}
where $P|\psi_n\rangle$ is an $N$-dimensional column vector.
We can always reproduce the vector in the whole space from the solution $P|\psi_n\rangle$ of Eq.~\eqref{eq100} as follows:\cite{Livshits56,Feshbach58,Feshbach62,Rotter91,Albeverio96,Petrosky96,Petrosky97,Fyodorov97,Dittes00,Pichugin01,Sadreev03,Okolowicz03,Kunz06,Kunz08,Rotter09,Sasada11}
\begin{align}\label{eq300}
|\psi_n\rangle&=P|\psi_n\rangle+Q|\psi_n\rangle
\nonumber\\
&=P|\psi_n\rangle+\frac{1}{E_n-QHQ}QHP|\psi_n\rangle
\nonumber\\
&=\left(P+\frac{1}{E_n-QHQ}QHP\right)(P|\psi_n\rangle),
\end{align}
where we used Eq.~\eqref{eqa40}.
Using Eq.~\eqref{eqb240}, we can compute it explicitly for the specific open quantum system in Fig.~\ref{fig4}(a) in the form~\cite{Sasada08,Sasada11}
\begin{align}\label{eq320}
\langle x_{i\alpha} | \psi_n \rangle &= \langle x_{i\alpha} | \frac{1}{E_n-QHQ}QHP | d_i \rangle \langle d_i | \psi_n \rangle
\nonumber\\
&=t_{i\alpha}e^{ik_nx_{i\alpha}}\langle d_i|\psi_n\rangle
=t_{i\alpha}{\lambda_n}^{x_{i\alpha}}\langle d_i|\psi_n\rangle,
\end{align}
where $k_n$ and $\lambda_n$ are given in Eq.~\eqref{eq170}.
This is consistent with the fact that all eigenstates with point spectra are given under the Siegert boundary condition~\eqref{eq2-30}.\cite{Gamow28,Siegert39,Peierls59,leCouteur60,Zeldovich60,Hokkyo65,Romo68,Berggren70,Gyarmati71,Landau77,Romo80,Berggren82,Berggren96,Madrid05,Hatano08,GarciaCalderon10}
To summarize this section, we have mapped the infinite-dimensional eigenvalue problem $H|\psi\rangle=E|\psi\rangle$ to the $N$-dimensional eigenvalue problem~\eqref{eq100}.
We, however, cannot use Eq.~\eqref{eq100} for the expansion of the Green's function as it is;
it is not a standard eigenvalue problem because the effective Hamiltonian $H_\mathrm{eff}(E)$ itself depends on the energy.
Indeed, we will see below that the eigenvalue problem has $2N$ pieces of eigenvalues, not $N$ pieces.
\section{Quadratic eigenvalue problem for the effective Hamiltonian}
\label{sec6}
In the present section, we will formulate the eigenvalue problem of the effective Hamiltonian, Eq.~\eqref{eq100}, as a quadratic eigenvalue problem in the $N$-dimensional space.
We then map the problem into a generalized (linear) eigenvalue problem in an expanded $2N$-dimensional space.
Let us rewrite the eigenvalue equation~\eqref{eq100} using $\lambda$ defined in Eq.~\eqref{eq145}.
Since the particle energy now is given by Eq.~\eqref{eq146},
we have\cite{Klaiman11}
\begin{align}\label{eq120}
\left(-\lambda I_N-\frac{1}{\lambda}I_N-H_\mathrm{d}+\lambda\Theta\right)P|\psi\rangle=0,
\end{align}
where
\begin{align}
\Theta:= -\frac{1}{\lambda}\Sigma(E)=PHQHP
\end{align}
is an $N\times N$ diagonal matrix with the constant diagonal elements
\begin{align}\label{eq350}
\Theta_{ii}:=\sum_{\alpha=1}^{n_i}\left(t_{i\alpha}\right)^2
\end{align}
for $1\leq i\leq N$.
It is indeed related to the matrix $\Gamma$ in Ref.~\cite{Sasada11} as
\begin{align}\label{eq135}
\Gamma=\Theta\sin k.
\end{align}
We can further rewrite Eq.~\eqref{eq120} as
\begin{align}\label{eq150}
Z(\lambda) P|\psi\rangle=0
\end{align}
with
\begin{align}\label{eq160}
Z(\lambda):= \lambda^2(I_N-\Theta)+\lambdaH_\mathrm{d}+I_N,
\end{align}
which is a quadratic eigenvalue problem~\cite{Tisseur01} in the sense that $Z(\lambda)$ is quadratic in $\lambda$ instead of a linear function in the standard eigenvalue problem.
Note that the term in the original eigenvalue function~\eqref{eq100} is recovered by
\begin{align}\label{eq180}
E(\lambda)I_N-H_\mathrm{eff}(E(\lambda))=-\frac{1}{\lambda} Z(\lambda).
\end{align}
Writing down the eigenvalue equation~\eqref{eq100} in the form of the quadratic eigenvalue equation~\eqref{eq150} as well as further algebra leading to the expansion of the Green's function, Eq.~\eqref{eq800} below, was first done in Ref.~\cite{Klaiman11} for a one-dimensional open quantum system.
Following the standard treatment of the quadratic eigenvalue problem~\cite{Tisseur01}, we linearize Eq.~\eqref{eq150} as follows:
\begin{align}\label{eq190}
\left(A-\lambda B\right)|\Psi\rangle=0,
\end{align}
where $A$ and $B$ are $\lambda$-independent $2N\times 2N$ real symmetric matrices given by
\begin{align}\label{eq410}
A&:=
\begin{pmatrix}
0 & I_N \\
I_N & H_\mathrm{d}
\end{pmatrix},
\\\label{eq420}
B&:=
\begin{pmatrix}
I_N & 0 \\
0 & -I_N+\Theta
\end{pmatrix},
\end{align}
and $0$ here means the $N\times N$ zero matrix, yielding
\begin{align}
A-\lambda B&=\begin{pmatrix}
-\lambda I_N & I_N \\
I_N & H_\mathrm{d}+\lambda(I_N-\Theta)
\end{pmatrix},
\end{align}
while $|\Psi\rangle$ is a $\lambda$-dependent $2N$-dimensional column vector given by
\begin{align}
|\Psi\rangle:=
\begin{pmatrix}
P|\psi\rangle \\
\lambda P|\psi\rangle
\end{pmatrix}.
\end{align}
Note that the matrix $B$ is a diagonal matrix.
Equation~\eqref{eq190} is a generalized linear eigenvalue problem.
It is called `generalized' because we have the matrix $B$ in place of the identity matrix for the standard linear eigenvalue problem.
The important point here is that now the matrices $A$ and $B$ are both independent of the energy.
We can therefore use Eq.~\eqref{eq190} to expand the Green's function.
We can confirm Eq.~\eqref{eq190} as follows.
The first row gives the trivial identity
\begin{align}
-\lambda P|\psi\rangle+\left(\lambda P|\psi\rangle\right) =0,
\end{align}
whereas the second row gives
\begin{align}
P|\psi\rangle+\left[H_\mathrm{d}+\lambda(I_N-\Theta)\right]\left(\lambda P|\psi\rangle\right)=0,
\end{align}
which is equivalent to Eq.~\eqref{eq150}.
This is analogous to the technique of splitting a second-order differential equation into a set of two first-order differential equations.\cite{Tisseur01}
We can see the equivalence between Eqs.~\eqref{eq150} and~\eqref{eq190} more clearly using the two matrices\cite{Tisseur01}
\begin{align}
X(\lambda)&:=
\begin{pmatrix}
-H_\mathrm{d}-\lambda(I_N-\Theta) & I_N \\
I_N & 0
\end{pmatrix},
\\
Y_1(\lambda)&:=
\begin{pmatrix}
I_N & 0 \\
\lambda I_N & I_N
\end{pmatrix},
\\
Y_2(\lambda)&:=
\begin{pmatrix}
I_N & \lambda I_N \\
0 & I_N
\end{pmatrix}.
\end{align}
Straightforward algebra shows
\begin{align}
X(\lambda)(A-\lambda B)Y_1(\lambda)
=Y_2(\lambda)(A-\lambda B)X(\lambda)
=\begin{pmatrix}
Z(\lambda) & 0 \\
0 & I_N
\end{pmatrix}.
\end{align}
Since the determinants of the matrices $X(\lambda)$, $Y_1(\lambda)$ and $Y_2(\lambda)$ are nonzero constant, we can invert them to obtain
\begin{align}
X(\lambda)^{-1}&=\begin{pmatrix}
0 & I_N \\
I_N & H_\mathrm{d}+\lambda(I_N-\Theta)
\end{pmatrix},
\\
Y_1(\lambda)^{-1}&=\begin{pmatrix}
I_N & 0 \\
-\lambda I_N & I_N
\end{pmatrix},
\\
Y_2(\lambda)^{-1}&=\begin{pmatrix}
I_N & -\lambda I_N \\
0 & I_N
\end{pmatrix}.
\end{align}
We then have
\begin{align}
Y_1(\lambda)^{-1}(A-\lambda B)^{-1}X(\lambda)^{-1}
=X(\lambda)^{-1}(A-\lambda B)^{-1}Y_2(\lambda)^{-1}
=\begin{pmatrix}
Z(\lambda)^{-1} & 0 \\
0 & I_N
\end{pmatrix},
\end{align}
which confirms that the singularities of $(A-\lambda B)^{-1}$ are the same as those of $Z(\lambda)^{-1}$.
Note that because of Eq.~\eqref{eq180}, the Green's function of the effective Hamiltonian is now given by\cite{Tisseur01}
\begin{align}\label{eq340}
PG(E(\lambda))P=G_\mathrm{eff}(E(\lambda))
&=\frac{1}{E(\lambda)I_N-\Ham_\mathrm{eff}(E(\lambda))}
\nonumber\\
&=-\lambda Z(\lambda)^{-1}
\nonumber\\
&=-\lambda\begin{pmatrix}
I_N & 0
\end{pmatrix}
Y_1(\lambda)^{-1}(A-\lambda B)^{-1}X(\lambda)^{-1}
\begin{pmatrix}
I_N\\
0
\end{pmatrix}
\nonumber\\
&=-\lambda\begin{pmatrix}
I_N & 0
\end{pmatrix}
(A-\lambda B)^{-1}
\begin{pmatrix}
0\\
I_N
\end{pmatrix}.
\end{align}
To summarize the present section, we can solve the $N$-dimensional eigenvalue equation~\eqref{eq100} of the energy-dependent effective Hamiltonian $\Ham_\mathrm{eff}(E)$ by solving the $2N$-dimensional generalized eigenvalue equation~\eqref{eq190} of the energy-independent matrices $A$ and $B$.
Since the $2N$-dimensional generalized eigenvalue problem generally yields $2N$ eigenstates, we have $2N$ eigenstates for the $N$-dimensional energy-dependent eigenvalue problem~\eqref{eq100}.
(Some eigenvalues can become infinite for special values of the system parameters; see Appendix~H of Ref.~\cite{Sasada11})
We will use these $2N$ eigenstates to expand the Green's function.
Note that although the matrices $A$ and $B$ are both Hermitian (more precisely, real symmetric), the present generalized eigenvalue problem is still non-Hermitian (more precisely, real asymmetric) because $A$ and $B$ do not commute with each other in general: $[A,B]\neq 0$.
If $B$ is invertible in particular, the eigenvalue equation~\eqref{eq190} reduces to the standard eigenvalue equation $(B^{-1}A-\lambda I)|\Psi\rangle=0$ for the non-Hermitian (real asymmetric) matrix $B^{-1}A$.
\section{Generalized eigenvalue problem}
\label{sec7}
In order to see how we can use the $2N$ eigenstates for the Green's function expansion, let us give a tutorial review of the generalized eigenvalue problem in the context of Eq.~\eqref{eq190}.
In the present section, we will drop the projection operator $P$ for brevity;
we always work in the $P$ subspace here.
Suppose that the eigenvalue equation~\eqref{eq150} has an eigenvalue $\lambda_n$ with the right-eigenvector $|\psi_n\rangle$ and the left-eigenvector $\langle\tilde{\psi}_n|$:
\begin{align}
Z(\lambda_n)|\psi_n\rangle= \langle\tilde{\psi}_n|Z(\lambda_n)=0
\end{align}
for $n=1,2,\ldots,2N$.
Then the $2N$-dimensional column vector
\begin{align}\label{eq570}
|\Psi_n\rangle:=
\begin{pmatrix}
|\psi_n\rangle \\
\lambda_n|\psi_n\rangle
\end{pmatrix}
\end{align}
and the $2N$-dimensional row vector
\begin{align}\label{eq580}
\langle\tilde{\Psi}_n|:=
\begin{pmatrix}
\langle\tilde{\psi}_n| & \lambda_n\langle\tilde{\psi}_n|
\end{pmatrix}
\end{align}
are the right- and left-eigenvectors of the generalized eigenvalue equation~\eqref{eq190} with the eigenvalue $\lambda_n$:
\begin{align}\label{eq380}
(A-\lambda_n B)|\Psi_n\rangle=\langle\tilde{\Psi}_n|(A-\lambda_nB)=0
\end{align}
for $n=1,2,\ldots,2N$.
We can indeed confirm Eq.~\eqref{eq380} by using
\begin{align}
A-\lambda B
=X(\lambda)^{-1}
\begin{pmatrix}
Z(\lambda) & 0 \\
0 & I
\end{pmatrix}
Y_1(\lambda)^{-1}
=Y_2(\lambda)^{-1}
\begin{pmatrix}
Z(\lambda) & 0 \\
0 & I
\end{pmatrix}
X(\lambda)^{-1}
\end{align}
because
\begin{align}
Y_1(\lambda_n)^{-1}|\Psi_n\rangle&=\begin{pmatrix}
|\psi_n\rangle \\
0
\end{pmatrix},
\\
\langle\tilde{\Psi}_n|Y_2(\lambda_n)^{-1}&=\begin{pmatrix}
\langle\tilde{\psi}_n| &
0
\end{pmatrix}.
\end{align}
We now show that the vectors $\{|\Psi_n\rangle|n=1,2,\ldots,2N\}$ and $\{\langle\tilde{\Psi}_n||n=1,2,\ldots,2N\}$ constitute a bi-orthonormal set under the metric given by $B$.
Let us normalize the vector $|\psi_n\rangle$ and $\langle\tilde{\psi}_n|$ so that $|\Psi_n\rangle$ and $\langle\tilde{\Psi}_n|$ may satisfy
\begin{align}\label{eq630}
\langle\tilde{\Psi}_n|B|\Psi_n\rangle=1.
\end{align}
Equation~\eqref{eq380} then is followed by
\begin{align}
0=\langle\tilde{\Psi}_n|(A-\lambda_nB)|\Psi_n\rangle=\langle\tilde{\Psi}_n|A|\Psi_n\rangle-\lambda_n,
\end{align}
or
\begin{align}
\langle\tilde{\Psi}_n|A|\Psi_n\rangle=\lambda_n.
\end{align}
Assume that the eigenvalues have no degeneracy $\lambda_n\neq\lambda_m$ for $n\neq m$.
Then we have
\begin{align}
\langle\tilde{\Psi}_m|(A-\lambda_nB)|\Psi_n\rangle=
\langle\tilde{\Psi}_m|(A-\lambda_mB)|\Psi_n\rangle=0,
\end{align}
or
\begin{align}
(\lambda_m-\lambda_n)\langle\tilde{\Psi}_m|B|\Psi_n\rangle=0,
\end{align}
and therefore
\begin{align}\label{eq680}
\langle\tilde{\Psi}_m|B|\Psi_n\rangle=0
\end{align}
for $m\neq n$.
Equations~\eqref{eq630} with~\eqref{eq680} indicate that the vectors $\{|\Psi_n\rangle|n=1,2,\ldots,2N\}$ and $\{\langle\tilde{\Psi}_n||n=1,2,\ldots,2N\}$ constitute a bi-orthonormal pair under the metric $B$.
By constructing the $2N\times 2N$ matrices
\begin{align}\label{eq690}
U&:=
\begin{pmatrix}
|\Psi_1\rangle &
|\Psi_2\rangle &
\cdots &
|\Psi_{2N}\rangle
\end{pmatrix}
\nonumber\\
&=\begin{pmatrix}
|\psi_1\rangle &
|\psi_2\rangle &
\cdots &
|\psi_{2N}\rangle
\\
\lambda_1|\psi_1\rangle &
\lambda_2|\psi_2\rangle &
\cdots &
\lambda_{2N}|\psi_{2N}\rangle
\end{pmatrix},
\\\label{eq700}
\tilde{U}&:=
\begin{pmatrix}
\langle\tilde{\Psi}_1| \\
\langle\tilde{\Psi}_2| \\
\vdots \\
\langle\tilde{\Psi}_{2N}|
\end{pmatrix}
=\begin{pmatrix}
\langle \tilde{\psi}_1| & \lambda_1\langle \tilde{\psi}_1| \\
\langle \tilde{\psi}_2| & \lambda_2\langle \tilde{\psi}_2| \\
\vdots & \vdots \\
\langle \tilde{\psi}_{2N}| & \lambda_{2N}\langle \tilde{\psi}_{2N}| \\
\end{pmatrix},
\end{align}
we have
\begin{align}
\tilde{U}AU&=\Lambda,
\\\label{eq530}
\tilde{U}BU&=I_{2N},
\\\label{eq540}
\tilde{U}(A-\lambda B)U&=\Lambda-\lambda I_{2N},
\end{align}
where $\Lambda$ is the $2N\times 2N$ diagonal matrix with the diagonal element $\Lambda_{nn}=\lambda_n$ and $I_{2N}$ here is the $2N\times 2N$ identity matrix.
In the present case, in particular, the matrices $A$ and $B$ are real symmetric as given in Eq.~\eqref{eq410} and~\eqref{eq420};
note that both $H_\mathrm{d}$ in Eq.~\eqref{eq40} and $\Theta$ in Eq.~\eqref{eq350} can be taken to be real symmetric because there are no magnetic fields and hence we do not break the time-reversal symmetry explicitly.
The first equation of Eq.~\eqref{eq380} is therefore followed by its transpose
\begin{align}
(|\Psi_n\rangle)^T(A-\lambda_nB)=0,
\end{align}
which, compared with the second equation of Eq.~\eqref{eq380}, yields the relation
\begin{align}
\langle\tilde{\Psi}_n|=|\Psi_n\rangle^T
\end{align}
and hence
\begin{align}\label{eq755}
\langle\tilde{\psi}_n|=|\psi_n\rangle^T.
\end{align}
The last relation is the standard one found in the literature;\cite{Zeldovich60,Hokkyo65,Romo68,Moiseyev11} see Eq.~\eqref{eq201}.
\section{Eigenstate expansion of the Green's function}
\label{sec8}
We now have $2N$ eigenstates in the $2N$-dimensional space which are bi-orthogonal to each other with a metric $B$.
Using the eigenstates as bases, we will first expand the Green's function in the $2N$-dimensional space.
We will then map the expansion back into the $N$-dimensional space.
Equation~\eqref{eq540} is followed by
\begin{align}
U^{-1}(A-\lambda B)^{-1}\tilde{U}^{-1}=(\Lambda-\lambda I_{2N})^{-1},
\end{align}
or
\begin{align}\label{eq760}
(A-\lambda B)^{-1}&=U(\Lambda-\lambda I_{2N})^{-1}\tilde{U}.
\end{align}
This is the expansion of the Green's function in the $2N$-dimensional space.
Substituting Eq.~\eqref{eq760} into Eq.~\eqref{eq340}, we have\cite{Klaiman11}
\begin{align}\label{eq800}
PG(E(\lambda))P=
G_\mathrm{eff}(E(\lambda))&=
-\lambda
\begin{pmatrix}
I_N & 0
\end{pmatrix}
U(\Lambda-\lambda I_{2N})^{-1}\tilde{U}
\begin{pmatrix}
0\\
I_N
\end{pmatrix}
\nonumber\\
&
\sum_{n=1}^{2N}
P|\psi_n\rangle
\frac{\lambda\lambda_n}{\lambda-\lambda_n}
\langle \tilde{\psi}_n|P,
\end{align}
where we used the expressions~\eqref{eq690} and~\eqref{eq700} in the last line.
This is the result that we presented in Eq.~\eqref{eq140} above and was first given in Ref.~\cite{Klaiman11} for a one-dimensional open quantum system.
We stress again that the expansion is given only by the eigenstates with point spectra.
It is remarkable that we do not have any background integrals.
We also emphasize that this expansion is time-reversal symmetric;
the resonant and anti-resonant states contribute in a time-reversal symmetric way.
Appendix~\ref{appC} shows that the new expansion~\eqref{eq800} leads to the expansion in our previous work (Eqs.~(4) and~(56) in Ref.~\cite{Sasada11}):
\begin{align}\label{eq810}
P(G^\mathrm{R}(E)+G^\mathrm{A}(E))P=\sum_{n=1}^{2N}
P|\phi_n\rangle\frac{1}{E-E_n}\langle\tilde{\phi}_n|P,
\end{align}
where $G^R(E)$ and $G^A(E)$ are the retarded and advanced Green's functions of the whole Hamiltonian,
\begin{align}
G^\mathrm{R/A}(E)=\frac{1}{EI_\infty-H\pm i\delta}
\end{align}
with an infinitesimal $\delta$.
This is another piece of evidence that the expansion~\eqref{eq800} is time-reversal symmetric;
it contains both the retarded and advanced components of the Green's function.
We will see in Sec.~\ref{sec11} that the retarded component is chosen when we consider the initial-condition problem, while the advanced component is chosen when we consider the terminal-condition problem.
Incidentally, we denoted in Eq.~\eqref{eq810} the eigenstates by $|\phi_n\rangle$ intentionally because its normalization is different from the one of $|\psi_n\rangle$ in Eq.~\eqref{eq800} as follows:
\begin{align}\label{eq841}
|\phi_n\rangle&=\sqrt{1-{\lambda_n}^2}|\psi_n\rangle,
\\\label{eq842}
\langle\tilde{\phi}_n|&=\sqrt{1-{\lambda_n}^2}\langle\tilde{\psi}_n|;
\end{align}
see Appendix~\ref{appD} for the derivation.
Although the expansion~\eqref{eq800} is done in the $P$ subspace, we can always relate the eigenstates in the $P$ subspace with those in the $Q$ subspace as shown in Eqs.~\eqref{eq300} and~\eqref{eq320}.
More specifically, we can utilize Eqs.~\eqref{eqa110}--\eqref{eqa130} in order to obtain expansions outside the $P$ subspace.
For example, we have
\begin{align}\label{eq820}
QG(E(\lambda))P&=Q\frac{1}{E-QHQ}QHPG_\mathrm{eff}(E(\lambda))
\nonumber\\
&=-\sum_{n=1}^{2N}\sum_{i\alpha}{t_{i\alpha}}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi} Q|k_{i\alpha}\rangle
\frac{\sqrt{2}\sin k_{i\alpha}}{E(\lambda)+2\cos k_{i\alpha}}
\langle d_i|\psi_n\rangle
\frac{\lambda\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P,
\end{align}
where we used the expression~\eqref{eqb200}.
We can thereby express the Green's function in the whole space in the form
\begin{align}\label{eq825}
G(E(\lambda)) &:=\frac{1}{E(\lambda)I_\infty-H}
\nonumber\\
&= Q\frac{1}{E(\lambda)-QHQ}Q \nonumber\\
&+\left(I_N+Q\frac{1}{E(\lambda)-QHQ}QHP\right)G_\mathrm{eff}(E(\lambda))\left(I_N+PHQ\frac{1}{E(\lambda)-QHQ}Q\right)
\nonumber\\
&=\sum_{i\alpha} \int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi} Q|k_{i\alpha}\rangle\frac{1}{E(\lambda)+2\cos k_{i\alpha}}\langle k_{i\alpha}|Q \nonumber\\
&+\sum_{n=1}^{2N}
\left(I_N-\sum_{i\alpha}{t_{i\alpha}}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi} Q|k_{i\alpha}\rangle
\frac{\sqrt{2}\sin k_{i\alpha}}{E(\lambda)+2\cos k_{i\alpha}}\langle d_i|P
\right)
\nonumber\\
&\times\left(P|\psi_n\rangle\frac{\lambda\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P\right)
\nonumber\\
&\times
\left(I_N
-\sum_{i\alpha}t_{i\alpha}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi} P|d_i\rangle
\frac{\sqrt{2}\sin k_{i\alpha}}{E(\lambda)+2\cos k_{i\alpha}}\langle k_{i\alpha}|Q
\right)
\end{align}
The first term on the right-hand side is the expansion~\eqref{eqb190} of the Green's function of the \textit{un}perturbed lead Hamiltonian in the $Q$ subspace.
The second term is an expansion of the contributions that involve the $P$ subspace.
The expansion~\eqref{eq825} is different from the conventional expansion~\eqref{eq230} in the following two notable points:
(i) the second term is still written as a sum over all discrete eigenstates;
(ii) the integrals are taken over the \textit{un}perturbed states and hence can be carried out rigorously for specific matrix elements;
see Sec.~\ref{sec13}, for example.
The remarkable difference is due to the fact that we have essentially factorized the $P$ subspace from the $Q$ subspace.
Thanks to this factorization, the expansion~\eqref{eq825} is given in terms of the perturbed states $|\psi_n\rangle $ in the $P$ subspace and the \textit{un}perturbed states $|k_{i\alpha}\rangle$ in the $Q$ subspace.
\section{New resolution of unity}
\label{sec9}
We now prove the new resolution of unity presented in Eq.~\eqref{eq179}.
We begin the proof with the conventional resolution of unity for the whole system, namely Eq.~\eqref{eq220}.
We can cast the right-hand side of Eq.~\eqref{eq220} into the form
\begin{align}\label{eq845}
I_\infty=\frac{1}{2\pi i}\int_{C_1}\frac{1}{EI_\infty-H}dE,
\end{align}
where the integration contour $C_1$ is specified in Fig.~\ref{fig7}(a).
\begin{figure}
\includegraphics[width=\textwidth]{fig7.eps}
\caption{(a) The integration contour $C_1$ in the complex energy plane. (b) The contour mapped onto the complex wave-number plane. (c) The contour mapped onto the complex $\lambda$ plane, which is referred to as $C_2$.
The contours are all marked by yellow curves. The blue crosses indicate the bound states, the red crosses the anti-bound states and the purple crosses the resonant and anti-resonant states. The green curves indicate the scattering states.
In (a), the contour is on the first Riemann sheet, whereas the anti-bound states, the resonant states and the anti-resonant states are marked as thin crosses to indicate that they are all on the second Riemann sheet.
The scattering states double back in (a).
The black cross in (c) is the pole at the origin, which corresponds to a point infinitely far away in (a) and (b).}
\label{fig7}
\end{figure}
We project Eq.~\eqref{eq845} onto the $P$ subspace, having
\begin{align}\label{eq830}
I_N&=\frac{1}{2\pi i}\int_{C_1}P\frac{1}{EI_\infty-H}PdE
\nonumber\\
&=\frac{1}{2\pi i}\int_{C_1}G_\mathrm{eff}(E)dE,
\end{align}
where we used Eq.~\eqref{eq260}.
We further transform this integral over $E$ to an integral over $\lambda$, which yields
\begin{align}\label{eq840}
I_N&=\frac{1}{2\pi i}\int_{C_2}G_\mathrm{eff}(E(\lambda))\left(-1+\frac{1}{\lambda^2}\right)d\lambda
\nonumber\\
&=\frac{1}{2\pi i}\sum_{n=1}^{2N}\int_{C_2}P|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P
\left(-\lambda+\frac{1}{\lambda}\right)d\lambda,
\end{align}
where the integration contour $C_2$ is specified in Fig.~\ref{fig7}(c) and we used the expansion~\eqref{eq800} of the Green's function in the second line.
We can indeed obtain the same expression by putting $t=0$ in Eq.~\eqref{eq880} below in Sec.~\ref{sec10}.
The only pole that contributes to the contour $C_2$ is the one at $\lambda=0$.
Since we circle around the pole in the clockwise direction, we have
\begin{align} \label{eqIN}
I_N=\sum_{n=1}^{2N}P|\psi_n\rangle\langle\tilde{\psi}_n|P,
\end{align}
which proves Eq.~\eqref{eq220}.
It is again remarkable that this resolution of unity is free of any background integrals.
We also stress again that the summation contains the resonant and anti-resonant states in a time-reversal symmetric way.
We can extend the argument to the resolution of unity in the whole space, $I_\infty$.
Using the same procedure for the Green's function in the whole space, Eq.~\eqref{eq825}, or putting $t=0$ in Eq.~\eqref{eq895} below in Sec.~\ref{sec10}, we have
\begin{align}\label{eq881}
I_\infty
&=\frac{1}{2\pi i}\int_{C_1}dE\,Q\frac{1}{E-QHQ}Q
\nonumber\\
&+\sum_{n=1}^{2N} \frac{1}{2\pi i}
\int_{C_2}d\lambda\left(-\lambda+\frac{1}{\lambda}\right) \nonumber\\
&\times\left(I_N+Q\frac{1}{E(\lambda)-QHQ}QHP\right)
\nonumber\\
&\times\left(P|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P\right)
\nonumber\\
&\times \left(I_N+PHQ\frac{1}{E(\lambda)-QHQ}Q\right)
\end{align}
In the second term on the right-hand side, the poles that appear in addition to the ones in Eq.~\eqref{eq840} are at $\lambda=e^{\pm ik_0}$ as can be read off from Eq.~\eqref{eq820}.
These poles are on the unit circle in the complex $\lambda$ plane, and hence do not contribute to the integral along the contour $C_2$.
Only the contribution due to the pole at $\lambda=0$ remains again, which gives the same expansion as Eq.~\eqref{eqIN}.
The first term, on the other hand, gives the resolution of unity in the $Q$ subspace, which can be written in terms of the scattering eigenstates $|k_{i\alpha}\rangle$ of the \textit{un}perturbed Hamiltonian $QHQ$, and hence can be generally carried out for specific matrix elements.
We therefore arrive at
\begin{align}\label{eqIinfty}
I_\infty
= \sum_{n=1}^{2N}P|\psi_n\rangle\langle\tilde{\psi}_n|P
+ \sum_{i\alpha}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi} \, Q|k_{i\alpha}\rangle\langle k_{i\alpha} |Q,
\end{align}
which is Eq.~\eqref{eq245}.
To summarize this section, we have expressed the unit operator in terms of the perturbed discrete eigenstates on the dot and the unperturbed eigenstates on the lead;
this is the critical difference of Eq.~\eqref{eqIinfty} from the conventional resolution of unity, Eq.~\eqref{eq220}, as we emphasized above.
\section{Eigenstate expansion of the time-evolution operator}
\label{sec10}
In the present section, we show the expansion of the time-evolution operator $\exp(-iHt)$ only with respect to the eigenstates with point spectra, without any background integrals.
We first cast the time-evolution operator into the integral form
\begin{align}
e^{-iHt}=\frac{1}{2\pi i}\int_{C_1}e^{-iEt}\frac{1}{EI_\infty-H}dE,
\end{align}
where the integration contour $C_1$ is again specified in Fig.~\ref{fig7}(a).
We then consider the operator in the $P$ subspace, having
\begin{align}
Pe^{-iHt}P&=\frac{1}{2\pi i}\int_{C_1}e^{-iEt}P\frac{1}{EI_\infty-H}PdE
\nonumber\\
&=\frac{1}{2\pi i}\int_{C_1}e^{-iEt}G_\mathrm{eff}(E)dE.
\end{align}
Following the same transformation from Eq.~\eqref{eq830} to Eq.~\eqref{eq840}, we have
\begin{align}\label{eq880}
Pe^{-iHt}P&=
\frac{1}{2\pi i}\sum_{n=1}^{2N}\int_{C_2}
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
P|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P
\left(-\lambda+\frac{1}{\lambda}\right)d\lambda,
\end{align}
where the integration contour $C_2$ is again specified in Fig.~\ref{fig7}(c).
This is the result that we presented in Eq.~\eqref{eq141} above.
We will calculate the matrix element explicitly in Sec.~\ref{sec11}.
We will then see that the time-reversal symmetry is broken upon choosing $t>0$ or $t<0$.
The time-evolution operator in the other subspaces can be obtained from Eqs.~\eqref{eqa110}--\eqref{eqa130}.
For example, we have
\begin{align}\label{eq890}
Qe^{-iHt}P
=&\frac{1}{2\pi i}\int_{C_1}e^{-iEt}Q\frac{1}{EI_\infty-H}PdE
\nonumber\\
=&\frac{1}{2\pi i}\int_{C_1}e^{-iEt}Q\frac{1}{E-QHQ}QHPG_\mathrm{eff}(E)dE
\nonumber\\
=&\frac{1}{2\pi i}\sum_{n=1}^{2N}\int_{C_2}d\lambda\left(-\lambda+\frac{1}{\lambda}\right)
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
&\qquad
\sum_{i\alpha}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi}Q|k_{i\alpha}\rangle\frac{-\sqrt{2}t_{i\alpha}\sin k_{i\alpha}}{E(\lambda)+2\cos k_{i\alpha}}
\langle d_i|\psi_n\rangle
\frac{\lambda_n}{\lambda-\lambda_n}\langle \psi_n|P,
\end{align}
where we used Eq.~\eqref{eqb200} in the transformation from the second line to the third.
The time-evolution operator in the whole space is therefore expressed in the form
\begin{align}\label{eq895}
e^{-iHt}=&\frac{1}{2\pi i}
\int_{C_2}d\lambda\left(-1+\frac{1}{\lambda^2}\right) \exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
&\qquad\sum_{i\alpha}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi}Q|k_{i\alpha}\rangle\frac{1}{E(\lambda)+2\cos k_{i\alpha}}\langle k_{i\alpha}|Q
\nonumber\\
&+ \frac{1}{2\pi i}
\sum_{n=1}^{2N}
\int_{C_2}d\lambda\left(-\lambda+\frac{1}{\lambda}\right)
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
&\times\left(I_N
-\sum_{i\alpha}t_{i\alpha}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi}Q|k_{i\alpha}\rangle\frac{\sqrt{2}\sin k_{i\alpha}}{E(\lambda)+2\cos k_{i\alpha}}
\langle d_i|P
\right)
\nonumber\\
&\times\left(P|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|P\right)
\nonumber\\
&\times \left(I_N
-\sum_{i\alpha}t_{i\alpha}\int_{-\pi}^\pi \frac{dk_{i\alpha}}{2\pi}P|d_i\rangle
\frac{\sqrt{2}\sin k_{i\alpha}}{E(\lambda)+2\cos k_{i\alpha}}\langle k_{i\alpha}|Q
\right).
\end{align}
Once again, the integrals on the right-hand side are taken over the \textit{un}perturbed scattering eigenstates in the $Q$ subspace and hence can be carried out rigorously, whereas the contributions involving the $P$ subspace are given by a sum over all discrete eigenstates in a time-reversal symmetric way.
\section{Time evolution of a dot state: Survival amplitude}
\label{sec11}
We here compute the survival amplitude and more generally the matrix element
\begin{align}\label{eq901}
\langle d_j|e^{-iHt}|d_i\rangle,
\end{align}
using Eq.~\eqref{eq880}.
We will show that for $t>0$, this has exponentially decaying terms due to the resonant states, oscillatory terms due to the bound and anti-bound states and power-law decaying terms due to integrals.
We stress that the integrals are not background integrals but do appear in the coefficients of the resonant-state expansion.
For $t<0$, the exponentially decaying terms are replaced by the exponentially growing terms due to the anti-resonant states.
This is exactly where the time-reversal symmetry is broken for the first time in the sense that we have to take contributions of resonant-state poles for $t>0$ and those of anti-resonant-state poles for $t<0$;
we will discuss this point further at the end of the present section.
We compute Eq.~\eqref{eq901} from Eq.~\eqref{eq880} in the form
\begin{align}\label{eq910}
\langle d_j|e^{-iHt}|d_i\rangle=&\frac{1}{2\pi i}\sum_{n=1}^{2N}\int_{C_2}\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\langle d_j|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle \left(-\lambda+\frac{1}{\lambda}\right)d\lambda.
\end{align}
The integration has two essential singularities because of the exponential factor in the integrand, one at $\lambda=0$ and the other at $\lambda=\infty$.
In order to avoid the contributions from the essential singularities, we have to modify the contour $C_2$ in the ways specified in Fig.~\ref{fig8}(a) and~(b) for $t>0$ and $t<0$, respectively, and thereby taking different poles in the complex $\lambda$ plane.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{fig8.eps}
\caption{The contours for the integration~\eqref{eq910}: (a) the contour $C_{2+}$ for $t>0$; (b) the contour $C_{2-}$ for $t<0$.}
\label{fig8}
\end{figure}
This is where the time-reversal symmetry of the expansion is broken for the first time.
We can thus pinpoint the occurrence of the time-reversal symmetry breaking.
By the time-reversal symmetry breaking, we mean here that for $t>0$ the integration contour includes the poles of Green's function only in the upper half $\lambda$ plane (the lower half energy plane) and for $t<0$ only those in the lower half $\lambda$ plane (the upper half energy plane).
For $t>0$, the half-circle part of the contour $C_{2+}$ far away from the origin vanishes because the imaginary part of $\lambda$ diverges positively in the exponent while the imaginary part of $1/\lambda$ vanishes negatively.
On the other hand, the small half-circle part of the contour around the essential singularity at the origin $\lambda=0$ also vanishes because the imaginary part of $1/\lambda$ diverges positively in the exponent while the imaginary part of $\lambda$ vanishes negatively.
We thereby eliminate the contributions from the two essential singularities at $\lambda=0$ and $\lambda=\infty$.
Therefore, we have the contributions from all the resonant-state poles $\lambda_n$ in the upper half plane, the half-circle contributions from all bound and anti-bound states on the real axis, and the principal part of the integration over the real axis:
\begin{align}\label{eq920}
\lefteqn{
\langle d_j|e^{-iHt}|d_i\rangle
}\nonumber\\
=&\sum_{n\in\mbox{\scriptsize res.}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\lambda_n(-2i\sin k_n)
\\\label{eq921}
+\frac{1}{2}&\sum_{\substack{n\in\mbox{\scriptsize bound}\\ n\in\mbox{\scriptsize anti-bound}}}
e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\lambda_n(-2i\sin k_n)
\\\label{eq940}
-\frac{1}{2\pi i}&\sum_{n=1}^{2N}\mathop{\mathrm{P}}\int_{-\infty}^\infty d\lambda
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\langle d_j|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle \left(-\lambda+\frac{1}{\lambda}\right).
\end{align}
Note that each term in the the first line~\eqref{eq920} exponentially decays in time because the summation is restricted to the resonant states, whereas each term in the second line~\eqref{eq921} oscillates in time because the summation is restricted to the states with real eigenvalues.
Let us evaluate each integration in the third line~\eqref{eq940} in the saddle-point approximation, which will produce the power law $t^{-3/2}$.
The saddle points of the exponent $it(\lambda+1/\lambda)$ are at $\lambda=\pm 1$, which correspond to the band edges $E=\mp 2$ in Fig.~\ref{fig5}(a).
Indeed, the band edges are branch points in the complex $E$ plane, which are known to produce non-Markovian dynamics without a characteristic time scale and hence cause a power-law decay in the long-time limit.\cite{Khalfin57}
The saddle points of the integral in Eq.~\eqref{eq940} always correspond to the band edges because these are the minimum and maximum values of the energy $E(\lambda) =-( \lambda+\lambda^{-1})$, for which $dE/d\lambda=0$.
We can expand the exponent around each saddle point in the form
\begin{align}\label{eq950}
it\left(\lambda+\frac{1}{\lambda}\right)=\pm2it\pm it(\lambda\mp1)^2+\mathop{\mathrm{O}}\left((\lambda\mp1)^3\right),
\end{align}
where the upper signs correspond to the lower band edge $E=-2$ and the lower ones to the upper band edge $E=+2$.
We can convert the exponential function in the line~\eqref{eq940} into the Gaussian form $e^{-s^2}$ by choosing a new integration variable $s$ around each saddle point in the form
\begin{align}\label{eq960}
s=&\sqrt{\mp it}(\lambda\mp1) =e^{\mp i\pi/4}\sqrt{t}(\lambda \mp 1),
\end{align}
or by rotating the integration contour around the saddle points as shown in Fig.~\ref{fig9}.
\begin{figure}
\centering
\includegraphics[width=0.3\textwidth]{fig9.eps}
\caption{The rotation of the integration contour around the saddle points $\lambda=\pm 1$.}
\label{fig9}
\end{figure}
This is legitimate only when the range of the Gaussian function, which is $\sim 1/\sqrt{t}$, is narrow enough not to include any bound or anti-bound states.
In other words, the evaluation here is correct in a long time scale or if any bound or anti-bound states are not close to the band edges.
It may not be correct in a shorter time scale or when a bound or anti-bound state approaches a band edge, which was indeed observed in Ref.~\citep{Garmon13}.
Coming back to the saddle-point approximation, we can approximate the integrand in the term~\eqref{eq940} around the two saddle points as
\begin{align}
\lefteqn{
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\frac{\lambda_n}{\lambda-\lambda_n}\left(-\lambda+\frac{1}{\lambda}\right)
}
\nonumber\\
= & -\frac{\lambda_n e^{-s^2}}{(\lambda_n\mp1)-e^{\pm i\pi/4} s/\sqrt{t}}
\left(-2e^{\pm i\pi/4} \frac{s}{\sqrt{t}}+\mathop{\mathrm{O}}(s^2)\right)
\nonumber\\
= & \frac{2e^{\pm i\pi/4}\lambda_n}{\lambda_n\mp 1}\left(1+\frac{e^{\pm i\pi/4}}{\lambda_n\mp1}\frac{s}{\sqrt{t}}\right) \frac{s}{\sqrt{t}}e^{-s^2}
+\mathop{\mathrm{O}}(s^3e^{-s^2}).
\end{align}
in the long-time limit $t\to\infty$.
Because the integral of $s e^{-s^2}$ vanishes, the greatest contribution in the long-time limit comes from
\begin{align}\label{eq270}
&e^{\pm 2it}\frac{2e^{\pm i\pi/4}\lambda_n}{\lambda_n \mp 1}\frac{e^{\pm i\pi/4}}{\lambda_n \mp 1}
\int_{-\infty}^\infty e^{-s^2}\left(\frac{s}{\sqrt{t}}\right)^2
\frac{ds}{e^{\mp i\pi/4}\sqrt{t}}
\nonumber\\
=&e^{\pm 2it}\frac{\sqrt{\pi}e^{\pm 3 i\pi/4}}{t^{3/2}}\frac{\lambda_n}{(\lambda_n \mp 1)^2}
\nonumber\\
=&-e^{\pm 2it}\frac{\sqrt{\pi}e^{\pm 3 i\pi/4}}{t^{3/2}}\frac{1}{E_n\pm 2}.
\end{align}
We thereby summarize the terms~\eqref{eq920}--\eqref{eq940} as
\begin{align}\label{eq275}
\lefteqn{
\langle d_j|e^{-iHt}|d_i\rangle
}\nonumber\\
=&-\sum_{n\in\mbox{\scriptsize res.}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n}\sin k_n
\\
&-\frac{1}{2}\sum_{\substack{n\in\mbox{\scriptsize bound}\\ n\in\mbox{\scriptsize anti-bound}}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n}\sin k_n
\\
&+\frac{1}{t^{3/2}}\sum_{n=1}^{2N}\sum_{\sigma=\pm 1}e^{2\sigma it}\frac{\sqrt{\pi}e^{3\sigma i\pi/4}}{2\pi i}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\frac{1}{2\sigma+E_n}
\end{align}
in the long-time limit $t\to\infty$, where $\sigma=1$ indicates the contribution from the lower band edge $E=-2$ and $\sigma=-1$ from the upper band edge $E=+2$.
We thus have exponentially decaying terms, oscillatory terms and power-law decaying terms.
The exponent $3/2$ of the power-law decay coincides with the one given in Ref.~\cite{Garmon13} for the long-time limit.
Finally for $t<0$, we modify the contour $C_2$ into the one in Fig.~\ref{fig8}(b).
The contributions from the half-circle contour far away from the origin as well as the one close to the origin vanish.
We thus eliminate the contributions from the two essential singularities again.
We therefore have the contributions from all the anti-resonant states this time instead of the resonant states.
Employing the same argument as above, we end up with
\begin{align}\label{eq276}
\lefteqn{
\langle d_j|e^{-iHt}|d_i\rangle
}\nonumber\\
=&-\sum_{n\in\mbox{\scriptsize anti-res.}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n}\sin k_n
\\
&-\frac{1}{2}\sum_{\substack{n\in\mbox{\scriptsize bound}\\ n\in\mbox{\scriptsize anti-bound}}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n}\sin k_n
\\\label{eq278}
&-\frac{1}{t^{3/2}}\sum_{n=1}^{2N}\sum_{\sigma=\pm 1}e^{2\sigma it}\frac{\sqrt{\pi}e^{3\sigma i\pi/4}}{2\pi i}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\frac{1}{2\sigma+E_n}
\end{align}
in the long-time limit $t\to-\infty$.
An advantage in the present framework of computing the survival probability is that we can clearly see which contribution produces which time dependence;
the resonant and anti-resonant states cause the exponentially decaying and growing terms, respectively, the bound and anti-bound state cause the oscillatory terms, and the branch points cause the power-law terms.
We again stress that choosing $t>0$ or $t<0$ breaks the time-reversal symmetry in the sense that we have to take different poles for $t>0$ and $t<0$;
to be able to pinpoint the instance of time-reversal symmetry breaking is another advantage of the present framework.
The two cases of $t>0$ and $t<0$ above correspond to the initial condition problem and the terminal condition problem for the Schr\"{o}dinger equation, respectively.
The computation of the time-evolution operator is equivalent to integrating the Schr\"{o}dinger equation.
Since the Schr\"{o}dinger equation is a first-order differential equation with respect to time, we need to specify one boundary condition in order to obtain a physical solution.
The boundary condition in time can be either the initial condition or the terminal condition.
In the initial-condition problem, we seek a solution for positive times after the initial condition.
This is equivalent to computing the time evolution operator for $t>0$ and applying it to an initial ket vector.
In the terminal-condition problem, on the other hand, we seek a solution for negative times before the terminal condition.
This is equivalent to computing the time evolution operator for $t<0$ and applying it to a final bra vector.
In short, choosing $t>0$ or $t<0$ respectively corresponds to setting the boundary condition either as the initial condition or the terminal condition, and thereby breaks the time-reversal symmetry between the decaying resonant states and the growing anti-resonant states.
This view is to some extent shared by Peierls\cite{Peierls79} and Price\cite{Price06}.
We will discuss the time-reversal symmetry breaking more quantitatively in Sec.~\ref{sec13}. We will show that the broken symmetry between the resonant states and the anti-resonant states becomes exact only in the case in which the boundary condition (the initial and terminal condition) itself is time-reversal symmetric.
We will demonstrate that without the symmetry of the boundary condition, we can even see pole contributions of growing ant-resonant states in the time-evolution from an initial condition.
\section{Time evolution of a dot state: Escaping amplitude}
\label{sec12}
We next compute the escaping amplitude from a dot state,
$\langle k_{j\beta}|e^{-iHt}|d_i\rangle$
and $\langle x_{j\beta}|e^{-iHt}|d_i\rangle$.
We will show that there is an additional oscillatory term in the former matrix element due to the plane wave $|k_{j\beta}\rangle$.
For the calculation of the matrix element $\langle k_{j\beta}|e^{-iHt}|d_i\rangle$, we use Eq.~\eqref{eq890} to have
\begin{align} \label{eqEscapingA}
\langle k_{j\beta}|e^{-iHt}|d_i\rangle
=&\frac{1}{2\pi i}\sum_{n=1}^{2N}
\int_{C_2}\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
&\times
\frac{-\sqrt{2}t_{j\beta}\sin k_{j\beta}}{E(\lambda)+2\cos k_{j\beta}}
\langle d_j|\psi_n\rangle
\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle
\left(-\lambda+\frac{1}{\lambda}\right)d\lambda
\nonumber\\
=&
\frac{1}{2\pi i}\frac{t_{j\beta}}{\sqrt{2}i}\sum_{n=1}^{2N}
\int_{C_2}\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
&\times
\left(\frac{\lambda}{\lambda-e^{ik_{j\beta}}}-\frac{\lambda}{\lambda-e^{-ik_{j\beta}}}\right)
\langle d_j|\psi_n\rangle
\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle
\left(-\lambda+\frac{1}{\lambda}\right)d\lambda.
\end{align}
We therefore have an additional pole contribution from either $\lambda=e^{ik_{j\beta}}$ or $e^{-ik_{j\beta}}$ in the integration over $\lambda$.
Let us assume $k_{j\beta}>0$, which means an out-going wave.
The pole $\lambda=e^{ik_{j\beta}}$ is on the upper half of the unit circle $|\lambda|=1$ and the other pole $\lambda=e^{-ik_{j\beta}}$ on the lower half.
The former contributes for $t>0$ and the latter contributes for $t<0$.
For $t>0$, we therefore have
\begin{align}
\lefteqn{
\langle k_{j\beta}|e^{-iHt}|d_i\rangle
}
\\
=
&\sqrt{2} t_{j\beta}\sin k_{j\beta}\sum_{n:\mathrm{res}}e^{-iE_nt}\frac{2i\lambda_n\sin k_n}{E_n+2\cos k_{j\beta}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\
+
&\frac{1}{\sqrt{2}} t_{j\beta}\sin k_{j\beta}\sum_{n:\mathrm{b,ab}}e^{-iE_nt}\frac{2i\lambda_n\sin k_n}{E_n+2\cos k_{j\beta}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\
-&\sqrt{2} t_{j\beta}\sin k_{j\beta}\sum_{n=1}^{2N}
e^{2it\cos k_{j\beta}}\frac{e^{ik_{j\beta}} \lambda_n}{e^{ik_{j\beta}}-\lambda_n}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\\label{eq1160}
-\frac{1}{2\pi i}&\sqrt{2} t_{j\beta}\sin k_{j\beta}\sum_{n=1}^{2N}\mathrm{\mathop{P}}\int_{-\infty}^\infty\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
&\qquad\times\frac{1}{\lambda+1/\lambda-2\cos k_{j\beta}}
\langle d_j|\psi_n\rangle
\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle
\left(-\lambda+\frac{1}{\lambda}\right)d\lambda.
\end{align}
We evaluate the integral in the last line~\eqref{eq1160} again in the saddle-point approximation.
After the transformation to the new integration variable in Eq.~\eqref{eq960}, the integrand takes the form
\begin{align}
&
\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\frac{1}{\lambda+1/\lambda-2\cos k_{j\beta}}
\frac{\lambda_n}{\lambda-\lambda_n}\left(-\lambda+\frac{1}{\lambda}\right)
\nonumber\\
&=\frac{e^{-s^2}}{\pm 2-2\cos k_{j\beta}+\mathop{\mathrm{O}}(s^2)}
\nonumber\\
&\times\left[\frac{2e^{\pm i\pi/4}\lambda_n}{\lambda_n\mp 1}\left(1+\frac{e^{\pm i\pi/4}}{\lambda_n\mp1}\frac{s}{\sqrt{t}}\right) \frac{s}{\sqrt{t}}+\mathop{\mathrm{O}}(s^3)\right]
\end{align}
in the long-time limit $t\to\infty$.
The greatest contribution is similar to Eq.~\eqref{eq270}:
\begin{align}
-\frac{1}{\pm 2-2\cos k_{j\beta}}\frac{\sqrt{\pi}e^{\pm 3 i\pi/4}}{t^{3/2}}\frac{1}{E_n\pm 2}.
\end{align}
We thereby summarize all lines as
\begin{align}
\lefteqn{
\langle k_{j\beta}|e^{-iHt}|d_i\rangle=\sqrt{2} t_{j\beta}\sin k_{j\beta}
}
\\\label{eq1218}
&\times\left[\sum_{n:\mathrm{res}}e^{-iE_nt}\frac{2ie^{ik_n}\sin k_n}{E_n+2\cos k_{j\beta}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle \right.
\\\label{eq1219}
&+
\frac{1}{2}\sum_{n:\mathrm{b,ab}}e^{-iE_nt}\frac{2ie^{ik_n}\sin k_n}{E_n+2\cos k_{j\beta}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\\label{eq1220}
&-\sum_{n=1}^{2N}
e^{2it\cos k_{j\beta}}\frac{e^{ik_{j\beta}} e^{ik_n}}{e^{ik_{j\beta}}-e^{ik_n}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\\label{eq1221}
&\left.+\frac{1}{t^{3/2}}\sum_{n=1}^{2N}\sum_{\sigma=\pm 1}e^{2\sigma it}
\frac{\sqrt{\pi}e^{3\sigma i\pi/4}}{2\pi i}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\frac{1}{2\sigma-2\cos k_{j\beta}}\frac{1}{2\sigma+E_n} \right]
\end{align}
in the long-time limit $t\to\infty$.
For $t<0$, we similarly have
\begin{align}
\lefteqn{
\langle k_{j\beta}|e^{-iHt}|d_i\rangle=\sqrt{2} t_{j\beta}\sin k_{j\beta}
}
\\\label{eq1268}
&\times\left[\sum_{n:\mathrm{ar}}e^{-iE_nt}\frac{2ie^{ik_n}\sin k_n}{E_n+2\cos k_{j\beta}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle \right.
\\\label{eq1269}
&+
\frac{1}{2}\sum_{n:\mathrm{b,ab}}e^{-iE_nt}\frac{2ie^{ik_n}\sin k_n}{E_n+2\cos k_{j\beta}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\\label{eq1270}
&-\sum_{n=1}^{2N}
e^{2it\cos k_{j\beta}}\frac{e^{-ik_{j\beta}}e^{ik_n}}{e^{-ik_{j\beta}}-e^{ik_n}}\langle d_j|\psi_n\rangle\langle\tilde{\psi}_n|d_i\rangle
\\\label{eq1271}
&\left.-\frac{1}{t^{3/2}}\sum_{n=1}^{2N}\sum_{\sigma=\pm 1}e^{2\sigma it}
\frac{\sqrt{\pi}e^{3\sigma i\pi/4}}{2\pi i}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\frac{1}{2\sigma-2\cos k_{j\beta}}\frac{1}{E_n+2\sigma}\right]
\end{align}
in the long-time limit $t\to-\infty$.
The terms~\eqref{eq1218} and~\eqref{eq1268} decay and grow exponentially, respectively, the terms~\eqref{eq1219} and~\eqref{eq1269} oscillate in time, and the terms~\eqref{eq1221} and~\eqref{eq1271} decay as $t^{-3/2}$, which were also present in the survival amplitude~\eqref{eq275}--\eqref{eq278}.
The terms~\eqref{eq1220} and~\eqref{eq1270} are the additional oscillatory terms due to the plane wave $|k_{j\beta}\rangle$.
For the calculation of the matrix element $\langle x_{j\beta}|e^{-iHt}|d_i\rangle$, we use Eq.~\eqref{eqEscapingA} to have
\begin{align}\label{eq14000}
\langle x_{j\beta}|e^{-iHt}|d_i \rangle
&= \int_{-\pi}^{\pi}\frac{dk_{j\beta}}{2\pi}\langle x_{j\beta} |k_{j\beta}\rangle\langle k_{j\beta}|e^{-iHt}|d_i\rangle
\nonumber\\
&= \frac{1}{2\pi i} \frac{t_{j\beta}}{i}\sum_{n=1}^{2N} \int_{C_2}d\lambda\left(-\lambda+\frac{1}{\lambda}\right)\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\nonumber\\
& \times\langle d_j|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle
\nonumber\\
& \times \int_{-\pi}^\pi \frac{dk_{j\beta}}{2\pi} \sin(k_{j\beta} x_{j\beta})
\left(\frac{\lambda}{\lambda-e^{ik_{j\beta}}}-\frac{\lambda}{\lambda-e^{-ik_{j\beta}}}\right),
\end{align}
where we used $\langle x_{j\beta} |k_{j\beta}\rangle = \sqrt{2}\sin(k_{j\beta} x_{j\beta})$ in Eq.~\eqref{plainwave}.
We can easily carry out the integral with respect to $k_{j\beta}$ by changing the integration variable to $\lambda_0=e^{ik_{j\beta}}$:
\begin{align}\label{eq14100}
&\int_{-\pi}^\pi \frac{dk_{j\beta}}{2\pi} \sin(k_{j\beta} x_{j\beta})
\left(\frac{\lambda}{\lambda-e^{ik_{j\beta}}}-\frac{\lambda}{\lambda-e^{-ik_{j\beta}}}\right)
\nonumber\\
&=\int_{C_0} \frac{d\lambda_0}{2\pi i\lambda_0} \frac{{\lambda_0}^{x} - {\lambda_0}^{-x}}{2 i}
\left(\frac{\lambda}{\lambda-\lambda_0}-\frac{\lambda}{\lambda-\lambda_0^{-1}}\right),
\end{align}
where $C_0$ is the contour of the counterclockwise unit circle and we left out the subscript of $x_{j\beta}$ for brevity.
Since $\lambda$ runs on the contour $C_2$ specified in Fig.~\ref{fig7}(c), which is inside $C_0$ in the complex $\lambda_0$ plane, the pole at $\lambda_0=\lambda$ is inside the unit circle but the one at $\lambda_0=\lambda^{-1}$ is outside it.
For evaluating the term ${\lambda_0}^x$ in Eq.~\eqref{eq14100}, we take the pole $\lambda_0=\lambda$, while for the term ${\lambda_0}^{-x}$, we take the pole $\lambda_0=\lambda^{-1}$ because $x\geq 1$, each of which yields the residue $i\lambda^x/2$.
Equation~\eqref{eq14100} therefore reduces to $i\lambda^x$, which gives the following simplified form of the matrix element~\eqref{eq14000}:
\begin{align} \label{eq14200}
\langle x_{j\beta}|e^{-iHt}|d_i \rangle
=\frac{t_{j\beta}}{2\pi i}
\sum_{n=1}^{2N} \int_{C_2}\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right]
\langle d_j|\psi_n\rangle\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_i\rangle\left(-\lambda+\frac{1}{\lambda}\right)\lambda^{x_{j\beta}}d\lambda.
\end{align}
This is indeed the same as Eq.~\eqref{eq910} except for the factor $t_{j\beta}\lambda^{x_{j\beta}}$, which is consistent with Eq.~\eqref{eqb240}.
Since the factor $\lambda^x$ reduces to $\sigma^x$ in the saddle-point approximation~\eqref{eq950}--\eqref{eq270}, we accordingly modify Eqs.~\eqref{eq275}--\eqref{eq278} to have
\begin{align}
\lefteqn{
\langle x_{j\beta}|e^{-iHt}|d_i\rangle
}\nonumber\\
=&-t_{j\beta}\sum_{n\in\mbox{\scriptsize res.}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n(x_{j\beta}+1)}\sin k_n
\\
&-\frac{t_{j\beta}}{2}\sum_{\substack{n\in\mbox{\scriptsize bound}\\ n\in\mbox{\scriptsize anti-bound}}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n(x_{j\beta}+1)}\sin k_n
\\
&+\frac{t_{j\beta}}{t^{3/2}}\sum_{n=1}^{2N}\sum_{\sigma=\pm 1}e^{2\sigma it}\frac{\sqrt{\pi}e^{3\sigma i\pi/4}}{2\pi i}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\frac{\sigma^{x_{j\beta}}}{2\sigma+E_n}
\end{align}
in the limit $t\to\infty$ and
\begin{align}
\lefteqn{
\langle x_{j\beta}|e^{-iHt}|d_i\rangle
}\nonumber\\
=&-t_{j\beta}\sum_{n\in\mbox{\scriptsize anti-res.}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n(x_{j\beta}+1)}\sin k_n
\\
&-\frac{t_{j\beta}}{2}\sum_{\substack{n\in\mbox{\scriptsize bound}\\ n\in\mbox{\scriptsize anti-bound}}}e^{-iE_nt}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle 2ie^{ik_n(x_{j\beta}+1)}\sin k_n
\\
&-\frac{t_{j\beta}}{t^{3/2}}\sum_{n=1}^{2N}\sum_{\sigma=\pm 1}e^{2\sigma it}\frac{\sqrt{\pi}e^{3\sigma i\pi/4}}{2\pi i}\langle d_j|\psi_n\rangle \langle\tilde{\psi}_n|d_i\rangle\frac{\sigma^{x_{j\beta}}}{2\sigma+E_n}
\end{align}
in the limit $t\to-\infty$.
\section{Time-reversal symmetry breaking and growth from the initial condition}
\label{sec13}
We have shown in the two preceding sections that:
(i) for the time evolution from the initial condition, the resonant states contribute, giving decays;
(ii) for the time evolution to the terminal condition, the anti-resonant states contribute, giving growths.
One may then pose the following question.
Suppose that we trace back the time evolution from a terminal condition $|\psi(0)\rangle$ to a moment in the past, $t=-t_0<0$, and find a state $|\psi(-t_0)\rangle$.
The time evolution from this new initial condition $|\psi(-t_0)\rangle$ then must be a growth into the state $|\psi(0)\rangle$.
This would seem to contradict the above statement.
The answer to the question is as follows.
The above statement exactly applies only to the case in which the boundary condition chosen as the initial and terminal condition is time-reversal symmetric.
Because we go back in time from the state $|\psi(0)\rangle$, the state $|\psi(-t_0)\rangle$ consists of anti-resonant states of exponentially large amplitudes and resonant states of exponentially small amplitudes.
In other words, it is not a time-reversal symmetric state but an asymmetric state which was engineered so that it may be dominated by anti-resonant states.
This is why the anti-resonant states give exponentially growing contributions to the time evolution from the initial condition $|\psi(-t_0)\rangle$.
Mathematically, these contributions are originated when we perform an integration over $k_{j\beta}$ in terms such as Eqs.~\eqref{eq1220} and~\eqref{eq1270} with a suitable function of $k_{j\beta}$.
We will show this explicitly in the present section.
Consider the survival amplitude~\eqref{eq901}, $\langle d_j|e^{-iHt}|d_i\rangle$, again with $t>0$.
According to the expansions~\eqref{eq910}, we have terms of the form
\begin{align}\label{eq1410}
\langle d_j|e^{-iHt}|d_i\rangle&=
\sum_{n=1}^{2N}
p_n \langle d_j|\psi_n\rangle \langle \tilde{\psi}_n|d_i\rangle
\end{align}
with appropriate numbers $\{p_n\}$.
We showed in Sec.~\ref{sec11} that for $t>0$, the contributions of the resonant-state poles take over and the contributions of the anti-resonant-state poles vanish in the summation.
Let us then take the time reversal (the complex conjugate) of Eq.~\eqref{eq1410}.
Noting that the Hamiltonian is a real matrix because of the time-reversal symmetry, we have
\begin{align}\label{eq1420}
\langle d_j| e^{-iH(-t)}|d_i\rangle&=
\langle d_j|e^{-iHt}|d_i\rangle^\ast
\nonumber\\
&=
\sum_{n=1}^{2N}
{p_n}^\ast \langle d_j|\psi_n\rangle^\ast \langle \tilde{\psi}_n|d_i\rangle^\ast
\nonumber\\
&=
\sum_{n=1}^{2N}
p_{\bar{n}} \langle d_j|\psi_{\bar{n}}\rangle \langle \tilde{\psi}_{\bar{n}}|d_i\rangle,
\end{align}
where $\psi_{\bar{n}}$ is the corresponding anti-resonant state if $\psi_n$ is a resonant state;
we used Eq.~\eqref{eq:n-nbar} here.
We showed in Sec.~\ref{sec11} too that for $(-t)<0$, the contributions of the anti-resonant-state poles take over instead of the resonant-state poles.
In particular, Eq.~\eqref{eq1420} indicates that if one resonant state has a specific contribution in Eq.~\eqref{eq1410}, the corresponding anti-resonant state has a contribution of the same magnitude in Eq.~\eqref{eq1420}.
Note that this is because the initial and terminal conditions are time-reversal symmetric states: $|d_i\rangle^\ast=|d_i\rangle$ and $\langle d_j|^\ast=\langle d_j|$.
In other words, if we choose a time-reversal symmetric state as initial and terminal conditions, the following statement becomes exact:
the initial-condition problem features only the decaying states while the terminal-condition problem features the growing states, and the solutions are time reversal to each other.
An amendment to this statement is in order if we choose a time-reversal \textit{asymmetric} state as the initial condition.
We will show in the following that if we choose a specifically engineered state for the initial condition, we can even observe anti-resonant contributions in the time evolution from it.
The basic argument is as follows.
Consider an initial state where the particle is located at a site $d_i$.
This state evolves as $|\psi(t) \rangle = e^{-i Ht} |d_i\rangle$.
Say that at $t=t_0>0$ we perform a time inversion to obtain the state $|\psi(-t_0) \rangle = e^{i Ht_0} |d_i\rangle$.
The time inversion can be obtained by reversing the velocity of the particle.
We can alternatively regard this state as the solution at $t=-t_0$ of the terminal-condition problem.
We will hereafter take the state $|\psi(-t_0) \rangle$ as a new initial state and consider its time evolution from it.
In other words, we use a time-reversal asymmetric initial condition: $|\psi(-t_0)\rangle\neq|\psi(-t_0)\rangle^\ast$.
Specifically, we will consider the `T-shaped' quantum dot model shown in Fig.~\ref{fig:Tdot} and compute the amplitude $\langle x_{2{\rm R}}|e^{-i H t}|\psi(-t_0)\rangle$ that the particle is found on the right lead $x_{2\mathrm{R}}$ for $t>0$ as it is being absorbed by the dot site $d_1$.
\begin{figure}
\centering
\includegraphics[width=0.6\textwidth]{fig10}
\caption{The T-shaped quantum dot model.
The $P$ subspace is a two-dimensional space formed by the two sites $d_1$ and $d_2$ encircled by the circle, and therefore the model has four discrete eigenvalues.
The site $d_1$ is not connected to any leads directly, while the site $d_2$ is connected to two leads, which we refer to as $x_{2\mathrm{R}}$ and $x_{2\mathrm{L}}$.
In some parameter regions, there are two bound states, a resonant state and the corresponding anti-resonant state.
For the numerical demonstration here, we chose the parameter values at $t_{12}=t_{21}=t_{2\mathrm{R}}=t_{2\mathrm{L}}=1$, $\varepsilon_1=-0.85$ and $\varepsilon_2=0$.}
\label{fig:Tdot}
\end{figure}
As we argue now, this amplitude in fact contains contributions growing exponentially for $t>0$ due to the anti-resonant states.
Indeed, we have
\begin{align} \label{Tinv}
\langle x_{2{\rm R}}|e^{-i H t}|\psi(-t_0)\rangle &=\langle x_{2{\rm R}}|e^{-i H t} e^{i H t_0}|d_1\rangle
\nonumber\\
&= \langle x_{2{\rm R}}|e^{-i H (t-t_0)} |d_1\rangle.
\end{align}
As discussed in section \ref{sec12}, when time $t$ in the amplitude $\langle x_{2{\rm R}}|e^{-i H t}|d_1\rangle$ is negative, the anti-resonant states give contributions that grow exponentially in the form $\exp(-i E_\mathrm{ar} t)$ with $\mathop{\mathrm{Im}} E_\mathrm{ar}>0$ as $t$ increases. In the amplitude of Eq.~\eqref{Tinv}, this occurs for $t<t_0$. Therefore, even though we have $t>0$, the anti-resonant states give exponentially growing contributions to this amplitude until $t = t_0$.
For comparison, let us first compute the escaping probability
\begin{align}\label{eq1450}
\left|\langle x_{2\mathrm{R}} | e^{-iHt} |d_1\rangle\right|^2
\end{align}
for the T-shaped model, using Eq.~\eqref{eq14200}.
The model, in an appropriate parameter region, has only one resonant state and one anti-resonant state forming a complex-conjugate pair, along with two bound states.
We show in Fig.~\ref{fig:emitted}(a) a numerical evaluation of both the resonant-state contribution $\lambda_n=\lambda_\mathrm{res}$ and the anti-resonant-state contribution $\lambda_n = \lambda_\mathrm{ar}$ to the amplitude in Eq.~\eqref{eq14200}.
\begin{figure}
\centering
\includegraphics[width=.5\textwidth]{fig11a}\\
(a)\\
\includegraphics[width=.5\textwidth]{fig11b}\\
(b)
\caption{
(a) Resonant and anti-resonant pole contributions to the escaping probability~\eqref{eq1450} for $t=15$ (dashed lines) and $t=30$ (solid lines).
The wave packet was emitted from the dot site $d_1$ at time $t=0$ and moves to the right as time $t$ increases.
The resonant-state contributions are the blue lines (larger peaks) and the anti-resonant-state contributions are the red lines (smaller peaks).
(b) Resonant and anti-resonant pole contributions to the probability~\eqref{eq1460} with $t_0=30$ for $t=0$ (dashed lines) and $t=15$ (solid lines).
As time $t$ increases, both wave packets move to the left towards the contact site $d_2$ at $x_{2\mathrm{R}}=0$.
The resonant-state contributions are the blue lines (smaller peaks) and the anti-resonant-state contributions are the red lines (larger peaks).
For the parameter values specified in Fig.~\ref{fig:Tdot}, we have the resonant and anti-resonant poles at $\lambda_\mathrm{res} = 0.502834 - 1.21680 i$ and $\lambda_\mathrm{ar} = 0.502834 + 1.21680 i $, respectively.}
\label{fig:emitted}
\end{figure}
We see that for $t>0$ the resonant-state contribution is far greater than the anti-resonant-state contribution.
The resonant-state contribution forms most of the wave packet emitted from the dot site $d_1$.
As time increases, the wave packet shifts in the positive $x_{2\mathrm{R}}$ direction away from the dot site $d_1$ and the contact site $d_2$ at $x_{2\mathrm{R}}=0$.
For any given specific location through which the wave packet passes, there is a time period during which the amplitude decreases exponentially;
this is directly related to the exponential shape of the wave packet shown in Fig.~\ref{fig:emitted} and corresponds to the exponential decay due to the resonant-state pole.
In contrast, in the probability
\begin{align}\label{eq1460}
\left|\langle x_{2\mathrm{R}} | e^{-iHt} |\psi(-t_0)\rangle\right|^2
\end{align}
for the time-inverted state $|\psi(-t_0)\rangle=e^{+iHt_0}|d_1\rangle$, the roles of the resonant-state and anti-resonant-state contributions are exchanged as shown in Fig.~\ref{fig:emitted}(b).
The wave packet now moves in the negative $x_{2\mathrm{R}}$ direction towards the contact site $d_2$ at $x_{2\mathrm{R}}=0$.
As the wave packet passes through a given location on the lead, the amplitude grows exponentially.
All exponential growth stops at $t=t_0$ when the wave packet has been absorbed by the dot site $d_1$.
Subsequently the wave packet is re-emitted as shown in Fig.~\ref{fig:emitted}(a).
\section{Resonant scattering of a wave packet}
\label{sec14}
In the present section we describe time-reversal symmetry breaking in resonant scattering of a wave packet.
At $t=0$ we specify a time-reversal symmetric wave packet located on a lead.
We will show that again time-reversal symmetry is broken depending on whether we regard the wave packet at $t=0$ as either a terminal or an initial condition.
We will again consider the T-shaped quantum dot model described in the previous section.
At $t=0$ we have a Gaussian wave packet (Fig.~\ref{fig:Gaussian1}) on the left lead $x_{2\mathrm{L}}$, given by
\begin{align}\label{init-term-cond}
\langle x_{2\beta} |\varphi(0)\rangle\ = \delta_{\beta, \mathrm{L}} A e^{-(x_{2\mathrm{L}}-x_0)^2/\sigma^2} e^{i x_{2\mathrm{L}} k_0}\theta(x_{2\mathrm{L}})
\end{align}
where $A$ is the normalization constant, $x_0$ is the location of the peak, $\sigma$ is the width, $k_0$ is the initial momentum and $\theta(x)$ is the step function, equal to $1$ for $x\ge 1$ and $0$ otherwise. We will consider the case $k_0=0$, for which the wave packet is time-reversal symmetric.
\begin{figure}
\centerline{\includegraphics[width=.5\textwidth]{fig12}}
\caption{Gaussian wave packet
$|\langle x|\varphi(t)\rangle|^2$
at $t=0$ for the T-shaped model of Fig.~\ref{fig:Tdot}.
In this figure as well as in Fig.~\ref{fig:Gaussian2} the positions $x_{2\mathrm{L}}$ on the left lead are represented by negative values of $x=-x_{2\mathrm{L}}$, while the positions $x_{2\mathrm{R}}$ on the right lead by positive $x=x_{2\mathrm{R}}$.
The contact site $d_2$ is at $x=0$.}
\label{fig:Gaussian1}
\end{figure}
The wave packet evolves as
\begin{align}
\langle x_{2\beta} |\varphi(t)\rangle\ = \sum_{x_{2\mathrm{L}}'=1}^\infty \langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle\langle x_{2\mathrm{L}}'|\varphi(0)\rangle.
\end{align}
In order to isolate the contribution from each discrete eigenvalue (the point spectra) of the Hamiltonian,
we will decompose the transition amplitude $\langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle$ into the following terms:
\begin{align}
\langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle = \langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle_0 + \sum_{n=1}^{2 N} \langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle_n,
\end{align}
where the first term is the free time-evolution (involving only the left lead, not the dot), corresponding to the first term in the right-hand side of Eq.~\eqref{eq895}, whereas the other terms correspond to the states with $n=1,2,\cdots, 2N$ in Eq.~\eqref{eq895}.
The free-evolving term is only non-zero when $x_{2\beta}$ is on the left lead $(\beta=\mathrm{L})$ and is given by
\begin{align}\label{eq:free-evol}
\langle x_{2\beta} | e^{-i H t}|x_{2\mathrm{L}}'\rangle_0 =\delta_{\beta,\mathrm{L}} \int_{-\pi}^{\pi} \frac{dk}{2\pi} 2 \sin(k x_{2\beta} ) \sin(k x_{2\mathrm{L}}') e^{2 i t \cos k}.
\end{align}
The other terms are given by
\begin{align}
\langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle_n
=&\frac{1}{2\pi i}
\int_{C_2}\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right] \left(-\lambda+\frac{1}{\lambda}\right)
\nonumber\\
&\times \int_{-\pi}^{\pi} \frac{dk_{2\beta}}{2\pi} \int_{-\pi}^{\pi} \frac{dk_{2\mathrm{L}}'}{2\pi} 2 \sin(k_{2\beta} x_{2\beta} ) \sin(k_{2\mathrm{L}}' x_{2\mathrm{L}}') \nonumber\\
&\times
\frac{-\sqrt{2}t_{2\beta}\sin k_{2\beta}}{E(\lambda)+2\cos k_{2\beta}}
\langle d_2|\psi_n\rangle
\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_2\rangle \frac{-\sqrt{2}t_{2L}\sin k_{2\mathrm{L}}'}{E(\lambda)+2\cos k_{2\mathrm{L}}'}
d\lambda.
\nonumber\\
\end{align}
We can evaluate the integrals over $k_{2\beta}$ and $k_{2\mathrm{L}}'$ similarly to Eq.~\eqref{eq14200} as follows:
\begin{align} \label{eq:wpevol}
\langle x_{2\beta}| e^{-i H t}|x_{2\mathrm{L}}'\rangle_n
=&\frac{1}{2\pi i}
\int_{C_2}\exp\left[i\left(\lambda+\frac{1}{\lambda}\right)t\right] \left(-\lambda+\frac{1}{\lambda}\right) \lambda^{(x_{2\beta}+x_{2\mathrm{L}}')}
\nonumber\\
&\times
\langle d_2|\psi_n\rangle
\frac{\lambda_n}{\lambda-\lambda_n}\langle\tilde{\psi}_n|d_2\rangle\ d\lambda.
\end{align}
Note that this expression is independent of the lead ($\beta =\mathrm{L}$ or $\beta =\mathrm{R}$).
Therefore the $n$ components of the transition amplitude (with $n\ne 0$) are symmetric around the dot site $d_1$;
they are either incoming or outgoing scattered wave packets.
We show in Fig.~\ref{fig:Gaussian2} the components of the wave packets
\begin{align} \label{eq:wp_n}
\langle x_{2\beta}|\varphi(t)\rangle_n \equiv \sum_{x_{2\mathrm{L}}'} \langle x_{2\beta}|e^{-iHt}|x_{2\mathrm{L}}' \rangle_n \langle x_{2\mathrm{L}}'| \varphi(0)\rangle
\end{align}
at different times.
The negative times represent the time evolution towards the terminal condition \eqref{init-term-cond} at $t=0$.
The positive times represent the time evolution starting at $t=0$, which is now regarded as an initial condition.
\begin{figure}[p]
\begin{center}
\includegraphics[width=.49\textwidth]{fig13a}
\hfill
\includegraphics[width=.49\textwidth]{fig13b}
(a)\hspace{0.48\textwidth}(b)
\vspace{\baselineskip}
\includegraphics[width=.49\textwidth]{fig13c}
\hfill
\includegraphics[width=.49\textwidth]{fig13d}
(c)\hspace{0.48\textwidth}(d)
\end{center}
\caption{\textit{Continued to the next page.}}
\end{figure}
\addtocounter{figure}{-1}
\begin{figure}[p]
\begin{center}
\includegraphics[width=.49\textwidth]{fig13e}
\hfill
\includegraphics[width=.49\textwidth]{fig13f}
(e)\hspace{0.48\textwidth}(f)
\vspace{\baselineskip}
\includegraphics[width=.49\textwidth]{fig13g}
\hfill
\includegraphics[width=.49\textwidth]{fig13h}
(g)\hspace{0.48\textwidth}(h)
\end{center}
\caption{We here show dominant components of the time-evolved wave packet of Fig.~\ref{fig:Gaussian1} at eight different times: (a) $t=-20$, (b) $t=-15$, (c) $t=-10$, (d) $t=-5$, (e) $t=5$, (f) $t=10$, (g) $t=15$ and (h) $t=20$. The large (blue) wave packets are due to the free time-evolution ($\langle x_{2\beta}|\varphi(t)\rangle_0$ in Eq.~\eqref{eq:wp_n}).
The small red wave packets are due to the anti-resonant state for negative times and the green ones are due to the resonant state for positive times.
The anti-resonant wave packets move inward toward $x=0$ as $t$ increases, while the resonant ones move outward. To obtain the resonant and anti-resonant wave packets we numerically evaluated $\langle x_{2\beta}|\varphi(t)\rangle_n$ in Eq.~\eqref{eq:wp_n} for $n=\mbox{ res.}$ and $n=\mbox{anti-res.}$, respectively.
The anti-resonant wave packets at any time $-t$ are identical to the resonant wave packets at time $t$;
for example, compare (b) and (g).}
\label{fig:Gaussian2}
\end{figure}
The free time-evolution for $t>0$ is as follows (see Fig.~\ref{fig:Gaussian2}(e)--(h)).
At $t=0$ the free wave packet of the initial condition in Fig.~\ref{fig:Gaussian1} starts spreading out in both left and right directions as $t$ increases (Fig.~\ref{fig:Gaussian2}(e)). The right-hand side of the wave packet is then reflected by the dot site at $x=0$, producing interference with the portion of the wave packet that is not reflected yet (Fig.~\ref{fig:Gaussian2}(f)).
The interference pattern further continues to spread towards negative $x$ as shown in Fig.~\ref{fig:Gaussian2}(g)--(h).
The free time-evolution for $t<0$ towards $t=0$ is the exact inverse process; see Fig.~\ref{fig:Gaussian2}(a)--(d).
Starting with the interference pattern, the portion of reflected wave packet moves towards the dot site at $x=0$ as in Fig.~\ref{fig:Gaussian2}(a)--(c), bounces off the dot, forms a spread Gaussian wave packet in Fig.~\ref{fig:Gaussian2}(d), and ends up with the wave packet of the terminal condition in Fig.~\ref{fig:Gaussian1} at $t=0$.
Out of the other contributions to the time evolution, the anti-resonant-state contribution dominates for $t<0$ (small (red) wave packets) and the resonant-state contribution dominates for $t>0$ (small (green) wave packets). Note that both the resonant and anti-resonant wave packets obey causality. For example, for $t>0$ the resonant wave packets only appear after the incident wave packet reaches the dot site $d_1$ (within the quantum uncertainty), because the wave packet needs to be absorbed by the quantum dot before emission can occur. Similarly, the anti-resonant wave packets exist only {\textit{before}} the dot site $d_1$ ejects the reflected portion of the free wave packet to form the Gaussian wave packet at $t=0$.
\section{Summary and discussion}
\label{sec15}
We presented a new resolution of unity for a class of tight-binding open quantum systems and used it for expansions of the Green's function and the time evolution operator.
All of our expansions in the $P$ subspace are expressed in terms of the discrete states (the states with point spectra), not containing integrals over the continuum states (the states with a continuum spectrum).
Although they contain integrals over the continuum states outside the $P$ subspace, the integrations are taken over \textit{un}perturbed states.
This makes possible to factor out the expansion in the $P$ subspace and to keep the form of the summation over all discrete states.
Because of this feature of the expansion, we can clearly see which contribution produces which time dependence in time evolution;
the resonant states cause exponential decay, the anti-resonant states cause exponentially growth, the bound and anti-bound states cause oscillations, and the branch points cause power-law decay.
The most remarkable feature of the present expansions is that they observe the time-reversal symmetry because the resonant and anti-resonant states always come into the expansions as a pair.
The time-reversal symmetry is broken only as we try to compute matrix elements specifying the sign of $t$.
We can conceptually express this in the following way:
when we track the time evolution from an initial condition, we are forced to choose the resonant states and hence we observe decays;
when we track the time evolution towards a terminal condition, we are forced to choose the anti-resonant states and hence we observe growths.
The present expansion therefore symbolically shows that the time evolution itself does not break the time-reversal symmetry, but the choice of whether we solve the initial-condition problem or the terminal-condition problem does,
together with the condition's symmetry or anti-symmetry under time inversion.
We will detail this point in a separate publication.\cite{Ordonez14}.
The present argument is general in the sense that it does not depend on the scattering potential, but specific in the sense that the quadratic eigenvalue problem~\eqref{eq150} holds only for the tight-binding models.
A possibly related expansion in the one-dimensional continuum space has been formulated by Garc\'{i}a-Calder\'{o}n.\cite{GarciaCalderon10}
The coordinate representation of the Green's function in that study may be upgraded to an operator representation by extending the present formulation.
Directions of further possible generalizations include consideration of systems with massless linear dispersions as well as interacting systems.
The dispersion of light, $E\propto |k|$, has a singularity at $k=0$, which can yield an anomaly in the expansion.
Another interesting linear dispersion is the Dirac dispersion $E\propto k$, which does not have a singularity at $k=0$.
Particle-particle interactions will be essential in equilibration of the system and hence will be particularly important in discussing the entropy production and its connection to spontaneous time-reversal symmetry breaking.
Another ambitious generalization is the possible extension to the dynamics of the density matrix, which is governed by the Liouville-von Neumann equation
\begin{align}
i\frac{d}{dt}\rho(t)=[H,\rho(t)].
\end{align}
We may then be able to argue the monotonic time dependence of the entropy
\begin{align}
S(t)=-\mathop{\mathrm{Tr}}\rho(t)\ln\rho(t).
\end{align}
This approach can be quite different from a widely spread view of the time-reversal symmetry breaking.
When we reduce the microscopic degrees of freedom and specify the state of a system only in terms of macroscopic variables, we cannot trace back the time evolution of the system from a terminal condition.
This (possibly spuriously) suggests that coarse-graining is the reason of the time-reversal symmetry breaking.
Our approach may indicate that the time-reversal symmetry can be broken even in the level of microscopic description of the time-evolution.
\acknowledgments
One of the authors (N.H.) is deeply indebted to Dr.~S.~Klaiman for introducing to him the concept of the quadratic eigenvalue problem and to Prof.~A.~Leggett for introducing to him the articles by H.~Price.
He also greatly appreciates the Clark Way Harrison Visiting Professorship of Washington University in St.~Louis.
Both authors express sincere gratitude to Dr.~T.~Petrosky for helpful comments.
The paper is partially supported by JSPS Grant-in-Aid for Scientific Research (B) No.~22340110, a Research Grant from the Yamada Science Foundation, a Research Grant in the Natural Sciences from the Mitsubishi foundation, as well as the Holcomb Awards Committee and Woods Lecture Series at Butler University.
|
\section{INTRODUCTION}
Glitches are sudden increases in rotation
rates of pulsars, with $\Delta\Omega/\Omega \sim10^{-9}-10^{-6}$,
usually accompanied by jumps in the spin-down rate, $\Delta \dot
\Omega /\dot \Omega \sim 10^{-4}-10^{-2}$ \citep{espinoza11,yu13}.
These changes tend to relax fully or partially on long timescales
(days to years), attributed to superfluid components of the
neutron star \citep{baym69}. The electromagnetic signals of pulsars
do not change at glitches, indicating that there is no change in the external torque, so that glitches reflect angular
momentum exchange between the observed crust and interior components
of the neutron star (see \citet{weltevrede11} for a notable
exception). The energy source of large glitches is rotational kinetic energy, which
is the minimal free energy source available for the large and
frequent exchanges of angular momentum. If additional free energy
sources like elastic or magnetic energy were involved,
the accompanying energy dissipation would exceed the observational
bounds on glitch associated thermal radiation \citep{alpar98}.
Starquake models can account for the smaller
glitches typified by the Crab pulsar. Starquakes also act as triggers for the
large glitches \citep{alpar96}. A superfluid with quantized vortices
which can be pinned will explain the exchange of angular momentum
discontinuously as seen in the glitches \citep{packard72,anderson75},
if large numbers of vortices unpin in an avalanche which can be
self-organized \citep{melatos08}, or triggered by a starquake.
The vortex pinning and creep model \citep{alpar84a} explains glitches and postglitch
response in terms of moments of inertia and relaxation times of the
neutron superfluid in the neutron star crust's crystal lattice,
where vortex lines can pin to nuclei. Pinning leads
to a lag $\omega=\Omega_{s}-\Omega_{c} > 0$ between superfluid and
crustal angular velocities $\Omega_{s}$ and $\Omega_{c}$. As vortex
lines pin and unpin continually by thermal activation, the lag
$\omega$ drives an average vortex current radially outward from
the rotation axis. This "vortex creep" allows the superfluid to spin down. The system evolves towards a steady state at which
superfluid and the crust spin down at the same rate,
$\dot\Omega_{s}=\dot\Omega_{c}=\dot\Omega_{\infty}$, achieved at the steady state lag $\omega_{\infty}$. In
addition to the continual spindown by vortex creep, if $\omega$
reaches a critical value $\omega_{cr}$ beyond which pinning forces
can no longer sustain the lag, a sudden discharge of the pinned
vortices occurs. The resulting angular momentum transfer to the
crust is observed as a glitch. The superfluid rotation rate
decreases by $\delta\Omega_{s}$ and the crust rotation rate
increases by $\Delta\Omega_{c}$, so that the lag decreases by
$\delta\omega=\delta\Omega_{s}+\Delta\Omega_{c}$ at the glitch. This
glitch induced change in $\omega$ offsets the creep, leading to very
slow relaxation of the spindown rate by creep as thermal activation
has a nonlinear dependence on $\omega$. There is also a linear
regime of creep leading to prompt exponential relaxation from some
parts of the superfluid.
The superfluid core of the star is already
coupled to the crust tightly \citep{alpar84b, easson79}, on timescales short compared to the glitch rise
time, which is less than 40 seconds for the Vela pulsar \citep{dodson02}. When the interaction between vortex lines and flux lines is included the crust-core coupling timescale becomes even shorter \citep{sidery09}. The core superfluid is thus effectively included in the observed spindown of the outer
(normal matter) crust and magnetosphere. The effective crust moment
of inertia $I_c$ includes the core superfluid, so that $I_{c}\cong I$, the total moment of inertia of the star. The jump and
relaxation in the observed spindown rate of the crust indicates that the
moment of inertia fraction in crustal superfluid participating
in the glitch and postglitch relaxation is $\Delta\dot\Omega_c/
\dot\Omega_c \sim I_{\rm cr-sf}/I$. The observed $\Delta\dot\Omega_c/
\dot\Omega_c\sim10^{-3}-10^{-2}$ is consistent with the crustal
superfluid moment of inertia fraction for neutron stars. This was
proposed as a potential constraint for the equation of state
(\citet{datta93, lattimer07} and references therein).
Superfluid neutrons in the inner crust are in Bloch states of the crust lattice.
Their effective mass $m_{n}^{*}$ is larger than the bare
neutron mass $m_{n}$ \citep{chamel05,chamel12}.
This "entrainment" leaves only a fraction
of the neutron superfluid to be effectively free to store and
exchange angular momentum with the lattice
\citep{chamel06a,andersson12,chamel13}. The fractional change in the
observed spindown rate must be
multiplied by the enhancement factor $m_{n}^{*}/m_{n}> 1$. The total moment of inertia in
pinned superfluid sustaining vortex creep, $I_{\rm creep}$, must be large enough, such that $I_{\rm creep}/I \sim
(m_{n}^{*}/m_{n})\Delta\dot\Omega/\dot\Omega$.
The required moment of inertia in
components of the star with pinning/creep then exceeds the
moment of inertia of the crustal superfluid, $I_{\rm creep}> I_{\rm cr-sf}$, for
reasonable neutron star equations of state \citep{andersson12,chamel13}. This suggests the involvement of the core
superfluid in glitches and postglitch relaxation.
In the core, protons are expected to form a type II superconductor with a
dense array of flux lines \citep{baym69}. If present at all, type I superconductivity exists near the star's center, at $\rho > 2\rho_0$ \citep{jones06}. Vortices can pin to
flux lines by minimization of condensation and magnetic energies
when vortex and flux line cores overlap \citep{sauls89, ruderman98}. Arguments for type I
superconductivity based on putative precession \citep{link03} are invalidated by the possibility of vortex creep (see \citet{alpar05} and references therein).
The work of \citet{ haskell13} based on Vela glitches, concluding for either weak flux-vortex pinning or type I superconductivity also does not take creep into account. Type II superconductivity with flux-vortex pinning and creep will accommodate the observed glitch and postglitch behavior.
The bulk of the core proton superconductor-neutron superfluid region
is likely to carry a uniform or poloidal array of flux lines. The
associated moment of inertia fraction is too large, beyond the
requirement of the entrainment effect. Furthermore, a uniform or
poloidal arrangement of the flux lines offers easy directions for
vortex line motion without pinning or creep. This will make the
effect of pinning and creep in the core dependent on the angle
between the rotation and magnetic axes, making the moment of inertia
fractions involved highly variable among different pulsars. A
toroidal arrangement of flux lines, by contrast, provides a
topologically unavoidable site for pinning and creep, and can have
conditions similar to those of pinning against nuclei in the crustal
lattice \citep{sidery09}. We discuss the toroidal arrangement of flux
lines in neutron stars as a site of vortex pinning and creep and
its implications for pulsar glitches.
\section{THE TOROIDAL MAGNETIC FLUX IN NEUTRON STARS}
In normal
(non-superconducting) stars, like the progenitors of neutron stars,
purely toroidal \citep{tayler73} or poloidal \citep{wright73}
magnetic fields are unstable. \citet{spruit99} has found that for stability of magnetic fields in stratified stars, the toroidal $B_{\phi}$ to poloidal
$B_{p}$ field ratio satisfies
\begin{equation}
\frac{B^{2}_{\phi}}{B_{p}}< \frac{N r^{2}\rho^{1/2}}{l_{h}},
\label{criteria}
\end{equation}
where $\rho$ is density, $l_{h}$ is the horizontal
length scale of the perturbations which can be as large as the
stellar radius $R$, and $r$ is the cylindrical radial coordinate.
$N$, the buoyancy frequency of the stratified medium, has a typical
value of 500 s$^{-1}$ in neutron stars \citep{reisenegger92}. For a
very young neutron star which has not yet cooled below the
superconducting-superfluid transition temperatures, we obtain
$B_{\phi}\lesssim 10^{14}$ G by taking $l_{h}\sim r \sim R \sim10^{6}$
cm, $\rho\sim10^{14}$ g/cm$^{3}$ and $B_{p}\sim10^{12}$ G.
\citet{braithwaite09} has shown that stable equilibrium configurations in upper main sequence stars (neutron star progenitors) have strong toroidal fields surrounding the poloidal field. A qualitatively similar field configuration is likely to be maintained as the neutron star core
cools down and the core protons make the transition into the
type II superconductor phase. In a neutron star with a
superconducting core, a purely poloidal magnetic field in
hydromagnetic equilibrium at the crust-core boundary, though not
stable, is found to have a field strength of 10$^{14}$ G
corresponding to a surface magnetic field of
$B_{p}\sim3\times10^{12}$G, typical for radio pulsars
\citep{henriksson13}. Simulations of upper main sequence stars
\citep{braithwaite09} and neutron stars with superconducting cores
\citep{lander12,lander14} have common features. The toroidal field
component is confined within closed field lines of the poloidal
field. The poloidal field strength is maximum at the stellar center,
while toroidal field attains its largest value in the outer regions,
at $r >0.5 R$. The toroidal field is confined within the neutron
star crust for poloidal fields $\lesssim5\times10^{13}$ G \citep{lander14};
but electron differential rotation in the crust will wind the
poloidal field to generate strong toroidal flux
\citep{gourgouliatos14}, which is not likely to remain confined to
the crust, and will extend into the core. For a stable
configuration, the ratio of the toroidal and total magnetic field
energies, $E_{\rm tor}/E_{\rm mag}$ cannot be less than about 10 percent \citep{braithwaite09}. In
a model with superconducting core and proton fluid crust, this
energy ratio is found to be as large as 90 percent when crustal
toroidal fields are included \citep{ciolfi13}. Thus, simulations indicate a strong
toroidal magnetic field of 10$^{14}$ G \citep{lander14, ciolfi13},
which will be carried by the flux lines. The toroidal field is maximum at $r \sim 0.8R$,
confined within an equatorial belt of radial extension $\sim 0.1R$
\citep{lander12,lander14}. Plausible neutron star models with relatively hard equations of state, have radii $R\cong 12$ km, insensitive to the mass in the $M\sim (1 -2 )M_\odot$ range. The density in the outer core is approximately uniform, $\rho \sim \rho_0 = 2.8 \times 10^{14}$ g/ cm$^{3}$. The moment of
inertia fraction controlled by vortex lines passing through the
toroid, as shown in Fig.\ref{model} is estimated to be $I_{\rm tor}/I\approx 5\times 10^{-2}$. Depending on the
radial extent of the toroidal field, the moment of inertia of the
associated region can be comparable to and even larger than that of
the inner crust superfluid.
\begin{figure}
\centering
\includegraphics*[width=1.0\linewidth]{fig1.PNG}
\caption{Sketch showing the toroidal field (gray). The black shading marks the superfluid region, with moment of inertia $I_{\rm tor}$, effected by vortices creeping against the toroidal flux. For simplicity the magnetic and rotation axes are taken to be aligned.}
\label{model}
\end{figure}
\section{POSTGLITCH RELAXATION ACCORDING TO THE VORTEX CREEP MODEL}
The observed spindown rate $\dot\Omega_{c}$ typically displays several distinct
postglitch relaxation terms with different moments of inertia and
relaxation modes, including exponentially decaying transients and
permanent changes in rotation and spindown rates.
Depending on the pinning energy $E_{p}$ and the interior temperature $T$, vortex creep can
operate in linear or nonlinear regimes \citep{alpar89}. In the linear regime, the steady state lag $\omega_{\infty}$ is much smaller than $\omega_{cr}$. A linear creep region with moment of inertia $I_{l}$ contributes an exponentially relaxing term to the postglitch response \citep{alpar93}:
\begin{align}
\Delta\dot{\Omega}_{c}(t)=-\frac{I_{l}}{I}\frac{\delta\omega}{\tau_{l}}{\rm e}^{-t/\tau_{l}},
\label{lcreep}
\end{align}
with a relaxation time,
\begin{equation}
\label{taulin}
\tau_{l} \equiv \frac{k T }{E_{p}} \frac{R \omega_{cr}}{4 \Omega_{s} v_{0}} \exp \left( \frac{E_{p}}{kT} \right),
\end{equation}
where $v_{0}\approx 10^{7}$ cm/s is a microscopic vortex velocity. In a region where no glitch induced vortex motion takes place, $\delta\omega = \Delta\Omega_{c}$. The Vela pulsar, the
best studied glitching pulsar with glitches every $\sim 2 - 3$
years, typically exhibits three exponential transients, four
transients being resolved if the glitch is observed immediately
\citep{dodson02}. Other glitching pulsars show one or two transients
\citep{espinoza11, yu13}.
In the nonlinear creep regime $\omega_{\infty}$ is very close to $\omega_{cr}$. The contribution of a nonlinear creep region of
moment of inertia $I_{nl}$ to the postglitch
response of the observed crust spindown rate is \citep{alpar84a}:
\begin{equation}
\Delta\dot\Omega_{c}(t)=-\frac{I_{nl}}{I}\vert\dot{\Omega}\vert \left[1-\frac{1}{1+({\rm e}^{t_0/\tau_{nl}}-1){\rm e}^{-t/\tau_{nl}}}\right],
\label{ncreep}
\end{equation}
with the nonlinear creep relaxation time
\begin{equation}
\tau_{nl}\equiv \frac{kT}{E_{p}}\frac{\omega_{cr}}{\vert\dot{\Omega}\vert}.
\label{taun}
\end{equation}
We have omitted the subscript $\infty$ from $\vert\dot{\Omega}\vert$ as variations in the spindown rate do not exceed a few percent.
Vortices unpinned at a glitch move through some nonlinear creep regions. These parts of the superfluid are deeply affected by the resulting sudden decrease in the superfluid rotation rate with $\delta\omega \cong \delta\Omega_{s} \gg
\Delta\Omega_{c}$.
Creep temporarily stops, decoupling these regions from angular momentum
exchange with the crust, so that the external torque now acts on
less moment of inertia. Creep restarts after a waiting time $t_{0}
=\delta\omega/\vert\dot{\Omega}\vert$. When $t_{0}\gg \tau_{nl}$, Eq
(\ref{ncreep}) reduces to a Fermi function recovery within a time
interval of width $\sim \tau_{nl}$ around $t_{0}$. The combined
response for a distribution of waiting times $t_{0}(r) =
\delta\omega(r)/\vert\dot{\Omega}\vert$, which depends on the number
of unpinned vortices that move through each superfluid
region, can be integrated using Eq (\ref{ncreep}). If the density of
unpinned and repinned vortices is taken to be uniform throughout some superfluid
regions of total moment of inertia $I_A$, representing a mean field approach, then the integrated contribution to $\Delta\dot\Omega_{c}(t)$
is characterized by a constant second derivative $\ddot\Omega_{c}$
with which $\dot\Omega_{c}(t)$ recovers its preglitch value after a waiting
time $t_0$ corresponding to the maximum initial postglitch offset
$\delta\omega$ in these unpinning-repinning regions \citep{alpar84a}.
When initial transients are over, this slower response takes over. This behavior
prevails in the interglitch timing of the Vela pulsar, and its healing signals
the return to preglitch conditions, providing an estimate of the
time of occurrence for the next glitch. Such constant
$\ddot\Omega_{c}$ is common in older pulsars \citep{yu13}, and scales
with the parameters of the vortex creep model \citep{alpar06}. Part
of the glitch in $\Omega_{c}$, associated with moment of inertia
$I_B$, never relaxes back. This corresponds to vortex free regions
$B$ interspersed with the unpinning-repinning creep regions $A$. The
vortex free regions $B$ are analogous to capacitors in a circuit:
they do not support continuous vortex currents and do not contribute
to the spindown, transferring angular momentum only at glitches
when the unpinned vortices pass through. The glitch magnitude is
given by the angular momentum balance \citep{alpar93}
\begin{equation}
I_{c}\Delta\Omega_{c}=(I_{A}/2+I_{B})\delta\Omega_{s}.
\label{magnitude}
\end{equation}
\section{VORTEX CREEP AGAINST TOROIDAL FLUX LINES}
Junctions with toroidal flux lines inevitably constrain motion of the vortex lines.
Entrainment of the neutron and proton mass currents in the core
endows a vortex with a magnetic field of $B_v=[(m_p-m_{p}^*)/m_n
[\Phi_0/\pi\Lambda_*^2]\sim10^{14}$ G, while the magnetic field in a
flux line is $B_\Phi=[\Phi_0/\pi\Lambda_*^2]\sim10^{15}$ G
\citep{alpar84b}. The pinning energy due to magnetic interaction
between a vortex and a flux line is of the order of $E_p = (B_{v}
B_{\Phi}/4\pi) \times V$, where $V\cong 2\pi\Lambda_{*}^{3}$ is the
overlap volume with the interaction range given by the London length
$\Lambda_*=29.5[(m_p^*/m_p){x_p}^{-1}\rho_{14}^{-1}]^{1/2}$ fm
\citep{alpar84b}. In the above expressions $\Phi_0\equiv hc/2e$ is
flux quantum, $m_{p}^{*}$ and $m_{p}$ are effective and bare mass of
the proton, $x_{p}\sim 0.05$ is the proton fraction in
the outer core and $\rho_{14}$ is density in units of 10$^{14}$
g/cm$^{3}$. \citet{chamel06b} find $m_{p}^{*}/m_{p}\sim0.5-0.9$, with
$m_{p}^{*}/m_{p}\cong 0.5$ indicated by limits on crust-core coupling
from the resolution of Vela glitches \citep{dodson02}. A rough
estimate gives $E_{p}\sim 6$ MeV, though there is a wide range of
estimates $E_{p}\sim 0.1-10$ MeV \citep{sauls89, chau92}. Taking the
range of the pinning force as $\sim \Lambda_*$ and the average
length between junctions as the spacing between flux lines,
$l_\Phi=(B_{\phi}/\Phi_0)^{-1/2}$, the maximum lag $\omega_{cr}$ that can
be sustained by pinning forces is given by the Magnus equation $
\rho\kappa R\omega_{cr}= E_p/ l_\Phi\Lambda_* $. The temperature at
the crust-core boundary can be estimated for
cooling via the modified Urca process \citep{yakovlev11}, or by
relating the inner crust temperature to surface temperature
measurements \citep{gudmundsson83}. Both methods give interior
temperatures of $10^{8}-10^{9}$ K. With these ranges of $E_{p}$ and
$kT$, vortex creep will be in the
nonlinear regime. The nonlinear creep relaxation time does not have
the uncertainties of the $E_p$ estimate when divided by
$\omega_{cr}$, giving, scaling with Vela pulsar parameters,
\begin{align}
\tau&\simeq60\left(\frac{\vert\dot{\Omega}\vert}{10^{-10}\mbox{
rad/s$^{2}$}}\right)^{-1}\left(\frac{T}{10^{8}\mbox{
K}}\right)\left(\frac{R}{10^{6}\mbox{ cm}}\right)^{-1}
x_{p}^{1/2}\times\nonumber\\&\left(\frac{m_{p}^*}{m_{p}}\right)^{-1/2}\left(\frac{\rho}{10^{14}\mbox{ gr/cm$^{3}$}}\right)^{-1/2}\left(\frac{B_{\phi}}{10^{14}\mbox{ G}}\right)^{1/2}\mbox{days,}
\label{tau}
\end{align}
with $\rho = 2 \times 10^{14}$ g/cm$^{3}$ and $x_{p}= 0.05$ we
obtain $\tau \cong 30$ days. The toroidal flux line region has
no obvious structures to provide vortex traps. The crust lattice with its domains and dislocations,
can provide vortex trap regions $A$ and vortex free
regions $B$ interspersed with them, and is the locus of crust
breaking to trigger vortex unpinning. Thus it is likely
that vortices are unpinned from traps in the crust superfluid. As these vortices move outwards, they do not traverse the toroidal flux
region which lies further in. There is therefore no change in the superfluid rotation rate in the toroidal flux
region. The offset time here is determined by the glitch in the
observed rotation rate of the crust:
\begin{equation}
t_0=\frac{\Delta\Omega_c}{\vert\dot{\Omega}\vert}=7\left(\frac{t_{sd}}{10^4\mbox{
yr}}\right)
\left(\frac{\Delta\Omega_c/\Omega_c}{10^{-6}}\right)\mbox{days},
\label{offset}
\end{equation}
where $t_{sd}=\Omega/2\vert\dot{\Omega}\vert$ is pulsar spindown age. Expanding Eq.(\ref{ncreep}) in $t_{0}/\tau < 1$, we obtain
\begin{align}
\Delta\dot{\Omega}_{c}(t)=-\vert\dot{\Omega}\vert\frac{I_{\rm tor}}{I}\frac{t_{0}}{\tau}{\rm e}^{-t/\tau}.
\end{align}
We omit the mass entrainment correction $m_{p}^{*}/m_{p} < 1 $ in
the core superfluid. Its effect on estimating the moment of inertia
of the superfluid controlled by the toroidal field region will be
within the uncertainties in the actual extent of
the toroidal region. Taking into account
$m_{p}^{*}/m_{p} < 1 $ will decrease rather than increase the value
of $I_{\rm tor}$ to be inferred from $\Delta\dot{\Omega}_{c}$. This
response of the nonlinear creep against toroidal flux lines is of
the same form as the linear creep response of inner crust superfluid
associated with postglitch exponential relaxation,
Eq.(\ref{lcreep}), but with the nonlinear relaxation time and offset
time given by Eqs.(\ref{tau}) and (\ref{offset}).
\section{CONCLUSIONS}
The entrainment effect for the crustal superfluid
requires more moment of inertia in extra-crustal superfluid
regions with pinning and creep in order to account for the observed glitch related changes in the spindown rates of pulsars. The toroidal configuration of flux
lines in the outer core can provide the site for this. Creep response in this region provides an exponentially relaxing contribution to the glitch in the spindown rate.
For Vela \citep{alpar93} and Crab
\citep{alpar96} glitches, the crustal superfluid with exponential
relaxation makes up the largest part of the moment of inertia involved, $\sim
10^{-2}I$, without taking entrainment into account.
There is a particular exponential relaxation component with
$\tau \cong 32.7$ days in agreement with our estimate for the toroidal
flux line region for the first nine Vela glitches. The amplitudes of
this exponential relaxation are in the range
$\Delta\dot{\Omega}_{l}\cong(0.58-1.21)10^{-2}\dot\Omega$
\citep{chau93}. The nonlinear creep response of the toroidal flux
line region, as well as the linear creep response of crustal
superfluid employed in earlier work can contribute to the observed
$\Delta\dot{\Omega}_{l}$, as both components relax with similar
relaxation times and commensurate moments of inertia. Taking into
account vortex creep against toroidal flux lines, the moment of
inertia fraction $I_l/I$ in the crustal superfluid involved in
exponential relaxation leads to a new constraint on the total
crystalline crust moment of inertia $I_{\rm crust}$
\begin{equation}
\frac{I_l}{I} =\left(\frac{\Delta\dot{\Omega}_{l}}{\dot\Omega} -\frac{I_{\rm tor}}{I}\right) \frac{m_{n}^{*}}{m_{n}} \sim 10^{-3}- 10^{-2}< \frac{I_{\rm crust}}{I},
\end{equation}
which in principle can lead to constraints on the equation of state
\citep{lattimer07}, if uncertainties in $I_{tor}/I$, $m_{n}^{*}/m_{n}$ and
the location of the crust-core boundary are resolved. With
entrainment in the crustal superfluid, the angular momentum balance,
Eq.(\ref{magnitude}), becomes
\begin{equation}
\frac{\Delta\Omega_{c}}{\delta\Omega_{s}}=\frac{m_{n}}{m_{n}^{*}}\frac{I_{A}/2+I_{B}}{I_{c}}\lesssim\frac{m_{n}}{m_{n}^{*}}\frac{I_{\rm cr-sf}}{I-I_{\rm cr-sf}-I_{\rm tor}}.
\end{equation}
Using the
analysis of Vela pulsar glitches with the vortex creep model
\citep{alpar93,chau93} and estimate of $I_{\rm cr-sf}/I\simeq
4\times10^{-2}$ \citep{lattimer07}, we obtain $m_{n}^{*}/m_{n}\lesssim
2.2-4$. This range accounts for a density range
$\rho \gtrsim 6.4\times10^{13}$ g/cm$^{3}$ in the inner crust
\citep{chamel12,chamel13}. It should be noted that calculations of the
enhancement factor assume a bcc lattice that may not be valid
\citep{kobyakov13}; uncertainties about defects and impurities as well
as "pasta" structures may also lead
to smaller enhancement factors \citep{chamel13}. Recent work explores if plausible neutron star equations of state allow for a thicker crust to accommodate large enhancement factors \citep{steiner14, piekarewicz14}.
The magnetar 1RXS J170849.0-4000910 \citep{kaspi03} and the radio
pulsar PSR B2334+61 \citep{yuan10} underwent glitches with
exponential relaxation for both of which
$\Delta\dot\Omega_{c}/\dot\Omega_{c}\sim 0.1$, indicating moments of
inertia larger than the crustal superfluid even without entrainment.
The response of the toroidal field region can account for these as
well: In regions without glitch associated vortex motion the
response would still be exponential relaxation, and the toroidal
flux line region would contribute a similar response, providing the
extra moment of inertia. In older pulsars, the linear creep regions
of the crustal superfluid progressively become nonlinear creep
regions, and relaxation times calculated by Eq.(\ref{tau}) become longer.
In this case glitches would be step like
increases with no significant relaxation. Such behavior is indeed
observed \citep{espinoza11,yu13}. The
exponential relaxation time $\tau$ in Eq.(\ref{tau}), if identified
from pulsars of different ages as corresponding to the toroidal flux
region, can yield information on microphysical parameters and the
location of the crust-core boundary.
We have given a proof of principle about the role of vortex pinning
and creep response from the toroidal flux region in the outer core
of the neutron star. The superfluid controlled
by pinning and creep in this region can complement the crust superfluid to
accommodate the moment of inertia requirements of entrainment. The vortex
creep relaxation times are consistent with analysis of
postglitch response in the Vela glitches and scaling of the model to
other pulsars.
\acknowledgments We thank the referee for constructive comments. This work is supported by the Scientific and Technological Research
Council of Turkey (T\"{U}B\.{I}TAK) under the grant 113F354. M.A.A.
is a member of the Science Academy (Bilim Akademisi), Turkey.
|
\section{Introduction}\label{sec:introduction}
In order to provide near real-time wide area monitoring and control of power systems, synchrophasor data from PMUs are provided from across the power system by electric utilities. Typically, the PMUs are designed to record up to 30 - 60 measurements/second, and phasor data are transmitted to a centrally located wide area monitoring system (WAMS) server, where they are archived and recovered for several applications \cite{Phadke2008}. The high frequency of measurements by the PMUs and the applications involving synchrophasor data, together with the proposition to populate the future grid by a large number of PMUs, necessitates reliable and robust communications infrastructure within the power network \cite{Phadke2007}, \cite{Kirti2007}. In this paper, we make progress in this direction by devising a simple yet reliable method for transmission of phasor data from the PMUs to the WAMS server over dedicated direct communications links. We begin by summarizing the problem setup and the methodology developed to achieve the desired objective.
Consider $N$ PMUs installed on the power network. Let time be divided into frames with the duration of each frame equal to $t$ units. A time frame is further divided into $N$ slots, each of duration $\frac{t}{N}$ time units. Within a time frame, the $N$ PMUs transmit phasor data to the WAMS server via dedicated channels of finite capacity. We illustrate this setup in \figref{fig:pmu_trans1}. In the communications theory literature, this is commonly referred to as \emph{time division multiplexing}, and the scheme incurs a delay of $(N-1)t/N$ time units per frame for each PMU. Given this setup, a fundamental question that arises is the following: What is the order in which these $N$ PMUs transmit to the WAMS server? In other words, what is the transmission schedule for the $N$ PMUs, so that the WAMS server can use the received data from the ordered set of PMUs to more quickly and more reliably determine changes in the system state. We investigate this question in this paper.
\begin{figure}[t]
\includegraphics[height=1.25in,width=3.5in]{pmu_trans1.eps}
\caption{Time division multiplexing of PMU transmission. }
\label{fig:pmu_trans1}
\end{figure}
Scheduling policies intended for collecting data from PMUs should take into account the electrical properties of the power network being monitored. More precisely, the policy should be governed by the measure of electrical influence or connectedness between various network components. One way to measure the connectivity is to characterize the electrical coupling between buses in the network; the coupling can be obtained by computing the magnitude of the entries of the singular vectors obtained from the singular value decomposition (SVD) of the network matrices \cite{Wang2010}.
We answer the aforementioned question in two steps:
\begin{enumerate}[(1)]
\item The first step is the classic PMU placement/seclection problem of obtaining the optimal number ($N$) of PMUs with the goal to have either complete or incomplete network observability either in the presence/absence of zero injection measurements. Here, we consider two cases:
\begin{enumerate}[(a)]
\item the topology-based PMU placement \cite{Nuqui2005} - \nocite{Gou2008}\nocite{Gou2008a}\nocite{Baldwin1993}\nocite{Milosevic2003}\nocite{Zhang2010}\nocite{Xu2004}\nocite{Azizi2012}
\nocite{Kekatos2012}\nocite{Li2013}\nocite{Fesharaki2013}\cite{Anderson2014}, where the optimal number of $N$ PMUs is obtained from the node degree distribution of the grid; \label{case:topology} and
\item the electrical structure-based approach to PMU placement, which was first adopted in \cite{Nagananda2014}. \label{case:electrical}
\end{enumerate}
\item Next, we devise the scheduling scheme for the $N$ PMUs. If $B$ is the number of buses in the network, we construct the $B\times B$ bus admittance and resistance distance matrices (see \cite[Section III]{Cotilla-Sanchez2012}). For case (\ref{case:topology}) stated above, we pick the $N \times N$ sub-matrix of the $B\times B$ bus admittance matrix with the rows and columns corresponding to bus locations where the PMUs are installed. We perform SVD of this sub-matrix; the absolute values of elements of the resulting $N$ singular vectors are central to devising the ordering strategy for PMU transmissions. For (\ref{case:electrical}), this procedure is repeated on the resistance distance matrix.
\end{enumerate}
In order to quantify the performance of the scheduling scheme, we couple it with a fault detection algorithm in which changes in the bus susceptance parameters are detected. The detection problem is formulated using a linear errors-in-variables model, and a generalized likelihood ratio test (GLRT) based on the total least squares (TLS) methodology is presented. The performance of TLS-GLRT is analyzed with and without the proposed scheduling policy. Results demonstrate that scheduling PMU transmissions leads to an improvement in the probability of fault detection.
Some advantages of the proposed scheduling policy are:
\begin{enumerate}[1)]
\item The topology-based approach to PMU placement (case (\ref{case:topology})) incorporates less known information, since it neglects the sensitivity between power injections and nodal phase angle differences, while case (\ref{case:electrical}) is based on the complex networks perspective of the power grid, and was shown to provide a more comprehensive characterization of the electrical influence between network components (see \cite{Cotilla-Sanchez2012}). However, the general framework of the scheduling policy derived in this work remains unchanged for both cases.
\item In practice, there is a layer of phasor data concentrators (PDCs) between the PMUs and the WAMS server. Our scheduling policy is unaffected by the presence of PDCs, since it is devised from a transmitter-centric viewpoint.
\end{enumerate}
\vspace{-0.2in}
\subsection{Scheduling in the power grid}
There are different types of scheduling in power networks which have been widely examined in the literature. For instance, there are architectures for power scheduling, algorithms for traffic ({\eg}, multimedia data) scheduling on the grid, user-access scheduling procedures for smart power appliances, {\etc}. In this following we point to a few references, where each paper concerns a specific type of scheduling on the grid. In the interest of space we restrict ourselves to four references, with due credit to other valuable contributions.
In \cite{Zhou2012}, the authors proposed a power scheduling scheme for the smart grid from the perspectives of architecture, strategy and methodology based on the quality of experience (QoE); the QoE metric quantifies the customers' degree of satisfaction. A novel approach to QoE modeling was proposed, and an automatic proactive in-service strategy was employed to estimate the end user's QoE. In \cite{He2013}, a multi-time scheduling scheme, in the framework of Markov decision processes, was proposed for two classes of energy users, namely, traditional and opportunistic energy users. The reliability of the power system operation was analyzed under supply uncertainty as a result of variable and non-stationary wind generation, demand uncertainty owing to the stochastic behavior of a large number of opportunistic users and the coupling between sequential decisions across multiple timescales.
In \cite{Huang2013}, the advantages provided by wireless multimedia sensor networks in conjunction with the benefits of cognitive radio technology were exploited to devise a priority-based scheduling scheme for smart grid traffic. The traffic types included control commands, multimedia sensing data and meter readings. A joint access and scheduling approach for in-home appliances (both schedulable and critical) was devised in \cite{Chen2013} to coordinate the power usage to keep the total energy demand for the home below a target value. Uncertainties in the variations of electricity prices and distributed wind power were incorporated into the scheduling scheme to optimize the performance of the energy management controller.
The scheduling scheme devised in this paper is different from the above mentioned works in that our scheme is aimed at ordering the transmission of PMUs for transfer of phasor data to the WAMS server. Our work is concerned with improving the communications efficiency of the set of PMUs installed on the grid to quickly and reliably detect changes in the system state; this paper does not deal with the power and/or traffic scheduling that have been addressed in the aforementioned references. To the best of our knowledge, this is the first instance where a scheduling scheme for PMU transmissions has been reported in the literature.
The remainder of this paper is organized as follows. In \secref{sec:pmu_placement}, we review the PMU placement problem, employing the topology- and electrical structure-based approaches. The PMU scheduling scheme is developed in \secref{sec:schedulingPMUs}. In \secref{sec:fault_detection}, we present the fault detection framework. \secref{sec:results} includes simulation results and related discussion. \secref{sec:conclusion} concludes the paper. The advantages of using the electrical structure of the grid over its topological structure, and the construction of the resistance distance matrix are relegated to \appref{app:electrical_structure}.
\section{The PMU placement problem}\label{sec:pmu_placement}
In this section, we revisit the PMU placement problem from two different perspectives: (a) topology-based approach and (b) electrical structure-based approach.
In the general setting, for a power network with $B$ buses and $K$ branches, and for complete network observability without zero injection measurements\footnote{Zero injection measurements are present when the power system has nodes without generation or load.}, the PMU placement problem is formulated as an integer linear program as follows \cite{Gou2008}, \cite{Gou2008a}:
\begin{eqnarray}\label{eq:pmuplacement}
\nonumber \min \sum_{i=1}^{B}d_i\\
\text{such that}~ \bm{C}\bm{d} &\geq& \bm{1},
\end{eqnarray}
with
\begin{eqnarray}\label{eq:binarydecision_vector}
d_{i} = \begin{cases}
1, ~\text{if a PMU is installed at bus}~i,\\
0, ~\text{otherwise}.
\end{cases}
\end{eqnarray}
$\bm{C}$ is the $B\times B$ binary connectivity matrix of the grid, $\bm{1}$ denotes a $B\times 1$ vector of 1's. The solution to \eqref{eq:binarydecision_vector} gives the optimal number $(N)$ of PMUs to be installed on the grid.
\begin{enumerate}[(1)]
\item For PMU placement based on the topology of the grid, the entries of the bus admittance matrix are transformed into binary form and used in the problem setup \eqref{eq:pmuplacement}. In this case, $\bm{C}$ is given by
\begin{eqnarray}\label{eq:topological_A}
\bm{C}:
\begin{cases}
c_{ij} = 1, ~\text{if}~i=j,\\
c_{ij} = 1, ~\text{if}~i~\text{and}~j~\text{are connected},\\
c_{ij} = 0, ~\text{if}~i~\text{and}~j~\text{are not connected}.
\end{cases}
\end{eqnarray}
The entries $c_{ij}$ of $\bm{C}$ characterize the electrical connections between network buses $i$ and $j$.
\item For the electrical structure-based PMU placement, matrix $\bm{C}$ is derived as shown in \appref {app:electrical_structure} (see \eqref{eq:adjcency_matrix}), and this will be used in the formulation \eqref{eq:pmuplacement}. The entries $c_{ij}$ of $\bm{C}$ are obtained taking into account the sensitivity between power injections and nodal phase angles differences between various buses in the grid.
\end{enumerate}
\vspace{-0.1in}
\section{Scheduling policy for PMU transmission}\label{sec:schedulingPMUs}\vspace{-0.05in}
In this section, we present the scheduling scheme so that the optimal $N$ PMUs have a predefined order to transmit phasor data to the WAMS server. In this work, we only consider complete network observability, and without zero injection measurements. In the following, $b = 1,\dots,B$ is the bus number index, while $n = 1,\dots,N$ is the index of the optimal number of PMUs. We devise the algorithm for the $N$ PMUs obtained from the topological structure-based placement (case (\ref{case:topology}) in the previous section). The same scheme is readily applicable for the case where PMU placement is solved by employing the electrical structure-based approach (case (\ref{case:electrical})).
The following is a step-by-step procedure for the proposed PMU scheduling policy:
\begin{enumerate}[1.]
\item Obtain the optimal number $(N)$ of PMUs by solving the PMU placement problem \eqref{eq:pmuplacement}. \label{step1}
\item In the $B\times B$ bus admittance matrix, pick those rows and columns which correspond to the bus numbers where PMUs are installed. We, therefore, have an $N \times N$ sub-matrix. \label{step2}
\item Perform the SVD of the $N \times N$ sub-matrix to obtain the singular values and singular vectors. The $N \times 1$ left and right singular vectors are denoted $\bm{u}_n$ and $\bm{v}_n$, respectively, while the singular values are denoted $\sigma_n$. \label{step3}
\item Compute the magnitude of the elements of the vectors $\sigma_n\bm{u}_{n}$. Note that, the index of each entry of the vector $\sigma_n\bm{u}_{n}$ corresponds to a bus location where a PMU is installed.
\item In the vector $\sigma_1\bm{u}_{1}$, {\ie}, the first column of the $N \times N$ sub-matrix, the PMU placed on the entry with the highest magnitude transmits first. Note that, $\bm{u}_{1}$ is the eigenvector corresponding to the largest eigenvalue.
\label{transmit_first}
\item The procedure in Step \ref{transmit_first} is repeated for the remaining vectors $\sigma_n\bm{u}_{n}$, $n = 2,\dots,N$, where the $\bm{u}_{n}$s are picked in the decreasing order of the corresponding eigenvalues. \label{conflict}
\end{enumerate}
In Step \ref{conflict}, there is a possibility of conflict, which is explained via an example. Consider two vectors $\sigma_1\bm{u}_{1}$ and $\sigma_3\bm{u}_{3}$ used to schedule the transmission of PMUs during the first and third time slots, respectively. Suppose the entry having the largest magnitude in vector $\sigma_1\bm{u}_{1}$ is the same as the entry having the largest magnitude in vector $\sigma_3\bm{u}_{3}$. Then, the scheduling scheme picks the same PMU for both (first and third) time slots. To resolve this conflict, we propose the following modification to the scheduling scheme: for the third time slot, pick the entry in the vector $\sigma_3\bm{u}_{3}$ having the \emph{second largest magnitude}. If this entry is not the same as the one in vector $\sigma_2\bm{u}_{2}$ (used to schedule a PMU transmission for the second time slot), then the PMU placed on that entry is scheduled to transmit in the third time slot. This simple procedure is implemented for all the vectors $\sigma_n\bm{u}_{n}$. Note that, the priority given to PMU transmissions is solely based on the electrical connectedness of buses in the network, making it different from scheduling schemes devised for typical communications networks. In the next subsection, we explain the scheduling scheme via an illustration.
\vspace{-0.15in}
\subsection{An illustration}\label{subsec:illustration}
\begin{figure}[h]
\centering
\includegraphics[height=2.5in,width=3.25in]{pmu_schedule_14bus.eps}
\caption{Scheduling PMUs for transmission for the IEEE 14-bus network.}
\label{fig:pmu_schedule_14bus}
\end{figure}
We consider the IEEE 14-bus system to illustrate the scheduling scheme. We obtain the optimal number of PMUs to be placed on the network employing the electrical structure-based approach to PMU placement. For sake of brevity, we consider a single time frame, and implement the following steps to schedule the transmission of PMUs:
\begin{enumerate}[(1)]
\item For the 14-bus network, solving \eqref{eq:pmuplacement} yields an optimum of $N$ = 7 PMUs to be placed on buses numbered 1, 3, 8, 11, 12, 13 and 14 for complete network observability without zero injection measurements.
\item In the 14 $\times$ 14 resistance distance matrix, pick the rows and columns numbered 1, 3, 8, 11, 12, 13 and 14, thereby yielding a 7 $\times$ 7 sub-matrix.
\item Perform the SVD of the 7 $\times$ 7 sub-matrix to obtain the 7 $\times$ 1 right and left singular vectors $\bm{u}_n$ and $\bm{v}_n$, respectively, and the singular values $\sigma_n$, $n$ = 1, \dots, 7. Compute the magnitude of the elements of the vectors $\sigma_n\bm{u}_{n}$. The index of each entry of the vector $\sigma_n\bm{u}_{n}$ corresponds to a bus location where a PMU is installed. The seven vectors $\sigma_n\bm{u}_{n}$, $n$ = 1, \dots, 7 are depicted in \figref{fig:pmu_schedule_14bus}, where a column denotes a vector, while a box in each column denotes an entry of the vector. The number of boxes in each column equals the number $N$ of PMUs installed on the bus system.
\item In the vector $\sigma_1\bm{u}_{1}$ (the first column in \figref{fig:pmu_schedule_14bus}), the entry having the largest magnitude appears in the last row - marked in blue. Thus, the PMU placed on bus numbered 14 is scheduled to transmit in the first time slot.
\item In the vector $\sigma_2\bm{u}_{2}$ (the second column), the entry having the largest magnitude appears in the last row, similar to that in the vector $\sigma_1\bm{u}_{1}$, again allocating the PMU placed on bus numbered 14 to transmit in the second time slot. However, as described in the scheduling policy, this conflict is resolved by scheduling the PMU placed on the bus numbered 8, which has the second largest magnitude in the vector $\sigma_2\bm{u}_{2}$, to transmit in the second time slot.
\item Continuing in this fashion, and employing the conflict-resolution strategy, the PMUs installed on buses 14, 8, 12, 11, 3, 13 and 1 are scheduled to transmit in time slots numbered 1, 2, 3, 4, 5, 6 and 7, respectively.
\end{enumerate}
As mentioned in \secref{sec:introduction}, the scheduling policy is independent of (i) the approach (topology-based or electrical structure-based) taken to address the PMU placement problem, and (ii) the dynamic nature of the power system states. In the next section, we analyze the impact of the scheduling scheme when it is incorporated into a change detection framework within the power network.
\vspace{-0.1in}
\section{A fault detection procedure}\label{sec:fault_detection}
The power system under test is modeled using the direct current (DC) power flow model with a linear relation between the active power flow on a transmission line and the difference of the voltage angles on the two corresponding buses \cite{Grainger2003}:
\begin{eqnarray}
\text{power flow} = \text{susceptance} \times \text{voltage angle difference}. \label{eq:dc_model}
\end{eqnarray}
Given the noisy measurements of voltage angle differences and power flows across various lines in the power network, we seek to detect changes in the susceptance parameters, {\ie}, whether the susceptance are equal to some nominal known values, or have changed. We pose the change detection as a hypothesis testing problem employing a linear error-in-variables (EIV) model, which allows for noise in both sides of the linear relationship \eqref{eq:dc_model}. The standard approach to parameter estimation
in such problems is known as total least squares (TLS) \cite{Huffel2002}, while hypothesis testing in EIV models have been addressed by deriving the generalized likelihood ratio tests (GLRT) \cite{Huang2001}. The TLS-GLRT to detect changes in the susceptance parameters of a grid was first proposed in \cite{Wei2012}.
The PMUs are typically used to obtain the noisy measurements of voltage angle differences and power flows across the network. In this section, we analyze the performance of TLS-GLRT to detect changes in the susceptance parameters when these PMUs are constrained to follow the scheduling scheme to transmit the noisy measurements to a fusion center, where the detection algorithm is implemented. \secref{subsec:tlsglrt_problem} comprises the problem formulation, while the TLS-GLRT solution is presented in \secref{subsec:tlsglrt_test}.
\vspace{-0.15in}
\subsection{Problem formulation}\label{subsec:tlsglrt_problem}\vspace{-0.05in}
Let us consider a power system with $B$ buses and $K$ branches in the network, which can be modeled as an undirected graph $\mathcal{G} = (\mathcal{B}, \mathcal{K})$ with $\mathcal{B} \triangleq \{1,\dots,B\}$ and $\mathcal{K} \triangleq \{(i_1, j_1),\dots,(i_K, j_K)\}$ denoting the sets of buses and branches, respectively. The $B \times 1$ vector $\bm{\theta}(t)$ and the $B \times B$ skew-symmetric matrix $\bm{Y}(t)$ models the voltage angles and the active power flow between the buses at time slot $t$, respectively. The $(i, j)^{\text{th}}$ entry of the matrix $\bm{S}$ describes the susceptance between buses $i$ and $j$: $\bm{S}_{ij} = \bm{S}_{ji}$ if $(i, j) \in \mathcal{K}$ and $\bm{S}_{ij} = 0$ otherwise. From \eqref{eq:dc_model}, we have
\begin{eqnarray}
\bm{Y}_{ij}(t) = \bm{S}_{ij}\left(\theta_i(t) - \theta_j(t)\right); t = 1,\dots,T. \label{eq:dc_powerflow}
\end{eqnarray}
Let $\bm{s}$ be a $K \times 1$ vector with elements $\bm{S}_{ij}$ and let $K \times 1$ vector $\bm{z}(t)$ be defined to collect $\bm{Y}_{ij}(t)$, for $(i, j) \in \mathcal{K}$ and $i > j$. Thus, \eqref{eq:dc_powerflow} can be written as
\begin{eqnarray}
\bm{z}(t) = \text{diag}(\bm{s})\bm{D}\bm{\theta}(t), \label{eq:dc_alternative}
\end{eqnarray}
where the $K \times B$ matrix $\bm{D}$ is defined as follows: for the $k^{\text{th}}$ branch $(i_k, j_k) \in \mathcal{K}$ and $i_k > j_k$, $\bm{D}_{k, i_k} = 1$ and $\bm{D}_{k, j_k} = -1$. The other elements in the $k^{\text{th}}$ row of $\bm{D}$ are zero. In practice, the noisy power flow and voltage angle measurements are given by
\begin{eqnarray}
\tilde{\bm{z}}(t) &=& \bm{z}(t) + \bm{w}_{z}(t), \label{eq:noisy_powerflow} \\
\tilde{\bm{\theta}}(t) &=& \bm{\theta}(t) + \bm{w}_{\theta}(t), \label{eq:noisy_voltageangle}
\end{eqnarray}
where the noise processes are given by
\begin{eqnarray}
\bm{w}_{z}(t) &\sim& \mathcal{N}(\mathbf{0}, \sigma_{z}^2\mathbf{I}), \\
\bm{w}_{\theta}(t) &\sim& \mathcal{N}(\mathbf{0}, \sigma_{\theta}^2\mathbf{I}).
\end{eqnarray}
In matrix notation, for $T$ time instants, we have
\begin{eqnarray}
\bm{Z} &=& \text{diag}(\bm{s})\bm{D}\bm{\Theta},\\
\tilde{\bm{Z}} &=& \bm{Z} + \bm{W}_z,\\
\tilde{\bm{\Theta}} &=& \bm{\Theta} + \bm{W}_{\theta}, \label{eq:matrix_notation}
\end{eqnarray}
where $\bm{Z}$ and $\bm{W}_z$ are $K \times T$ matrices used to collect $T$ samples of $\bm{z}(t)$ and $\bm{w}_z(t)$, respectively, while $\bm{\Theta}$ and $\bm{W}_{\theta}$ are of dimension $B \times T$ used to collect $T$ samples of $\bm{\theta}(t)$ and $\bm{w}_{\theta}(t)$, respectively.
The problem is to detect changes in the susceptance vector $\bm{s}$ based on the noisy observations $\tilde{\bm{Z}}$ and $\tilde{\bm{\Theta}}$. Towards this end, we assume knowledge of a vector $\bm{s}_0$ corresponding to the nominal behavior of the grid and test the following hypotheses:
\begin{eqnarray}
\begin{cases}
H_0: \bm{s} = \bm{s}_0\\
H_1: \bm{s} \neq \bm{s}_0.
\end{cases}\label{eq:hypothesis_test}
\end{eqnarray}
Under both hypotheses, $\bm{Z}$ and $\bm{\Theta}$ are unknown and have to be estimated.
\vspace{-0.15in}
\subsection{TLS-GLRT solution}\label{subsec:tlsglrt_test}
The TLS-GLRT solution to the aforementioned hypothesis testing problem assumes $\bm{\Theta}$, and therefore $\bm{Z}$, are deterministic unknown vectors. The maximum likelihood (ML) estimation of $\bm{s}$, $\bm{\Theta}$ and $\bm{Z}$ is therefore known as TLS \cite{Huffel2002}. The TLS-GLRT is given by
\begin{eqnarray}
\text{t}_{\text{TLS}} = \log \frac{\max_{\bm{s},\bm{\Theta}}\prod_{t=1}^{T}p\left(\tilde{\bm{z}}(t), \tilde{\bm{\theta}}(t); \bm{s}, \bm{\theta}(t)\right)}{\max_{\bm{\Theta}}\prod_{t=1}^{T}p\left(\tilde{\bm{z}}(t), \tilde{\bm{\theta}}(t); \bm{s}_0, \bm{\theta}(t)\right)} \stackrel[H_0]{H_1}{\gtrless} \rho, \label{eq:tls_glrt}
\end{eqnarray}
where $\rho$ is a fixed threshold. We choose $H_0$ if the statistic is smaller than $\rho$, and H1 otherwise. The joint distribution of the observations is
\begin{eqnarray}
p\left(\tilde{\bm{z}}(t), \tilde{\bm{\theta}}(t); \bm{s}, \bm{\theta}(t)\right)\!\! &=&\!\! p(\tilde{\bm{z}}(t); \bm{s}, \bm{\theta}(t))p(\tilde{\bm{\theta}}(t);\bm{\theta}(t)),\\
p(\tilde{\bm{z}}(t); \bm{s}, \bm{\theta}(t)) &\sim& \mathcal{N}\left(\text{diag}(\bm{s})\bm{D}\bm{\theta}(t), \sigma_z^2\bm{I}\right),\\
p(\tilde{\bm{\theta}}(t);\bm{\theta}(t)) &\sim& \mathcal{N}\left(\bm{\theta}(t), \sigma_{\theta}^2\bm{I}\right).
\end{eqnarray}
In simplified form, the TLS-GLRT is given by
\begin{eqnarray}
\nonumber\text{t}_{\text{TLS}} &=& \frac{1}{2}\text{Tr}\left\{\bm{A}^{\mathrm{T}}(\bm{s}_0)\bm{H}^{-1}(\bm{s}_0)\bm{A}(\bm{s}_0) \right\} \\ &&
- \frac{1}{2}\min_{\bm{s}}\text{Tr}\left\{\bm{A}^{\mathrm{T}}(\bm{s})\bm{H}^{-1}(\bm{s})\bm{A}(\bm{s}) \right\} \stackrel[H_0]{H_1}{\gtrless} \rho, \label{eq:tls_glrt_simplified}\\
\bm{A}(\bm{s}) &=& \tilde{\bm{Z}} - \text{diag}\{\bm{s}\}\bm{D}\tilde{\bm{\Theta}},\\
\bm{H}(\bm{s}) &=& \sigma_{z}^2\mathbf{I} + \sigma_{\theta}^2\text{diag}\{\bm{s}\}\bm{D}\bm{D}^{\mathrm{T}}\text{diag}\{\bm{s}\}
\end{eqnarray}
The threshold $\rho$ is chosen as follows: given enough samples, the asymptotic performance of TLS-GLRT under $H_0$ is:
\begin{eqnarray}
2 \text{t}_{\text{TLS}} \sim \mathcal{X}^2_K, \label{eq:chisquared}
\end{eqnarray}
where $\mathcal{X}^2_K$ is a Chi-squared random variable with $K$ degrees of freedom, and $K$ is the dimension of $\bm{s}$ \cite{Kay2001}. This result is independent of the specific value of the unknown $\bm{\Theta}$. A reasonable approach to choosing the threshold for a given
false alarm rate $\alpha$ is
\begin{eqnarray}
\rho_{\alpha} = \frac{1}{2}F^{-1}_{\mathcal{X}^2_K}(\alpha), \label{eq:false_alarm}
\end{eqnarray}
where $F^{-1}_{\mathcal{X}^2_K}(.)$ is the inverse cumulative distributive function of the Chi-squared distribution with $K$ degrees of freedom.
Some comments on \eqref{eq:false_alarm} are in order. The physical meaning of \eqref{eq:false_alarm} is that if $T$ is large and the assumptions in \eqref{eq:dc_powerflow} - \eqref{eq:matrix_notation} hold, then the test statistic $2\text{t}_{\text{TLS}}$ in \eqref{eq:chisquared} can be shown to be a Chi-squared random variable with $K$ degrees of freedom. Even for moderately large $T$, the approximation is often quite accurate. A Chi-squared random variable has an inverse cumulative distributive function $F^{-1}_{\mathcal{X}^2_K}(.)$ that is a common and extensively tabulated function available in many software packages and whose values are tabulated in books \cite[Chapter 2.2]{Kay2001}. Numerous efficient algorithms are available to compute this function, however, the computational complexity is not really an issue. In practice, one can chose a set of desirable false alarm probabilities, $\alpha$ in \eqref{eq:false_alarm}, and the corresponding thresholds, $\rho_{\alpha}$ in \eqref{eq:false_alarm}, can be computed off-line and stored in a look-up table. Then the system can choose any of these false alarm probabilities and this will determine the threshold to employ. Fixing the false alarm probability is accepted practice in hypothesis testing \cite{Kay2001}. Given the fixed false alarm probability, the test in \eqref{eq:tls_glrt_simplified} is chosen to optimize the probability of detection ($P_d$) as described in \cite{Wei2012}. Note that in detection theory literature, the probability of false alarm is the probability of incorrectly choosing $H_1$ when $H_0$ is actually true, while the probability of deciding on hypothesis $H_1$ when $H_1$ is indeed true is referred to as the probability of detection \cite[Chapter 3]{Kay2001}.
\vspace{-0.1in}
\section{Simulation results and discussion}\label{sec:results
In this section, we present simulation results to demonstrate the performance improvement of the TLS-GLRT fault detection scheme when the transmission of PMUs follow the scheduling policy described in \secref{sec:schedulingPMUs} compared to its performance when the PMUs transmit in a round-robin fashion without a predefined order. Results also enable us to compare the performance of the detection scheme offered by scheduling the PMU transmissions, when the PMUs are placed employing the topology- and electrical structure-based approaches. We first describe the experimental setup, followed by simulation results and related discussion.
\vspace{-0.1in}
\subsection{Experimental setup}\label{subsec:setup}
We consider the IEEE 14-bus system with the number of buses $B$ = 14 and the number of branches $K$ = 20. The noise variances are given by $\sigma_{z}^2$ = $\sigma_{\theta}^2$ = 0.01. We perform 10$^4$ Monte Carlo simulations. At each realization, the elements of $\bm{\Theta}$ are generated independently as standard normal random variables. The active power flow $\bm{Y}$ between the buses were obtained using the power flow algorithm in MATPOWER \cite{Zimmerman2011}; the susceptance vector $\bm{s}$ was then computed using $\bm{Y}$. Under $H_0$, we use the susceptance parameters given by the test profile. Under $H_1$, we apply a change of -2\% to every element of $\bm{s}$. Here, -2\% essentially means that every element of $\bm{s}$ is made smaller by a factor of 2\%; this number can be arbitrarily chosen without loss of generality.
Simulations are conducted for one frame lasting 20 time units. This corresponds to $T = 20$ in \eqref{eq:dc_powerflow}. Furthermore, each frame is subdivided into $N$ slots, where $N$ is the solution to the PMU placement problem. The setup comprises solving the hypothesis testing problem \eqref{eq:hypothesis_test} after each PMU transmission. Note that, we need phasor data from all the PMUs to form the test statistic. In light of the previous statement, we do not induce a change in the susceptance vector at the first time instant; however, for subsequent time instants, every element of the susceptance vector $\bm{s}$ will be changed by -2\%. After each transmission, we use the phasor data that we previously had from each PMU to form the test statistic.
To analyze the performance of the fault detection scheme presented in \secref{sec:fault_detection}, we compute the variation of probability of detection $(P_d)$ of change in the susceptance parameter versus time, using a fixed value of false alarm rate ($\alpha$) uniformly picked between [0, 0.2] for which the corresponding threshold is calculated using \eqref{eq:false_alarm}. We then pick a new value of $\alpha \in$ [0, 0.2] and repeat the calculations to obtain the corresponding variation of $P_d$ with time. This procedure is repeated for different values of $\alpha \in$ [0, 0.2], resulting in a set of $P_d$s versus time. Lastly, we plot the average (over the number of $\alpha$s) $P_d$ versus time for two scenarios: (a) when the PMUs transmit in a random manner without a prescribed scheduling policy and (b) when the PMUs transmissions are allowed to follow the scheduling scheme. For both scenarios, we consider the separate cases of the PMUs being placed employing the topology- and electrical structure-based approaches.
\vspace{-0.125in}
\subsection{Simulation results}\label{subsec:simulation_results}
We first present the results for the PMU placement problem for the 14-bus system. For the topology-based approach to the PMU placement problem, the connectivity matrix is given by \eqref{eq:topological_A}. Solving \eqref{eq:pmuplacement}, we obtain 4 as the optimum number of PMUs to be installed on buses 2, 6, 7 and 9. For the electrical structure-based approach to the PMU placement problem, the connectivity matrix is given by \eqref{eq:adjcency_matrix} (see \appref{app:electrical_structure}). Solving \eqref{eq:pmuplacement}, we obtain 7 as the optimum number of PMUs to be installed on buses 1, 3, 8, 11, 12, 13 and 14. For both topology- and electrical structure-based approaches, the objective is to achieve complete network observability.
Following the scheduling policy described in \secref{sec:schedulingPMUs}, for the topology-based approach (where an optimal number of 4 PMUs are placed on buses numbered 2, 6, 7 and 9), the transmission schedule is as follows: the PMU placed on bus 7 transmits first, followed by those placed on buses 2, 6 and lastly 9, in that order.
Similarly, for the electrical structure-based approach (where an optimal number of 7 PMUs are placed on buses numbered 1, 3, 8, 11, 12, 13 and 14), the transmission scheduled is the same as presented in \secref{subsec:illustration}, {\ie}, the PMUs installed on buses 14, 8, 12, 11, 3, 13 and 1 transmit in time slots numbered 1, 2, 3, 4, 5, 6 and 7, respectively.
\begin{figure}[t]
\includegraphics[height=2.5in,width=3.5in]{Pd_time_elec_14bus.eps}
\caption{Probability of detection versus time, for one frame lasting 20 time units, when the PMUs are placed employing the electrical structure-based approach.}
\label{fig:Pd_time_elec_14bus}
\end{figure}
\begin{figure}[t]
\includegraphics[height=2.5in,width=3.5in]{Pd_time_topo_14bus.eps}
\caption{Probability of detection versus time, for one frame lasting 20 time units, when the PMUs are placed employing the topology-based approach.}
\label{fig:Pd_time_topo_14bus}
\end{figure}
The plots of $P_d$ versus time for one frame lasting 20 time units when the PMUs are placed employing the electrical structure- and topology-based approaches are shown in \figref{fig:Pd_time_elec_14bus} and \figref{fig:Pd_time_topo_14bus}, respectively. In each figure, we also plot $P_d$ versus time when the PMU-transmissions are random, without a predefined order/schedule. As seen from the plots, scheduling the PMU-transmissions results in better detection performance compared to PMUs transmitting at random. Note that, at the end of the time frame, after the vector of phasor data from all the PMUs are updated, the $P_d$ for both scheduled and random transmissions become the same since we have phasor data from all the PMUs in both cases.
For the topology-based approach to PMU placement, we have installed 4 PMUs on the grid, while for the electrical structure-based approach, there are 7 PMUs. To compare the scheduling performance between the topology- and electrical structure-based PMU placements, we modify the experimental setup as follows. We let the same number of PMUs transmit for both the topology-based and electrical structure-based approaches to analyze the detection performance. Specifically, we allow only 4 PMUs transmit their phasor data to the control center for both PMU placement approaches before applying the scheduling scheme. More precisely, for the topology-based approach we allow PMUs on buses numbered 2, 6, 7 and 9 to transmit as before. For the electrical structure-based approach, we now allow only 4 PMUs on buses numbered 8, 11, 12 and 14 transmit their phasor data. This choice was based on the magnitudes of the entries in the vectors $\sigma_n\bm{u}_{n}$, $n=1,\dots,4$, {\ie}, the first four column vectors in the 7 $\times$ 7 sub-matrix described in \secref{subsec:illustration}.
\begin{figure}[t]
\includegraphics[height=2.5in,width=3.5in]{Pd_time_elec_topo.eps}
\caption{Probability of detection versus time, for one frame lasting 20 time units, to compare the topology- and electrical structure-based approaches to PMU placement.}
\label{fig:Pd_time_elec_topo}
\end{figure}
The scheduling policy is now applied to the above setup. Similar to previous experiments, for the topology-based approach the transmission schedule is as follows: the PMU placed on bus 7 transmits first, followed by those placed on buses 2, 6 and lastly 9, in that order. For the electrical structure-based approach, the PMUs installed on buses 14, 8, 12 and 11 transmit, in that order. For both approaches, each frame (lasting 20 time units) is divided into 4 slots. The plot of $P_d$ versus time is shown in \figref{fig:Pd_time_elec_topo}, where we see that the electrical structure-based approach has a better detection performance compared to the topology-based approach.
Finally, we conduct an experiment in which we allow lesser number of PMUs for the electrical structure-based approach to transmit compared to the topology-based approach. Specifically, we allow only 3 PMUS (on buses 14, 8 and 12, based on the the magnitudes of the entries in the vectors $\sigma_n\bm{u}_{n}$, $n=1,\dots,3$) to transmit for the electrical structure-based approach, while continuing with 4 PMUs for the topology-based approach. The scheduling scheme is applied to both these cases and the resulting plots are shown in \figref{fig:Pd_time_elecless_topo}. As shown in the \figref{fig:Pd_time_elecless_topo}, the probability of detection for the electrical structure-based approach is in fact slightly better than the topology-based approach. It is important to note that, the above two modifications in the experimental setup are especially useful when there are stringent constraints ({\eg}, bandwidth limitation) for communications on the grid.
\begin{note}\label{note:fig5_6}
For \figref{fig:Pd_time_elec_topo} and \figref{fig:Pd_time_elecless_topo}, it should be noted that we have installed the optimum number of 7 PMUs for the electrical structure-based approach, so as to achieve complete network observability. However, we allowed only 4 PMUs to transmit their phasor data to the control center. Though the remaining 3 PMUs record phasor data across the lines on which they are installed, they do not transmit them to the control unit. So long as phasor data from the ``high priority'' buses ({\ie}, buses having high electrical influence on the remaining buses in the network) are obtained, the test statistic can be computed and the hypotheses test can be tested.
\end{note}
In summary, since the scheduling is based on the electrical connectedness between buses in the grid, PMUs placed on buses having a strong electrical influence with other buses are given a higher priority. Furthermore, as discussed in \secref{sec:introduction}, the electrical structure-based approach provides a stronger characterization of the electrical influence between network components compared to the topology-based approach. Therefore, the probability of detection of change in the susceptance parameter is higher when the PMUs are installed based on the electrical structure-based approach compared to the case when they are installed using the topology-based approach. However, both approaches outperform the case when phasor data from PMUs are transmitted without prior scheduling.
\begin{figure}[t]
\includegraphics[height=2.5in,width=3.5in]{Pd_time_elecless_topo.eps}
\caption{Probability of detection versus time, for one frame lasting 20 time units, with lesser number of PMUs for the electrical structure-based approach compared to the topology-based approach.}
\label{fig:Pd_time_elecless_topo}
\end{figure}
The scheduling scheme proposed in this paper assumes dedicated direct communications link between the individual PMUs and the WAMS control center. However, in practice the communications infrastructure for the power grid encompasses many complicated and interrelated operations \cite{Shahidehpour2003}. For instance, the integrated control center system (ICCS) provides user interfaces to view the power grid information and distributed computing environment to monitor and coordinate the security of transmission system. Then there is the paradigm of common information model (CIM) using which information sharing of power system applications can be achieved using a common markup language. Such communications overheads could be termed as ``constraints'', since they impede the rate of transfer of synchrophasor data from the PMUs to the WAMS server. However, as mentioned in \secref{sec:introduction}, since our scheduling policy is transmitter-centric, it is unaffected by the communications constraints. It should also be noted that integration of PMU schedulers into existing standards ({\eg}, C37.118.2-2011 \cite{2011}) could have a bearing on communication protocols, data types and formats for phasor data transmission on the grid. These aspects are relegated to future work.
\vspace{-0.1in}
\section{Conclusion}\label{sec:conclusion}
We proposed a time division multiplexing scheme for transmission of phasor data from the PMUs to a central server, where time was divided into frames and the optimal set of PMUs within a given time frame take turns to transmit to the control center. The proposed scheduling policy was governed by the measure of electrical connectedness between buses in the power grid. We presented the PMU placement problem from two different perspectives, namely, topology- and electrical structure-based approaches. For both these approaches, the scheduling scheme was coupled with a fault detection algorithm, which was posed as hypothesis testing problem. The performance of the fault detection, which was formulated to detect changes in the susceptance parameters of the network, was shown to improve due to scheduling transmissions from PMUs compared to transmitting phasor data in a random manner. Future work would involve scheduling algorithms with incomplete network observability and WAMS servers with multiuser reception/detection capability.
\vspace{-0.1in}
\appendices
\section{On the electrical structure of the grid}\label{app:electrical_structure}
The electric power grid has received considerable attention from the perspective of complex networks \cite{Dorfler2010}. In the following we briefly present this perspective, which promotes the electrical structure of the grid over its topological structure. Then, the binary connectivity matrix is derived using the resistance distance between buses in the network.
The study of the electrical structure of the grid was motivated by the following drawbacks suffered by its topological structure:
\begin{enumerate}
\item In \cite{Cotilla-Sanchez2012} (see Section I and references therein), it was reported that electric grids
in different geographical locations had different degree distributions leading to varied topological structures.
\item It was also pointed out that the same grid had different topological structures by carrying out different model-based analyses. This discrepancy was attributed to the weaker characterization of the electrical connections between network components as provided by the topological structure.
\item Related reports supporting this line of argument were found in \cite{Wu1995} - \nocite{Wu2005}\cite{Atay2006}, where it was shown that, for many classes of complex networks, characterizing the network structure using degree distribution alone was suboptimal and had implications on node synchronization and performance of the network.
\end{enumerate}
In the context of PMU placement, for the topology-based approach, the bus admittance matrix plays a central role in solving the placement problem. Though the admittance matrix characterizes the electrical behavior of the network, the sensitivity between power injections and nodal phase angle differences can be utilized to better characterize the electrical influence between network components. Towards this end, we derive the resistance distance matrix, which provides a strong characterization of the electrical influence between various network components.
Consider a network with $B$ buses, described by the conductance matrix $\bm{G}$. Let $V_j$ and $g_{ij}$ denote the voltage magnitude at bus $j$ and the conductance between buses $i$ and $j$, respectively. The current injection at bus $i$ is then given by
\begin{eqnarray}\label{eq:current_injection}
I_i = \sum_{j=1}^{B}g_{ij}V_j.
\end{eqnarray}
$\bm{G}$ acts as a Laplacian matrix to the network, provided there are no connections to the ground, {\ie}, if $\bm{G}$ has rank $B-1$. The singularity of $\bm{G}$ can be overcome by letting a bus $r$ have $V_r = 0$. The conductance matrix associated with the remaining $B-1$ buses is full-rank, and thus we have
\begin{eqnarray}\label{eq:nonreferencenodes}
\bm{V}_k = \bm{G}^{-1}_{kk}\bm{I}_k, k \neq r.
\end{eqnarray}
Let the diagonal elements of $\bm{G}^{-1}_{kk}$ be denoted $g^{-1}_{kk}$, $\forall k$, indicating the change in voltage due to current injection at bus $k$ which is grounded at bus $r$. The voltage difference between a pair of buses $(i,j)$, $i\neq j\neq r$, is computed as follows:
\begin{eqnarray}\label{eq:voltage_difference}
e(i,j) = g^{-1}_{ii} + g^{-1}_{jj} - g^{-1}_{ij} - g^{-1}_{ji},
\end{eqnarray}
indicating the change in voltage due to injection of $1$ Ampere of current at bus $i$ which is withdrawn at bus $j$. $e(i,j)$ is called the resistance distance between buses $i$ and $j$, and describes the sensitivity between current injections and voltage differences. In matrix form, letting $\boldsymbol{\Gamma} \triangleq \text{diag}(\bm{G}^{-1}_{kk})$, we have $\forall k \neq r$
\begin{eqnarray}
\bm{E}_{kk} &=& \boldsymbol{1}\boldsymbol{\Gamma}^{\mathrm{T}} + \boldsymbol{\Gamma}\boldsymbol{1}^{\mathrm{T}} - \bm{G}^{-1}_{kk} - \left[\bm{G}^{-1}_{kk}\right]^{\mathrm{T}},\label{eq:resistance_distance1}\\
\bm{E}_{rk} &=& \boldsymbol{\Gamma}^{\mathrm{T}},\label{eq:resistance_distance2}\\
\bm{E}_{kr} &=& \boldsymbol{\Gamma}.\label{eq:resistance_distance3}
\end{eqnarray}
The resistance distance matrix $\bm{E}$, thus defined, possesses the properties of a metric space \cite{Klein1993}.
To derive the sensitivities between power injections and phase angles, we start with the upper triangular part of the Jacobian matrix obtained from the power flow analysis, for the distance matrix to be real-valued:
\begin{eqnarray}\label{eq:upper_jacobian}
\Delta \bm{P} = \left[\frac{\partial P}{\partial \theta}\right]\Delta \theta + \left[\frac{\partial P}{\partial |V|}\right]\Delta |V|.
\end{eqnarray}
The matrix $\left[\frac{\partial P}{\partial \theta}\right]$ will be used to form the distance matrix, by assuming the voltages at the buses to be held constant, {\ie}, $\Delta|V|=0$. It was observed that $\left[\frac{\partial P}{\partial \theta}\right]$ possesses most of the properties of a Laplacian matrix. By letting $\bm{G} = \left[\frac{\partial P}{\partial \theta}\right]$, the resulting distance matrix $\bm{E}$ measures the incremental change in phase angle difference between two buses $i$ and $j$, $(\theta_i - \theta_j)$, given an incremental average power transaction between those buses, assuming the voltage magnitudes are held constant. It was proved in \cite[Appendix]{Cotilla-Sanchez2012} that $\bm{E}$, thus defined, satisfies the properties of a distance matrix, as long as all series branch reactance are nonnegative.
For a power grid with $B$ buses and $K$ branches, the distance matrix $\bm{E}$ translates into an undirected graph with $B(B-1)$ weighted branches. In order to compare the grid with an undirected network without weights, one has to retain the $B$ buses, but replace the $K$ branches with $K$ smallest entries in the upper or lower triangular part of $\bm{E}$. This results in a graph of size $\{B, K\}$ with edges representing electrical connectivity rather than direct physical connections. The adjacency matrix $\bm{C}$ of this graph is obtained by setting a threshold, $\lambda$, adjusted to produce exactly $K$ branches in the network:
\begin{eqnarray}\label{eq:adjcency_matrix}
\bm{C}:
\begin{cases}
\tilde{c}_{ij} = 1, ~\forall e(i,j) < \lambda,\\
\tilde{c}_{ij} = 0, ~\forall e(i,j) \geq \lambda.
\end{cases}
\end{eqnarray}
To obtain the threshold $\lambda$ present in \eqref{eq:adjcency_matrix}, we first consider the upper triangular part of the matrix $\bm{E}$, and sort the elements in descending (or ascending) order. We then pick a number of elements equal to the number of branches in the given power network. For instance, given the IEEE-14 bus network having 20 branches, we pick the top 20 sorted elements from the upper triangular part of the matrix $\bm{E}$.
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Cosmology has indeed undergone a dramatic change over the past couple of decades. The availability of high precision data regarding our universe and its indication towards an expanding universe with an accelerated rate invoked all sorts of modifications of Einstein's equations. That the acceleration must have set in a not too distant past is a theoretical requirement as well as has been observationally supported. There are excellent reviews regarding the accelerated expansion\citep{sahni2000,paddy2003,copeland2006}.
\par The theoretical investigation towards finding a viable option which can drive this acceleration is done in two distinct ways. One way is to modify the matter sector by adding an exotic field giving rise to an effective negative pressure. The most talked about agent capable of doing this is certainly the cosmological constant $\Lambda$. A scalar field with some potential, known as the quintessence matter, is also amongst the most favourite candidates as a "dark energy", the agent driving this alleged expansion. We refer to the reviews \citep{sahni2000,paddy2003,copeland2006} and the references therein.
\par The second option is to look for a theory of gravity where the Einstein-Hilbert action is modified. One way to do that is to consider a scalar field nonminimally coupled to the geometry sector \citep{nbdp2001a,nbdp2001b,ss2001,bertolami2000,elizalde2004,onemliwood2002,onemliwood2004,brunier2005,dasbaneree2006} or to the matter sector \citep{khoury2004a,khoury2004b,mota2004a,mota2004b,dass2006,nb2010} or both \citep{dasbaneree2008}. The other popular way is to use an analytic function $f=f(R)$ in place of $R$ in the action, where $R$ is the Ricci scalar \citep{capo2003,brofra2004,nojiri2003,dolgov2003,carr2004,dsnbnd2006}. It had already been noted that $f(R) \propto R^2$ kind of theories could successfully generate inflationary universe scenario for the early universe \citep{starobaa1980,kerner1982,duruisseau1986}. As the curvature $R$ decreases with time, inverse power of $R$ in an $f(R)$ theory might be expected to generate a late time acceleration. For a detailed description of $f(R)$ theories and their application in cosmological models, we refer to some recent reviews \citep{sotiriou2010,felice2010,nojiriod2011}. The $f(R)$ gravity models, available in the literature mostly deal with only the present acceleration and hardly talk about the smooth transition from a decelerated to an accelerated regime. There are anyway a few investigations regarding this signature flip in the deceleration parameter $q$. For instance, we refer to \cite{nojiodin2006} and \cite{nojiodinstef2006}. Nojiri and Odintsov \citep{nojiri2007a} also reconstructed an $f(R)$ gravity model from a $\Lambda$CDM one.
\par Das, Banerjee and Dadhich \citep{dsnbnd2006} indeed discussed models that show such a smooth transition analytically, but the models do not contain matter. However, it has been shown that, along with a matter field, any such model could behave in quite a different manner\citep{amendola1,amendola2}. Some models where the Ricci scalar is non-minimally coupled to the matter sector are also there in literature \citep{thakur2011}. Some of the modifications of the geometry sector also involves a function of $G$, the Gauss-Bonnet scalar. Nojiri and Odintsov discussed $f(G)$ and $f(R,G)$ models in connection with the recent accelerated expansion of the universe \citep{nojiri2007d}. A reconstruction of $f(R)$ gravity model can be found in reference\citep{nojiri2007e}. As the $\Lambda$CDM model does well regarding the fits with the observational data, there are attempts to distinguish between models that mimic the $\Lambda$CDM model and those which do not. For example, the models given by Hu and Sawicki \citep{hu2007} and Starobinsky \citep{starob2007} are distinct from the $\Lambda$CDM, whereas those given by He and Wang \citep{he2013} and Dunsby et al \citep{dunsby2010} are consistent with that. Constraining the model parameters of an $f(R)$ gravity model has been discussed recently by Nojiri and Odintsov \citep{nojiri2007b, nojiri2007c}, Giron$\acute{e}s$ {\it et al} \citep{giro2010} and Basilakos {\it et al} \citep{basila2013}.
\par One general problem with $f(R)$ theories is that they either give an early inflation or a late time acceleration. There have been recent attempts to find some form of $f(R)$ which would yield accelerated expansion in two phases, one in an early epoch and the other in the late stage of evolution. Cognola {\it et al} \citep{cognola2009} and Elizalde {\it et at} \citep{elizalde2011} made such attempts with an $f(R)$ which is an exponential function of $R$. Nojiri and Odintsov \citep{odin2003} made an attempt to unify the two phases of accelerated expansion in the realm of a single $f(R)$ gravity model by combining positive and negative powers of $R$. Possible impacts of the existence of nonlinear terms involving $R$ in the action on the structure formation has been discussed by Thakur and Sen\citep{anjan2013}.
\par In the present work, a straightforward way to facilitate the investigation of the dynamics of the universe is discussed. The net conservation equation results from the contracted Bianchi identity. We assume that the matter content obeys its own conservation, which, in tandem with the net conservation equation yields an equation for the contribution from the geometry sector to the evolution of the universe. This equation is a second order differential equation in the Hubble parameter $H$. The equation is highly nonlinear and it is difficult to get an analytic solution. However, with proper boundary conditions, one can plot the relevant cosmological parameters like the deceleration parameter $q$, the effective equation of state parameter $w_{eff}$ such that the qualitative behaviour of the model is understood. We deal with two simple examples to elucidate the method, a two-parameter model ($f(R) \propto (\lambda +R)^{n}$) and a one-parameter model ($f(R) \propto exp(\alpha R)$).
\section{$f(R)$ gravity and the conservation equation}
The generalized Einstein-Hilbert action for $f(R)$ gravity is
\begin{equation}
{\cal A}=\int \Bigg[\frac{1}{16\pi G}f(R)+{\cal L}_m\Bigg]\sqrt{-g}d^4x,
\label{action}
\end{equation}
where $R$ is replaced by $f(R)$ in the Einstein-Hilbert action, $f(R)$ being an analytic function of $R$. Here ${\cal L}_m$ is the usual matter field Lagrangian. A variation of this action, with respect to the metric, yields the field equations as
\begin{equation}
f'(R)R_{\mu\nu}-\nabla_\mu\nabla_\nu f'(R)+\big[\Box f'(R)-\frac{1}{2}f(R)\big]g_{\mu\nu}=T^{(m)}_{\mu\nu},
\label{fRfieldeq}
\end{equation}
where a prime indicates differentiation with respect to the Ricci scalar $R$ and $T^{(m)}_{\mu\nu}$ represents the contribution to the energy momentum tensor from matter fields with a choice of unit as $8\pi G=1$. This variation is popularly dubbed as the metric $f(R)$ gravity as opposed to the Palatini formulation where the variation is carried out with respect to both the metric and the affine connections.
\par The present endeavour is to study the dynamics of the universe in the background of the spatially flat FRW metric, which is written as
\begin{equation}
ds^2=dt^2-a^2(t)[dr^2+r^2d\theta^2+r^2sin^2\theta d\phi^2],
\label{metric}
\end{equation}
where $a(t)$ is the scale factor. The field equations take the form
\begin{equation}
3\frac{\dot{a}^2}{a^2}=\frac{\rho_m}{f'}+\frac{1}{f'}\Big[\frac{1}{2}(f-Rf')-3\dot{R}f''\frac{\dot{a}}{a}\Big],
\label{frd1}
\end{equation}
\begin{equation}
2\frac{\ddot{a}}{a}+\frac{\dot{a}^2}{a^2}=-\frac{1}{f'}\Big[\ddot{R}f''+\dot{R}^2f'''+2\dot{R}f''\frac{\dot{a}}{a}-\frac{1}{2}(f-Rf')\Big],
\label{frd2}
\end{equation}
where dots are the derivatives with respect to cosmic time and a prime denotes derivative with respect to $R$. Also, $\rho_m$ is the dark matter density and, consistent with the cold dark matter, the corresponding pressure $p_m$ is taken to be zero.
\par We write
\begin{equation}
\Big[\frac{1}{2}(f-Rf')-3\dot{R}f''\frac{\dot{a}}{a}\Big]=\rho_c,
\label{rhoc}
\end{equation}
and
\begin{equation}
\Big[\ddot{R}f''+\dot{R}^2f'''+2\dot{R}f''\frac{\dot{a}}{a}-\frac{1}{2}(f-Rf')\Big]=p_c,
\label{pc}
\end{equation}
as $\rho_c$ and $p_c$ determine the contribution by the curvature to the density and pressures sectors respectively. For $f(R) = R$, both $\rho_c$ and $p_c$ would vanish as expected. In terms of the Hubble parameter ($H=\frac{\dot{a}}{a}$), the modified field equations (\ref{frd1}) and (\ref{frd2}) will read as
\begin{equation}
3H^2=\frac{\rho_m+\rho_c}{f'},
\label{frdhubble1}
\end{equation}
and
\begin{equation}
2\dot{H}+3H^2=-\frac{p_c}{f'},
\label{frdhubble2}
\end{equation}
respectively.
Finally, the contracted Biacchi identity will yield
\begin{equation}
\frac{d}{dt}\Big(\frac{\rho_m+\rho_c}{f'}\Big)+3H\Big(\frac{\rho_m+\rho_c+p_c}{f'}\Big)=0.
\label{BianchiID}
\end{equation}
Considering the matter conservation, i.e., $\dot{\rho_m}+3H\rho_m=0$, is satisfied independently, equation (\ref{BianchiID}) will be simplified to the form
\begin{equation}
\frac{d}{dt}\Big(\frac{\rho_c}{f'}\Big)+3H\Big(\frac{\rho_c+p_c}{f'}\Big)=\rho_m\Big(\frac{\dot{R}f''}{f'}\Big).
\label{BianchiID2}
\end{equation}
This equation can be written in the form
\begin{equation}
18\frac{f''}{f'}H(\ddot{H}+4H\dot{H})+3(\dot{H}+H^2)+\frac{f}{2f'}+\frac{\rho_m}{f'}=0,
\label{conservationequ}
\end{equation}
where use has been made of the expressions $R=-6(\dot{H}+2H^2)$ and $\dot{R}=-6(\ddot{H}+4H\dot{H})$. It is important to note that equation (\ref{BianchiID}) yields equation (\ref{conservationequ}) under the condition $f'' \neq 0$. So one cannot arrive at the corresponding equation for $f(R)=R$, the usual Einstein-Hilbert action. Now we write the equation (\ref{conservationequ}) with the redshift $z$ (given by $1+z=\frac{a_0}{a}$) as the argument. The equation now looks like
\begin{equation}
\begin{split}
\frac{d^2H}{dz^2}=\frac{3}{(1+z)}\frac{dH}{dz}-\frac{1}{H}\Bigg(\frac{dH}{dz}\Bigg)^2 \\
-\frac{3f'\Big(H^2-(1+z)H\frac{dH}{dz}\Big)+\frac{f}{2}+\rho_{m0}(1+z)^3}{18(1+z)^2H^3f''}.
\label{conservationequ2}
\end{split}
\end{equation}
This is the key equation in our attempt to study the dynamics of the universe in $f(R)$ gravity models. Though the equation is highly non-linear, we can at least investigate the the redshift dependence of the Hubble parameter $H(z)$ and other important parameters numerically when the form of $f(R)$ is given.
\par Now from equation (\ref{frdhubble1}), the present matter density $\rho_{m0}$ can be expressed as
\begin{equation}
\rho_{m0}=3H_0^2f'_o-\rho_{c0}.
\end{equation}
where a subscript 0 indicates the values of the functions at the present epoch, namely at $z=0$. Now $R=-6H^2(1-q)$ and $\dot{R}=-6H^3(j-q-2)$ where $q=-\frac{\ddot{a}}{aH^{2}}$ is the deceleration and $j=\frac{\ddot{a}}{a^3 H^3}$ is the jerk parameter. Hence present value of Ricci scalar and its derivative can be estimated from the knowledge of present deceleration parameter $q_0$ and jerk $j_0$. We scale $H$ as $\frac{H}{H_0}$ so that the present value of Hubble parameter $H_0$ is unity. A simple dimensional consideration shows that this can be done without any loss of generality in the equation (\ref{conservationequ2}) by dividing both sides by $H_0$. There are observational estimates for the parameters $q_0$ and $j_0$. In the present work, we pick up the relevant values from the work of Rapetti {\it et al}\citep{rapetti2007}. The relevant values are $q_0=-0.81\pm0.14$ and $j_0=2.16_{-0.75}^{+0.81}$.
\section{$f(R)$ gravity in a spatially flat FRW universe}
With a functional form of $f(R)$, equation (\ref{conservationequ2}), a second order differential equation in $H(z)$, can be numerically integrated. In this work, two $f(R)$ gravity models have been discussed. The aim is to find the parameters of the models that would be in agreement with the observed values of the relevant cosmolgical parameters, namely the deceleration parameter $q$ and the effective equation of state parameter $w_{eff}$ given by $w_{eff}=\frac{p_c}{\rho_c + \rho_m}$.
\subsection{CaseI: $f(R)= {\lambda_0}(\lambda+R)^n$}
We choose $f(R)={\lambda_0}(\lambda+R)^n$, where $\lambda_0$, $\lambda$ and $n$ are constants and they actually are the model parameters.As $f(R)$ should have the dimension of $R$, the constant $\lambda_0$ is there to take care of the dimension. In all subsequent discussion, the value of the constant $\lambda_0$ is taken to be unity. From equation (13), the numerical plots of deceleration parameter $q(z)$ and the effective equation of state parameter $w_{eff}(z)$ are obtained using the present values $q_0=-0.81\pm0.14$ and $j_0=2.16_{-0.75}^{+0.81}$ as mentioned in the previous section. There are two parameters in the model, namely $\lambda$ and $n$. The plots have been generated taking four sets of values of these two parameters. Each set has been adjusted in such a way that the recent acceleration starts around $z=0.5$. Figures 1-4 show these plots. The present value of the parameter $w_{eff}$ in all cases is between -0.8 to -1.0. This also in conformity with the observational estimate. It deserves mention that most of the examples of the power law type $f(R)$ gravity models leading to present acceleration involves some negative power for the Ricci scalar $R$ in the action. But in all the examples in this work, $n$ is positive and so there is no singularity in $f(R)$ for $R$ going to zero.
\begin{figure}[hbtp]
\centering
\includegraphics[width=\columnwidth]{qwmfr11.eps}
\caption{\small Plots of deceleration parameter $q$ (left panel) and effective equation of state $w_{eff}$ (right panel) against redshift $z$ for $f(R)=\lambda_0(\lambda+R)^n$ with $\lambda_0=1$, $n=0.5$ and $\lambda=13.5\pm0.5$. The central dark line is for $\lambda=13.5$ and $\lambda=13.0$ is the upper and $\lambda=14.0$ is the lower bounds of the plots. }
\end{figure}
\begin{figure}[hbtp]
\centering
\includegraphics[width=\columnwidth]{qwmfr12.eps}
\caption{\small Plots of deceleration parameter $q$ (left panel) and effective equation of state $w_{eff}$ (right panel) against redshift $z$ for $f(R)=\lambda_0(\lambda+R)^n$ with $\lambda_0=1$, $n=0.1$ and $\lambda=13.0\pm0.25$. The central dark line is for $\lambda=13.0$ and $\lambda=12.75$ is the upper and $\lambda=13.25$ is the lower bounds of the plots.}
\end{figure}
\begin{figure}[hbtp]
\centering
\includegraphics[width=\columnwidth]{qwmfr13.eps}
\caption{\small Plots of deceleration parameter $q$ (left panel) and effective equation of state $w_{eff}$ (right panel) against redshift $z$ for $f(R)=\lambda_0(\lambda+R)^n$ with $\lambda_0=1$, $n=1.5$ and $\lambda=16.0\pm3.0$. The central dark line is for $\lambda=16.0$ and $\lambda=19.0$ is the upper and $\lambda=13.0$ is the lower bounds of the plots.}
\end{figure}
\begin{figure}[hbtp]
\centering
\includegraphics[width=\columnwidth]{qwmfr14.eps}
\caption{\small Plots of deceleration parameter $q$ (left panel) and effective equation of state $w_{eff}$ (right panel) against redshift $z$ for $f(R)=\lambda_0(\lambda+R)^n$ with $\lambda_0=1$, $n=2.0$. The central dark line is for $\lambda=12.0$ and $\lambda=15.0$ is the upper and $\lambda=1.0$ is the lower bounds of the plots.}
\end{figure}
\subsection{Case II: $f(R)=R_0exp(\alpha R)$}
This exponential form of $f(R)$ had already been discussed in \citep{dsnbnd2006}. However, that was done with no matter content of the universe. In the present work, the same single parameter exponential form of $f(R)$ has been introduced along with the matter content. Like the previous example, the constant $R_0$ takes care of the dimensional requirement and is chosen to be unity in the subsequent discussion. The numerical plots for the deceleration parameter $q$ and the effective equation of state parameter $w_{eff}$ are obtained for a range of values of $\alpha$ (between 0.5 and 15.0) with the similar boundary conditions used for the previous model. The range of values of $\alpha$ are chosen so as to get the signature flip in $q$ close to $z=0.5$. Figure 5 clearly shows that this model also successfully generates late time acceleration accompanied by the decelerated expansion era that prevailed earlier. The central curve is for $\alpha = 1.5$ for both of $q$ and $w_{eff}$. If the valu
e of $\alpha$ is raised to 15, the lower curve is obtained. But almost similar amount of deviation is seen for the higher curve for a much smaller variation of the parameter. The upper curve is obtained when $\alpha$ is changed to 0.5. So the amount of acceleration is much more sensitive to a decrease of the parameter of the model.
\begin{figure}[hbtp]
\begin{center}
\includegraphics[width=\columnwidth]{qwmfr2.eps}
\end{center}
\caption{\small Plots of deceleration parameter $q$ (left panel) and effective equation of state $w_{eff}$ (right panel) against redshift $z$ for $f(R)=R_0exp(\alpha R)$. The central dark line is for $R_0=1$, $\alpha=1.5$ and $\alpha=0.5$ is the upper and $\alpha=15.0$ is the lower bounds of the plots.}
\end{figure}
\vskip 2.5 cm
\section{Discussion}
A straightforward way for the discussion of the dynamics of the much talked about $f(R)$ gravity models along with a cold dark matter content has been presented in this work.As it has been shown that an $f(R)$ gravity model could behave in a dramatically different manner in the presence of matter\citep{amendola1,amendola2}, it is imperative that the models are discussed in the presence of matter. Equation (\ref {conservationequ2}) sets up a basic framework for that. Both the models presented here work well in the presence of matter.
\par Two examples have been worked out, one of them, namely the case I is apparently new, and the second case has already been discussed, although without the requisite matter content. The parameters of the model are reconstructed from the observational values of some cosmological parameters. However, no rigorous statistical analysis has been employed for the estimation of the model parameters.
\par It deserves mention at this stage that the two models presented here do not have the same degree of stability. If a quantity, $m^2=\frac{1}{3}\Big[\frac{f'(R)}{f''(R)}-R\Big]$, is defined at $R=R_0$, the present value of the Ricci curvature, one can show that $m^2<0$ leads to a tachyonic instability \citep{nojiri2007b}. The second model of the present work (section 3.2) has this instability. Our first model $f(R)=\lambda_0(\lambda+R)^n$, on the other hand, passes this fitness test.
\par The primary motivation is to set up a general framework, but both the examples discussed can produce a signature flip at the right epoch and can reproduce the total effective equation of state parameter $w_{eff}$ close to its expected present value. This basic observational requirement is met for actually quite a wide range of the model parameters.
\par It also deserves mention that according to the criterion discussed by Basilakos\citep{basila2013}, none of the two models presented here would actually converge to the $\Lambda CDM$ model. The first example would do that only for the trivial case of $n=1$.
\noindent
\nocite{*}
\bibliographystyle{spr-mp-nameyear-cnd}
|
\section{Chernoff Bound for Limited Independence}
\begin{lemma}[Chernoff Bound for Limited Independence] \label{LemmaChernoff}
Let $X_1 \cdots X_\ell$ be $\delta$-almost $k$-wise independent random variables with $X_i \in \{0,1\}$ for all $i$. Set $X = \sum_i X_i$ and $\mu = \sum_i \mu_i = \sum_i \ex{X}{X_i}$, and suppose $\mu_i \leq 1/2$ for all $i$. If $k \leq \mu /10$ is even, then, for all $\alpha \in (0,1)$, $$\pr{X}{\left|X - \mu \right| \geq \alpha \mu } \leq \left( \frac{20k}{\alpha^2 \mu } \right)^{\lfloor k/2 \rfloor} + 2\delta \cdot \left(\frac{\ell}{\alpha\mu}\right)^k.$$
\end{lemma}
The following proof is based on \cite[Theorem 4]{SSS}. The only difference is that we extend to almost $k$-wise independence from $k$-wise independence.
\begin{proof
Assume, without loss of generality, that $k$ is even. It is well-known \cite{SSS,JelaniMO,BRchernoff} that, if the $X_i$s are fully independent, then $$\ex{X_i}{(X-\mu)^k} \leq (20k \mu)^{k/2}.$$ This also holds when the $X_i$s are only $k$-wise independent, as $(X-\mu)^k$ is a degree-$k$ polynomial in the $X_i$s. Here the $X_i$s are $\delta$-almost $k$-wise independent, which gives $$\ex{X_i}{(X-\mu)^k} \leq (20k \mu)^{k/2} + 2 \delta \ell^k.$$
Thus we can apply Markov's inequality to obtain the result:
$$\pr{X}{\left|X - \mu \right| \geq \alpha \mu } = \pr{X}{(X - \mu )^k \geq (\alpha \mu)^k } \leq \frac{\ex{X_i}{(X-\mu)^k}}{(\alpha \mu)^k} \leq \left( \frac{20 k \mu}{\alpha^2 \mu^2} \right)^{k/2} + 2 \delta \left(\frac{\ell}{\alpha \mu} \right)^k.$$
\end{proof}
\section{First-Order Fourier Coefficients of Branching Programs} \label{AppendixFirstOrder}
\begin{theorem}
Let $B$ be a width-$w$, length-$n$, read-once, oblivious branching program. Then $$\sum_{i \in [n]} \norm{\widehat{B}[\{i\}]}_2 \leq O(\log n)^{w-2}.$$
\end{theorem}
The proof is similar to the proof of the Coin Theorem by Steinberger \cite{Steinberger}. The main difference is that we need a new proof of the collision lemma:
We call a layer $B_i$ of a branching program \textbf{trivial} if $L(B_i)=0$ and nontrivial otherwise. We say that a layer $B_i$ has a \textbf{collision} if there exist two edges with the same label and the same endpoint, but different start points. All non-permutation layers have a collision. If there is a collision in layer $i$, then with probability at least $1/2$ a random restriction of layer $i$ reduces the width.
\begin{lemma}[Collision Lemma] \label{lemcollision}
Let $f: \{0,1\}^n \to \{0,1\}$ be a function computed by a width-$w$ ordered branching program $B$. Then there exists a function $g$ computed by a width-$w$ ordered branching program $B'$ such that every nontrivial layer of $B'$ has a collision and $$\sum_i |\widehat{f}[\{i\}]| \leq \sum_i |\widehat{g}[\{i\}]|.$$
\end{lemma}
\begin{proof}
We will construct $B'$ by flipping edge labels in $B$.
Begin by ordering the vertices in each layer by their acceptance probability -- that is, the probability that $f(U)=1$ conditioned on passing through that vertex. (If two vertices have the same acceptance probability order them arbitrarily.) We will flip the edge labels such that for every vertex the 0-edge leads to a higher-ranked vertex than the 1-edge (or they lead to the same vertex).
Note that flipping the edges in layer $i$ only affects the corresponding Fourier coefficient. Thus we need only show that flipping the edges in layer $i$ does not decrease $|\widehat{f}[i]|$.
Fix a layer $i$. For a vertex $u$ on the left of layer $i$ of $B$, let $u_0$ and $u_1$ be the vertices led to by the 0- and 1-edges respectively and let $p_u$ be the probability that a random walk in $B$ reaches $u$. For a vertex $v$ on the right of layer $i$ of $B$, let $q_v$ be the acceptance probability of $B$ under uniform input conditioned on passing through vertex $v$. Then $$\widehat{f}[\{i\}] = \frac12 \sum_u p_u (q_{u_0} - q_{u_1}).$$ Flipping edge labels corresponds to flipping the signs of the terms in the above sum. Clearly, $|\widehat{f}[\{i\}]|$ is maximised if every term has the same sign. Our choice of flips ensures this is the case, as $q_{u_0} \geq q_{u_1}$.
Every nontrivial layer of $B$ must have a collision, as a result of the ordering of edge labels: Consider a layer $i$ and let $v$ be the highest ranked vertex on its right such that the incoming edges are from different vertices on the left. Suppose, for the sake of contradiction, that the incoming edges have different labels. Pick the edge labelled 1 and let u be its start point. Let $u_0$ be the vertex reached from $u$ by the edge labelled $0$. Then, by our choice of labels $u_0$ is ranked higher than $u$ -- a contradiction, as $u$ is the highest ranked vertex with distinct incoming edges.
\end{proof}
Define $\xi(n,w)$ to be the maximal first-order Fourier mass of any function computed by a length-$n$, width-$w$, ordered branching program. We follow the structure of Steinberger's proof \cite{Steinberger} to bound $\xi$.
\begin{lemma}
For all $n$ and $w \geq 3$, $\xi(n,w) \leq (2 + 2 \log_2(n)) \cdot (\xi(n,w-1)+1)$.
\end{lemma}
\begin{proof}
Let $B$ be a length-$n$, width-$w$, ordered branching program that maximises the first-order Fourier mass of the function $f$ it computes. By Lemma \ref{lemcollision}, we may assume that every nontrivial layer of $B$ has a collision. We may assume that there are no trivial layers: otherwise we can remove them without affecting the Fourier mass.
Let $m = \lceil 1 + 2 \log_2 n \rceil$. Split the first-order Fourier coefficients into $m$ groups of the form $$G_{i'} = \{ i \in [n] : i ~\mathrm{mod}~ m = i' \}.$$ We bound the first-order Fourier mass of each group separately and sum them together. i.e. $\sum_{i \in [n]} |\widehat{f}[\{i\}]| = \sum_{i' \in [m]} \sum_{i \in G_{i'}} |\widehat{f}[\{i\}]|$. Fix one group $G = G_{i'}$.
We apply a random restriction to $B$ to obtain the function $f|_{\overline{G} \leftarrow U}$ computed by the branching program $B|_{\overline{G} \leftarrow U}$. We have $$\sum_{i \in G} |\widehat{f}[\{i\}] \leq \ex{U}{\sum_{i \in G} |\widehat{f|_{\overline{G} \leftarrow U}}[\{i\}]|}.$$ So if suffices to bound the first-order Fourier mass of $f|_{\overline{G} \leftarrow U}$.
We claim that $B|_{\overline{G} \leftarrow U}$ is a width-$(w-1)$, ordered branching program with probability at least $1 - n \cdot 2^{1-m}$. This implies that
$$\ex{U}{\sum_{i \in G} |\widehat{f|_{\overline{G} \leftarrow U}}[\{i\}]|} \leq \xi(n,w-1) + n \cdot 2^{1-w} \cdot n \leq \xi(n,w-1) + 1.$$
Thus $$\sum_{i \in [n]} |\widehat{f}[\{i\}]| \leq \sum_{i' \in [m]} \ex{U}{\sum_{i \in G_{i'}} |\widehat{f|_{\overline{G_{i'}} \leftarrow U}}[\{i\}]|} \leq \sum_{i' \in [m]} \xi(n,w-1) + 1 \leq m (\xi(n,w-1) + 1),$$ as required.
Now to prove the claim: Fix an unrestricted layer $i$ of $B|_{\overline{G} \leftarrow U}$ other than the last layer (which can always be assumed to have width 2 anyway). Layer $i$ is followed by $m-1$ restricted layers. With probability at least $1-2^{1-m}$ at least one of these layers will contain a collision, thus reducing the number of vertices on the right of layer $i$. A union bound gives the required probability.
\end{proof}
\begin{lemma}
$\xi(n,2) \leq 10$.
\end{lemma}
Solving the recurrance for $\xi$ gives $\xi(n,w) \leq O(\log n)^{w-2}$, as required.
\omitted
\section{Bootstrapping}
The following proposition shows that if we can bound the Fourier mass up to level $O(\log n)$, then we can bound the Fourier mass at all levels.
\begin{proposition} \label{PropositionBootstrapping}
Let $B$ be a length-$n$, ordered branching program such that, for all $i,j,k \in [n]$ with $k \leq 2k^*$ and $i \leq j$, we have $L^k(B_{i \cdots j}) \leq a \cdot b^k$. Suppose $a \cdot n \leq 2^{k^*}$. Then, for all $i,j,k \in [n]$ with $i \leq j$, we have $L^k(B_{i \cdots j}) \leq a \cdot (2b)^k$.
\end{proposition}
The proof is similar to that of Lemma \ref{LemmaWellOrder}.
\begin{proof}
Suppose the proposition is false and fix the smallest $k$ such that the statement does not hold. Clearly $k>2k^*$. Let $k'=k-k^*$. By minimality $L^{k'}(B_{i \cdots j}) \leq a \cdot (2b)^{k'}$ for all $i \leq j$. Now $$L^k(B_{i \cdots j}) \leq \sum_{\ell=i}^{j+1} L^{k'}(B_{i \cdots \ell-1}) \cdot L^{k^*}(B_{\ell \cdots j}) \leq \sum_{\ell=i}^{j+1} a \cdot (2b)^{k'} \cdot a \cdot b^{k^*} \leq n \cdot a \cdot (2b)^{k'} \cdot a \cdot b^{k^*} = a \cdot (2b)^k \cdot \left( \frac{n a}{2^{k^*}} \right).$$ Since $ n a \leq 2^{k^*}$, we have a contradiction, as we assumed $L^{k^*}(B_{\ell \cdots j}) > a \cdot (2b)^k$.
\end{proof}
\section{Optimality of Result} \label{AppendixOptimal}
The following result shows that Theorem \ref{TheoremLow} is close to optimal.
\omitted
\begin{proposition} \label{PropositionTight}
There exists a constant $c>0$ and an infinite family of functions $f_n: \{0,1\}^n \to \{0,1\}$ and that are computed by 3OBPs such that the following holds. There is no polynomial $q$ such that $$L^k(f_n) \leq q(n) \cdot \left(\frac{c \cdot \log n}{\log \log n}\right)^k$$ for all $n$ and $k \in [n]$.
\end{proposition}
Our main result shows that $L^k(f_n) = \mathrm{poly}(n) \cdot (O(\log n))^k$. Hence this proposition shows that the base $O(\log n)$ cannot be improved by more than a $\log \log n$ factor. The $\log \log n$ factor comes from the fact that we allow a polynomial factor $q(n)$ in the bound. If we remove $q(n)$, the $\log \log n$ factor also disappears in Proposition \ref{PropositionTight}.
\begin{proof}
Let $n = m \cdot 2^m$, where $m$ is an integer. Define $f_{n} : \{0,1\}^{n} \to \{0,1\}$ by $$f_{n}(x) = \prod_{i \in [2^m]} \left( 1 - \prod_{j \in [m]} x_{i,j} \right),$$ where we view $x \in \{0,1\}^n$ as a $2^m \times m$ matrix $x \in \{0,1\}^{2^m \times m}$. This is (up to a negation) the Tribes function \cite{tribes}. This function can be computed by a 3OBP. Now we show that it has large Fourier growth.
The Fourier coefficients of $f_{n}$ are given as follows. For $s \subset [n]$ (which we identify with $s \in \{0,1\}^{2^m \times m}$),
\begin{align*}
\widehat{f_{n}}[s] =& \ex{U}{f(U)\chi_s(U)}\\
=& \ex{U}{ \prod_{i \in [2^m]} \left( 1 - \prod_{j \in [m]} U_{i,j} \right) \chi_{s_i}(U_i)}\\
=&\prod_{i \in [2^m]} \ex{U}{\chi_{s_i}(U_i) - \prod_{j \in [m]} U_{i,j} \chi_{s_i}(U_i)}\\
=&\prod_{i \in [2^m]} \left( \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right),
\end{align*}
where $s_i = (s_{i,1},s_{i,2},\cdots,s_{i,m})$.
The damped Fourier mass is also easy to compute. For $p \in (0,1)$,
\begin{align*}
L_p(f_{n}) + |\widehat{f_{n}}[0]|=& \sum_{s \in \{0,1\}^{2^m \times m}} p^{|s|} |\widehat{f_{n}}[s]|\\
=& \sum_{s \in \{0,1\}^{2^m \times m}} \prod_{i \in [2^m]} p^{|s_i|} \left| \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right|\\
=& \prod_{i \in [2^m]} \sum_{s_i \in \{0,1\}^m} p^{|s_i|} \left| \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right|\\
=& \prod_{i \in [2^m]} \left( 1-2^{-m} + \sum_{s_i \ne 0} p^{|s_i|} 2^{-m} \right)\\
=& \prod_{i \in [2^m]} \left( 1- 2^{-m} + 2^{-m} (1+p)^m - 2^{-m} \right)\\
=& \left( 1 + \frac{(1+p)^m - 2}{2^m} \right)^{2^m}.
\end{align*}
Set $p=10 \log (m) / m$. We have $$(1+p)^m = \left( 1 + \frac{10 \log m}{m}\right)^m = e^{10 \log m} (1 \pm o(1)) \geq 10 m^2 + 10$$ for sufficiently large $m$. Thus, for sufficiently large $m$, $$L_p(f_{n}) \geq \left( 1 + \frac{(1+p)^m - 2}{2^m} \right)^{2^m} -1 \geq \left( 1 + \frac{10m^2+8}{2^m} \right)^{2^m}-1 = e^{10m^2+8}(1 \pm o(1)) -1 \geq 2^{m^2} = n^{\omega(1)}.$$
However, if there exists a polynomial $q$ satisfying the conditions of the proposition with $c=1/20$, we have $$L_p(f_{n}) = \sum_{k \in [n]} p^k L^k(f_{n}) \leq \sum_{k \in [n]} \left(\frac{10 \log (m)}{ m}\right)^k \cdot q(n) \cdot \left(\frac{2 c m}{\log m}\right)^k = \sum_{k \in [n]} q(n) = n^{O(1)},$$ which is a contradiction.
\end{proof}
\begin{proposition} \label{PropositionTight}
There exists an infinite family of functions $f_n: \{0,1\}^n \to \{0,1\}$ and that are computed by 3OBPs such that the following holds. Let $q : \mathbb{N} \to \mathbb{R}$ be an increasing function with $20 \leq q(n) \leq \exp(\exp(o(\sqrt{\log n})))$. For all sufficiently large $n$, there exists $k \in [n]$ such that $$L^k(f_n) > q(n) \cdot \left(\frac{\log n}{20 \log \log q(n)}\right)^k.$$.
\end{proposition}
Our main result shows that $L^k(f_n) \leq \mathrm{poly}(n) \cdot (O(\log n))^k$. Setting $q(n) = \mathrm{poly}(n)$, this proposition shows that the base $O(\log n)$ cannot be improved by more than a $\log \log n$ factor. The $\log \log n$ factor comes from the fact that we allow a polynomial factor $q(n)$ in the bound. If we demand $q(n) = O(1)$, the base $O(\log n)$ is optimal.
\begin{proof}
Let $n = m \cdot 2^m$, where $m$ is an integer. Define $f_{n} : \{0,1\}^{n} \to \{0,1\}$ by $$f_{n}(x) = \prod_{i \in [2^m]} \left( 1 - \prod_{j \in [m]} x_{i,j} \right),$$ where we view $x \in \{0,1\}^n$ as a $2^m \times m$ matrix $x \in \{0,1\}^{2^m \times m}$. This is (up to a negation) the Tribes function \cite{tribes}. This function can be computed by a 3OBP. Now we show that it has large Fourier growth.
The Fourier coefficients of $f_{n}$ are given as follows. For $s \subset [n]$ (which we identify with $s \in \{0,1\}^{2^m \times m}$),
\begin{align*}
\widehat{f_{n}}[s] =& \ex{U}{f(U)\chi_s(U)}\\
=& \ex{U}{ \prod_{i \in [2^m]} \left( 1 - \prod_{j \in [m]} U_{i,j} \right) \chi_{s_i}(U_i)}\\
=&\prod_{i \in [2^m]} \ex{U}{\chi_{s_i}(U_i) - \prod_{j \in [m]} U_{i,j} \chi_{s_i}(U_i)}\\
=&\prod_{i \in [2^m]} \left( \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right),
\end{align*}
where $s_i = (s_{i,1},s_{i,2},\cdots,s_{i,m})$.
The damped Fourier mass is also easy to compute. For $p \in (0,1)$,
\begin{align*}
L_p(f_{n}) + |\widehat{f_{n}}[0]|=& \sum_{s \in \{0,1\}^{2^m \times m}} p^{|s|} |\widehat{f_{n}}[s]|\\
=& \sum_{s \in \{0,1\}^{2^m \times m}} \prod_{i \in [2^m]} p^{|s_i|} \left| \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right|\\
=& \prod_{i \in [2^m]} \sum_{s_i \in \{0,1\}^m} p^{|s_i|} \left| \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right|\\
=& \prod_{i \in [2^m]} \left( 1-2^{-m} + \sum_{s_i \ne 0} p^{|s_i|} 2^{-m} \right)\\
=& \prod_{i \in [2^m]} \left( 1- 2^{-m} + 2^{-m} (1+p)^m - 2^{-m} \right)\\
=& \left( 1 + \frac{(1+p)^m - 2}{2^m} \right)^{2^m}.
\end{align*}
Set $p=(1+\log (3 + \log q(n))) / m$. We have $$(1+p)^m = \left( 1 + \frac{1 + \log (3 + \log q(n))}{m}\right)^m = e^{1 + \log (3 + \log q(n))} (1 - o(1)) \geq 3 + \log q(n)$$ for sufficiently large $m$. Thus, for sufficiently large $m$, $$L_p(f_{n}) \geq \left( 1 + \frac{(1+p)^m - 2}{2^m} \right)^{2^m} -1 \geq \left( 1 + \frac{1 + \log q(n)}{2^m} \right)^{2^m}-1 = e^{1 + \log q(n)}(1 - o(1)) -1 \geq q(n).$$
Suppose for the sake of contradiction that $L^k(f_n) \leq q(n) \cdot (\log n / 20 \log \log q(n))^k$ for all $k \in [n]$. We have $$L_p(f_{n}) = \sum_{k \in [n]} p^k L^k(f_{n}) \leq \sum_{k \in [n]} \left(\frac{1 + \log (3 + \log q(n))}{m}\right)^k \cdot q(n) \cdot \left(\frac{\log n}{20 \log \log q(n)}\right)^k \leq \sum_{k \in [n]} q(n) 2^{-k} < q(n),$$ which is a contradiction.
\end{proof}
A more careful analysis gives the following bound.
\begin{proposition}
There exists an infinte family of functions $f_n : \{0,1\}^n \to \{0,1\}$ that are computed by 3OBPs such that, for all $k \in [n]$, $$L^k(f) \geq \Omega\left(\frac{\log n}{\log k}\right)^k.$$
\end{proposition}
\begin{proof}
Let $n$, $m$, and $f_n$ be as in the proof of Proposition \ref{PropositionTight}. For $s \in \{0,1\}^{2^m \times m}$, denote $$\ell(s) = |\{i \in [2^m] : s_i \ne 0 \}| = |\{ i \in [2^m] : \exists j \in [m] ~~ s_{i,j}=1\}|.$$ Then, for all $s \in \{0,1\}^{n \times m}$, we have $$|\widehat{f_{n}}[s]| = \prod_{i \in [2^m]} \left| \mathbb{I}(s_i = 0) - 2^{-m} (-1)^{|s_i|} \right|= (1-2^{-m})^{2^m-\ell(s)} \cdot (2^{-m})^{\ell(s)}.$$
Fix $\ell$ with $k/\ell \leq m$. Set $h=\lfloor k/\ell \rfloor$. Choose $i,j \geq 0$ with $i+j=\ell$ and $ih+j(h+1) = k$. Then
\begin{align*}
L^k(f_n) \geq& \sum_{|s|=k \wedge \ell(s)=\ell} |\widehat{f_{n}}[s]|\\
\geq& {2^m \choose \ell} {m \choose h}^i {m \choose h+1}^j \cdot (1-2^{-m})^{2^m-\ell} \cdot (2^{-m})^{\ell}\\
\geq& \left( \frac{2^m}{\ell} \right)^\ell \left(\frac{m}{h}\right)^{hi} \left(\frac{m}{h+1}\right)^{(h+1)j} \cdot \left(1-\frac{1}{2^m}\right)^{2^m} \cdot \left(\frac{1}{2^m}\right)^{\ell}\\
\geq& \frac{1}{4} \left( \frac{1}{\ell} \right)^\ell \left(\frac{m}{h}\right)^{hi} \left(\frac{m}{h+1}\right)^{(h+1)j}\\
\geq& \frac{1}{4} \left( \frac{1}{\ell} \right)^\ell \left(\frac{m}{h+1}\right)^{hi+(h+1)j}\\
\geq& \frac{1}{4} \left( \frac{1}{\ell} \right)^\ell \cdot \left( \frac{m}{k/\ell+1}\right)^k.
\end{align*}
Suppose $k \leq 2^{m-1}$. Setting $\ell = \lceil k/\log_2 (2k) \rceil$, we have
\begin{align*}
L^k(f_n) \geq& \frac{1}{4} \left( \frac{1}{\ell} \right)^\ell \left(\frac{m}{k/\ell + 1}\right)^{k}\\
\geq& \frac{1}{4} \cdot \frac{1}{2^{2k+2}} \cdot \left( \frac{m}{\log_2 k + 2}\right)^k,
\end{align*}
as $$\log_2(\ell^\ell) = \ell \log_2 \ell < \frac{k + \log_2 k}{\log_2 k} \log_2 (k + \log_2 k) \leq 2k+2.$$
Since $m = \Theta(\log n)$, this gives the result for $k \leq 2^{m-1}$. If $k > 2^{m-1}$, then $\log k = \Theta(\log n)$ and the result is trivial.
\end{proof}
\section{Further Work} \label{SectionConclusion}
Our results hinge on the fact that ``mixing'' is well-understood for regular branching programs \cite{BRRY,ReingoldSteinkeVadhan2013,KNP,De,Steinke12} and for (non-regular) width-2 branching programs \cite{BogdanovDvVeYe09}. We are able to use random restrictions to reduce from width 3 to width 2 (Section \ref{subsect:width-reduction}), where we can exploit our understanding of mixing (Section \ref{SubSectionMixing}). Indeed, this understanding underpins most results for these restricted models of branching programs.
What about width 4 and beyond? Using a random restriction we can reduce analysing width 4 to ``almost'' width 3 -- that is, Proposition \ref{PropositionPart1} generalises. Unfortunately, the reduction does not give a true width-3 branching program and thus we cannot repeat the reduction to width 2. Moreover, we have a poor understanding of mixing for non-regular width-3 branching programs, which means we cannot use the same techniques that have worked for width-2 branching programs.
Our results provide some understanding of mixing in width-3. We hope this understanding can be developed further and will lead to proving Conjecture \ref{conj:fouriergrowth} and other results.
The biggest obstacle to extending our techniques to $w>3$ is Lemma \ref{LemmaLL}.
The problem is that the
parameter $\lambda(D)$ is no longer a useful measure of mixing for width-3 and above. In particular, $\lambda(D)>1$ is possible if $\ex{U}{D[U]}$ is a $3 \times 3$ matrix.
To extend our techniques, we need a better notion of mixing. Using $\lambda(D)$ is useful for regular branching programs (it equals the second eigenvalue for regular programs), but is of limited use for non-regular branching programs. Our proof uses a different notion of mixing -- collisions: To prove Proposition \ref{PropositionPart1}, we used the fact that a random restriction of a non-regular layer will with probability at least 1/2 result in the width of the right side of the layer being reduced. This is a form of mixing, but it is not captured by $\lambda$. Ideally, we want a notion of mixing that captures both $\lambda$ and width-reduction under restrictions.
Our proofs combine the techniques of Braverman et al.~\cite{BRRY} and those of Brody and Verbin \cite{BrodyVerbin} and Steinberger \cite{Steinberger}. We would like to combine them more cleanly -- presently the proof is split into two parts (Proposition \ref{PropositionPart1} and Lemma \ref{LemmaLL}). This would likely involve developing a deeper understanding of the notion of mixing.
Our seed length $\tilde{O}(\log^3 n)$ is far from the optimal $O(\log n)$. Further improvement would require some new techniques:
We could potentially relax our notion of Fourier growth to achieve better results. Rather than bounding $L^k(f)$, it suffices to bound $L^k(g)$, where $g$ \emph{approximates} $f$:
\begin{proposition}[{\cite[Proposition 2.6]{DETT}}] \label{PropositionSandwich}
Let $f, f_+, f_- : \{0,1\}^n \to \mathbb{R}$ satisfy $f_-(x) \leq f(x) \leq f_+(x)$ for all $x$ and $\ex{U}{f_+(U) - f_-(U)} \leq \delta$. Then any $\varepsilon$-biased distribution $X$ gives $$\left| \ex{X}{f(X)} - \ex{U}{f(U)} \right| \leq \delta + \varepsilon \cdot \max \left\{ L(f_+), L(f_-) \right\}.$$
\end{proposition}
The functions $f_+$ and $f_-$ are called sandwiching polynomials for $f$. This notion of sandwiching is in fact a tight characterisation of small bias \cite[Proposition 2.7]{DETT}. That is, any function $f$ fooled by all small bias generators has sandwiching polynomials satisfying the hypotheses of Proposition \ref{PropositionSandwich}.
Gopalan et al.~\cite{GMRTV} use sandwiching polynomials in the analysis of their generator for CNFs. This allows them to set a constant fraction of the bits at each level of recursion ($p = \Omega(1)$), while we set a $1/O(\log n)$ fraction at each level. We would like to similarly exploit sandwiching polynomials for branching programs to improve the seed length of the generator.
A further avenue for improvement would be to modify the generator construction to have $\Theta(1/p)$ levels of recursion, rather than $\Theta(\log(n)/p)$. This would require a significantly different analysis.
\omitted
\begin{itemize}
\item Generalise to width $>3$. Discuss issues: the partition lemma (L \ref{LemmaPartition}) is interesting. This is where width 2 is most crucial. If we can find a different measure of mixing and prove a version of the partition lemma, we might be able to generalise proof.
\item Find the proof from the book: This is ugly. I think the `right' proof of this result will be highly insightful. In particular, we use techniques from both Braverman \cite{BRRY} and Brody-Verbin \cite{BrodyVerbin} to prove the main lemma. Properly unifying these techniques would be awesome.
\item Improve seed length. Using this approach (low-order bound + this generator) we cannot hope to do better than $O(log^3 n)$ seed length.
\end{itemize}
One open problem is to extend the main result (Theorem \ref{TheoremPRG}) to regular or even non-regular branching programs while maintaining $\mathrm{polylog}(n)$ seed length. As discussed in Section \ref{SubSectionPRG}, the only part of our analysis that fails for regular branching programs is the recursive analysis. The problem is that regular branching programs are not closed under restriction---that is, setting some of the bits of a regular branching program does not necessarily yield a regular branching program. In particular, we cannot bound $$\norm{\ex{\tilde{U}}{B[\text{Select}(t,x,\tilde{U})]}-\ex{U}{B[\text{Select}(t,x,{U})]}}$$ for fixed $t$ and $x$ by the distinguishability of $\tilde{U}$ and $U$ by another read-once, oblivious, regular branching program $\overline{B}_{x,t}$. We have two options:
\begin{itemize}
\item Find another way to bound $\norm{\ex{T,X,\tilde{U}}{B[\text{Select}(T,X,\tilde{U})]}-\ex{T,X,}{B[\text{Select}(T,X,{U})]}}$.
\item Extend the main lemma (Theorem \ref{TheoremMainLemma}) to non-regular branching programs.
\end{itemize}
Towards the latter option, we have the following conjecture.
\begin{conjecture} \label{Conjecture}
For every constant $w$, the following holds. Let $B$ be a length-$n$, width-$w$, read-once, oblivious branching program. Then $$L_2^k(B) = \sum_{s \in \{0,1\}^n : |s| = k} \norm{\widehat{B}[s]}_2 \leq n^{O(1)}(k \log n )^{O(k)}$$
for all $k \in [n]$.
\end{conjecture}
This conjecture relates to the Coin Theorem of Brody and Verbin (see the discussion in Section \ref{SubSectionCoin}). Specifically, if we remove the $n^{O(1)}$ factor, this conjecture implies the Coin Theorem.
Conjecture \ref{Conjecture} would suffice to construct a pseudorandom generator for constant-width, read-once, oblivious branching programs with seed length $\mathrm{polylog}(n)$.
\bigskip
The seed length of our generators is worse than that of generators for ordered branching programs. Indeed, for ordered \emph{permutation} branching programs of constant width, it is known how to achieve seed length $O(\log n)$ \cite{KNP}, whereas we only achieve seed length $O(\log^2 n)$ in Theorem \ref{TheoremPRG}. For general ordered branching programs, Nisan \cite{Nisan} obtains seed length $O(\log(nw) \log(n))$, whereas Theorem \ref{TheoremPRGgeneral} gives seed length $\tilde{O}(\sqrt{n} \log(w))$. It would be interesting to close these gaps.
\section{Fourier Analysis of Width-3 Branching Programs} \label{SubSectionLowOrder}
In this section we prove a bound on the low-order Fourier mass of width-3, read-once, oblivious branching programs. This is key to the analysis of our pseudorandom generator. Improvements to this result directly translate to improvements in our final result.
\begin{theorem} \label{TheoremLow}
Let $f : \{0,1\}^n \to \{0,1\}$ be computed by a width-3, read-once, oblivious branching program. Then, for all $k \in [n]$, $$L^k(f) \leq 8 n^2 \cdot \left( C \cdot \log_2(3 n) \right)^k = n^2 \cdot \left( O(\log n) \right)^k,$$ where $C$ is a universal constant.\footnote{We have not optimised any constants and only show $C \leq 10^7$.}
\end{theorem}
To prove Theorem \ref{TheoremLow} we will consider the matrix valued function $B$ of the branching program computing $f$. Note that $|\hat{f}[s]| \leq ||\widehat{B}[s]||_2$ for all $s$ so $L^k(f) \leq L^k(B)$. We may also assume without loss of generality that the first and last layers of the program have width 2 (there is only one start state, and there are at most 2 accept states otherwise the program is trivial).
The proof proceeds in two parts. The first part reduces the problem to one about branching programs of a special form, namely ones where many layers have been reduced to width-2. The second part uses the mixing properties of width-2 programs to bound the Fourier mass.
\subsection{Part 1 -- Reduction of Width by Random Restriction} \label{subsect:width-reduction}
Our reduction can be stated as follows.
\begin{proposition} \label{PropositionPart1}
Let $B$ be a length-$n$ width-3 ordered branching program (abbreviated \textbf{3OBP}), $m \geq k$, and $k \in [n]$ with the first and last layers having width at most 2.
Then
$$L^k(B) \leq n\cdot \binom{m}{k} \sum_{\ell \geq 0} 2^{-\ell(m-k)} L^k(D^{6(\ell+1)k})$$
where each $D^{6(\ell+1)k} = D_1^{6(\ell+1)k} \circ D_2^{6(\ell+1)k} \circ \cdots \circ D_r^{6(\ell+1)k}$, where $r \in [n]$, each $D_i^{6(\ell+1)k}$ is a 3OBP with at most $6(\ell+1) k$ non-regular layers, and the first and last layers of each $D_i$ have width at most 2.
\end{proposition}
In Section \ref{SubSectionMixing}, we will prove $L^k(D^{6(\ell+1)k}) \leq n \cdot O(\ell)^k$. Taking $m=2k$, this implies $L^k(B) \leq n^2 \cdot O(k)^k$. Finally, in Section \ref{SubSectionBootstrapping}, we show that we may assume $k \leq O(\log n)$, so we get a Fourier growth bound of $L^k(B) \leq n^2 \cdot O(\log n)^k$.
Here we focus on the proof of Proposition \ref{PropositionPart1}.
First some definitions:
For $g \subset [n]
-- and $x \in \{0,1\}^n$, define the \textbf{restriction of $B$ to $g$ using $x$ -- denoted $B|_{\overline{g} \leftarrow x}$ --} to be the branching program obtained by setting the inputs (layers of edges) of $B$ outside $g$ to values from $x$ and leaving the inputs in $g$ free. More formally, $$B|_{\overline{g} \leftarrow x}[y] = B[\mathrm{Select}(g,y,x)],$$ where $$\mathrm{Select}(g,y,x)_i = \left\{ \begin{array}{cl} y_i & i \in g \\ x_i & i \notin g \\ \end{array} \right\}.$$
We prove Proposition \ref{PropositionPart1} by considering a restriction $B|_{\overline{g} \leftarrow x}$ for a carefully chosen $g$ and a random $x$. We show (Lemma \ref{LemmaInterwoven}) that is suffices to bound the Fourier growth of the restricted program $B|_{\overline{g} \leftarrow x}$ and (Lemma \ref{LemmaRestrict}) that the restricted program is of the desired form $D^{6(\ell+1)k}$ with high probability.
Define a \textbf{chunk} to be a 3OBP with exactly one non-regular layer. An \textbf{$l$-chunk} 3OBP is a 3OBP $B$ such that $B = C_1 \circ C_2 \circ \cdots \circ C_l$, where each $C_i$ is a chunk. Equivalently, an $l$-chunk 3OBP is a 3OBP with exactly $l$ non-regular layers. The partitioning of $B$ into chunks is not necessarily unique. But we fix one such partitioning for each 3OBP and simply refer to the $i^\text{th}$ chunk $C_i$. If $B$ is an $l$-chunk length-$n$ 3OBP, let $c_i \subset [n]$ be the coordinates corresponding to $C_i$.
We will compute a bound on the level-$k$ Fourier weight of $B$ via a series of ``interwoven'' restrictions similar to Steinberger's technique \cite{Steinberger}. Lemma \ref{LemmaInterwoven} below tells us that we may obtain a bound by bounding, in expectation, the level-$k$ weight of a restricted branching program. We then argue that with high probability over this restriction, the width of the resulting program will be essentially reduced. In particular, there is a layer of width 2 after every $O(k)$ non-regular layers.
We now describe some notation that will be used for the interwoven restrictions.
For $t \subset [m]$, define $$g_t := \bigcup_{(i \text{ mod } m)+1 \in t} c_i$$ and $$G_t^k := \{s \subset g_t : |s| = k\}.$$ We refer to $g_t$ as the \textbf{$t^\text{th}$ group of indices} and $G_t^k$ as the \textbf{$t^\text{th}$ group of (order $k$) Fourier coefficients}.
The following lemma tells us that we may bound the level-$k$ Fourier weight by considering a fixed subset $t\subset [m]$ of size $k$ and the level-$k$ Fourier weight of the branching program that results by randomly restricting the variables outside of $g_t$:
\begin{lemma}\label{LemmaInterwoven} Let $B$ be a length-$n$ 3OBP, $k\in [n]$, $m\geq k$ and $g_t$ as above. Then
$$ L^k(B)\leq \binom{m}{k} \cdot ~\max_{t \subset [m] : |t| = k}~ \ex{U}{L^k(B|_{\overline{g_t} \leftarrow U})}.$$
\end{lemma}
\begin{proof}
Note that some Fourier coefficients appear in multiple $G_t$s, but every coefficient at level $k$ appears in at least one $G_t$. Thus
\begin{align*}
L^k(B) & \leq \sum_{t : |t| = k} \sum_{s \in G_t^k} \norm{\widehat{B}[s]}_2 \\
& = \sum_{t : |t| = k} \sum_{s \in G_t^k} \norm{\ex{U}{\widehat{B_{\overline{g_t} \leftarrow U}}[s]}}_2 \\
& \leq \sum_{t : |t| = k} \sum_{s \in G_t^k} \ex{U}{\norm{\widehat{B_{\overline{g_t} \leftarrow U}}[s]}_2} \\
& = \sum_{t : |t| = k} \ex{U}{L^k(B|_{\overline{g_t} \leftarrow U})} \\
& \leq \binom{m}{k} \max_{t \subset [m] : |t| = k} \ex{U}{L^k(B|_{\overline{g_t} \leftarrow U})},
\end{align*}
where the second inequality follows from the convexity of the norm
\end{proof}
Given Lemma \ref{LemmaInterwoven}, we may now
prove Proposition \ref{PropositionPart1} by giving an upper bound on
$\ex{U}{L^k(B|_{\overline{g_t} \leftarrow U})}$ for any fixed $t \subset [m]$ with $|t|=k$.
To do this, we prove that a random restriction to $\overline{g_t}$ will, with high probability, result in a branching program of the desired form.
\begin{lemma}\label{LemmaRestrict}
Let $B$ be a length-$n$ 3OBP, $k,\ell\in [n]$, $m\geq k$ and fix $t\subseteq [m]$ with $|t|=k$.
Then with probability at least $1-n\cdot 2^{-\ell \cdot (m-k)}$ over a random choice of $x \in \{0,1\}^n$, $$ B|_{\overline{g_t}\leftarrow x} = D_1 \circ D_2\circ \cdots \circ D_r,$$
where $r\in [n]$ and each $D_i$ is a 3OBP with at most $6\ell k$ non-regular layers and the layer of vertices between $D_{i-1}$ and $D_i$ have width at most $2$.
\end{lemma}
\begin{proof}
Let $t = \{t_1, t_2, \cdots, t_k\}$ where $t_1 < t_2 < \cdots < t_k$. Define $t_{a+k} = t_a + m$ for all $a$. Let $a'$ be the largest value of $a$ such that $t_a \leq n$. We redefine $t_{a'+1}=n+1$. We can write $$B|_{\overline{g_t} \leftarrow x} = C_{t_1}' \circ C_{t_2}' \circ \cdots \circ C_{t_{a'}}',$$ where each $C_{t_a}'$ corresponds to $C_{t_a}$. However, the chunks $C_{t_a+1}, C_{t_a+2}, \cdots C_{t_{a+1}-1}$ have been restricted and $C_{t_a}'$ reflects that. Formally, for all $a \in [l'] \backslash \{1\}$ and $y \in \{0,1\}^{|c_{t_a}|}$, we have $$C_{t_a}'[y] = C_{t_a}[y] \cdot C_{t_{a}+1}[x_{c_{t_{a}+1}}] \cdot \cdots \cdot C_{t_{a+1}-1}[x_{c_{t_{a+1}-1}}],$$ and, for all $y \in \{0,1\}^{|c_{t_{1}}|}$, we have $$C_{t_{1}}'[y] = C_{1}[x_{c_{1}}] \cdot \cdots \cdot C_{t_{1}-1}[x_{c_{t_{1}-1}}] \cdot C_{t_{1}}[y] \cdot C_{t_{{1}}+1}[x_{c_{t_{1}+1}}] \cdot \cdots \cdot C_{t_2-1}[x_{c_{t_2-1}}],$$
where $x_{c_t}$ is the coordinates of $x$ contained in $c_t$. Moreover, we remove any unreachable vertices. That is, if $\ex{U}{C'_{t_a}[U]}$ has a column of zeros, we remove that column (making the matrix non-square) and remove the corresponding row from $C'_{t_{a+1}}$. This reduces the width of the layer of vertices between $C'_{t_a}$ and $C'_{t_{a+1}}$.
It should be clear that each $C_{t_a}'$ is a 3OBP with between one and three non-regular layers. (One non-regular layer comes from $C_{t_a}$ and the first and last layers may become non-regular after the restriction.)
Consider $\ell\cdot k+1$ consecutive $C_{t_a}'$s in the restricted program $C_{t_a}' \circ C_{t_{a+1}}' \circ \cdots \circ C_{t_{a+\ell\cdot k}}'$ and the corresponding subprogram before restriction $C_{t_a}\circ\cdots \circ C_{t_{a+\ell\cdot k}}$ which contains $\ell\cdot m+1$ chunks (recall that $t_{a+k}=t_a+m$). There are exactly $\ell\cdot k+1$ chunks $C_{t_a}',\dots,C_{t_a+k}',\dots, C_{t_{a+\ell\cdot k}}'$ which remain free after the restriction, i.e., $\ell (m-k)$ chunks have been restricted. Furthermore, each such chunk contains a non-regular layer.
Each non-regular layer that is restricted has at least a $1/2$ probability of reducing the width: Being non-regular implies that there are two edges with the same label going to the same vertex. There is a 1/2 probability of the restriction picking that label. If this happens, then one of the vertices on the right side of the layer becomes unreachable and therefore the width is reduced. This is the same argument that is used by Brody and Verbin \cite{BrodyVerbin} and Steinberger \cite{Steinberger}.
So, with probability at least $1-2^{-\ell(m-k)}$ over the choice of $x$, there exists a layer
between $C_{t_{a}}'$ and $C_{t_{a+\ell\cdot k}}'$ that has width at most 2. Call such a layer of vertices a \textbf{bottleneck}
By a union bound, with probability at least $1 - n \cdot 2^{-\ell(m-k)}$ over the choice of $x$, every $\ell\cdot k+1$ successive $C_{t_a}'$s have one such bottleneck. We split the $C_{t_a}'$s into groups that are separated by bottlenecks to write
$B|_{\overline{g_t} \leftarrow x} = D_1 \circ D_2 \circ \cdots \circ D_{r}$ (one $D_i$ per group).
Since each remaining chunk has at most 3 non-regular layers, and there can be at most $2\ell k$ chunks in each group, each $D_i$ is a 3OBP with at most $6\ell k$ non-regular layers and the layer between $D_i$ and $D_{i+1}$ has width at most 2 (i.e., is a bottleneck).
\end{proof}
We may now complete the proof of Proposition \ref{PropositionPart1}.
\begin{proof}[Proof of Proposition \ref{PropositionPart1}]
By Lemma \ref{LemmaInterwoven}, it suffices to bound, for every fixed $t$, the quantity
$\ex{U}{L^k(B|_{\overline{g_t} \leftarrow U})}.$
We compute the expectation of $L^k(B|_{\overline{g_t}\leftarrow U})$ by conditioning on the how far apart bottlenecks are in the restricted program and applying Lemma \ref{LemmaRestrict}. Let $\beta_x$ be the largest number of non-regular layers occuring in $B|_{\overline{g_t} \leftarrow x}$ that are not separated by a bottleneck (i.e., a width-2 layer).
\begin{align*}
\ex{U}{L^k(B|_{\overline{g_t} \leftarrow U})} & \leq \sum_{\ell \geq 0}
\pr{x}{ 6\ell k < \beta_x \leq 6(\ell+1)k} \cdot \ex{x}{L^k(B|_{\overline{g_t} \leftarrow x}) \mid \beta_x \leq 6(\ell+1)k}\\
& \leq \sum_{\ell \geq 0} n\cdot 2^{-\ell \cdot (m-k)} \cdot L^k(D^{6(\ell+1)k})
\end{align*}
where $D^{6(\ell+1)k}$ is the branching program that maximizes $L^k(B|_{\overline{g_t} \leftarrow x})$ over $x$ such that $\beta_x \leq 6(\ell+1)k$.
\end{proof}
\subsection{Part 2 -- Mixing in Width-2} \label{SubSectionMixing}
Now it remains to bound the Fourier mass of 3OBPs of the form given by Proposition \ref{PropositionPart1}.
\begin{proposition} \label{PropositionDmass}
Let $D^\ell$ be a length-$n$ 3OBP such that $$D^\ell = D_1^\ell \circ D_2^\ell \circ \cdots \circ D_r^\ell,$$ where each $D_i^\ell$ is a 3OBP with at most $\ell$ non-regular layers and width 2 in the first and last layers. Then $L^k(D^\ell) \le 2n \cdot (6000(\ell+1))^k = n \cdot O(\ell)^k$ for all $k,\ell \geq 1$.
\end{proposition}
Before we prove Proposition \ref{PropositionDmass}, we show how it implies a bound on Fourier mass of general 3OBPs.
\begin{proposition} \label{PropositionLow}
Let $B$ be a length-$n$, width-3, read-once, oblivious branching program with width 2 in the first and last layers. Then, for all $k \in [n]$,
$$L^k(B) := \sum_{s \in \{0,1\}^n : |s|=k} |\widehat{B}[s]| \leq 8 n^2 \cdot \left(200000 k \right)^k = n^2 \cdot O(k)^k.$$
\end{proposition}
\begin{proof}
Let $B$ be a 3OBP computing $f$ and assume $B$ has width 2 in the first and last layers. By Propositions \ref{PropositionPart1} and \ref{PropositionDmass}, we have, setting $m=2k+1$,
\begin{align*}
L^k(B) \leq& n\cdot \binom{m}{k} \sum_{\ell \geq 0} 2^{-\ell(m-k)} L^k(D^{6(\ell+1)k})\\
\leq& n\cdot \binom{m}{k} \sum_{\ell \geq 0} 2^{-\ell(m-k)} 2n \cdot (6000(6(\ell+1)k+1))^k\\
\leq& 2n^2 \binom{2k+1}{k} \sum_{\ell \geq 0} 2^{-\ell (k+1)} (6000(6(\ell+1)k+1))^k\\
\leq& 4n^2 4^k \sum_{\ell \geq 0} 2^{-\ell} \cdot \left( 6000\frac{6(\ell+1)k+1}{2^\ell} \right)^k\\
\leq& 4n^2 4^k \left(\sum_{\ell \geq 0} 2^{-\ell}\right) \cdot \left( 6000 \cdot 7k \right)^k\\
\leq& 8n^2 \cdot \left( 6000 \cdot 4 \cdot 7k \right)^k,
\end{align*}
as required.
\end{proof}
A key notion in our proof is a measure of the extent to which a branching program (or subprogram) mixes, and the way this is reflected in the Fourier spectrum.
For an ordered branching program $D$ of width $w$, define
$$\lambda(D) = \max_{x \in \mathbb{R}^w : \sum_i x_i = 0} \frac{\norm{x\ex{U}{D[U]}}_2}{\norm{x}_2}$$
The quantity $\lambda(D)$ is a measure of the \textbf{mixing} of $D$. If $D$ is regular, we have $\lambda(D) \in [0,1]$, where $0$ corresponds to perfect mixing and $1$ to no mixing. If $D$ is not regular, it is possible that $\lambda(D)>1$. However, for width-$2$ -- where $\ex{U}{D[U]}$ is a $2\times 2$ matrix -- it turns out that $\lambda(D) \leq 1$ even if $D$ is non-regular. In particular, $$\text{if}~~\ex{U}{D[U]} = \left( \begin{array}{cc} 1-\alpha & \alpha \\ \beta & 1- \beta \end{array} \right), ~~\text{then}~~ \lambda(D) = \frac{\norm{ (1, -1) \ex{U}{D[U]} }_2}{\norm{(1,-1)}_2} = |1-\alpha-\beta|.$$ The rows of $\ex{U}{D[U]}$ must sum to $1$ and have non-negative entries (as they are a probability distribution). So $\alpha, \beta \in [0,1]$, which implies $\lambda(D) \leq 1$. This fact is crucial to our analysis and is the main reason our results do not extend to higher widths.
Note that for any $s \ne 0$, the rows of $\widehat{D}[s]$ sum to zero.
Thus for any branching program $D=D_1\circ D_2$ and coefficient $\widehat{D}[s]$ with $s=(s_1,s_2)$ satisfying $s_2=0$, we have
\begin{equation}\label{eqn:mix-coeff}
\norm{\widehat{D}[s]}_2\leq \norm{\widehat{D_1}[s_1]}_2\cdot \lambda(D_2).
\end{equation}
For branching programs $B$ in which every layer is mixing -- that is $\lambda(B_i) \leq C < 1$ for all $i$ -- this fact can be used with an inductive argument (simpler than the proof below) to obtain a $1/(1-C)^{O(k)}$ bound on the level-$k$ Fourier mass. We show that any $D_i$ in the branching program of the form given by Proposition \ref{PropositionPart1} will either mix well or have small Fourier mass after restriction.
More precisely,
define the \textbf{$p$-damped Fourier mass} of a branching program $B$ as $$L_p(B) = \sum_{k > 0} p^k L^k(B) = \sum_{s \ne 0} p^{|s|} \norm{\widehat{B}[s]}_2.$$ Note that $L^k(B) \leq L_p(B) p^{-k}$ for all $k$ and $p$.
The main lemma we prove in this section is the following.
\begin{lemma} \label{LemmaLL}
If $D$ is a length-$d$ 3OBP with $k \geq 1$ non-regular layers that has only two vertices in the first and last layers, then $$\lambda(D) + L_p(D) \leq 1$$ for any $p \leq 1/6000(k+1)$.
\end{lemma}
First, we show that Lemma \ref{LemmaLL} implies Proposition \ref{PropositionDmass}:
\begin{proof}[Proof of Proposition \ref{PropositionDmass}]
We inductively show that $$L_p(D_1^\ell \circ \cdots\circ D_i^\ell) \leq 2i,$$
and hence $L_p(D) \leq 2r \leq 2n$.
For $i=0$ this is trivial. Now suppose it holds for $i$. By decomposition (Lemma \ref{LemmaFourier}), we have
\begin{align*}
L_p(D_1^\ell \cdots D_i^\ell \circ D_{i+1}^\ell) =& \sum_{(s,t) \ne 0} p^{|s|+|t|} \norm{\widehat{D_1^\ell \cdots D_i^\ell}[s] \cdot \widehat{D_{i+1}^\ell}[t]}_2\\
\leq& \sum_{s \ne 0} p^{|s|} \norm{\widehat{D_1^\ell \cdots D_i^\ell}[s] }_2 \sum_{t \ne 0} p^{|t|} \norm{\widehat{D_{i+1}^\ell}[t]}_2 \\&+ \sum_{s \ne 0} p^{|s|} \norm{\widehat{D_1^\ell \cdots D_i^\ell}[s] \cdot\widehat{D_{i+1}^\ell}[0]}_2 \\&+ \norm{\widehat{D_1^\ell \cdots D_i^\ell}[0] }_2 \sum_{t \ne 0} p^{|t|} \norm{\widehat{D_{i+1}^\ell}[t]}_2\\
\leq& L_p(D_1^\ell \cdots D_i^\ell) \cdot L_p(D_{i+1}^\ell) + L_p(D_1^\ell \cdots D_i^\ell) \lambda(D_{i+1}^\ell) \\&+ \norm{\widehat{D_1^\ell \cdots D_i^\ell}[0] }_2 \cdot L_p(D_{i+1}^\ell)\\
\leq& L_p(D_1^\ell \cdots D_i^\ell) \cdot 1 + \sqrt{2} L_p(D_{i+1}^\ell)\\
\leq& 2i+2.
\end{align*}
The second inequality follows from Equation \ref{eqn:mix-coeff} and the third from Lemma \ref{LemmaLL}.
Thus, we have that
$L^k(D^\ell)\leq p^{-k} L_p(D^\ell) \leq 2n \cdot (6000(\ell+1))^k,$ as required
\end{proof}
Now we turn our attention to Lemma \ref{LemmaLL}. We split into two cases: If $\lambda(D)$ is far from $1$ i.e. $\lambda(D) \leq 0.99$, then we need only ensure $L_p(D) \leq 1/100$. This is the `easy case' which proceeds much like the analysis of regular branching programs \cite{ReingoldSteinkeVadhan2013}.
If $\lambda(D)=1$, then $D$ is trivial -- i.e. $L_p(D)=0$ -- and we are also done. The `hard case' is when $\lambda(D)$ is very close to $1$. i.e. $0.99 \leq \lambda(D) < 1$.
\subsubsection*{Easy Case -- Good Mixing}
We consider the case where $\lambda(D) < 0.99$. We use the following result as a black box.
\begin{lemma}[{\cite[Lemma 4]{BRRY}, \cite[Lemma 3.1]{ReingoldSteinkeVadhan2013}}] \label{LemmaBRRY}
Let $B$ be a length-$n$, width-$w$, ordered, regular branching program. Then $$\sum_{1 \leq i \leq n} \norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}] }_2 \leq 2w^2.$$
\end{lemma}
The quantity $\norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}] }_2$ measures the correlation between the $i^\text{th}$ input bit and the final state of the program, which we call the \textbf{weight} of bit $i$. The entry in the $u^\text{th}$ row and $v^\text{th}$ column of $2\widehat{B_{i \cdots n}}[1 \circ 0^{n-i}] $ is the the probability of reaching vertex $v$ in layer $n$ given that we reached vertex $u$ in layer $i-1$ and the $i^\text{th}$ input bit is $0$ minus the same probability given that the $i^\text{th}$ input bit is $1$. Braverman et al.~\cite{BRRY} proved this result for a different measure of weight. Their result was translated into the above Fourier-analytic form by Reingold et al.~\cite{ReingoldSteinkeVadhan2013}.
We can add some non-regular layers to get the following.
\begin{lemma} \label{LemmaBRRYnonregular}
Let $B$ be a length-$n$, width-$w$, ordered branching program with at most $k$ non-regular layers. Then $$\sum_{1 \leq i \leq n} \norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}] }_2 \leq (2w^2 + 1)\sqrt{w} (k+1).$$
\end{lemma}
\begin{proof}
The proof proceeds by induction on $k$. If $k=0$, the result follows from Lemma \ref{LemmaBRRY}. Suppose the result holds for some $k$ and let $B$ be a length-$n$, width-$w$ ordered branching program with $k+1$ non-regular layers. Let $i^*$ be the index of the first non-regular layer. Then
\begin{align*}
\sum_{1 \leq i \leq n} \norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}] }_2
=& \sum_{1 \leq i < i^*} \norm{ \widehat{B_{i \cdots (i^*-1)}}[1 \circ 0^{i^*-i-1}] \cdot \widehat{B_{i^* \cdots n}}[0] }_2\\
&+ \norm{ \widehat{B_{i^* \cdots n}}[1 \circ 0^{n-i^*}] }_2\\
&+\sum_{i^* < i \leq n} \norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}] }_2\\
\leq& \left( \sum_{1 \leq i < i^*} \norm{ \widehat{B_{i \cdots (i^*-1)}}[1 \circ 0^{i^*-i-1}]}_2 \right) \cdot \norm{\widehat{B_{i^* \cdots n}}[0] }_2 \\&+ \sqrt{w} + (2w^2 + 1)\sqrt{w} (k+1)\\
\leq& 2 w^2 \cdot \sqrt{w} + \sqrt{w} + (2w^2 + 1)\sqrt{w} (k+1)\\
\leq& (2w^2 + 1)\sqrt{w} (k+2),\\
\end{align*}
where we use the fact that $\norm{\widehat{B}[s]}_2 \leq \sqrt{w}$ for any $s$ and width-$w$ branching program $B$.
\end{proof}
This gives us the following bound on the Fourier mass.
\begin{theorem
Let $B$ be a length-$n$, width-$w$, orderd branching program with at most $k$ non-regular layers. Then, for all $k' \in [n]$, $$L^{k'}(B) := \sum_{s \in \{0,1\}^n : |s|=k} \norm{\widehat{B}[s]}_2 \leq \sqrt{w} \cdot ((2w^2 + 1)\sqrt{w} (k+1))^{k'} \leq \sqrt{w} \cdot (3w^{2.5}(k+1))^{k'}.$$
\end{theorem}
This result is proved using Lemma \ref{LemmaBRRYnonregular} analogously to how \cite[Theorem 3.2]{ReingoldSteinkeVadhan2013} is proved using Lemma \ref{LemmaBRRY}.
\begin{proof}
We perform an induction on $k'$. If $k'=0$, then there is only one Fourier coefficient to bound---namely, $\widehat{B}[0^n] = \ex{U}{B[U]}$. Since $\ex{U}{B[U]}$ is stochastic, $\norm{\ex{U}{B[U]}}_2 \leq \sqrt{w}$ and the base case follows. Now suppose the bound holds for $k'$ and consider $k'+1$. We split the Fourier coefficients based on where the last $1$ is:
\begin{align*}
\lefteqn{\sum_{s \in \{0,1\}^n : |s|=k'+1} \norm{\widehat{B}[s]}_2}~~~~~~~~~~&\\ =& \sum_{1 \leq i \leq n} \sum_{s \in \{0,1\}^{i-1} : |s|=k'} \norm{\widehat{B}[s \circ 1 \circ 0^{n-i}]}_2\\
=& \sum_{1 \leq i \leq n} \sum_{s \in \{0,1\}^{i-1} : |s|=k'} \norm{\widehat{B_{1 \cdots i-1}}[s] \cdot \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}]}_2~~~\text{(by Lemma \ref{LemmaFourier} (Decomposition))}\\
\leq& \sum_{1 \leq i \leq n} \sum_{s \in \{0,1\}^{i-1} : |s|=k'} \norm{\widehat{B_{1 \cdots i-1}}[s]}_2 \cdot \norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}]}_2\\
\leq& \sum_{1 \leq i \leq n} \sqrt{w} \cdot ((2w^2 + 1)\sqrt{w} (k+1))^{k'} \cdot \norm{ \widehat{B_{i \cdots n}}[1 \circ 0^{n-i}]}_2~~~\text{(by the induction hypothesis)}\\
\leq& \sqrt{w} \cdot ((2w^2 + 1)\sqrt{w} (k+1))^{k'} \cdot (2w^2 + 1)\sqrt{w} (k+1) ~~~\text{(by Lemma \ref{LemmaBRRYnonregular})}\\
=&\sqrt{w} \cdot ((2w^2 + 1)\sqrt{w} (k+1))^{k'+1},\\
\end{align*}
as required.
\end{proof}
\begin{lemma} \label{LemmaEasy}
Let $D$ be a 3OBP with at most $k$ non-regular layers. If $p \leq 1/6000 (k+1)$, then $L_p(D) \leq 1/100$.
\end{lemma}
\begin{proof}
We have
\begin{align*}
L_p(D) =& \sum_{k' \geq 1} p^{k'} L^{k'}(D)\\
\leq& \sum_{k' \geq 1} p^{k'} \sqrt{3} ((2\cdot 3^2 + 1)\sqrt{3} (k+1))^{k'}\\
\leq& \sqrt{3} \sum_{k' \geq 1} \left( \frac{19 \sqrt{3} (k+1)}{6000 (k+1)} \right)^{k'}\\
\leq& 1/100.\\
\end{align*}
\end{proof}
It immediately follows that $\lambda(D) + L_p(D) \leq 1$ when $p \leq 1/6000 (k+1)$, assuming $\lambda(D) < 0.99$. This covers the `easy' case of Lemma \ref{LemmaLL}.
\subsubsection*{Hard Case -- Poor Mixing}
Now we consider the case where $\lambda(D) \in [0.99, 1]$.
\begin{lemma} \label{LemmaHard}
Let $D$ be a 3OBP with at most $k$ non-regular layers where the first and last layers of vertices have width 2. Suppose $\lambda(D) \in [0.99, 1]$. If $p\leq 1/(24k+12)$, then $L_p(D) + \lambda(D) \leq 1$.
\end{lemma}
This covers the `hard' case of Lemma \ref{LemmaLL} and, along with Lemma \ref{LemmaEasy} completes the proof of Lemma \ref{LemmaLL}.
Since $D$ has width 2 in the first and last layers, we view $D[x]$ as a $2 \times 2$ matrix. We can write the expectation (which is stochastic) as $$\ex{U}{D[U]} = \left( \begin{array}{cc} 1-\alpha & \alpha \\ \beta & 1-\beta \end{array} \right).$$ We can assume (by permuting rows and columns) that $\lambda(D) = 1 - \alpha - \beta$ and $\alpha, \beta \in [0,1/100]$.
Now write $$D[x] = \left( \begin{array}{cc} 1-f(x) & f(x) \\ g(x) & 1-g(x) \end{array} \right),$$ where $f,g : \{0,1\}^d \to \{0,1\}$. Then $\alpha = \ex{U}{f(U)}$ and $\beta = \ex{U}{g(U)}$. We can view $D$ has having two \emph{corresponding} start and end states. The probability that, starting in the first start state, we end in the first end state is $1-\alpha \geq 0.99$. Likewise, the probability that, starting in the second start state, we end in the second end state is $1-\beta \geq 0.99$. The function $f$ is computed by starting in the first start state and accepting if we end in the second end state -- that is, we ``cross over''. Likewise, $g$ computes the function telling us whether we will cross over from the second start state to the first end state. Intuitively, there is a low ($1/100$) probability of crossing over, so the program behaves like two disjoint programs.
We will show that $L_p(f) \leq (12k+6) p \alpha$ and $L_p(g) \leq (12k+6) p \beta$ for $p \leq 1/(6k+3)$, from which the result follows by choosing $p$ such that $L_p(f) \leq \alpha/2$ and $L_p(g) \leq \beta/2$.
The plan is as follows.
\begin{itemize}
\item[1.] Show that we can partition the vertices of $D$ into two sets with $O(k)$ edges crossing between the sets such that each layer has at least one vertex in each set. Intuitively, this partitions $D$ into two width-2 branching programs with a few edges going between them.
\item[2.] Using this partition, show that we can write $f(x) = \sum_s \prod_j f_{s,j}(x_j)$, where each $f_{s,j}$ is a $\{0,1\}$-valued function computed by a \emph{regular} width-2 branching program, the product is over $O(k)$ terms and the $x_j$s are a partition of $x$.
\item[3.] Let $f_s(x) = \prod_j f_{s,j}(x_j)$ and $\alpha_s = \ex{U}{f_s(U)}$. Show that $L_p(f_s) \leq (12k+6) p \alpha_s$ for $p \leq 1/(6k+3)$. Then $$L_p(f) \leq \sum_s L_p(f_s) \leq \sum_s (12k+6) p \alpha_s \leq (12k+6) p \alpha,$$ as required.
\end{itemize}
The same holds for $g$, which gives the result.
\subsubsection*{Step 1.}
\begin{lemma} \label{LemmaPartition}
Let $D$ be a 3OBP with at most $k$ non-regular layers and width-2 in the first and last layers of vertices. Suppose $\lambda(D) \in [0.99,1]$. Then there is a partition of the vertices of $D$ such that each layer has at least one vertex in each side of the partition and there are at most $2k+1$ layers with an edge that crosses the partition.
\end{lemma}
\begin{proof}
We assign each vertex of $D$ a \textbf{charge}: The two vertices $v_+$ and $v_-$ in the first layer are assigned charges $1$ and $-1$ respectively. Each edge is assigned a charge that is half the charge of the vertex it originates from and the charge of each subsequent vertex is the sum of the charges of the incoming edges. In other words, the charge of a vertex $u$ is the probabililty that a random walk from $v_+$ reaches $u$ minus the probability that a random walk from $v_-$ reaches $u$.
The partition is given by the sign of the charge: Let $Q$ be the set of vertices with positive or zero charge and let $\overline{Q}$ be the set of vertices with negative charge. Now we must prove that there are $O(k)$ edges crossing between $Q$ and $\overline{Q}$.
Define the \textbf{total charge} of a layer to be the sum of the absolute values of the charges in that layer. Clearly the total charge cannot increase from one layer to the next (by the triangle inequality). Moreover, the total charge in the final layer equals $((1-\alpha)-\alpha) + \left( (1-\beta)-\beta \right)=2\lambda(D)$.
By assumption ($\lambda(D) \geq 0.99$) the total charge decreases by at most $1/50$. The total charge only decreases when an edge crosses the $(Q,\overline{Q})$ partition, as this is when positive and negative charges cancel. In fact, it decreases by precisely the charge of the crossing edge.
So there is very little charge crossing the partition. However, it is possible that many edges with little charge cross the partition. To preclude this possibility, we also track the \textbf{minimum charge} of each layer, which is the minimum absolute value of a charge of a vertex in that layer.
Now we use the fact that there are at most $k$ non-regular layers in $D$. Call a layer a \textbf{crossing layer} if it contains an edge that crosses the partition i.e. where the signs of the charges of the endpoints of the edge are different. Clearly there are at most $k$ non-regular crossing layers. We need only account for regular crossing layers.
Consider a regular crossing layer. We will argue that the minimum charge must go from `small' to `large'. Then we will argue that in order for the minimum charge to go from large back to small, a non-regular layer is needed. So each such regular crossing layer must have a corresponding non-regular layer. This ensures that there are at most $k+1$ regular crossing layers, as required.
Let $B_i$ be a regular crossing layer. Let $a \leq b \leq c$ be the charges on the left vertices of the layer. Since $a+b+c=0$ and $|a|+|b|+|c| \geq 1.98$, we have that $a \leq -0.49$ and $c \geq 0.49$. The vertices corresponding to $a$ and $c$ cannot have a common neighbour, as otherwise the total charge would decrease by at least $0.2$. Up to permuting vertices, this leaves three possibilities for the layer, which we depict in Figure \ref{fig:charge}.
\begin{figure}[h]
\centering
\includegraphics[scale=.5]{threepossibilities.eps}
\caption{The bold arrows indicate double edges.}
\label{fig:charge}
\end{figure}
Possibility (iii) does not have a crossing, so can be ignored. Possibilities (i) and (ii) are essentially the same up to flipping signs. So let's analyse possibility (i).
For there to be a crossing, we must have $b<0$. The total charge then decreases by $|b|$. So $|b| \leq 1/50$. Now $|b|$ is the minimum charge of the vertices on the left. So the minimum charge on the left of a regular crossing layer is at most $1/50$. On the right, the minimum charge is $\min\{|a|,|b+c|/2\} \geq \min\{0.49, (0.49-1/50)/2\} >1/5$. So this layer increases the minimum charge by at least $1/5-1/50 > 0.1$.
Now we will show that any two regular crossing layers must have a non-regular layer between them. For the sake of contradiction, let $B_i$ and $B_j$ ($i<j$) be two regular crossing layers with no non-regular layers between them. We can assume that there are no crossing layers between $B_i$ and $B_j$: If not, replace $B_j$ with the first crossing layer after $B_i$.
Now consider the vertices in $B_{i+1}$, call them $v_1,v_2$ and $3$. We assume without loss of generality that one vertex has positive charge (say $v_1$), while $v_2$ and $v_3$ have negative charge. Because there are no crossing layers, no path from $v_1$ can share a vertex with any path starting from $v_2$ or $v_3$. Thus, up to permutation on the vertices (determining which vertex is ``isolated''), the first column and first row of the matrix $\ex{U}{B_{i+1,\ldots,j-1[U]}}$ are equal to $(1,0,0)$. Because every layer of
$B_{i+1,\ldots,j-1}$ must be regular,
up to permuting vertices, we have that $\ex{U}{B_{i+1 \cdots j-1}[U]}$ is of the form $$\left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & 1 \end{array} \right) ~~~~~~\text{or}~~~~~~ \left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & 1/2 & 1/2 \\ 0 & 1/2 & 1/2 \end{array} \right).$$ The first possibility cannot decrease the minimum charge at all. The second possibility can only decrease the minimum charge by cancellation: Let $a$, $b$, and $c$ be the charges on the left. Then the charges on the right are $a$, $(b+c)/2$, and $(b+c)/2$. The only way the minimum charge can decrease is if $b$ and $c$ have opposite signs. By symmetry, we can assume that $c \geq -b \geq 0$. Thus the new minimum charge is either $|a|$ or $(c-|b|)/2 = (c+|b|)/2-|b|$. So the minimum charge decreases by at most $|b|$. However, the total charge decreases by $|b|+|c|-|b+c| = c+|b| - (c-|b|) = 2|b|$. Since the total charge can decrease by at most $1/50$, we have $|b|\leq1/100$ and the minimum charge can decrease by at most $1/100$ -- a contradiction, as it must decrease from at least $1/5$ to at most $1/50$.
Thus we have shown that each pair of regular crossing layers has a non-regular layer between then. Since there are at most $k$ non-regular layers, there are at most $2k+1$ crossing layers, as required.
\end{proof}
\subsubsection*{Step 2.}
\begin{lemma} \label{LemmaSumProduct}
Let $D$ be a length-$d$ 3OBP with at most $k$ non-regular layers and width-2 in the first and last layers of vertices. Suppose $\lambda(D) \geq 0.99$. If $f : \{0,1\}^n \to \{0,1\}$ is the function computed by $D$, then we can write $f(x) = \sum_s \prod_j f_{s,j}(x_j)$, where each $f_{s,j}$ is computed by a regular width-2 ordered branching program and the $x_j$'s are a partition of $x$ into at most $6k+3$ parts.
\end{lemma}
\begin{proof}
Call a layer of edges of $D$ \textbf{critical} if it is either non-regular or it has an edge crossing the partition given by Lemma \ref{LemmaPartition}. Let $\Gamma$ be the set of critical layers. By Lemma \ref{LemmaPartition}, $D$ has at most $3k+1$ critical layers. Between critical layers, $D$ is partitioned into two width-2 regular branching programs. (To be more precise, it is partitioned into a width-2 regular branching program and a width-1 regular branching program.)
Define $\tilde{\Gamma}$ to be the set of `fixings' of edges in $\Gamma$, that is, $s \in \tilde{\Gamma}$ specifies for each $i \in \Gamma$ an edge $s(i)$ in layer $i$ (specifiing one of three states to the left of layer $i$ and the label, which is 0 or 1, of the edge taken). We can think of $s \in \tilde{\Gamma}$ as a function $s : \Gamma \to [3]\times \{0,1\}$.
For $s \in \tilde{\Gamma}$, define $f_s : \{0,1\}^d \to \{0,1\}$ to be the following indicator function: $$f_s(x)=1 \iff f(x)=1 \wedge \text{the path in $D$ given by $x$ uses all the edges in $s$}.$$
Clearly $f(x) = \sum_{s \in \tilde{\Gamma}} f_s(x)$, as each path in $D$ is consistent with exactly one $s$.
Now we claim that each $f_s$ can be written as the conjunction of at most $2|\Gamma|+1$ regular width-2 branching programs. In particular, there is one term for each critical layer and one term for each gap between critical layers.
Let $i_1 < i_2 < \cdots < i_{|\Gamma|}$ be an enumeration of $\Gamma$. (Also define $i_0=0$ and $i_{|\Gamma|+1}=d+1$.) We will write $$f_s(x) = \prod_{j=1}^{2|\Gamma|+1} f_{s,j}(x_j),$$ where the $x_j$s are a partition of $x$ as follows. For $j \in [|\Gamma|+1]$, $x_{2j-1} \in \{0,1\}^{i_j-i_{j-1}-1}$ contains the coordinates from $i_{j-1}+1$ to $i_j-1$ of $x$. For $j \in [|\Gamma|]$, $x_{2j} \in \{0,1\}$ is coordinate $i_j$ of $x$. The function $f_{s,j}$ checks the bits in $x_j$ and is 1 if and only if the path is consistent with $f_s=1$ (assuming the bits outside of $x_j$ are set consistently with $f_s=1$. Thus $f_{s,2j-1}(x_{2j-1})$ verifies that if started in the state $s(i_{j-1})_1$, the input $x_{2j-1}$ leads $D$ to state $s(i_{j})_1$ in layer $i_{j},$
and $f_{s,2j}(x_{2j})$ verifies that $x_{2j} = s(i_{2j})_2$, i.e., that the setting of $x_{2j}$ is the same as the label specified by $s(i_{2j})$. Note that for $f_{s,j}$ to be satisfied, there is only one correct vertex at each end of the path. The functions $f_{s,2j}$ are determined by a single literal and can be computed by width-$2$ branching programs, and since the functions $f_{s,2j-1}$ are computed over non-critical layers from a single starting vertex, they are also computed by width-2.
\end{proof}
\subsubsection*{Step 3.}
We use the following fact about regular width-2 ordered branching programs -- a very simple class of functions.
\begin{lemma} \label{LemmaWidth2}
Let $f : \{0,1\}^n \to \{0,1\}$ be computed by a width-2 regular ordered branching program. Then $\ex{U}{f(U)} \in \{0,1/2,1\}$ and $L_p(f) \leq p/2$ for all $p \in [0,1]$.
\end{lemma}
\begin{proof}
Every layer of a regular width-2 ordered branching program falls into one of three cases: trivial layers (the input bit does not affect the state), a (negated) XOR (flips the state depending on the input bit), or a (negated) dictator (sets the state based on the current input bit regardless of the previous state). Thus any such branching program is either a constant function, which gives $\ex{U}{f(U)}\in \{0,1\}$ and $L_p(f)=0$, or is a (possibly negated) XOR of a subset of the input bits. In the latter case $f$ has
one non-trivial coefficient of magnitude $1/2$, which implies $\ex{U}{f(U)}=1/2$ and $L_p(f) \leq p \cdot L(f) \leq p/2$.
\end{proof}
\begin{lemma} \label{LemmaDecomp}
Let $f :\{0,1\}^n \to \{0,1\}$ be of the form $f(x) = \prod_{j \in [k]} f_j(x_j)$, where the $x_j$s are a partition of $x$ and each $f_j$ is computed by a width-2 ordered regular branching program. Then $L_p(f) \leq 2 k p \cdot \ex{U}{f(U)}$ for any $p \leq 1/k$.
\end{lemma}
\begin{proof}
Define $\alpha_j = \ex{U}{f_j(U)}$. We have $\alpha = \prod_{j \in [k]} \alpha_j$. Now
\begin{align*}
\alpha + L_p(f)
=& \sum_s p^{|s|} |\widehat{f}[s]|\\
=& \sum_s p^{|s|} \prod_{j \in [k]} |\widehat{f_j}[s_j]|\\
=& \prod_{j \in [k]} \sum_{s_j} p^{|s_j|} |\widehat{f_{j}}[s_j]|\\
=& \prod_{j \in [k]} \left( \alpha_{j} + L_p(f_{j}) \right).
\end{align*}
Since $f_{j}$ is computed by a width-2 regular branching program, $\alpha_{j} \in \{0,1/2,1\}$. If $\alpha_{j}=0$, then $\alpha = 0$ and $L_p(f)=0$, so we can ignore this case. Moreover, if $\alpha_{j} =1$, then $f_{j}=1$ is constant and $L_p(f_{j})=0$, so we can ignore the terms with $\alpha_{j} = 1$. Let $J = \{j \in [k] : \alpha_{j}=1/2\}$. Thus we are left with
\begin{align*}
L_p(f) =& \prod_{j \in J} \left( \frac{1}{2} + L_p(f_{j}) \right) - 2^{-|J|}\\
=& \alpha\cdot \left( \prod_{j \in J} \left( 1 + 2 L_p(f_{j}) \right) - 1 \right)\\
\leq& \alpha \cdot \left( \prod_{j \in J} \left( 1 + p \right) - 1 \right)\\
\leq& \alpha \cdot \left( (e^{p})^{|J|} -1 \right)\\
\leq& \alpha \cdot 2 p |J|\\
\leq& \alpha 2 p k.\\
\end{align*}
as long as $p|J| \leq 1$.
\end{proof}
\begin{proof}[Proof of Lemma \ref{LemmaHard}]
By Lemma \ref{LemmaSumProduct}, we can write $f(x) = \sum_s \prod_j f_{s,j}(x_j)$. Let $f_s(x) = \prod_j f_{s,j}(x_j)$, where the product is over at most $6k+3$ terms. Then, by Lemma \ref{LemmaDecomp}, $$L_p(f) \leq \sum_s L_p(f_s) \leq \sum_s (12k+6)p \cdot \ex{U}{f_s(U)} = (12k+6)p \alpha,$$ as long as $p \leq 1/(6k+3)$. Likewise $L_p(g) \leq (12k+6)p \beta$ for $p \leq 1/(6k+3)$.
Now $L_p(D) \leq 2L_p(f)+2L_p(g) \leq (24k+12)\cdot p\cdot (\alpha+\beta)$. If $p\leq 1/(24k+12)$, then $L_p(D) + \lambda(D) \leq \alpha + \beta + 1 - \alpha - \beta = 1$, as required.
\end{proof}
\subsection{Bootstrapping} \label{SubSectionBootstrapping}
Proposition \ref{PropositionLow} gives a bound on the Fourier growth of width-3 branching programs of the form $L^k(B) \leq n^2 \cdot O(k)^k$. The $k^k$ term is inconvenient, but can be easily removed by ``bootstrapping'':
The following proposition shows that if we can bound the Fourier mass up to level $O(\log n)$, then we can bound the Fourier mass at all levels.
\begin{proposition} \label{PropositionBootstrapping}
Let $B$ be a length-$n$, ordered branching program such that, for all $i,j,k \in [n]$ with $k \leq 2k^*$ and $i \leq j$, we have $L^k(B_{i \cdots j}) \leq a \cdot b^k$. Suppose $a \cdot n \leq 2^{k^*}$. Then, for all $i,j,k \in [n]$ with $i \leq j$, we have $L^k(B_{i \cdots j}) \leq a \cdot (2b)^k$.
\end{proposition}
The proof is similar to that of Lemma \ref{LemmaWellOrder}.
\begin{proof}
Suppose the proposition is false and fix the smallest $k$ such that the statement does not hold. Clearly $k>2k^*$. Let $k'=k-k^*$. By minimality $L^{k'}(B_{i \cdots j}) \leq a \cdot (2b)^{k'}$ for all $i \leq j$. Now $$L^k(B_{i \cdots j}) \leq \sum_{\ell=i}^{j+1} L^{k'}(B_{i \cdots \ell-1}) \cdot L^{k^*}(B_{\ell \cdots j}) \leq \sum_{\ell=i}^{j+1} a \cdot (2b)^{k'} \cdot a \cdot b^{k^*} \leq n \cdot a \cdot (2b)^{k'} \cdot a \cdot b^{k^*} = a \cdot (2b)^k \cdot \left( \frac{n a}{2^{k^*}} \right).$$ Since $ n a \leq 2^{k^*}$, we have a contradiction, as we assumed $L^{k^*}(B_{\ell \cdots j}) > a \cdot (2b)^k$.
\end{proof}
Now we combine Propositions \ref{PropositionLow} with \ref{PropositionBootstrapping} to prove Theorem \ref{TheoremLow}
\begin{proof}[Proof of Theorem \ref{TheoremLow}]
Let $B$ be a length-$n$ 3OBP computing $f$ with width 2 in the first and last layers.
By Proposition \ref{PropositionLow}, we have $$L^k(B) \leq 8 n^2 \cdot \left(200000 k \right)^k $$ for any length-$n$ 3OBP $B$ and $k \in [n]$. Since a subprogram of a 3OBP is also a 3OBP, this bound also applies to $L^k(B_{i \cdots j})$ for all $i,j \in [n]$. If we set $k^* = \lceil \log_2 (8n^3) \rceil$, $a = 8n^2$, and $b=200000 \cdot 2k^*$, then the hypotheses of Proposition \ref{PropositionBootstrapping} are satisfied. Thus, for any $k \in [n]$, we have $$L^k(B) \leq 8n^2 \cdot \left( 2 \cdot 200000 \cdot 2 k^* \right)^k.$$ Since $L^k(f) \leq L^k(B)$, this gives the result.
\end{proof}
\section{Introduction} \label{SectionIntroduction}
\subsection{Pseudorandom Generators for Space-Bounded Computation}
A major open problem in the theory of pseudorandomness is to construct an ``optimal'' pseudorandom generator for space-bounded computation. That is,
we want an
explicit algorithm that stretches a uniformly random seed
of length $O(\log n)$ to $n$ bits that cannot be distinguished from uniform by any $O(\log n)$-space algorithm (which receives the pseudorandom bits one at a time, in a streaming fashion, and may be nonuniform).
Such a generator would imply that every randomized algorithm can be derandomized with only a constant-factor increase in space ($\textrm{RL}=\textrm{L}$), and would also have a variety of other applications, such as
in streaming
algorithms~\cite{Indyk},
deterministic dimension reduction and SDP rounding~\cite{Sivakumar,EngebretsonInOd01}, hashing~\cite{CReingoldSW}, hardness
amplification~\cite{HealyVaVi}, almost $k$-wise independent
permutations~\cite{KaplanNaRe}, and cryptographic pseudorandom generator constructions~\cite{HaitnerHaRe06}.
To construct a pseudorandom generator for space-bounded algorithms using space $s$, it suffices to construct a generator that is pseudorandom against ordered branching programs of width $2^s$.
A branching program\footnote{In this work and the definition we give here, we consider read-once, oblivious branching programs, and refer to them simply as branching programs for brevity.} $B$ is a non-uniform model of space-bounded computation that reads one input bit at a time, maintaining a state in $[w] = \{1,\dots,w\}$, where $w$ is called the width of $B$. At each time step $i=1,\ldots,n,$ $B$ can read a different input bit $x_{\pi(i)}$ (for some permutation $\pi$) and uses a different state transition function $T_i:[w]\times \{0,1\}\to [w]$. It is often useful to think of a branching program as a directed acyclic graph consisting of $n+1$ layers of $w$ vertices each, where the $i^\text{th}$ layer corresponds to the state at time $i$. The transition function defines a bipartite graph between consecutive layers, where we connect state $s$ in layer $i-1$ to states $T_i(s,0)$ and $T_i(s,1)$ in layer $i$ (labeling those edges 0 and 1, respectively).
Most previous constructions of pseudorandom generators for space-bounded computations consider \textit{ordered} branching programs, where the input bits are read in order -- that is, $\pi(i)=i$
The classic work of Nisan~\cite{Nisan} gave a generator with seed length $O(\log ^2 n)$ that is pseudorandom against ordered branching programs of polynomial width. Despite intensive study, this is the best known seed length for ordered branching programs even of width 3, but a variety of works have shown improvements for restricted classes such as
branching programs of width 2 \cite{SZ,BogdanovDvVeYe09}, and regular or permutation branching programs (of constant width)~\cite{BRRY,BrodyVerbin,KNP,De,Steinke12}. For width 3, hitting set generators (a relaxation of pseudorandom generators) have been constructed \cite{SimaZak, GMRTV}.
The vast majority of these works are based on Nisan's original generator or its variants by Impagliazzo, Nisan, and Wigderson~\cite{INW} and Nisan and Zuckerman~\cite{NZ}, and adhere to a paradigm that seems unlikely to yield generators against general logspace computations with seed length better than $\log^{1.99} n$ (see \cite{BrodyVerbin}).
All known analyses of Nisan's generator and its variants rely on the order in which the output bits are fed to the branching program (given by the permutation $\pi$). The search for new ideas leads us to ask: Can we construct a pseudorandom generator whose analysis does not depend on the order in which the bits are read? A recent line of work \cite{BPW,IMZ,ReingoldSteinkeVadhan2013} has constructed pseudorandom generators for unordered branching programs (where the bits are fed to the branching program in an arbitrary, fixed order); however, none match both the seed length and generality of Nisan's result.
For
unordered branching programs of length $n$ and width $w$, Impagliazzo, Meka, and Zuckerman \cite{IMZ} give
seed length $O((nw)^{1/2+o(1)})$
improving on the linear seed length $(1-\Omega(1))\cdot n$ of Bogdanov, Papakonstantinou, and Wan \cite{BPW}.\footnote{A generator with seed length $\tilde{O}(\sqrt{n} \log w)$ is given in \cite{ReingoldSteinkeVadhan2013}. The generator in \cite{IMZ} also extends to branching programs that read their inputs more than once and in an adaptively chosen order, which is more general than the model we consider. } Reingold, Steinke, and Vadhan \cite{ReingoldSteinkeVadhan2013} achieve seed length $O(w^2 \log^2 n)$ for the restricted class of \textit{permutation} branching programs, in which $T_i(\cdot, b)$ is a permutation on $[w]$ for all $i \in [n]$ and $b \in \{0,1\}$.
Recently, a new approach for constructing pseudorandom generators has been suggested in the work of Gopalan et al.~\cite{GMRTV}; they constructed pseudorandom generators for read-once CNF formulas and combinatorial rectangles, and hitting set generators for width-3 branching programs, all having seed length $\tilde{O}(\log n)$ (even for polynomially small error). Their basic generator (e.g. for read-once CNF formulas) works by pseudorandomly partitioning the bits into several groups and assigning the bits in each group using a small-bias generator~\cite{NaorNa93}.
A key insight in their analysis is that the small-bias generator only needs to fool the function ``on average,'' where the average is taken over the possible assignments to subsequent groups, which is a weaker
requirement than fooling the original function or even a random restriction of the original function.
(For a more precise explanation, see Section \ref{SubSectionPRrestriction}.)
The analysis of Gopalan et al.~\cite{GMRTV} does not rely on the order in which the output bits are read, and
the previously mentioned work by Reingold, Steinke, and Vadhan \cite{ReingoldSteinkeVadhan2013} uses Fourier analysis of branching programs to show that the generator of Gopalan et al.~fools unordered permutation branching programs.
In this work we further develop Fourier analysis of branching programs and show that the pseudorandom generator of Gopalan et al.~with seed length $\tilde{O}(\log^6 n)$ fools width-3 branching programs:
\begin{theorem}[Main Result] \label{thm:main-intro}
There is an explicit pseudorandom generator $G : \{0,1\}^{O(\log^3 n \cdot \log \log n)} \rightarrow \{0,1\}^n$ fooling oblivious, read-once (but unordered), branching programs of width $3$ and length $n$.
\end{theorem}
The previous best seed length for this model is the aforementioned length of $O(n^{1/2+o(1)})$ given in \cite{IMZ}.
The construction of the generator in Theorem \ref{thm:main-intro} is essentially the same as the generator of Gopalan et al.~\cite{GMRTV} for read-once CNF formulas, which was used by Reingold et al.~\cite{ReingoldSteinkeVadhan2013} for permutation branching programs. In our analysis, we give a new bound on the Fourier mass of
width-3 branching programs.
\subsection{Fourier Growth of Branching Programs}
For a function $f : \{0,1\}^n\rightarrow \mathbb{R}$, let
$\widehat{f}[s] = \ex{U}{f[U] \cdot (-1)^{s\cdot U}}$ be the standard Fourier transform over $\mathbb{Z}_2^n$, where $U$ is a random variable distributed uniformly over $\{0,1\}^n$ and $s\subseteq [n]$ or, equivalently, $s \in \{0,1\}^n$. The Fourier mass of $f$ (also called the spectral norm of $f$),
defined as $L(f):=\sum_{s\neq \emptyset} |\hat{f}[s]|$, is a fundamental measure of complexity for Boolean functions (e.g., see \cite{green2008boolean}), and its study has applications to learning theory \cite{kushilevitz1993learning,mansour1995nlog}, communication complexity \cite{grolmusz1997power,ada2012spectral,tsang2013fourier,shpilka2014structure}, and circuit complexity \cite{brandman1990spectral,Bruck90harmonicanalysis,bruck1992polynomial}. In the study of pseudorandomness, it is well-known that small-bias generators\footnote{A small-bias generator with bias $\mu$ outputs a random variable $X\in \{0,1\}^n$ such that $\left|\ex{X}{(-1)^{s\cdot X}}\right|\leq \mu$ for every $s\subset [n]$ with $s \ne \emptyset$.} with bias $\varepsilon/L$ (which can be sampled using a seed of length $O(\log (n \cdot L/\varepsilon))$ \cite{NaorNa93,AGHP}) will $\varepsilon$-fool any function whose Fourier mass is at most $L$.
Width-2 branching programs have Fourier mass at most $O(n)$
\cite{BogdanovDvVeYe09,SZ} and are thus fooled by small-bias generators with bias $\varepsilon/n$. Unfortunately, such a bound does not hold even for very simple width-3 programs. For example, the `mod 3 function,' which indicates when the hamming weight of its input is a multiple of 3 has Fourier mass exponential in $n$.
However, a more refined measure of Fourier mass is possible and often useful: Let $L^k(f) = \sum_{|s|=k} |\hat{f}[s]|$ be the level-$k$ Fourier mass of $f$. A bound on the Fourier growth of $f$, or the rate at which $L^k(f)$ grows with $k$, was used by Mansour \cite{mansour1995nlog} to obtain an improved query algorithm for polynomial-size DNF; the junta approximation results of Friedgut \cite{friedgut1998boolean} and Bourgain \cite{bourgain2002distribution} are proven using approximating functions that have slow Fourier growth.
This notion turns out to be useful in the analysis of pseudorandom generators as well: Reingold et al.~\cite{ReingoldSteinkeVadhan2013} show that the generator of Gopalan et al.~\cite{GMRTV} will work if there is a good bound on the Fourier mass of low-order coefficients.
More precisely, they show that for any class $\mathcal{C}$ of functions computed by branching programs that is closed under restrictions and decompositions and satisfies $L^k(f) \leq \mathrm{poly}(n) \cdot c^k$ for every $k$ and $f \in \mathcal{C}$,
there is a pseudorandom generator with seed length $\tilde{O}(c\cdot \log^2 n)$ that fools every $f\in \mathcal{C}$.
They then bound the Fourier growth of permutation branching programs (and the even more general model of ``regular'' branching programs, where each layer is a regular bipartite graph) to obtain a pseudorandom generator for permutation branching programs:
\begin{theorem}[{\cite[Theorem 1.4]{ReingoldSteinkeVadhan2013}}]\label{thm:rsv}
Let $f :\{0,1\}^n \to \{0,1\}$ be computed by a length-$n$, width-$w$, read-once, oblivious, \emph{regular} branching program. Then, for all $k \in [n]$, $L^k(f) \leq (2w^2)^k$.
\end{theorem}
In particular, the mod 3 function over $O(k)$ bits, which is computed by a permutation branching program of width 3, has Fourier mass $2^{\Theta(k)}$ a level $k$.
However, the Tribes function,\footnote{The Tribes function (introduced by Ben-Or and Linial \cite{tribes}) is DNF formula where all the terms are the same size and every input appears exactly once. The size of the clauses in this case is chosen to give an asymptotically constant acceptance probability on uniform input.} which is also computed by a width-3 branching program, has Fourier mass $\Theta_k(\log^k n)$ at level $k$,
so the bound in Theorem \ref{thm:rsv} does not hold for non-regular branching programs even of width 3.
The Coin Theorem of Brody and Verbin \cite{BrodyVerbin} implies a related result: essentially, a function computed by a width-$w$, length-$n$ branching program cannot distinguish product distributions on $\{0,1\}^n$ any better than a function satisfying $L^k(f) \leq (\log n)^{O(wk)}$ for all $k$. To be more precise, if $X \in \{0,1\}^n$ is $n$ independent samples from a coin with bias $\beta$ (that is, each bit has expectation $(1+\beta)/2$), then $\ex{X}{f[X]} = \sum_s \widehat{f}[s] \beta^{|s|}$. If $L^k(f) \leq (\log n)^{O(wk)}$ for all $k$, then $$\left|\ex{X}{f[X]}-\ex{U}{f[U]}\right| = \left| \sum_{s \ne 0} \widehat{f}[s] \beta^{|s|} \right| \leq \sum_{k \in [n]} L^k(f) |\beta|^{|s|} \leq O(|\beta| (\log n)^{O(w)}),$$ assuming $|\beta| \leq 1/(\log n)^{O(w)}$. Brody and Verbin prove that, if $f$ is computed by a length-$n$, width-$w$ branching program, then $|\ex{X}{f[X]}-\ex{U}{f[U]}| \leq O(|\beta| (\log n)^{O(w)})$.
Since distinguishing product distributions captures much of the power of a class of functions, this leads to the following conjecture.
\begin{conjecture}[\cite{ReingoldSteinkeVadhan2013}]\label{conj:fouriergrowth}
For every constant $w$, the following holds. Let $f : \{0,1\}^n \to \{0,1\}$ be computed by a width-$w$, read-once, oblivious branching program. Then $$L^k(f) \leq n^{O(1)}\cdot (\log n )^{O(k)},$$ where the constants in the $O(\cdot)$ may depend on $w$.
\end{conjecture}
In this work, we prove this conjecture for $w=3$:
\begin{theorem}[Fourier Growth of Width 3] \label{MainThmIntro}
Let $f : \{0,1\}^n \to \{0,1\}$ be computed by a width-3, read-once, oblivious branching program. Then, for all $k \in [n]$, $$L^k(f) := \sum_{s : |s|=k} |\widehat{f}[s]| \leq n^2 \cdot (O(\log n))^k .$$
\end{theorem}
This bound is the main contribution of our work and, when combined with the techniques of Reingold et al.~\cite{ReingoldSteinkeVadhan2013}, implies our main result (Theorem \ref{thm:main-intro}).
The Tribes function of \cite{tribes} shows that the base of $O(\log n)$ of the Fourier growth in Theorem \ref{MainThmIntro} is tight up to a factor of $\log \log n$. (See Appendix \ref{AppendixOptimal}.)
We also prove Conjecture \ref{conj:fouriergrowth} with $k=1$ for any constant width $w$:
\begin{theorem} \label{TheoremFirstOrder}
Let $f : \{0,1\}^n \to \{0,1\}$ be computed by a width-$w$, length-$n$, read-once, oblivious branching program. Then $$L^1(f) = \sum_{i \in [n]} |\widehat{f}[\{i\}]| \leq O(\log n)^{w-2}.$$
\end{theorem}
The proof is left to Appendix \ref{AppendixFirstOrder}.
\subsection{Techniques}
The intuition behind our approach begins with two extreme cases of width-3 branching programs: permutation branching programs and branching programs in which every layer is a non-permutation layer.
Permutation branching programs ``mix'' well: on a uniform random input, the distribution over states gets closer to uniform (in $\ell_2$ distance) in each layer.
We can use this fact with an inductive argument to achieve a bound of $2^{O(k)}$ on the level-$k$ Fourier mass (this is the bound of Theorem \ref{thm:rsv}).
For branching programs in which \textit{every} layer is a non-permution layer, we can make use of an argument from the work of Brody and Verbin \cite{BrodyVerbin}: when we apply a random restriction (where each variable is kept free with probability roughly $1/k \log n$) to such a branching program,
the resulting program is `simple' in that the width has collapsed to 2 in many of the remaining layers. This allows us to use arguments tailored to width-2 branching programs, which are well-understood. In particular, we can use the same concept of mixing as used for permutation branching programs.
To handle general width-3 branching programs, which may contain an arbitrary mix of permutation and non-permutation layers, we group the layers into ``chunks'' containing exactly one non-permutation layer each. Instead of using an ordinary random restriction, we consider a series of restrictions similar to those in Steinberger's ``interwoven hybrids'' technique \cite{Steinberger} (in our argument each chunk will correspond to a single layer in \cite{Steinberger}). In Section \ref{subsect:width-reduction}, we use such restrictions to show that the level-$k$ Fourier mass of an arbitrary width-3 program can be bounded in terms of the level-$k$ Fourier mass of a program $D$ which has the following ``pseudomixing'' form: $D$ can be split into $r\in [n]$ branching programs $D_1\circ D_2 \circ \cdots \circ D_r$, where each $D_i$ has at most $3k$ non-regular layers and the layer splitting consecutive $D_i$s has width 2.
We then generalize the arguments used for width-2 branching programs to ``pseudomixing'' branching programs. We can show that each chunk $D_i$ is either mixing or has small Fourier growth, which suffices to bound the Fourier growth of $D$.
\subsection{Organization}
In Section \ref{SectionTechniques} we introduce the definitions and tools we use in our proof. In Section \ref{SubSectionBP} we formally define branching programs and explain our view of them as matrix-valued functions. In Sections \ref{SubSectionFourier} and \ref{SectionFourierBounds} we define the matrix-valued Fourier transform and explain how we use it.
We prove the upper bound on Fourier mass of oblivious, read-once, width-3 branching programs
(i.e., Theorem \ref{MainThmIntro}) in Section \ref{SubSectionLowOrder} (Theorem \ref{TheoremLow}).
In Sections \ref{SubSectionPRrestriction}
and Section \ref{SubSectionPRG} we construct and analyse our pseudorandom generator, which proves the main result (Theorem \ref{thm:main-intro}).
The proof of Theorem \ref{TheoremFirstOrder} is left to Appendix \ref{AppendixFirstOrder}.
\section{Random Restrictions} \label{SubSectionCoin}
\section{Pseudorandom Restrictions} \label{SubSectionPRrestriction}
Our pseudorandom generator repeatedly applys pseudorandom restrictions. For the analysis, we introduce the concept of an averaging restriction as in Gopalan et al.~\cite{GMRTV} and Reingold et al.~\cite{ReingoldSteinkeVadhan2013}, which is subtly different to the restrictions in Section \ref{SubSectionLowOrder}.
\begin{definition}
For $t \in \{0,1\}^n$ and a length-$n$ branching program $B$, let $B|_t$ be the \textbf{(averaging) restriction} of $B$ to $t$---that is, $B|_t : \{0,1\}^n \to \mathbb{R}^{w \times w}$ is a matrix-valued function given by $B|_t [x] := \ex{U}{B[\mathrm{Select}(t,x,U)]}$, where $U$ is uniform on $\{0,1\}^n$.
\end{definition}
In this section we show that, for a \emph{pseudorandom} $T$ (generated using few random bits), $L(B|_T)$ is small. We will generate $T$ using an almost $O(\log n)$-wise independent distribution:
\begin{definition} \label{DefinitionLimitedIndependence}
A random variable $X$ on $\Omega^n$ is \textbf{$\delta$-almost $k$-wise independent} if, for every\\ $I=\{i_1, i_2, \cdots, i_k\} \subset [n]$ with $|I| = k$, the coordinates $(X_{i_1}, X_{i_2}, \cdots, X_{i_k}) \in \Omega^k$ are $\delta$-close (in statistical distance) to being independent---that is, for all $T \subset \Omega^k$, $$\left|\sum_{x \in T} \left( \pr{X}{(X_{i_1},X_{i_2},\cdots,X_{i_k}) = x} - \prod_{l \in [k]} \pr{X}{X_{i_l} = x_l} \right) \right| \leq \delta.$$ We say that $X$ is \textbf{$k$-wise independent} if it is $0$-almost $k$-wise independent.
\end{definition}
We can sample a random variable $X$ on $\{0,1\}^n$ that is $\delta$-almost $k$-wise independent such that each bit has expectation $p=2^{-d}$ using $O(kd+\log(1/\delta)+d\log(nd))$ random bits \cite[Lemma B.2]{ReingoldSteinkeVadhan2013}.
The following lemma, proven in essentially the same way as Lemma 5.3 in \cite{ReingoldSteinkeVadhan2013}, tells us that $L(B|_T)$ will be small for $T$ chosen from a $\delta$-almost $k$-wise distribution with appropriate parameters.
\begin{lemma} \label{TheoremMainLemma}
Let $B$ be a length-$n$, width-$w$, ordered branching program. Let $T$ be a random variable over $\{0,1\}^n$ where each bit has expectation $p$ and the bits are $\delta$-almost $2k$-wise independent. Suppose that, for all $i,j,k' \in [n]$ such that $k \leq k' < 2k$, we have $L^{k'}(B_{i \cdots j}) \leq a\cdot b^{k'}$. If we set
$p \leq 1/2b$ and $\delta \leq 1/(2b)^{2k},$ then
$$\pr{T}{L^{\geq k} (B|_T) > 1 } \leq n^4 \cdot \frac{2a}{2^k}.$$
(Recall that $L^{\geq k}(g) = \sum_{j=k}^n L^j(g)$.)
\end{lemma}
\begin{proof
Let $k \leq k' < 2k$. We have that for all $i$ and $j$, $$\ex{T}{L^{k'}(B_{i \cdots j}|_T)} = \sum_{s \subset \{i \cdots j\} : |s|=k'} \pr{T}{s \subset T} \norm{\widehat{B_{i \cdots j}}[s]}_2 \leq L^{k'}(B) (p^{k'} + \delta) \leq a b^{k'} \left(\frac{1}{(2b)^{k'}} + \frac{1}{(2b)^{2k}}\right)\leq \frac{2a}{2^k}.$$
Applying Markov's inequality and a union bound, we have that for all $\beta > 0$:
\begin{equation*}
\pr{T}{\forall 1 \leq i \leq j \leq n ~~ L_2^{k'}(B_{i \cdots j}|_T) \leq \beta} \geq 1-n^2\frac{2a}{2^k \beta}.\end{equation*}
Applying a union bound over values of $k'$ and setting $\beta =1/n$, we obtain:
$$\pr{T}{\forall k \leq k' < 2k ~ \forall 1 \leq i \leq j \leq n ~~ L^{k'}(B_{i \cdots j}|_T) \leq \frac{1}{n}} \geq 1- n^4 \cdot \frac{2a}{2^k}.$$
The result now follows from the following Lemma
\begin{lemma}[{\cite[Lemma 5.4]{ReingoldSteinkeVadhan2013}}] \label{LemmaWellOrder}
Let $B$ be a length-$n$, ordered branching program and $t \in \{0,1\}^n$. Suppose that, for all $i$, $j$, and $k'$ with $1 \leq i \leq j \leq n$ and $k \leq k' < 2k$, $L_2^{k'}(B_{i \cdots j}|_t) \leq 1/n$. Then, for all $k'' \geq k$ and all $i$ and $j$, $L_2^{k''}(B_{i \cdots j}|_t) \leq 1/n$.
\end{lemma}
\end{proof}
\section{The Pseudorandom Generator} \label{SubSectionPRG}
Our main result Theorem \ref{thm:main-intro} follows from plugging our Fourier growth bound (Theorem \ref{TheoremLow}) into the analysis of \cite{ReingoldSteinkeVadhan2013}. We include the proof and a general statement here for completeness:
\begin{theorem} \label{TheoremPRGformula}
Let $\mathcal{C}$ be a set of ordered branching programs of length at most $n$ and width at most $w$ that is closed under restrictions and subprograms -- that is, if $B \in \mathcal{C}$, then $B|_{t \leftarrow x} \in \mathcal{C}$ for all $t$ and $x$ and $B_{i \cdots j} \in \mathcal{C}$ for all $i$ and $j$. Suppose that, for all $B \in \mathcal{C}$ and $k \in [n]$, we have $L^k(B) \leq a b^k$, where $b \geq 2$. Let $\varepsilon>0$.
Then there exists a pseudorandom generator $G_{a,b,n,\varepsilon} : \{0,1\}^{s_{a,b,n,\varepsilon}} \to \{0,1\}^n$ with seed length $s_{a,b,n,\varepsilon} = O\left(b \cdot \log(b) \cdot \log(n) \cdot \log\left(\frac{abwn}{\varepsilon}\right)\right)$ such that, for any length-$n$, width-$w$, read-once, oblivious (but unordered) branching program $B$ that corresponds to an ordered branching program in $\mathcal{C}$,\footnote{That is, there exists $B' \in \mathcal{C}$ and a permutation of the bits $\pi : \{0,1\}^n \to \{0,1\}^n$ such that $B[x] = B'[\pi(x)]$ for all $x$.} $$\norm{\ex{U_{s_{a,b,n, \varepsilon}}}{B[G_{a,b,n,\varepsilon}(U_{s_{a,b,n, \varepsilon}})]}-\ex{U}{B[U]}}_2 \leq \varepsilon.$$ Moreover, $G_{a,b,n, \varepsilon}$ can be computed in space $O(s_{a,b,n,\varepsilon})$.
\end{theorem}
To prove Theorem \ref{thm:main-intro} we set $\mathcal{C}$ to be the class of all 3OBPs of length at most $n$. Theorem \ref{TheoremLow} gives a bound corresponding to $a = O(n^2)$ and $b = O(\log n)$. This gives the required generator. The statements of Theorems \ref{thm:main-intro} and \ref{TheoremPRGformula} differ in that Theorem \ref{TheoremPRGformula} bounds the error of the pseudorandom generator with respect to a matrix-valued function, while Theorem \ref{thm:main-intro} bounds the error with respect to a $\{0,1\}$-valued function. These statements are equivalent as the $\{0,1\}$-valued function is simply one entry in the matrix-valued function.
The following lemma gives the basis of our pseudorandom generator.
\begin{lemma} \label{LemmaOneStep}
Let $a$, $b$, and $\mathcal{C}$ be as in Theorem \ref{TheoremPRGformula} and $B \in \mathcal{C}$. Let $\varepsilon \in (0,1)$. Let $T$ be a random variable over $\{0,1\}^n$ that is $\delta$-almost $2k$-wise independent and each bit has expectation $p$, where we require $$p \leq 1/(2b),~~~~ k \geq \log_2 \left( 8 a n^4 w / \varepsilon \right), ~~~~\text{and}~~~~\delta \leq 1/(2b)^{2k}.$$ Let $U$ be uniform over $\{0,1\}^n$. Let $X$ be a $\mu$-biased random variable over $\{0,1\}^n$ with $\mu \leq \varepsilon /2ab^k.$ Then $$\norm{\ex{T,X,U}{B[\mathrm{Select}(T,X,U)]} - \ex{U}{B[U]}}_2 \leq \varepsilon.$$
\end{lemma}
\begin{proof}
For a fixed $t \in \{0,1\}^n$, we have
\begin{align*}
\norm{\ex{X,U}{B[\mathrm{Select}(t,X,U)]} - \ex{U}{B[U]}}_2 =& {\norm{\ex{X}{B|_t[X]} - \ex{U}{B[U]}}_2}\\
=& {\norm{\sum_{s \ne 0} \widehat{B|_t}[s] \widehat{X}(s)}_2}\\
\leq& {\sum_{s \ne 0} \norm{\widehat{B|_t}[s]}_2 |\widehat{X}(s)|}\\
\leq& {L(B|_t) \mu},\\
\end{align*}
Conditioning on whether or not $L^{\geq k}(B|_t)>1$, we have
\begin{align*}
\norm{\ex{T,X,U}{B[\mathrm{Select}(T,X,U)]} - \ex{U}{B[U]}}_2
& \leq \pr{T}{L^{\geq k}(B|_T) > 1} \max_{t} \norm{\ex{X,U}{B[\mathrm{Select}(t,X,U)]} - \ex{U}{B[U]}}_2 \\&+ \pr{T}{L^{\geq k} (B|_T) \leq 1} \mu \ex{T}{L(B|_T) \mid L^{\geq k}(B|_T) \leq 1}. \\
\end{align*}
We have $$\L^{< k}(B) \leq \sum_{1 \leq k' < k} a b^{k'} = ab \frac{b^{k-1}-1}{b-1} \leq ab^k-1.$$ Thus $\ex{T}{L(B|_T) \mid L^{\geq k}(B|_T) \leq 1} \leq ab^k$.
Lemma \ref{TheoremMainLemma} gives
$$
\pr{T}{L^{\geq k}(B|_T) > 1} \leq n^4 \cdot \frac{2a}{2^k}.
$$
For all $t,x,y$, we have $\norm{B[\mathrm{Select}(t,x,y)] - \ex{U}{B[U]}}_2 \leq 2w$. Thus
\begin{align*}
\norm{\ex{T,X,U}{B[\mathrm{Select}(T,X,U)]} - \ex{U}{B[U]}}_2 \leq& n^4 \cdot \frac{2a}{2^k} \cdot 2w + 1 \cdot \mu \cdot ab^k\\
\leq& \frac{4 a n^4 w}{8 a n^4 w / \varepsilon} + a b^k \frac{\varepsilon}{2 a b^k}\\
\leq& \varepsilon.
\end{align*}
\end{proof}
Now we use the above results to construct our pseudorandom generator
The pseudorandom generator is formally defined as follows.
\begin{quote}
\begin{center}\textbf{Algorithm for $G_{a,b,n, \varepsilon} : \{0,1\}^{s_{a,b,n, \varepsilon}} \to \{0,1\}^n$.}\end{center}
\begin{itemize}
\item[Parameters:] $n \in \mathbb{N}$, $\varepsilon>0$.
\item[Input:] A random seed of length $s_{a,b,n, \varepsilon}$.
\item[1.] Compute appropriate values of $p \leq 1/2b$, $\varepsilon' = \varepsilon p / 14w \log_2(n)$, $k \geq \log_2 \left( 8 a n^4 w / \varepsilon' \right)$, $\delta \leq \varepsilon' (p/2)^{2k}$, and $\mu \leq \varepsilon' /2ab^k.$ \footnote{For the purposes of the analysis we assume that $\varepsilon'$, $k$, $p$, $\delta$, and $\mu$ are the same at every level of recursion. So if $G_{a,b,n,w,\varepsilon}$ is being called recursively, use the same values of $\varepsilon'$, $p$, $k$, $\delta$, and $\mu$ as at the previous level of recursion. We pick values within a constant factor of these constraints.}
\item[2.] If $n \leq 320 \cdot \lceil \log_2(1/\varepsilon') \rceil/p$, output $n$ truly random bits and stop.
\item[3.] Sample $T \in \{0,1\}^n$ where each bit has expectation $p$ and the bits are $\delta$-almost $2k$-wise independent.
\item[4.] If $|T|<pn/2$, output $0^n$ and stop.
\item[5.] Recursively sample $\tilde{U} \in \{0,1\}^{\lfloor n(1-p/2) \rfloor}$. i.e. $\tilde{U}=G_{a,b,\lfloor n(1-p/2) \rfloor,\varepsilon}(U)$.
\item[6.] Sample $X \in \{0,1\}^n$ from a $\mu$-biased distribution.
\item[7.] Output $\mathrm{Select}(T,X,\tilde{U}) \in \{0,1\}^n$.\footnote{Technically, we must pad $\tilde{U}$ with zeros in the locations specified by $T$ (i.e. $\tilde{U}_i = 0$ for $i \in T$) to obtain the right length.}
\end{itemize}
\end{quote}
The analysis of the algorithm proceeds roughly as follows.
\begin{itemize}
\item We have $p = \Theta(1/b)$, $\varepsilon' = \Theta(\varepsilon/wb\log n)$, $k = \Theta(\log(abwn/\varepsilon))$, $\delta=1/b^{\Theta(k)}$, and $\mu=1/b^{\Theta(k)}$.
\item Every time we recurse, $n$ is decreased to $\lfloor n(1-p/2) \rfloor$. After $O(\log(n)/p)$ recursions, $n$ is reduced to $O(1)$. So the maximum recursion depth is $r=O(\log(n)/p) = O( b \log n )$.
\item The probability of failing because $|T|<pn/2$ is small by a Chernoff bound for limited independence. (This requires that $n$ is not too small and, hence, step 2.)
\item The output is pseudorandom, as $$\ex{U}{B[G_{a,b,n,\varepsilon}(U)]} = \ex{T,X,\tilde{U}}{B[\text{Select}(T,X,\tilde{U})]} \approx \ex{T,X,U}{B[\text{Select}(T,X,{U})]} \approx \ex{U}{B[U]}.$$ The first approximate equality holds because we inductively assume that $\tilde{U}$ is pseudorandom. The second approximate equality holds by Lemma \ref{LemmaOneStep}.
\item The total seed length is the seed length needed to sample $X$ and $T$ at each level of recursion and $O(\log(1/\varepsilon')/p) = O(b \log(bwn/\varepsilon))$ truly random bits at the last level. Sampling $X$ requires seed length $O(\log(n/\mu)) = O(k \log b)$ and sampling $T$ requires seed length $O(k \log(1/p) + \log(1/\delta) + \log(1/p) \cdot \log(n \log(1/p))) = O(k \log b)$ so the total seed length is
$$r \cdot O(k \log b) + O(b \log(bwn/\varepsilon)) = O\left(b \cdot \log(b) \cdot \log(n) \cdot \log\left(\frac{abwn}{\varepsilon}\right)\right).$$
\end{itemize}
\begin{lemma} \label{LemmaPRGfail}
The probability that $G_{a,b,n,\varepsilon}$ fails at step 4 is bounded by $3\varepsilon'$---that is, $\pr{T}{|T|<pn/2}\leq 3\varepsilon'$.
\end{lemma}
\begin{proof}
By a Chernoff bound for limited independence (Lemma \ref{LemmaChernoff}), $$\pr{T}{|T| < pn/2} \leq \left( \frac{20k'}{(1/2)^2pn} \right)^{\lfloor k'/2 \rfloor} + 2 \delta\cdot \left(\frac{n}{(1/2)pn}\right)^{k'},$$ where $k' \leq 2k$ even is arbitrary. Set $k'= 2 \lceil \log_2(1/\varepsilon') \rceil$. Step 2 ensures that
$n > 160 k'/p$ and our setting of $\delta$ gives that
$\delta \leq \varepsilon' (p/2)^{k'}.$
Thus we have $$\pr{T}{|T| < pn/2} \leq 2^{-\log_2(1/\varepsilon')} + 2 \varepsilon' \leq 3\varepsilon'.$$
\end{proof}
The following bounds the error of $G_{a,b,n,\varepsilon}$.
\begin{lemma} \label{LemmaPRGerror}
Let $B \in \mathcal{C}$. Then
$$\norm{\ex{U_{s_{n,\varepsilon}}}{B[G_{n,\varepsilon}(U_{s_{n,\varepsilon}})]}-\ex{U}{B[U]}}_2 \leq 7w r \varepsilon' ,$$ where $r=O(\log(n)/p)$ is the maximum recursion depth of $G_{a,b,n,\varepsilon}$.
\end{lemma}
\begin{proof}
For $0 \leq i < r$, let $n_i$, $T_i$, $X_i$, and $\tilde{U}_i$ be the values of $n$, $T$, $X$, and $\tilde{U}$ at recursion level $i$. We have $n_{i+1}=\lfloor n_i(1-p/2) \rfloor \leq n(1-p/2)^{i+1}$ and $\tilde{U}_{i-1} = \mathrm{Select}(T_i,X_i,\tilde{U}_i)$. Let $\Delta_i$ be the error of the output from the $i^\text{th}$ level of recursion---that is, $$\Delta_i := \max_{B' \in \mathcal{C}} \norm{\ex{T_i,X_i,\tilde{U}_i}{B'[\mathrm{Select}(T_i,X_i,\tilde{U}_i)]}-\ex{U}{B'[U]}}_2.$$
Since the last level of recursion outputs uniform randomness, $\Delta_r=0$. For $0 \leq i<r$, we have, for some $B' \in \mathcal{C}$,
\begin{align*}
\Delta_i \leq& \norm{\ex{T_i,X_i,\tilde{U}_i}{B'[\text{Select}(T_i,X_i,\tilde{U}_i)]}-\ex{U}{B'[U]}}_2 \cdot \pr{T}{|T|\geq pn/2} \\&+ 2w \cdot \pr{T}{|T|<pn/2}\\
\leq& \norm{\ex{T_i,X_i,\tilde{U}_i}{B'[\text{Select}(T_i,X_i,\tilde{U}_i)]}-\ex{T_i,X_i,{U}}{B'[\text{Select}(T_i,X_i,{U})]}}_2 \\&+ \norm{\ex{T_i,X_i,{U}}{B'[\text{Select}(T_i,X_i,{U})]}-\ex{U}{B'[U]}}_2 \\&+2w\cdot\pr{T}{|T|<pn/2}\\
\end{align*}
By Lemma \ref{LemmaOneStep}, $$\norm{\ex{T_i,X_i,{U}}{B'[\text{Select}(T_i,X_i,{U})]}-\ex{U}{B'[U]}}_2 \leq \varepsilon'.$$ By Lemma \ref{LemmaPRGfail}, $$\pr{T}{|T|<pn/2} \leq 3\varepsilon'.$$
We claim that $$\norm{\ex{T_i,X_i,\tilde{U}_i}{B'[\text{Select}(T_i,X_i,\tilde{U}_i)]}-\ex{T_i,X_i,{U}}{B'[\text{Select}(T_i,X_i,{U})]}}_2 \leq \Delta_{i+1}.$$ Before we prove the claim, we complete the proof: This gives $\Delta_i \leq \Delta_{i+1} + \varepsilon' + 2w \cdot 3\varepsilon'$. It follows that $\Delta_0 \leq 7w r \varepsilon'$, as required.
To prove the claim, consider \emph{any} fixed $T_i=t$ and $X_i=x$. We have $$\norm{\ex{\tilde{U}_i}{B'[\text{Select}(t,x,\tilde{U}_i)]}-\ex{U}{B'[\text{Select}(t,x,{U})]}}_2 \leq \Delta_{i+1}.$$
Consider $\overline{B}_{x,t}[y] := B'[\text{Select}(t,x,y)]$ as a function of $y \in \{0,1\}^{n_i-|t|}$. Then $\overline{B}_{x,t}$ is a width-3 read-once oblivious branching program of length-$(n_i-|t|)$.
We inductively know that $\tilde{U}_i $ is pseudorandom for $\overline{B}_{x,t}$---that is, $\norm{\ex{\tilde{U}_i}{\overline{B}_{x,t}[\tilde{U}_i]} - \ex{U}{\overline{B}_{x,t}[U]}}_2 \leq \Delta_{i+1}$. Thus $$\norm{\ex{\tilde{U}_i}{B'[\text{Select}(t,x,\tilde{U}_i)]}-\ex{U}{B'[\text{Select}(t,x,{U})]}}_2 = \norm{\ex{\tilde{U}_i}{\overline{B}_{x,t}[\tilde{U}_i]}-\ex{U}{\overline{B}_{x,t}[U]}}_2 \leq \Delta_{i+1},$$ as required.
\end{proof}
\begin{proof}[Proof of Theorem \ref{TheoremPRGformula}]
Since $\varepsilon' \leq \varepsilon/(7w r) $, Lemma \ref{LemmaPRGerror} implies that $G_{a,b,n,\varepsilon}$ has error at most $\varepsilon$.
The seed length is $$s_{a,b,n,\varepsilon}=O\left(b \cdot \log(b) \cdot \log(n) \cdot \log\left(\frac{abwn}{\varepsilon}\right)\right)$$ as required.
\end{proof}
\section{Preliminaries} \label{SectionTechniques}
\subsection{Branching Programs} \label{SubSectionBP}
We define a length-$n$, width-$w$ \textbf{program} to be a function $B : \{0,1\}^n \times [w] \to [w]$, which takes a start state $u \in [w]$ and an input string $x \in \{0,1\}^n$ and outputs a final state $B[x](u)$.
Often we think of $B$ as having a fixed \textbf{start state} $u_0$ and a set $S \subset [w]$ of \textbf{accept states}. Then $B$ \textbf{accepts} $x \in \{0,1\}^n$ if $B[x](u_0) \in S$. We say that $B$ \textbf{computes the function} $f : \{0,1\}^n \to \{0,1\}$ if $f(x)=1$ if and only if $B[x](u_0) \in S$.
In our applications, the input $x$ is randomly (or pseudorandomly) chosen, in which case a program can be viewed as a Markov chain randomly taking initial states to final states. For each $x \in \{0,1\}^n$, we let $B[x] \in \{0,1\}^{w \times w}$ be a matrix defined by $$B[x](u,v) = 1 \iff B[x](u)=v.$$
For a random variable $X$ on $\{0,1\}^n$, we have $\ex{X}{B[X]} \in [0,1]^{w \times w},$ where $\ex{R}{f(R)}$ is the {expectation} of a function $f$ with respect to a random variable $R$. Then the entry in the $u^\text{th}$ row and $v^\text{th}$ column $\ex{X}{B[X]}(u,v)$ is the probability that $B$ takes the initial state $u$ to the final state $v$ when given a random input from the distribution $X$---that is, $$\ex{X}{B[X]}(u,v) = \pr{X}{B[X](u)=v},$$ where $\pr{R}{e(R)}$ is the {probability} of an event $e$ with respect to the random variable $R$.
A branching program reads one bit of the input at a time (rather than reading $x$ all at once) maintaining only a state in $[w] = \{1,2, \cdots, w\}$ at each step. We capture this restriction by demanding that the program be composed of several smaller programs, as follows.
Let $B$ and $B'$ be width-$w$ programs of length $n$ and $n'$ respectively. We define the \textbf{concatenation} $B \circ B' : \{0,1\}^{n+n'} \times [w] \to [w]$ of $B$ and $B'$ by $$(B \circ B')[x \circ x'](u) := B'[x'](B[x](u)),$$ which is a width-$w$, length-$(n+n')$ program. That is, we run $B$ and $B'$ on separate inputs, but the final state of $B$ becomes the start state of $B'$. Concatenation corresponds to matrix multiplication---that is, $(B \circ B')[x \circ x'] = B[x] \cdot B'[x']$, where the two programs are concatenated on the left hand side and the two matrices are multiplied on the right hand side.
A length-$n$, width-$w$, \textbf{ordered branching program} (abbreviated \textbf{OBP}) is a program $B$ that can be written $B = B_1 \circ B_2 \circ \cdots \circ B_n$, where each $B_i$ is a length-$1$ width-$w$ program. We refer to $B_i$ as the $i^\text{th}$ \textbf{layer} of $B$. We denote the \textbf{subprogram} of $B$ from layer $i$ to layer $j$ by $B_{ i \cdots j} := B_i \circ B_{i+1} \circ \cdots \circ B_j$.
A length-$n$, width-$w$, ordered branching program can also be viewed as a directed acyclic graph. The vertices are arranged into $n+1$ layers each of size $w$. The edges go from one layer to the next. In particular, there is an edge labelled $b$ from vertex $u$ in layer $i-1$ to vertex $B[b](u)$ in layer $i$.
\begin{figure}[h]
\centering
\includegraphics[scale=.25]{bp.eps}
\end{figure}
We use the following notational conventions when referring to layers of a length-$n$ branching program. We need to distinguish between layers of vertices and layers of edges (although this is often clear from context). Layers of \emph{edges} are the $B_i$s and are numbered from $1$ to $n$. Layers of \emph{vertices} are the states between the $B_i$s and are numbered from $0$ to $n$. The edges in layer $i$ ($B_i$) go from vertices in layer $i-1$ to vertices in layer $i$.
General read-once, oblivious branching programs (a.k.a. unordered branching programs) can be reduced to the ordered case by a permutation of the input bits. Formally, a \textbf{read-once, oblivious branching program} $B$ is an ordered branching program $B'$ composed with a permutation $\pi$. That is, $B[x]=B'[\pi(x)]$, where the $i^\text{th}$ bit of $\pi(x)$ is the $\pi(i)^\text{th}$ bit of $x$
For a program $B$ and an arbitrary distribution $X$, the matrix $\ex{X}{B[X]}$ is \textbf{stochastic}---that is, $$\sum_v \ex{X}{B[X]}(u,v) = 1$$ for all $u$ and $\ex{X}{B[X]}(u,v) \geq 0$ for all $u$ and $v$. A program $B$ is called a \textbf{regular program} if the matrix $\ex{U}{B[U]}$ is \textbf{doubly stochastic}---that is, both $\ex{U}{B[U]}$ and its transpose $\ex{U}{B[U]}^*$ are stochastic. A program $B$ is called a \textbf{permutation program} if $B[x]$ is a permutation matrix for every $x$ or, equivalently, $B[x]$ is doubly stochastic. Note that a permutation program is necessarily a regular program and, if both $B$ and $B'$ are regular or permutation programs, then so is their concatenation.
A regular program $B$ has the property that the uniform distribution is a stationary distribution of the Markov chain $\ex{U}{B[U]}$, whereas, if $B$ is a permutation program, the uniform distribution is stationary for $\ex{X}{B[X]}$ for \emph{any} distribution $X$.
A \textbf{regular branching program} is a branching program where each layer $B_i$ is a regular program and likewise for a \textbf{permutation branching program}. We will refer to layer $i$ as regular if $B_i$ is a regular program and we say that layer $i$ is non-regular otherwise.
Equivalently, a regular branching program is one where, in the directed acyclic graph, each vertex has in-degree 2 (in addition to having out-degree 2) except those in the start layer -- that is, each layer of edges is a regular graph (hence the name). A permutation branching program has the additional constraint that the incoming edges have distinct labels.
We also consider branching programs of varying width -- some layers have more vertices than others. The overall width of the program is the maximum width of any layer. This means that the edge layers $B_i$ may give non-square matrices. For $i \in [n]$, if $B_i[x] \in \{0,1\}^{w \times w'}$, then we refer to $w$ as the width of layer $i-1$ and $w'$ as the width of layer $i$.
\subsection{Norms}
We are interested in constructing a random variable $X$ (the output of the pseudorandom generator) such that $\ex{X}{B[X]} \approx \ex{U}{B[U]}$, where $U$ is uniform on $\{0,1\}^n$. Throughout we use $U$ to denote the \textbf{uniform distribution}. The error of the pseudorandom generator will be measured by a norm of the matrix $\ex{X}{B[X]} - \ex{U}{B[U]}$.
For a matrix $A \in \mathbb{R}^{w \times w'}$, define the \textbf{$\rho$ operator norm} of $A$ by $$\norm{A}_\rho = \max_x \frac{\norm{xA}_\rho}{\norm{x}_\rho},$$ where $\rho$ specifies a vector norm (usually $1$, $2$, or $\infty$ norm). Define the \textbf{Frobenius norm} of $A \in \mathbb{R}^{w \times w'}$ by $$\frob{A}^2 = \sum_{u,v} A(u,v)^2 = \text{trace}(A^*A) = \sum_\lambda |\lambda|^2,$$ where $A^*$ is the (conjugate) transpose of $A$ and the last sum is over the singular values $\lambda$ of $A$. Note that $\norm{A}_2 \leq \frob{A}$ for all $A$.
\subsection{Fourier Analysis} \label{SubSectionFourier}
Let $B : \{0,1\}^n \to \mathbb{R}^{w\times w'}$ be a matrix-valued function (such as given by a length-$n$, width-$w$ branching program). Then we define the \textbf{Fourier transform} of $B$ as a matrix-valued function $\widehat{B} : \{0,1\}^n \to \mathbb{R}^{w \times w'}$ given by $$\widehat{B}[s] := \ex{U}{B[U] \chi_s(U)},$$ where $s \in \{0,1\}^n$ (or, equivalently, $s \subset [n]$) and $$\chi_s(x) = (-1)^{\sum_i x(i) \cdot s(i)} = \prod_{i \in s} (-1)^{x(i)}.$$ We refer to $\widehat{B}[s]$ as the $s^\text{th}$ \textbf{Fourier coefficient} of $B$. The \textbf{order} of a Fourier coefficient $\widehat{B}[s]$ is $|s|$---the \textbf{Hamming weight} of $s$, which is the size of the set $s$ or the number of $1$s in the string $s$. Note that this is equivalent to taking the real-valued Fourier transform of each of the $w \cdot w'$ entries of $B[x]$ separately, but we will see below that this matrix-valued Fourier transform is nicely compatible with matrix algebra.
For a random variable $X$ over $\{0,1\}^n$ we define its $s^\text{th}$ \textbf{Fourier coefficient} as $$\widehat{X}(s) := \ex{X}{\chi_s(X)},$$ which, up to scaling, is the same as taking the real-valued Fourier transform of the probability mass function of $X$.
We have the following useful properties.
\begin{lemma} \label{LemmaFourier}
Let $A: \{0,1\}^n \to \mathbb{R}^{w\times w'}$ and $B: \{0,1\}^n \to \mathbb{R}^{w'\times w''}$ be matrix valued functions. Let $X$, $Y$, and $U$ be independent random variables over $\{0,1\}^n$, where $U$ is uniform. Let $s,t \in \{0,1\}^n$. Then we have the following.
\begin{itemize}
\item Decomposition: If $C[x \circ y] = A[x] \cdot B[y]$ for all $x,y \in \{0,1\}^n$, then $\widehat{C}[s \circ t] = \widehat{A}[s] \cdot \widehat{B}[t]$.
\item Expectation: $\ex{X}{B[X]} = \sum_s \widehat{B}[s] \widehat{X}(s)$.
\item Fourier Inversion for Matrices: $B[x] = \sum_s \widehat{B}[s] \chi_s(x)$.
\item Fourier Inversion for Distributions: $\pr{X}{X=x} = \ex{U}{\widehat{X}(U) \chi_U(x)}$.
\item Convolution for Distributions: If $Z = X \oplus Y$, then $\widehat{Z}(s) = \widehat{X}(s) \cdot \widehat{Y}(s)$.
\item Parseval's Identity: $\sum_{s \in \{0,1\}^n} \frob{\widehat{B}[s]}^2 = \ex{U}{\frob{B[U]}^2}$.
\item Convolution for Matrices: If, for all $x \in \{0,1\}^n$, $C[x] = \ex{U}{A[U] \cdot B[U \oplus x]}$, then $\widehat{C}[s] = \widehat{A}[s] \cdot \widehat{B}[s]$.
\end{itemize}
\end{lemma}
The Decomposition property is what makes the matrix-valued Fourier transform more convenient than separately taking the Fourier transform of the matrix entries as done by Bogdanov et al.~\cite{BPW}. If $B$ is a length-$n$ width-$w$ branching program, then, for all $s \in \{0,1\}^n$, $$\widehat{B}[s] = \widehat{B}_1[s_1] \cdot \widehat{B}_2[s_2] \cdot \cdots \cdot \widehat{B}_n[s_n].$$
\subsection{Small-Bias Distributions} \label{SubSectionSmallBias}
The \textbf{bias} of a random variable $X$ over $\{0,1\}^n$ is defined as $$\text{bias}(X) := \max_{s \ne 0} |\widehat{X}(s)|.$$ A distribution is \textbf{$\varepsilon$-biased} if it has bias at most $\varepsilon$. Note that a distribution has bias $0$ if and only if it is uniform. Thus a distribution with small bias is an approximation to the uniform distribution. We can sample an $\varepsilon$-biased distribution $X$ on $\{0,1\}^n$ with seed length $O(\log(n/\varepsilon))$ and using space $O(\log(n/\varepsilon))$ \cite{NaorNa93,AGHP}.
Small-bias distributions are useful pseudorandom generators: A $\varepsilon$-biased random variable $X$ is indistinguishable from uniform by any linear function (a parity of a subset of the bits of $X$). That is, for any $s \subset [n]$, we have $\left| \ex{X}{\bigoplus_{i \in s} X_i} - 1/2\right|\leq 2\varepsilon$. Small bias distributions are known to be good pseudorandom generators for width-$2$ branching programs \cite{BogdanovDvVeYe09}, but not width-$3$. For example, the uniform distribution over $\{x \in \{0,1\}^n : |x| \mod{3} = 0 \}$ has bias $2^{-\Theta(n)}$, but does not fool width-$3$, ordered, permutation branching programs.
\subsection{Fourier Mass} \label{SectionFourierBounds}
We analyse small bias distributions as pseudorandom generators for branching programs using Fourier analysis. Intuitively, the Fourier transform of a branching program expresses that program as a linear combination of linear functions (parities), which can then be fooled using a small-bias space.
Define the \textbf{Fourier mass} of a matrix-valued function $B$ to be $$L(B) := \sum_{s \ne 0} \norm{\widehat{B}[s]}_2.$$ Also, define the \textbf{Fourier mass of $B$ at level $k$} as $$L^k(B) := \sum_{s \in \{0,1\}^n : |s|=k} \norm{\widehat{B}[s]}_2.$$ Note that $L(B) =\sum_{k \geq 1} L^k(B)$. We define $L^{\geq k}(B) := \sum_{k' \geq k} L^{k'}(B)$ and $L^{\leq k}(B)$, $L^{>k}(B)$, $L^{<k}(B)$ are defined analogously.
The Fourier mass is unaffected by order:
\begin{lemma} \label{LemmaFourierPermutation}
Let $B, B' : \{0,1\}^n \to \mathbb{R}^{w \times w}$ be matrix-valued functions satisfying $B[x]=B'[\pi(x)]$, where $\pi : [n] \to [n]$ is a permutation. Then, for all $s \in \{0,1\}^n$, $\widehat{B}[s]=\widehat{B'}[\pi(s)]$. In particular, $L(B)=L(B')$ and $L^k(B)=L^k(B')$ for all $k$ and $\rho$.
\end{lemma}
Lemma \ref{LemmaFourierPermutation} implies that the Fourier mass of any read-once, oblivious branching program is equal to the Fourier mass of the corresponding ordered branching program.
If $L(B)$ is small, then $B$ is fooled by a small-bias distribution:
\begin{lemma} \label{LemmaBiasMass}
Let $B$ be a length-$n$, width-$w$ branching program. Let $X$ be a $\varepsilon$-biased random variable on $\{0,1\}^n$. We have $$\norm{\ex{X}{B[X]}-\ex{U}{B[U]}}_2 = \norm{\sum_{s \ne 0} \widehat{B}[s] \widehat{X}(s)}_2 \leq L(B) \varepsilon.$$
\end{lemma}
In the worst case $L(B) = 2^{\Theta(n)}$, even for a length-$n$ width-$3$ permutation branching program $B$. For example, the program $B_{\text{mod 3}}$ that computes the Hamming weight of its input modulo $3$ has exponential Fourier mass.
We show that, using `restrictions,' we can ensure that $L(B)$ is small. |
\section{Introduction}
\label{intro}
Some of the most important discoveries in condensed matter physics over the last few decades have been about topological states of matter. Topological states are gapped states of matter that are not characterized by broken symmetries and local order parameters, but are still distinct phases from a trivial insulator. A subclass of topological states of matter is the topologically ordered states, which are stable against any local perturbations and have topologically protected properties including fractional quasi-particle statistics, ground state degeneracy determined by topology of the spatial manifold, etc\cite{Wen2004}. Since the discovery of integer and fractional quantum Hall effects\cite{klitzing1980,tsui1982}, 2D topologically ordered states have been characterized and constructed by many different approaches, such as ideal wavefunctions\cite{laughlin1983,Moore1991,wen2008a,bernevig2008}, topological field theories\cite{wen1992,zhang1992}, exact lattice models\cite{Kitaev2003,wen2003,kitaev2006,LevinWen} and tensor networks\cite{Verstraete2006,gu2009,zaletel2012,estienne2013,wahl2013,dubail2013}.
The general structure of topological states in 2D is governed by the mathematical frame work of tensor category theory\cite{kitaev2006,zhwang2010,nayak2008}.
Since the discovery of topological insulators\cite{hasan2010,qiRMP2011}, new types of topologically ordered states with nontrivial symmetry properties have also been studied, known as symmetry protected topological (SPT) states and symmetry enriched topological (SET) states.\cite{chen2011b,hung2013,vishwanath2013,senthil2014}
In contrast to the progress in 2D, much less is known about topologically ordered states in 3D. Since there is not yet a general framework for 3D topologically ordered states, building different models with topological order is an important method that can help us understand the general features of 3D topological order. The 3D topological insulators\cite{hasan2010,qiRMP2011} can be generalized to fractional topological insulators with topological order and fractionally charged partons\cite{bernevig2006a,levin2009,maciejko2010b,swingle2011b}. Some exact solvable models in 2D can be generalized to 3D. For example, the 3D generalization of the toric code model can capture the topological order of 3D lattice discrete gauge theory in which the ground state degeneracy is associated to the non-trivial $1$-cycles and the point particles have non-trivial mutual braiding statistics with flux string excitations. The Levin-Wen models can be generalized to the Walker-Wang models\cite{walker2012,simon2013} which describe 3D gapped states with interesting surface topological order and 3D bulk topological order that generalizes the 3D lattice gauge theory. In 3D, the statistics between particles is a representation of the permutation group instead of braiding group in 2D, so that there cannot be fractional statistics for point-like particles\cite{doplicher1971,doplicher1974} (although non-trivial statistics have been shown to be possible for point-like monopole defects\cite{teo2010,ran2011,freedman2011,freedman2011b}). However, 3D topologically ordered states can have 1D string-like excitations in addtion to particles. The existence of string excitation enables rich possibilities of braiding statistics. Recently, the topological properties of the lattice models of the 3+1D Dijkgraaf-Witten lattice gauge theory\cite{dijkgraaf1990,Levin2014StrBrd,Ran2014StrBrd,JuvenWenStrBrd2014} and 3D generalization of quantum double models\cite{Shor2011,Wen2014StrBrd} are studied. In these models, string excitations have non-trivial fusion and braiding statistics.
In this paper, we propose a different approach of constructing 3D topological states named as the layer construction. In this construction, 3D topological states can emerge in a system with stacked layers of 2D topological states and properly engineered inter-layer couplings. This type of construction is first proposed in an special example of 3D boson topological insulators in Ref. \onlinecite{wang2013}. It is reminiscent of the wire construction of 2D topological states\cite{sondhi2001,kane2002,teo2014}. It is shown that if we take an array of the 1D wires of Luttinger liquids and engineer the inter-wire coupling, this array of wires can effectively form a 2D quantum Hall liquid. Depending on the coupling, a large class of fractional quantum Hall states can be constructed. Motivated by the success of the wire construction for 2D topological states and by the example in Ref. \onlinecite{wang2013}, we study in this paper the layer construction of 3D states as a more general formalism of constructing and characterizing 3D topological states. In this general formalism, we consider coupling layers of 2D Abelian topological states where the consequence of the coupling is to induce anyon condensation of some composite particle formed by topological quasiparticles in two adjacent layers. By several explicit examples, we show that this approach can describe a large class of different 3D topological states, including those with only surface 2D topological order, and those with bulk and/or surface topological order.
For the systems with only surface topological order and no bulk topological order, we construct a series of examples which are 3D gapped states with interesting chiral 2D topological order on their surfaces, although each layer of 2D states in the construction is non-chiral.
This is similar to the observation of Ref. \onlinecite{simon2013,XieChen2013} in Walker-Wang models. We obtain general criteria for the absence of 3D topological order and the characterization of surface topological order. With these criteria satisfied, we show how to determine the surface topological order on top, bottom and side surfaces in a general situation and show the equivalence of different surfaces, which proves that the states obtained in this construction are topologically isotropic. For more general states with nontrivial 3D topological order, our layer construction enables description of the bulk point-particles and string-like excitations. We provide several examples of different topological orders, including a ``convectional" 3D topological order that resembles the lattice gauge theory, and more general systems with coexisting bulk and surface topological order. As a particularly interesting feature of this construction, we can construct some 3D topologically ordered states with non-trivial string-string braiding statistics, in addition to the usual particle-string braiding statistics. We provide a field theory description of the 3D topological order and the string-string and particle-string braiding statistics. We also discuss the general features of string braiding, and provide a more general proof of a constraint on string braiding statistics proposed recently by Ref.\cite{Levin2014StrBrd}. In the end,
give an example system with non-Abelian string braiding statistics, to show that more general topological orders are possible in 3D.
The rest of the paper is organized as follows. In Sec. \ref{generalconstrtuction}, we propose a general framework for the layer construction with layers of 2D Abelian topological states. We introduce the coupling between layers by turning on anyon condensations\cite{maissam2013ac1,maissam2013ac2,levin2013ac,Kong2013} for certain composite quasi-particles living in two adjacent layers. We provide general
requirements on such condensed particles. In Sec. \ref{pSTO}, we focus on the construction of 3D gapped states with only 2D topological order on their surfaces and no bulk topological order. Sec. \ref{EGpSTO} presents several examples, and Sec. \ref{GCpSTO} discusses the general criteria of surface-only topological order. Sec. \ref{surfaceOrder} studies the surface topological order on top, bottom and side surfaces and prove their equivalence.
In Sec. \ref{3DTO}, we discuss more general states with bulk topological order. Sec. \ref{Convent3DTO} discusses the conventional topological order of the type of discrete gauge theories. Sec. \ref{Coexist} discusses examples with bulk topological order coexisting with surface topological order. Sec. \ref{GenericState} discusses the most nontrivial types of system with nontrivial string-string braiding statistics.
In Sec. \ref{FieldTheory}, we provide a field theory description of the topological order and the string braiding statistics for the layer constructed models. Sec. \ref{GeneralStrBrd} is devoted to more general discussions on string braiding statistics. In Sec. \ref{GeneralA} we prove a general identity that should be satisfied by Abelian string braiding statistics. In Sec. \ref{GeneralB} we
we provide an example of non-Abelian strings and their statistics. The final section \ref{conclusion} concludes the paper with some discussions on open questions and future directions.
\section{General Setting of the Layer Construction}
\label{generalconstrtuction}
In this section, we will introduce the general formalism for the construction of $3$D topological states from layers of $2$D topological states. In general, one can start with any topological state in each $2$D layer. For concreteness of the discussion, we restrict ourselves to only consider layers of $2$D Abelian topological states. Before we introduce the general framework of the layer construction, we will first briefly review the theory of $2$D Abelian topological states and the condensation of certain quasi-particles in these states. $2$D Abelian topological states can be described by the Abelian Chern-Simons theory with a K-matrix, of which the Lagrangian\cite{wen1992kMatrix} is given by:
\begin{align}
L_{CS}=\int dx dy dt \frac{1}{4\pi}\epsilon^{\mu\nu\lambda} K_{IJ} a^I_\mu \partial_\nu a^J_\lambda,
\end{align}
where $a^I$ for $I=1,...,r$ are compact $U(1)$ gauge fields and $K$ is a non-singular integer symmetric matrix. Here We've denoted the dimension of the K-matrix as $r=\text{dim}(K)$. The quasi-particles in this theory are labelled by $r$-component integer-valued vectors. The topological spin of the quasi-particle labelled by the integer vector $l$ is given by $\theta_l=\pi l^T K^{-1} l$, and the mutual statistics of two quasiparticles $l$ and $l'$ is $\theta_{ll'}=2\pi l^T K^{-1} l'$. A particle labeled by $l=K v$, with $v$ an $r$-component integer vector, is considered as a local particle in the theory. In this theory, we can further consider the condensation of a subgroup of the quasiparticles $M_{\text{Lag}}$, called a ``Lagrangian subgroup", which has the following properties\cite{levin2013ac,maissam2013ac1,maissam2013ac2}:
\begin{enumerate}
\item $e^{i\theta_m}=e^{i\theta_{mm'}}=1$, for all $m,m'\in M_{\text{Lag}}$;
\item For all $l\notin M_{\text{Lag}}$, $e^{i\theta_{lm}}\neq 1$ for at least one $m\in M_{\text{Lag}}$.
\end{enumerate}
The condensation of a Lagrangian subgroup drives a transition from the Abelian topological state to a topologically trivial phase. For each Lagrangian subgroup, there is an equivalent description of the condensate using another set of particles $M_\text{null}$\cite{levin2013ac,maissam2013ac1,maissam2013ac2}.
We refer to the particles in this set as null particles. This set $M_\text{null}$ is closed under particle fusion and satisfies the following conditions:
\begin{enumerate}
\item $\theta_m=\theta_{mm'}=0$, for all $m,m'\in M_{\text{null}}$;
\item $M_\text{null}$ is generated by $\text{rank}(M_\text{null})=r/2$ null particles $m_i$, $i=1,...,r/2$.
\end{enumerate}
Notice that for Lagrangian subgroup or the null particle set to exist, the dimension of the K-matrix $r$ has to be even. It has been proven that the Lagrangian subgroups $M_{\text{Lag}}$ and the null particle sets $M_{\text{null}}$ are in one-to-one correspondence to each other. \footnote{More rigorously, this correspondence only applies when one is allowed to topologically deform the system by adding topologically trivial blocks to the $K$ matrix.
More details were discussed in Ref. \onlinecite{levin2013ac,maissam2013ac1,maissam2013ac2}. }
The condensation of the particles of the null set $M_{\text{null}}$ will induce the condensation of all the particles in its corresponding Lagrangian subgroup $M_{\text{Lag}}$ modulo the local particles that can be always thought of as condensed, and vice versa.
Now we can start introducing the general setting of the layer construction of the $3$D topological states. First, we consider stacked layers of the identical $2$D Abelian topological states with the K-matrix $K$ for each layer. Before we introduce the coupling between the layers, we can view the system as a $2$D state using the $2+1$D Chern-Simons theory with an extended K-matrix $\mathcal{K}$ given by
\begin{align}
\mathcal{K}=K \otimes \left(
\begin{array}{ccc}
1 & & \\ & 1 & \\ & & \ddots
\end{array}
\right)_{L\times L},
\end{align}
where $L$ stands for the number of layers in the system. We can turn on the ``local" coupling between the layers by introducing the condensation of composite quasi-particles that lives in finite range of layers. Without the loss of generality, we only consider the composite particles that live in two consecutive layers as is shown in Fig. \ref{nullparticle} (a). These particles, which we label as $n_i^{(m)}$, take the form of
\begin{align}
n_i^{(m)}=p_i \otimes e_m+q_i \otimes e_{m+1}, \label{nullparticles}
\end{align}
where $p_i$'s and $q_i$'s are $r$-component integer vectors that labels the quasi-particle type in one single layer and $e_m$ is a $L$ component unit vector with the $m^\text{th}$ entry $1$ and all the rest $0$. The superscript $m$ is effectively the layer index, which will later be identified as the ``lattice coordinate" in the $z$ direction. For a system that is periodic in the perpendicular $z$ direction, we need to consider $m=1,2,...,L$ with the identification that $e_{L+1}=e_1$ while, for the open boundary condition in the $z$ direction, we only need to consider the condensation of null particles with $m=1,2,...,L-1$. The subscript $i$ labels the different types of the condensed quasi-particles $n_i^{(m)}$. It should be noticed that the condensed particles are completely determined by the sets $\{p_i\}$ and $\{q_i\}$ that are independent from the layer index $m$. Thus, the condensate of these particles is invariant under the cyclic permutation between the layer, namely under $m\rightarrow m+1$, and therefore is translationally invariant in the $z$ direction. For the sets $\{p_i\}$ and $\{q_i\}$, we require them to satisfy the following null conditions:
\begin{align}
& p^T_i K^{-1} p_j+ q^T_i K^{-1} q_j=0, \nonumber \\
& p^T_i K^{-1} q_j=0,~~~ \forall~i \text{ and }j,
\label{nullcondition}
\end{align}
such that the mutual statistics and the topological spins of the condensed particles $n_i^{(m)}$ is guaranteed to be trivial:
\begin{align}
n_i^{(m)T}\mathcal{K}^{-1} n_{i'}^{(m')}=0,~~~~\forall i,i',m,m'.
\end{align}
This condition is similar to the null conditions of the null particle set $M_\text{null}$ for $2$D Abelian topological state introduced above. Moreover, in analogy to the requirement to the rank of the null set $\text{rank}(M_\text{null})$, here we also require that
\begin{align}
\text{number of } p_i\text{'s} = \text{number of } q_i\text{'s} = \text{dim}(K)/2=r/2. \label{nullcounting}
\end{align}
That is to say the index $i$ takes value $i=1,2,...,r/2$. In this counting requirement, we have implicitly assume that for a given $m$, the choice of $\{p_i\}$ and $\{q_i\}$ guarantees that all particles $n_i^{{m}}$ for $i=1,...,r/2$ are linearly independent. Again, we only consider cases with even $r$. Here the choice of the sets $\{p_i\}$ and $\{q_i\}$ is subject to an equivalence relation. For any integer matrix $W$ with $\det(W)=1$, we can define new sets of $\{p'_i\}$ and $\{q'_i\}$ by the linear transformation $p'_j=\sum_i W_{ij} p_i$ and $q'_j=\sum_i W_{ij} q_i$ so that the condensed particle $\{n'^{(m)}_{i}\}$ defined by the new sets satisfy $n'^{(m)}_j=\sum_i W_{ij} n^{(m)}_i$. It is obvious that the condensation of $\{n'^{(m)}_{i}\}$ is equivalent to that of $\{n^{(m)}_{i}\}$. Therefore, the choice of sets $\{p_i\}$ and $\{q_i\}$ is defined modular a linear transformation by
an integer uni-determinat matrix $W$.
To summarize the general frame work of this construction, we start with layers of 2D Abelian topological states with each layer described by a K-matrix theory. Then, in order to couple the layer together to form a 3D topological state, we turn on the condensation of the particles of the form given in Eq. (\ref{nullparticles}) which is determined by the sets $\{p_i\}$ and $\{q_i\}$ that satisfy the null conditions in Eq. (\ref{nullcondition}) and the counting requirements in Eq. (\ref{nullcounting}).
Before we move on to more detailed analysis of the layer construction, a couple of comments are in order. Firstly, we argue that the layer construction indeed generate a genuine 3D state. The reasons for this is the following: 1) It is shown in Ref. \onlinecite{levin2013ac,maissam2013ac1, maissam2013ac2} that the condensation of the null particles can be induced by introducing coupling using only local variables. Here the locality is understood in the 2D sense; 2) We only condense composite particles that live in two consecutive layers. Therefore, the locality in the $z$ direction is also guaranteed; 3) The condensation of each quasi-particle is a phase transistion from one gapped state to another gapped state. Therefore, the resulting states of the layer construction should also be a gapped 3D state; 4) The requirements, including Eq. (\ref{nullcondition}) and (\ref{nullcounting}) on the defining data $K$, $\{p_i\}$ and $\{q_i\}$ do not depend on the number of layers $L$, and the number of condensed null particles $n_i^{m}$ is always equal to $dim(\mathcal{K})/2$. The 3D states constructed this way is stable in the thermodynamical limit and the remaining degrees of freedom under the condensation will not increase as $L\rightarrow \infty$. Secondly, these layer constructed 3D states do have topologically non-trivial properties. It is straightforward to see (from Eq. (\ref{nullcondition})) that, for the open boundary condition, the particle set $\{q_i\}$ ($\{p_i\}$) is deconfined on the top (bottom) surface as in shown in Fig. \ref{nullparticle} (b). Such 3D states can host 2D topological order on its surface, which is similar to the Walker-Wang model with a modular tensor category\cite{walker2012,simon2013}. Later, we will see that depending on the choice of $\{p_i\}$ and $\{q_i\}$, we can also have deconfined particles in the 3D bulk that organize themselves to form 3D topological orders. In the following sections, we will introduce rigorous results and examples of 3D topological state using the layer construction.
\begin{figure}[tb]
\centerline{
\includegraphics[width=3
in]{nullparticle-01.pdf}
}
\caption{\label{nullparticle} (a) Each line represents a layer of 2D Abelian topological state, and each brown circle represents a composite particle, the condensation of which will introduce the coupling between layers. Here each composite particle lives only in two consecutive layers, and the component of the particle in upper and lower layers is given by $p_i$'s and $q_i$, respectively. (b) For open boundary system in the $z$ direction, $q_i$'s ($p_i$'s) are deconfined particles on the top (bottom) surface which will form surface topological order. Depending on the choice of $\{p_i\}$ and $\{q_i\}$, there can be deconfined particles (represented by ``$X$") in the 3D bulk that organize themselves to form 3D topological orders.
}
\end{figure}
\section{3D states with purely surface topological order}
\label{pSTO}
In this section, we will first introduce a series of examples of 3D states with purely surface topological order (and no bulk topological order). Then we will discuss the general criteria for the construction of 3D states with purely surface topological orders and a Hamiltonian formalism for their surface topological orders.
\subsection{Example: Coupled Layers of the $Z_p$ Toric Code with Non-Trivial Surface Topological Order}
\label{EGpSTO}
Ref. \onlinecite{wang2013} provides a layer constructed model of 3D bosonic topologial insulartor with time-reversal invariant gapped surface states. Motivated by this example, we study the generalization of it and obtain a series of layer constructed models with interesting topologically ordered surfaces. Let's consider the stacked layers of $Z_p$ toric code theory. Each layer of $Z_p$ toric code theory can be described by the K-matrix theory\cite{Kitaev2003, wen2003,ReadSachdev1991} with
\begin{align}
K_{Z_p}=\left(\begin{array}{cc}
0 & p \\ p & 0
\end{array}\right).
\end{align}
There are two types of elementary quasi-particles in the $Z_p$ toric code: the electric particle $(1,0)^T$ (which will also be labelled as the $e$ particles) and the magnetic particle $(0,1)^T$ (which will also be labelled as the $m$ particles). In the stacked layer theory, the null particles we choose to condense take the following form:
\begin{align}
(1,1)^T\otimes e_n + (1,0)^T\otimes e_{n+1} + (1,-1)^T\otimes e_{n+2},
\label{chiralsurfcondensate}
\end{align}
which is the composite particles of $e+m$, $e$ and $e-m$ in three consecutive layers, as is shown in Fig. \ref{chiralsurface} (a). Although we consider composite particles that live in three consecutive layers, we can still recast this type of theory back to the general setting described in Sec. \ref{generalconstrtuction} by viewing two layers of $Z_p$ toric code as one single layer of 2D Abelian topological state. For simplicity of the discussion in this case, we will stick with the current description. One can check that with the particles of the form in Eq. (\ref{chiralsurfcondensate}) condensed, there is no deconfined particle in the 3D bulk when $p\equiv1,2 (\text{mod} 3)$. This can be directly verified by exhausting the possible forms of bulk particles and by noticing that all the particles that are not confined can be viewed as composite particles of the condensed ones. In Sec. \ref{GCpSTO}, we will introduce a more general argument which leads to the same conclusion when applied to this case. When $p\equiv1,2 (\text{mod} 3)$, the only deconfined particles live on the open surface. The surface deconfined particles take the form:
\begin{align}
\alpha_1 (1,-1)^T\otimes e_1 + \alpha_2 ((1,0)^T\otimes e_1+(1,-1)^T\otimes e_2),
\end{align}
where $\alpha_{1,2}=1,2,...,p$. Therefore, we can use a 2-component integer vector $\alpha=(\alpha_1,\alpha_2)$. The two generators of these deconfined surface particles are shown in Fig. \ref{chiralsurface} (b). The topological spin of the particle $\alpha$ can be calculated using the original K-matrix:
\begin{align}
\theta_\alpha &= -\frac{2\pi }{p}(\alpha_1^2+\alpha_1\alpha_2+\alpha_2^2). \nonumber \\
&=\pi \alpha^T M \alpha,
\end{align}
where the second line is written in a matrix form with
\begin{align}
M=-\frac{1}{p}\left(
\begin{array}{cc}
2 & 1 \\ 1 & 2
\end{array}
\right).
\end{align}
The braiding statistics of particle $\alpha$ and $\beta$ is given by
\begin{align}
\theta_{\alpha\beta} &= -\frac{2\pi }{p}(2\alpha_1 \beta_1+\alpha_1\beta_2+\alpha_2\beta_1+2\alpha_2 \beta_2) \nonumber \\
&=2\pi \alpha^T M \beta.
\label{SurfDeQP}
\end{align}
Since all the surface particles are Abelian, their quantum dimensions are $d_\alpha=1$. There are $p^2$ deconfined particles on the surface. We expect the surface topological order to have total quantum dimension $\mathcal{D}=p$. For a 2D topological state, the chiral central charge $c$, is related to the quantum dimension and topological spins by the following formula \cite{Zhenghan2010}:
\begin{align}
\mathcal{D}e^{\frac{2\pi i c}{8}}=\sum_\alpha d_\alpha e^{i\theta_\alpha}.
\label{centralcharge}
\end{align}
Therefore, we can calculate the chiral central charge for these surface topological orders with $\mathcal{D}=p$:
\begin{align}
c\equiv0 (\text{mod}8),~~~\text{for } p\equiv1 (\text{mod} 3), \\
c\equiv4 (\text{mod}8),~~~\text{for } p\equiv2 (\text{mod} 3).
\end{align}
From this result, we see that, for $p\equiv2 (\text{mod} 3)$, the surface topological state has to have chiral topological order. This result is quite non-trivial because we start with layers of $Z_p$ toric code that are non-chiral.
Similar situation can also be found using the Walker-Wang construction \cite{walker2012, simon2013}. Also, for the special case with $p=2$, the resulting chiral surface topological state is the so-called $Z_2$ three-fermion state with chiral central charge $c=4$. When realized in $2D$, the corresponding the K-matrix is the Cartan matrix of $SO(8)$:
\begin{align}
K_{SO(8)}=\left(
\begin{array}{cccc}
2 &-1& -1& -1 \\ -1 & 2 & 0& 0 \\ -1 & 0 & 2& 0 \\-1 & 0 & 0& 2
\end{array}
\right).
\end{align}
This specific state is also a realization of the symmetry preserving surface topological state of a time-reversal invariant bosonic topological insulator\cite{wang2013}. To our knowledge, the chiral topological order for the cases with $p>2$ are not well-studied before.
The case with $ p\equiv0 (\text{mod} 3)$ requires a more careful treatment. The set of quasi-particles that are deconfined on the surface is still given by Eq. (\ref{SurfDeQP}), but the naive application of Eq. (\ref{centralcharge}) will fail in this case. That is because other than the trivial particle $\alpha=(0,0)$, there are two other particles $(p/3,p/3)$ and $(2p/3,2p/3)$ that do not have any non-trivial braiding statistics with any other surface deconfined particles. Therefore, the surface theory is not modular if we include all $p^2$ particles label by $(\alpha_1,\alpha_2)$ with $\alpha_{1,2}=1,...,p$. The simplest way to remedy this problem is to take the quotient of the $p^2$ particles by the three ``local" particles $(0,0)$, $(p/3,p/3)$ and $(2p/3,2p/3)$. Then, we end up with a theory of $p^2/3$ particles, namely the total quantum dimension $\mathcal{D}=p/\sqrt{3}$. Using the formula Eq. (\ref{centralcharge}), we obtain that
\begin{align}
c\equiv -2 (\text{mod}8),~~~\text{for } p\equiv0 (\text{mod} 3).
\end{align}
Again, we see that the surface topological order is chiral. For the simplest case with $p=3$, the surface topological order of the quotient theory is the time reversal copy of the $U(1)_3$ Chern-Simons theory with the K-matrix given by
\begin{align}
K^{\text{surf}}_{p=3}=-\left(
\begin{array}{cc}
2 & 1 \\ 1 & 2
\end{array}
\right).
\end{align}
In fact, rather than just a mathematical trick, the quotient does carry physical meanings. The particles that are mod out from the surface theory are a subgroup of particles generated by $(p/3,p/3)$. Written in the stacked-layer theory language, $(p/3,p/3)$ takes the form:
\begin{align}
(2p/3,-p/3)^T\otimes e_1+ (p/3,-p/3)^T\otimes e_2.
\label{neutralparticle}
\end{align}
In fact, we notice that the particles $(2p/3,-p/3)^T\otimes e_n+ (p/3,-p/3)^T\otimes e_{n+1}$ for $\forall n$ are deconfined. Therefore, we should identify them as one type of bulk deconfined point particle (at different positions in the $z$ direction). This deconfined bulk particle will become the one of the constituents of the 3D bulk topological order. Therefore when we consider the surface topological order, this type of particles should not be included. In the next section, we will come back to this example to study its bulk topological order. To summarize this series of examples, we have obtain the surface topological order with
\begin{align}
&\mathcal{D}=\frac{p}{\sqrt{3}} \text{ and } c\equiv -2 (\text{mod}8),~~~\text{for } p\equiv0 (\text{mod} 3); \nonumber \\
&\mathcal{D}=p \text{ and } c\equiv0 (\text{mod}8),~~~\text{for } p\equiv1 (\text{mod} 3); \nonumber \\
&\mathcal{D}=p \text{ and } c\equiv4 (\text{mod}8),~~~\text{for } p\equiv2 (\text{mod} 3).
\end{align}
For $p\equiv1,2 (\text{mod} 3)$, the 3D bulk is a trivial 3D gapped state. For $p\equiv0 (\text{mod} 3)$, the bulk state has 3D topological order. It would be interesting to work out the effective K-matrix for the surface topological states for all $p$, especially the chiral ones, but this is beyond the scope of this paper and will be left for future works.
\begin{figure}[tb]
\centerline{
\includegraphics[width=3
in]{chiralsurface-01.pdf}
}
\caption{\label{chiralsurface} (a) Illustration of the condensation of the composite particle of $e+m$, $e$ and $e-m$ in three consecutive layers, with each layer a $Z_p$ toric code state. (b) For $p\equiv 1,2 \text{ mod }3$, there is no deconfined particle in the 3D bulk. The deconfined particles only stay on the surface, which are generated by
two elementary particles shown in the green circles.
(The case of $p\equiv 0 \text{ mod }3$ is discussed later in Fig. \ref{coexTopo}.)
}
\end{figure}
\subsection{General Criteria for 3D States with Purely Surface Topological Order}
\label{GCpSTO}
In Sec. \ref{EGpSTO}, we've given a series of examples (for $p\equiv1,2 (\text{mod} 3)$) of 3D states with 2D surface topological order and trivial 3D bulk topological order using the layer construction. In this part of the discussion, we will provide a general criteria for the construction of 3D states with purely surface topological order and trivial bulk topological order.
We denote the quasi-particle lattices generated by $\{p_i\}$, $\{q_i\}$ and $\{p_i\}\cup\{q_i\}$ as $\Gamma_{\{p_i\}}$,$\Gamma_{\{q_i\}}$ and $\Gamma_{\{p_i\}\cup\{q_i\}}$, and the local-particle lattice generated by the column vectors of the K-matrix $K$ as $\Gamma_K$. Notice that the lattice of all quasi-particles is given by the integer lattice $\mathbb{Z}^r$. Then we have the following theorem:
\begin{itemize}
\item If $\Gamma_{\{p_i\}\cup\{q_i\}}=\mathbb{Z}^r \textnormal{ modulo }\Gamma_K$ and the overlap of $\Gamma_{\{p_i\}}$ and $\Gamma_{\{q_i\}}$ is trivial, i.e. $\Gamma_{\{p_i\}} \bigcap \Gamma_{\{q_i\}}=\emptyset \textnormal{ modulo }\Gamma_K $, the resulting 3D states from the layer construction have only 2D surface topological order and trivial 3D topological order.
\end{itemize}
One can check that the examples discussed in the previous subsection satisfy the condition of this theorem when $p=1,2$ and, therefore, only admits purely surface topological order. The proof of this theorem is the following. Suppose we start with a stack of $L$ layers and turn on the coupling between the layers (assuming open boundary condition in the $z$ direction) to form a 3D gapped state, as described in Sec. \ref{generalconstrtuction}. As is discussed in Sec. \ref{generalconstrtuction}, the 3D states from the layer construction always havej deconfined surface particle if we consider open boundary condition. Therefore we only need to prove that there is no non-trivial deconfined excitation in the bulk and thus no bulk topological order. If there is a deconfined particle living in the $m_1^\text{th}$ layer and the $m_2^\text{th}$ layer in the bulk, we can always denote it as $X_{m_1,m_2}= \sum_{m=m_1}^{m_2} x_m \otimes e_m$ with $1 \leq m_1 \leq m_2 \leq L$. Without the loss of generality, we can also assume that in this notation $x_m\neq 0$. There are 4 possible situations that need to be discussed separately: i) $1 < m_1 \leq m_2 < L$; ii) $1 = m_1 \leq m_2 < L$; iii) $1 < m_1 \leq m_2 = L$ and iv) $(m_1,m_2)=(1,L)$. We are going to rule out the 4 possibilities one by one.
For the situation i), if there exist a deconfined particle $X_{m_1,m_2}$ with $1<m_1,m_2<L$ for the open boundary (in the $z$ direction) system, this particle will stay deconfined even when one introduces the the coupling between the top and bottom layers by the condensation of the composite particles $p_i\otimes e_L+ q_i\otimes e_1$, $i=1,...,r/2$. However, since $\Gamma_{\{p_i\}} \bigcap \Gamma_{\{q_i\}}=\emptyset \textnormal{ modulo }\Gamma_K $, all condensed null particles $\{n_i^{(m)}\}_{i=1,..,r/2}^{m=1,...,L}$ for the periodic system are linearly independent. The number of these condensed null particles is equal to $dim(\mathcal{K})/2$. Thus, from the result in Ref. \onlinecite{levin2013ac,maissam2013ac1,maissam2013ac2}, there should be no deconfined particle (other than the condensed ones) left in the condensate. Therefore, deconfined particles $X_{m_1,m_2}$ with $1<m_1,m_2<L$ should not exist.
For the situation ii), we consider deconfined particles of the form $X_{1,m2}$ with $m_2<L$. The trivial statistics between $X_{1,m_2}$ and the condensed particles, especially $n_i^{(m_2)}$ (see Fig. \ref{braidproof}), implies that $x_{m_2}$, as a single layer particle, braids trivially with all the $p_i$'s. Since $\Gamma_{\{p_i\}\cup\{q_i\}}=\mathbb{Z}^r \textnormal{ modulo }\Gamma_K$, we can expand $x_{m_2}$ using in the basis of $\{p_i\}\cup\{q_i\}$, namely $x_{m_2}=\sum_i (a^{(m_2)}_i p_i + b^{(m_2)}_i q_i)$.
We will show that the $\sum_i a^{(m_2)}_i p_i$ part of this expansion is 0 modulo local particles. Since $x_{m_2}$ is a single layer particle that has trivial braiding statistics with all the $p_i$'s and all the $q_i$'s also braids trivially with the $p_i$ particles in the single theory (see Eq. (\ref{nullcondition})), the single layer particle $\sum_i a^{(m_2)}_i p_i$ should also braid trivially with all the particle $p_i$'s. Therefore, from Eq. (\ref{nullcondition}), we notice that the particles $(\sum_i a^{(m_2)}_i p_i)\otimes e_m$ with $\forall m$ are deconfined even for the system with periodic boundary condition. As is explained in the previous situation, this is impossible unless $(\sum_i a^{(m_2)}_i p_i)\otimes e_m$'s are local particles.
Given this, we can consider the fusion between the particle $X_{1,m_2}$ with the condensed particle $-\sum_i b^{(m_2)}_i n_i^{(m_2-1)}$, which does not change the nature of $X_{1,m_2}$ in the condensate. Now notice that $X_{1,m_2}-\sum_i b^{(m_2)}_i n_i^{(m_2-1)}$ is effectively a particle that lives in between the $1^\text{st}$ and $m_2-1^\text{th}$ layers, therefore can be identified as $\tilde{X}_{1,m_2-1}$. By iterating this argument, we can conclude that all deconfined quasi-particles for the situation ii) are equivalent to quasi-particles $X_{1,1}$ living at the top surface combined with a chain of condensed particles. Further, all the deconfined particles $X_{1,1}$ have to be decomposed into linear combinations of $q_i$'s on the top surface. Using similar analysis, we can also show, for the situation iii), that the only possible non-trivial deconfined quasi-particles for the situation ii) are the $X_{L,L}$'s and, therefore, are quasi-particles that live only on the bottom surface. Also, all the deconfined particles $X_{L,L}$ have to be decomposed into linear combinations of $p_i$'s on the bottom surface.
For the situation iv), we focus on the deconfined particles of the form $X_{1,L}$. As is discussed above, we can decompose the $L^\text{th}$ layer component $x_L$ of the $X_{1,L}$ into $x_{L}=\sum_i (a^{(L)}_i p_i + b^{(L)}_i q_i)$, so there exists a deconfined particle $X'_{1,L-1}$ such that $X_{1,L}=X'_{1,L-1}+ \sum_i b^{(L)}_i n_i^{(L-1)} + \sum_i a^{(L)}_i p_i \otimes e_L$. One can view this as the definition of $X'_{1,L-1}$. The deconfinement of $X'_{1,L-1}$ follows from the fact that $X_{1,L}$ and $\sum_i a^{(L)}_i p_i \otimes e_L$ deconfines in the open system, and $\sum_i b^{(L)}_i n_i^{(L-1)}$ is condensed. For the deconfined particle $X'_{1,L-1}$, the discussion on the situation ii) already shows that, modulo the condensed particles, $X'_{1,L-1}$ should be identified as $\tilde{X}'_{1,1}$, and therefore as a deconfined particle on the top surface. Thus, the deconfined particle $X_{1,L}$ can always be identified as a trivial composition of the a top surface deconfined particle $\tilde{X}'_{1,1}$ and a bottom surface deconfined particle $\sum_i a^{(L)}_i p_i \otimes e_L$.
Combining the discussion from situation i) to iv), we conclude that, when $\Gamma_{\{p_i\}} \bigcap \Gamma_{\{q_i\}}=\emptyset \textnormal{ modulo }\Gamma_K $ and $\Gamma_{\{p_i\}\cup\{q_i\}}=\mathbb{Z}^r \textnormal{ modulo }\Gamma_K$, the layer constructed 3D state only admits non-trivial deconfined quasi-particles on its surface, and therefore only admits surface topological order. The absence of deconfined bulk quasi-particle indicates the trivialness of the topological properties in the 3D bulk. Also, the discussion of situation ii) and iii) further shows that, with open boundary condition, the only possible excitation on the top (bottom) surface is given by the particle set $\Gamma_{\{q_i\}}$ ($\Gamma_{\{p_i\}}$). Therefore, the surface topological order on the top (bottom) surface is completely determined by the set $\Gamma_{\{q_i\}}$ ($\Gamma_{\{p_i\}}$) and the braiding statistics within it.
\begin{figure}[tb]
\centerline{
\includegraphics[width=1.5
in]{braidproof-01.pdf}
}
\caption{\label{braidproof} In the discussion of situation ii) (see text), we consider the braiding between a deconfined particle $X_{1,m2}$ (green circle) with a condensed null particle $n_i^{(m_2)}$ (brown circle).
}
\end{figure}
\subsection{Surface Topological Orders on Different Surfaces}
\label{surfaceOrder}
Having introduced the general criteria for 3D states with purely surface topological order from the layer construction, we will study the surface topological order on the top, bottom and, especially, the side surface in greater details in the following. As a method to build $3$D gapped states, the layer construction is strongly anisotropic. This anisotropy will manifest itself when we consider a layer-constructed system with finite size in both $z$ and $y$ directions (Remember that the $x$ and $y$ directions are equivalent in our description). The side surface appears to be significantly different from the top and bottom surfaces. We will show that, despite of the superficial distinction among this surfaces, the surface topological orders on them are, in fact, equivalent. The implication of this result will be that the 3D states resulted from the layer construction is topologically equivalent to an isotropic 3D gapped phase. Our discussion will still focus on the case with purely surface topological order.
We start by comparing the surface topological order on the top and bottom surfaces for the system with open boundary along the $z$ direction. As is discussed in Sec. \ref{GCpSTO}, quasi-particles of the surface topological order is given by $\Gamma_{\{q_i\}}$ for the top surface and by $\Gamma_{\{p_i\}}$ for the bottom surface. From the null condition $-p^T_i K^{-1} p_j= q^T_i K^{-1} q_j, \forall i,j$ in Eq. (\ref{nullcondition}), we can establish a one-to-one correspondence between the sets $\Gamma_{\{q_i\}}$ and $\Gamma_{\{p_i\}}$, such that the braiding statistics in the set $\Gamma_{\{q_i\}}$ is exactly the conjugate of that in $\Gamma_{\{p_i\}}$. The braiding statistics discussed here is calculated with respect to a certain normal direction of the layers, which is the $+z$ direction in this case. The $+z$ direction is a natural choice of the normal direction of the top surface, but is opposite to the normal direction of the bottom surface induced by the bulk. Consequently, taking the change of normal direction into account, we need to calculate the braiding statistics using the opposite K-matrix for the two surfaces. Therefore, the null condition
\begin{align}
-p^T_i K^{-1} p_j= q^T_i K^{-1} q_j, \forall i,j
\end{align}
is exactly the condition for the equivalence between the top and bottom surface topological orders.
Having establish the equivalence of top and bottom surfaces, we can move onto the discussion of the surface topological order on the side surface. First, let's briefly review the edge theory of a 2D Abelian topological state with a K-matrix $K$ and its behavior under quasi-particle condensation. The edge state of a 2D Abelian topological state with the K-matrix $K$ is described by the $1+1$D chiral Luttinger liquid theory\cite{wen1990edge}:
\begin{align}
\mathcal{L}_{\text{CL}}=\frac{K_{IJ}}{4\pi} \partial_x \phi^I \partial_t \phi^J- V_{IJ} \partial_x \phi^I \partial_x \phi^J,
\end{align}
where $V$ the velocity matrix. Suppose the 2D bulk is in the condensed phase of a set of particles $\{n_i\}$ that satisfies $n_i^T K^{-1} n_j=0, \forall i,j$. The condensation is induced by turning on extra coupling terms in chiral Luttinger liquid on the edge\cite{levin2013ac,maissam2013ac1,maissam2013ac2}:
\begin{align}
\delta\mathcal{L}_\text{cond}= \sum_i -g_i \cos(c_i n_i^T \phi),
\end{align}
where $g_i$'s are the coupling constants and $c_i\in \mathbb{Z}$ is the minimal integer, for each $i$, such that $e^{i c_i n_i^T \phi}$ only create/annihilate a local particle. Here we've already organized the $\phi_I$'s into a column vector $\phi$. Since $e^{i c_i n_i^T \phi}$ is a local operator, the $\delta\mathcal{L}_\text{cond}$ only involves local terms and, therefore, can be written as a local coupling using the microscopic degrees of freedom. Notice that the commutation relations between $n_i^T\phi$'s (with different $i$'s) are trivial in the chiral Luttinger liquid theory:
\begin{align}
[n_i^T\phi, n_j^T\phi]=0,~~~\forall~i,j.
\end{align}
Therefore, in the deep condensation limit with $g_i\rightarrow \infty$, the coupling term $\delta \mathcal{L}_\text{cond}$ force the fields $n_i^T\phi$ to develop non-zero vacuum expectation values:
\begin{align}
\langle n_i^T\phi\rangle \neq 0.
\end{align}
If the number of fields $n_i^T\phi$ equals $\text{dim}(K)/2$, then edge theory becomes fully gapped in the condensed phase.
Now we will use this language to describe the topological order on the side surface of the layer-constructed 3D state with purely surface topological order. Before we introduce the condensation that couples the layers, the side surface of the stacked layer is described by the Lagrangian density:
\begin{align}
\mathcal{L}_{\text{CL}}^\text{side}=
\sum_m\frac{1}{4\pi} \left(\partial_x \Phi^{(m)}\right)^T K \left( \partial_t \Phi^{(m)} \right) \nonumber \\
-\left(\partial_x \Phi^{(m)}\right)^T V \left( \partial_x \Phi^{(m)} \right),
\label{SurfChiralLuttinger}
\end{align}
where $\Phi^{(m)}$ is the $r$-component (remember $r=\text{dim}(K)$) chiral boson field on the edge of the $m^\text{th}$ layer and $V$ is the velocity matrix for each layer. By quantizing this theory, we obtain the commutation relation of the boson fields:
\begin{align}
[\partial_x \Phi^{(m)}_I(x), \Phi^{(n)}_J(y)]=-i2\pi (K^{-1})_{IJ} \delta(x-y) \delta_{n,m}.
\end{align}
The condensation of the particles $\{n^{(m)}_i\}$ that couple the layers also introduces the coupling in the chiral edge mode:
\begin{align}
\delta\mathcal{L}_\text{cond}= \sum_{i,m} - g_{m,i} \cos\left(l_i (p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)})\right),
\label{SurfCoupling}
\end{align}
where $g_{m,i}$'s are the coupling constants, and $l_i\in \mathbb{Z}$ the minimal integer which ensure the locality of the operator $e^{il_i (p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)})}$. That is to say $e^{il_i (p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)})}$ is the creation/annihilation operator of a local particle. This means that the layer construction can, in principal, be implemented by coupling the microscopic degrees of freedom without using non-local terms.
From the canonical quantization of the chiral Luttinger theory Eq. (\ref{SurfChiralLuttinger}) and the null condition Eq. (\ref{nullcondition}), we see that the fields in the $\cos$ terms have trivial commutation relations:
\begin{align}
[(p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)}),(p_j^T \Phi^{(n)} +q_j^T \Phi^{(n+1)})]=0,~~\forall i,j,m,n.
\end{align}
Therefore, at the strong coupling limit $g_{m,i}\rightarrow \infty$, the fields $(p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)})$ will develop non-trivial vacuum expectation value, namely
\begin{align}
\langle p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)} \rangle \neq 0.
\end{align}
In the following, we will denote the field $p_i^T \Phi^{(m)} +q_i^T \Phi^{(m+1)}$ as $\Lambda_i^{m+1/2}$.
Now we can use this formalism to study the topological order on the side surface. From the Lagrangian density in Eq. (\ref{SurfChiralLuttinger}) and Eq. (\ref{SurfCoupling}), we notice that we are in fact solving a problem of coupled array of 1D wires. Similar coupled wire systems that exhibits 2D topological order were studied in Ref. \onlinecite{sondhi2001, kane2002, teo2014}. The building blocks of these systems are 1D wires of normal Luttinger liquids. For our problem, we are considering the coupled arrays of fractionalized 1D wires which are the edge states of the 2D Abelian topological orders, as is shown in Fig. \ref{sideSurf} (a).
\begin{figure}[t]
\centerline{
\includegraphics[width=3.5
in]{sideSurf-01.pdf}
}
\caption{\label{sideSurf} (a) The condensation in the 3D bulk induces a coupling between the edge states of each layer. The side surface is effectively a coupled wire system. In the strong coupling limit, the vacuum expectation values $\langle \Lambda_i^{m+1/2}\rangle_0$ are pinned to the minima determined by the $\cos$ terms in Eq. (\ref{SurfCoupling}). The coloring on the right hand side indicates non-trivial, but uniform $\langle \Lambda_i^{m\pm 1/2}\rangle_0$ on the side surface. (b) The operator $e^{i p^T_j{\Phi^{m}(x)}}$ creates a collection of kinks (depicted as the change in color) in $\langle \Lambda_i^{m+1/2}(x)\rangle$, which will be identified as a topological quasi-particle $w_j$ on the $m+1/2^\text{th}$ layer of the side surface. (c) The operator $\chi_j^m$ creates a kink-anti-kink pair
that effectively tunnels
the quasi-particle $w_j$ from the $m-1/2\text{th}$ layer to the $m+1/2\text{th}$ layer.
}
\end{figure}
\begin{figure}[t]
\centerline{
\includegraphics[width=2.5
in]{surfbraid-01.pdf}
}
\caption{\label{surfbraid} This figure illustrates the braiding between two quasi-particles of the surface topological order on the side surface. The blue dot represent the quasi-particle $w_j$ created by the operator $e^{ip_j^T \Phi^(m)}$. The red line represents the trajectory of the second quasi-particle $w_k$ that the first one braids with. The operators that tunnel $w_k$ along the trajectory is explained in the main text. Especially, the vertical tunnel operator from the ${m'}^\text{th}$ layer to the ${m'+1}^\text{th}$ is given by $\chi_k^{m'}$.
}
\end{figure}
To identify the topological order on the side surface, we need to identify the non-trivial quasi-particles and their statistics. Deep in the condensed phase, the fields $\Lambda_i^{m+1/2}$ only fluctuate very weakly around its vacuum expectation value which are pinned by the $\cos$ terms in Eq. (\ref{SurfCoupling}). In fact, the $\cos$ terms create multiple degenerate minima for the fields $\Lambda_i^{m+1/2}$. The domain walls, or kinks, of $\langle \Lambda_i^{m+1/2}\rangle$ between different minima are topologically stable excitations and will be identified as topological quasi-particles on the side surface. We can start from the configuration with all $\langle \Lambda_i^{m+1/2}\rangle_0$ uniform. The operator $e^{i p^T_j{\Phi^{(m)}(x)}}$ creates a series of kinks in $\langle \Lambda_i^{m+1/2}\rangle$, $i=1,...,r/2$ (see Fig. \ref{sideSurf} (b)), such that the vacuum expectation values of $\langle \Lambda_i^{m+1/2} (x)\rangle$ on the right and left sides of $e^{i p^T_j{\Phi^{(m)}(x)}}$ differ by
\begin{align}
\Delta\Lambda_{ij}=2\pi p_i^T K^{-1} p_j.
\end{align}
We will identify the $e^{i p^T_j{\Phi^{(m)}(x)}}$ as the creation operator of a quasi-particle, denoted $w_j$, on the side surface. Since the kinks created by $e^{i p^T_j{\Phi^{m}(x)}}$ are the domain walls of $\langle \Lambda_i^{m+1/2} (x)\rangle$, we would like to think of this quasi-particle $w_j$ as residing in the $m+1/2^\text{th}$ layer. The quasi-particle $w_j$ can tunnel between the $m-1/2^\text{th}$ layer and $m+1/2^\text{th}$ layer, which is implemented by the tunneling operator:
\begin{align}
\chi_j^{m}=e^{i (p^T_j+q^T_j){\Phi^{m}(x)}}.
\end{align}
The operator $\chi_j^{m}$ creates simultaneously a kink in $\langle \Lambda_i^{m+1/2} (x)\rangle$ and an anti-kink in $\langle \Lambda_i^{m-1/2} (x)\rangle$ as is shown in Fig. \ref{sideSurf} (c). It is straightforward to show that, in the strong coupling limit,
\begin{align}
\chi_j^{m+1} e^{i p^T_j{\Phi^{(m)}(x)}} = e^{i \langle \Lambda_j^{m+1/2} (x)\rangle} e^{i p^T_j{\Phi^{(m+1)}(x)}},
\label{TunnelPhase}
\end{align}
which means that the operator $\chi_j^{m}$ tunnels the quasi-particle $w_j$ from the $m-1/2^\text{th}$ layer to the $m+1/2^\text{th}$ layer and, more importantly, with a phase factor $e^{i \langle \Lambda_j^{m+1/2}(x)\rangle} $. We will show that this phase factor will give rise to the braiding statistics of the quasi-particles on the side surface. We can consider the configuration with a quasi-particle $w_j$ on the $m+1/2^\text{th}$ layer created by the operator $e^{ip_j^T\Phi^{(m)}(x)}$, as is shown in Fig. \ref{surfbraid}. A close path of the quasi-particle $w_k$ is depicted as the red line. The horizontal tunneling of the quasi-particle $w_k$ in the uniform backgrounds of $\langle \Lambda_i^{m-1/2} (x)\rangle_0$ and $\langle \Lambda_i^{m+3/2} (x)\rangle_0$ does not produce any non-trivial phase factor. The vertical tunneling between the layers through the operator $\chi^{m'}_k$'s, as is shown in Eq. (\ref{TunnelPhase}), does carry non-trivial phase factors. Due the existence of the kinks in $\langle \Lambda_i^{m+1/2} (x)\rangle$, the net phase, which is viewed as the braiding phase between the quasi-particle $w_j$ and $w_k$, is non-zero and is given by:
\begin{align}
B_{jk}= e^{-i\Delta\Lambda_{jk}} =e^{-2\pi i p_j^T K^{-1} p_k}.
\end{align}
Therefore, by identifying the quasi-particle $w_i$ on the side surface with the deconfined particle $p_i$ on the bottom surface, we conclude that topological order on the side surface is the same as the one on the bottom surface. Furthermore, by noticing that the operators $e^{ip_j^T\Phi^{(m)}(x)}$ and $e^{-iq_j^T\Phi^{(m+1)}(x)}$ create exactly the same kink configuration on the side surface, we can also show the equivalence of the side surface and the top surface through the same analysis.
\section{3D Topologically Ordered States}
\label{3DTO}
In this section, we will focus on the layer construction of 3D topologically ordered states. In Sec. \ref{GCpSTO}, we have introduced the general criteria for the construction of a 3D gapped state with trivial bulk topological order. When this criterion is not satisfied, we will generically have bulk deconfined excitations, which will organize themselves to form 3D topological order. Unlike the topological order in 2D which is fully characterized by the modular tensor category theory, the general structure of 3D topological order is not yet known completely. The layer construction provides a method to construct 3D topological ordered gapped states that will shed light on the potential structure for a generic 3D topological order. In this section, we will consider three types of different examples of three dimensional topological order. Firstly, we will introduce the construction of conventional topological order with particle-string mutual statistics that resembles the $Z_p$ lattice gauge theory in 3 spatial dimensions. Secondly, we will construct an example with 3D topological order with coexisting surface topological order. Thirdly, we will consider a more non-trivial example with braiding statistics not only between particles and strings but also between different types of strings.
\subsection{Conventional 3D Topological Order}
\label{Convent3DTO}
To construct 3D gapped state with the conventional 3D topological order of 3D $Z_p$ lattice gauge theory, we start with layers of $Z_p$ toric code with the K-matrix:
\begin{align}
K_{Z_p}=\left(
\begin{array}{cc}
0 & p \\ p & 0
\end{array}
\right).
\end{align}
We choose
\begin{align}
p_1=(1,0)^T,~~~q_1=-(1,0)^T.
\end{align}
The condensed particles are illustrated in the Fig. \ref{LGT3d} (a). This condensation can be understood as an ``exciton condensate", in which the electric particle in each layer, i.e. $(1,0)\otimes e_m ~\forall m$, deconfined.
\begin{align}
\epsilon_e = (1,0)^T\otimes e_m.
\end{align}
This electric particle $\epsilon_e$ can hop between the layer through the fusion with the condensed ``excitons". Since spin of the electric particle $\epsilon_e $ is $0$, it is a boson in the 3D system. Besides the deconfined electric particle, the other type of deconfined excitation is the string excitation which is composed of the magnetic particle $m$ in the each layer, i.e
\begin{align}
\epsilon_{m\text{-string}}=\sum_m (0,1)^T\otimes e_m,
\end{align}
as is shown in Fig. \ref{LGT3d} (b). So far, we've only considered the string excitations that orient along the $z$-direction. String excitations can in fact take, especially in the continuum limit, any orientations with their excitation energy proportional to their length. We can consider contractible strings of finite size as depicted in Fig. \ref{LGT3d} (b).
From the K-matrix, it is easy to see that the braiding statistics between the electric particle and the magnetic string is
\begin{align}
\theta_{e,m-\text{string}}=2\pi/p.
\label{EMstat}
\end{align}
This statistical phase is independent from the location of the point-like particle $\epsilon_e$. By comparing with the 3D $Z_p$ lattice gauge theory, the deconfined electric particle can be identified as the electric particle in the $Z_p$ gauge theory, while the magnetic string can be identified as the magnetic flux string. These two types of deconfined excitation and their mutual statistics give rise to the topological order in 3D. In this example, the layer construction studies the anisotropic limit of the $Z_p$ gauge theory in which the $m$-string always orient itself along the $z$-direction.
The string excitation can actually take arbitrary shape, and its energy is always linearly proportional to its length. The statistics between the point particles and string excitations should not depend on the shape of the string. Therefore, the layer construction faithfully captures the topological properties of 3D lattice gauge theory.
Now, we can consider a variation of this case with a topological twist in each layer. For the convenience of the discussion, we will assume that $p$ is a prime number such that $Z_p$ is a field, namely $s^{-1}\in Z_p, \forall s \in Z_p $. In the following, we will use $s^{-1}$ to denote the integer whose product with $s$ is 1 modulo $p$. For a given $s\in Z_p$, we consider the condensate defined by
\begin{align}
p_1=(1,0)^T,~~~q_1=-(s,0)^T.
\end{align}
The condensed particles in this case are shown in Fig. \ref{LGT3d} (c). It is easy to see that the $e$ particle in each layer is still deconfined. It will be identified as a type of point particle in the 3D state. The hopping of $e$ particles between layers is still implemented through the fusion with the condensed particles. We see that in this hopping process an $e$ particle of the $m^\text{th}$ layer turns into a $se$ particle of $m+1^\text{th}$ layer. We can understand this process as a hopping of the $e$ particle followed by a symmetric twist. To be more precise, for $Z_p$ toric code theory, the topological properties of the theory, namely the braiding statistics, fusion rules, are invariant under the map generated by $e\rightarrow se, ~m\rightarrow s^{-1}m$. This mapping is a symmetry of the $Z_p$ toric code as far as the topological properties are concerned. We will call this symmetry action as the symmetric twist. The hopping process in this case is the hopping of the $e$ particle under one layer to the other followed by the symmetric twist $e\rightarrow se, ~m\rightarrow s^{-1}m$.
Based on this understanding, one should expect that the topological order obtained in the case is the same as the untwisted case. Indeed, in this twisted case, there is a twisted version of the $m$-string (see Fig. \ref{LGT3d} (d)) which is given by
\begin{align}
\epsilon'_{m\text{-string}}=\sum_m (0,s^{-m})\otimes e_m.
\end{align}
We pick the unit of point-like excitation and string-like excitation to be the ones shown in Fig. \ref{LGT3d} (d). Their mutual braiding statistics is given by
\begin{align}
\theta'_{e,m-\text{string}}=2\pi/p.
\end{align}
This is the same as the untwisted case. Also, we see that the hopping of the $e$-particle between the layers is compatible with its braiding with the twisted $m$-string. So far, the analysis indicates that the local properties of the twisted case are the same as those of the untwisted one. However, these analysis is done for an infinite system or a system with open boundaries. When we consider periodic boundary condition in the $z$-direction, the twisted case automatically generates a monodromy when we consider the hopping of the electric particle along the non-trivial cycle along the $z$-direction. To be more precise, an electric particle $e$ becomes the particle $s^L e$ after traversing along the $z$-direction cycle, where $L$ is the number of layers in the system. The twist generated by the condensed particle particle can be undone locally by considering the symmetric twist generated by $e\rightarrow s^{-l}e$ and $m\rightarrow s^{l}m$ for the $l^\text{th}$ layer. Then the system with a periodic boundary condition is equivalently described by the stacked layers of $Z_p$ toric code with following condensed particles: $(1,0)^T\otimes e_m+(1,0)^T\otimes e_m+1$ with $m=1,..,L-1$ and $(s^{-L},0)^T\otimes e_L+(1,0)^T\otimes e_1$. The system is locally equivalently to the untwisted case, and the twist along the $z$-direction cycle is generated by the coupling between the $1^\text{st}$ and the
$L^\text{th}$ layer induced by the condensation of the particle $(s^{-L},0)^T\otimes e_L+(-1,0)^T\otimes e_1$. Therefore we can think of the twisted model as the conventional 3D $Z_p$ gauge theory with 2D membrane twist defect between the $1^\text{st}$ and the $L^\text{th}$ layer. (The notation of twist defects in 2D and their properties are first studied in Ref. \onlinecite{KongTD2012,maissamTD2013}. Its generalization in 3D is first studied in Ref. \onlinecite{YingTD2013}.) This membrane defect implements the symmetry twist $e \rightarrow s^L e$ in the 3D $Z_p$ gauge theory. Let's look at the ground state degeneracy (GSD) of this twisted model on $T^3$. Assume that we also have periodic boundary conditions along $x$ and $y$ directions. For the untwisted case, we can study the Wilson loop operators $W_{x}$, $W_{y}$ and $W_{z}$ that tunnel the $e$ particle around the non-trivial cycles along the $x$, $y$ and $z$ directions. The Wilson loop operators measure the magnetic flux in the non-trivial cycles and can take $p$ different values $e^{i2\pi n/p}$, $n=1,...,p$ for $Z_p$ gauge field. Therefore, the untwisted system has $p^3$ degenerate ground states on $T^3$. For the twisted case, $W_z$ is not well-defined because the twist membrane render the tunneling a cycle around the $z$ direction un-closed unless $s^L\equiv 1 (\text{mod }p)$. $W_x$ and $W_y$ are still well-defined. However, if we adiabatically move the tunneling path of $W_{x(y)}$ all together one cycle along the $z$ direction, the twist membrane generates a map: $W_{x(y)} \rightarrow W^{s^L}_{x(y)} $. This implies that the flux along the $x(y)$ direction measured by $W_{x(y)}$ should be invariant under map, and therefore can only take the flux value 1 unless $s^L\equiv 1 (\text{mod }p)$. Therefore, when $s^L\not\equiv 1 (\text{mod }p)$, the GSD on $T^3$ is 1. For $s^L\equiv 1 (\text{mod }p)$, the twist membrane only perfoms a trivial twist which means the GSD on $T^3$ is the same as the untwisted case. These results are summaried as follows:
\begin{align}
&\text{GSD on } T^3=1, \text{ for } s^L\not\equiv 1 (\text{mod }p), \nonumber \\
&\text{GSD on } T^3=p^3, \text{ for } s^L\equiv 1 (\text{mod }p).
\end{align}
Therefore, we see that the twisted model has a different behavior of the ground state degeneracy on $T^3$ from the untwisted case.
\begin{figure}[t]
\centerline{
\includegraphics[width=3.5
in]{LGT3d-01.pdf}
}
\caption{\label{LGT3d} (a) The condensation of $e$ and $-e$ composite in the two consecutive layers in the system of stacked layers of $Z_p$ toric codes. (b) The electric particle in each layer is a deconfined particle that should be identified as the electric particle in the 3D $Z_p$ lattice gauge theory. The string of $m$ particles in every layer is a deconfined string excitation which can be viewed as the flux string in the $Z_p$ gauge theory. The flux strings don't have to align strictly along the $z$-direction. Contractible flux string of finite size, as is shown in the figure, can be considered. (c) A different set of condensed particles in layers of $Z_p$ toric code, the condensation of which leads to the twisted lattice gauge theory (see text). (d) The deconfined particles in the twisted theory, including an electric particle and the ``twisted" flux string.
}
\end{figure}
\subsection{Coexisting Bulk and Surface Toploogical Order}
\label{Coexist}
In Sec. \ref{EGpSTO}, we discussed a construction using layers of $Z_p$ toric codes with the condensed particles given by Eq. (\ref{chiralsurfcondensate}) (see Fig. \ref{chiralsurface} (a)). As is explained there, for $p$ dividable by $3$,
there are two particles which have trivial statistics (except the vacuum particle) with all the rest of surface deconfined particles. One of them can be written as
\begin{align}
(p/3,p/3)^T \otimes e_1 + (-p/3,p/3)^T \otimes e_2.
\end{align}
The other particle is a two-particle bound state of this one. (This form is essentially equivalent to the expression given in Eq. (\ref{neutralparticle}).) After the quotient of all the surface deconfined particles by these two particles, we can consistently calculate the central charge of the surface topological order. This implies that these two trivial particles shouldn't be considered as part of the surface topological order. In fact, one can check that particles of the form:
\begin{align}
\epsilon_\text{pt}=(p/3,p/3)^T \otimes e_m + (-p/3,p/3)^T \otimes e_{m+1}
\end{align}
are all deconfined. Therefore, this particle should be viewed as a bulk deconfined point particle, which is a generalization of the bulk deconfined electric particle in the previous example. Again, the spin of the particle $\epsilon_\text{pt}$ is 0, so it is a boson in the 3D system. This is consistent with the fact that any point particle in 3D should have either bosonic or fermionic statistics. Similar to the previous example, there is also a type of deconfined string excitation that takes the simple form of
\begin{align}
\epsilon_\text{str}=\sum_m (1,0)^T \otimes e_{2m},
\end{align}
which is shown in Fig. \ref{coexTopo} (b). Naively, there is another type of string-like deconfined excitation that take the form of $\sum_m (1,0)^T \otimes e_{2m+1}$. However, it can be identified as $\epsilon_\text{str}$ fused with a string of condensed particles and, therefore, is considered as topologically equivalent to $\epsilon_\text{str}$. The braiding statistical angle between the deconfined point particle and the string excitation is
\begin{align}
\theta_\text{pt,str}=2\pi/3.
\end{align}
Therefore, the bulk topological order of this case is equivalent to that of a 3D $Z_3$ lattice gauge theory. However, as is discussed in Sec. \ref{pSTO}, this system also hosts a surface topological order with chiral central charge $c\equiv 2 (\text{mod} 3)$ on its open boundary. Therefore, it is a layer-constructed system that has co-existing bulk and surface topological order.
This situation can also be compared with the Walker-Wang model\cite{walker2012}. The input data of the Walker-Wang model is a braided tensor category. If the braided tensor category (which describes a 2D topological order) is modular, namely only the vacuum particle in this tensor category braids trivially with all the rest of the quasi-particles, the Walker-Wang model provides a construction of a 3D gapped state with trivial bulk topological order but non-trivial surface topological order characterized by the input modular tensor category.
If the input data is only pre-modular, there will be, by definition, transparent particles (other than the vacuum particle) which braids trivially with all other particles. In the Walker-Wang model associated to a pre-modular tensor category, the transparent particles can propagate in the bulk and become the quasi-particles of the 3D topologically ordered states. This scenario seems similar to the case of our layer constructed model. For this model with $p\equiv1,2(\text{mod}3)$, the surface topological order studied in Sec. \ref{pSTO} is modular and the 3D bulk has trivial topological order. For $p\equiv0(\text{mod}3)$, the deconfined particles on the surface only forms a pre-modular tensor category. If we quotient out the surface particles that braid trivially with others, we get a consistent surface topological order. Those particles indeed penetrate in the bulk and form 3D topological order. Due to the similarity between these two situations, one might want to conjecture that the lattice model realizations of the layer construction in these cases can be formulated using the Walker-Wang model.
\begin{figure}[t]
\centerline{
\includegraphics[width=2.5
in]{coexTopo-01.pdf}
}
\caption{\label{coexTopo} (a) The condensed particle in the coupled $Z_p$ toric code system. This case is the same as the case studied in Sec. \ref{EGpSTO}, but here we focus on the cases with
$p\equiv 0 (\text{mod}3)$. (b) The bulk deconfined excitations in this sytem. The one on the left is a point particle while the one on the right is a string excitation. They have non-trivial mutual statistics.
}
\end{figure}
\subsection{More Generic 3D Tolopological States with String Braiding Statistics}
\label{GenericState}
Now we will consider a more generic situation. The starting point is still stacked layers of $Z_p$ toric code. Here we will first assume that $p$ is a multiple of 4. For the convenience of later discussions, we view two layers of $Z_p$ toric codes as one single layer described by the K-matrix:
\begin{align}
K_0=\left(\begin{array}{cccc}
0 & p & 0 & 0\\
p & 0 & 0 & 0\\
0 & 0 & 0 & p\\
0 & 0 & p & 0\\
\end{array}
\right)
\end{align}
For later convenience, we will denote the four generators of all quasi-particles $(1,0,0,0)^T$, $(0,1,0,0)^T$, $(0,0,1,0)^T$ and $(0,0,0,1)^T$ as $c_1$, $f_1$, $c_2$ and $f_2$. The condensed particles (see Fig. \ref{EGgeneric}) are given by the data:
\begin{align}
p_1=(1,0,2,0)^T,~~~q_1=(-1,0,-2,0); \nonumber\\
p_2=(1,2,0,-1)^T,~~~q_2=(1,-2,0,1).
\end{align}
In this condensate, there are two types of deconfined point particles
\begin{align}
&\epsilon_\text{c-pt}=(1,0,2,0)^T\otimes e_m, \nonumber \\
&\epsilon_\text{f-pt}=(0,0,0,p/2)^T\otimes e_m,~~~\forall m
\end{align}
as is shown in Fig. \ref{EGgeneric} (b). These point particles can hop between the layers by fusion with condensed particles. Their mutual statistics is trivial. There are also two different types of string excitations (Fig. \ref{EGgeneric} (b)) :
\begin{align}
\epsilon_\text{c-str}=\sum_m (0,0,1,0)^T \otimes e_m, \nonumber \\
\epsilon_\text{f-str}=\sum_m (0,0,0,1)^T \otimes e_m.
\end{align}
The braiding statistics between the point particles and string excitations are given by
\begin{align}
\theta_{I,J}=2\pi\left(
\begin{array}{cc}
2/p & 0 \\ 0 & 1/2
\end{array}
\right),
\label{pt_f-str}
\end{align}
where $I=1,2$ stand for point particles $\epsilon_\text{c-pt}$ and $\epsilon_\text{f-pt}$ respectively while $J=1,2$ stand for string excitations $\epsilon_\text{f-str}$ and $\epsilon_\text{c-str}$.
More interestingly, from the expression of $\epsilon_\text{c-str}$ and $\epsilon_\text{f-str}$, it seems that these string excitations can also have non-trivial mutual braiding statistics. Indeed, if the system is periodic along the $z$-direction, the two types of strings can wrap around the $z$-direction and their mutual braiding phase is given by $\omega_\text{c-str,f-str}=2\pi L/p$, where $L$ is the number of layers in the system. This result can simply be obtained by looking the system from the $z$-direction and viewing it as a quasi-2D system. From this point of view, the two string excitations are simply point particles in the 2D sense. Their braiding is also understood as the braiding of particles in 2D. This analysis only applies to vertical strings that wrap around the $z$ direction cycle. We need to inspect further the string braiding statistics for strings with finite size.
\begin{figure}[t]
\centerline{
\includegraphics[width=3
in]{EGgeneric2-01.pdf}
}
\caption{\label{EGgeneric} (a) The condensed particles for a more generic layer constructed 3D state. Each layer contains two identical copies of $Z_p$ toric code, with $p$ a multiple of $4$. $c_{1,2}$ and $f_{1,2}$ are the charge and flux particles in the two $Z_p$ toric code system, respectively. (b) The set of deconfined excitations, including two types of point particles and two types of string excitations with nontrivial mutual braiding.
}
\end{figure}
\begin{figure*}[t]
\centerline{
\includegraphics[width=7
in]{TwoLoops-01.pdf}
}
\caption{\label{TwoLoops} (a) The definition of $c$-string (blue) and $f$-string (orange). Each plane represents a layer of 2D topological state.
The convention on the orientation of the string is explained in the main text. (b1) The braiding process of the two strings. (b2-b4) The braiding process shown in decomposed steps.
}
\end{figure*}
The strings with finite size can be constructed in the same way as we did in Sec. \ref{Convent3DTO}. Fig. \ref{TwoLoops} (a) illustrates the construction of contractible strings of finite size for the example system. Each plane represents a layer in the layer construction. The particle contents of the string excitations in each layer are shown. For example, the blue string in Fig. \ref{TwoLoops} (a) represents the finite size version of the deconfined excitation $\epsilon_\text{f-str}$, which is given by a line with $c_2$ in each of the $5$ layers together with another line of $-c_2$ in each of the $5$ layers. In general, an oriented string is defined as a chain of particles by the following rule. When the string crosses a plane and the direction of the string is parallel (anti-parallel) to the normal direction of the plane (which is $\hat{z}$ direction in our example), a $c_2$ ($-c_2$) particle is assigned to the crossing point.
The same definition applies to the other type of string except for a substitution of $c_2$ with $f_2$. In the following, we will refer to the string consisting of $c_2$ particles as the $c$-string, and the other type as the $f$-string. It is straightforward to see that these string are deformable. The deformation of string along the $xy$ plane is simply the change of position of $c_2$'s or $f_2$'s within the layers. The deformation along $z$-direction is achieved by pair creation or pair annihilation of $c_2$'s or $f_2$'s. For example, if we fuse the $c_2$ and $-c_2$ of $f$-string in the bottom layer in Fig. \ref{TwoLoops} (a), the string effectively shrinks to a smaller size.
\begin{figure*}[t]
\centerline{
\includegraphics[width=7
in]{TwoLoopsDis-01.pdf}
}
\caption{\label{TwoLoopsDis} (a) The configuration of the $c$-string and the $f$-string that link with an edge dislocation (or a $d$-string) which is indicated by the purple line. (b1) The braiding process of the $c$-string and $f$-string that are both linked with the $d$-string. (b2-b4) show the decomposed steps in this braiding process.
}
\end{figure*}
The braiding between two types of strings is defined to be the process shown in Fig. \ref{TwoLoops} (b1-b4). Fig. \ref{TwoLoops} (b1) is the overall action to the $c$-string (blue) to complete the braiding with the $f$-string (orange). For the sake of clarity, we've shown the decomposed steps in this action in Fig. \ref{TwoLoops} (b2-b4). First, we expand the $c$-string and move it past the $f$-string such that the $f$-string effectively passes through the ``inside" of $c$-string (Fig. \ref{TwoLoops} (b2)). Then, we shrink the $c$-string and move it back through the inside of the $f$-string (Fig. \ref{TwoLoops} (b3)). Once the $c$-string passes through the $f$-string, we expand the $c$-string back to its original size and move it back to the original position (Fig. \ref{TwoLoops} (b4)). In fact, it is straightforward to show that these type of braiding is always trivial for contractible $c$-string and $f$-string. One can verify this by considering the total phase that results from the braiding of the particles $c_2$'s and $f_2$'s in each layer during the braiding of strings. The braiding phase between all the $c_2$'s with all the $f_2$'s is exactly canceled by that between all the $-c_2$'s and all the $-f_2$'s due to the opposite directions of the two braid. This is consistent with the observation of Ref. \onlinecite{Levin2014StrBrd} that the braiding between two contractible strings must be trivial.
In order to obtain some non-trivial string braiding statistics, we need to consider a system with an edge dislocation and the $c$-string and $f$-string that are linked with the edge dislocation (See Fig. \ref{TwoLoopsDis} (a)). In the following, we will refer to the edge dislocation with a unit Burgers vector as the $d$-string. The orientation of the $d$-string is determined through the right-hand rule by the normal direction of the surface that ends at the $d$-string. Now, we can study the braiding process between the $c$-string and the $f$-string, both of which are linked to a $d$-string.
Fig. \ref{TwoLoopsDis} (b1) shows the overall action to the $c$-string (blue) in the braiding process at presence of the $d$-string. Fig. \ref{TwoLoopsDis} (b2-b4) show the decomposed steps for the braid. (Similar three-string braiding processes in $Z_N^k$ or $Z_{N_1}\times Z_{N_2}\times Z_{N_3}\times Z_{N_4}$ gauge theory are considered in Ref. \onlinecite{Levin2014StrBrd, Ran2014StrBrd, JuvenWenStrBrd2014}.) In this process, the string braiding statistics, which is given by the total braiding phase between the quasi-particles that form these string, is
\begin{align}
\omega_{c,f}^d = \theta_{c_2,f_2}= 2\pi/p,
\label{cfbraid}
\end{align}
where $\theta_{c_2,f_2}$ is the braid statistics of the $c_2$ and $f_2$ particles in a singe layer of the 2D topological state defined by the K-matrix $K_0$. More generally, if we consider a composite string of $n_c$ $c$-strings and that of $n_f$ $f$-strings that are both linked with an edge dislocation with Burgers vector of length $b$, the string braiding statistics is given by
\begin{align}
b~n_c n_f \omega_{c,f}^d = 2\pi n_c n_f b/p.
\label{cfbraid}
\end{align}
From Fig. \ref{TwoLoopsDis} (a), it is easy to see that the net contribution to this phase $\omega_{c,f}^d $ is essentially given by the braiding between the $c_2$ particle and the $f_2$ particle in the defected layer. Compare to the braiding statistics between point particles and string excitations in Eq. (\ref{pt_f-str}), when $p$ is a multiple of $4$, the string braiding statistics $\omega_{c,f}^d $ is more fractionalized. Thus, it cannot be removed by attaching point particles to the strings. Also, the fact that the $d$-string is an extrinsic defect rather than a dynamical one does not undermine the topological protection of the string braiding statistics $\omega_{c,f}^d $ either. That is because throughout the braiding process, the $d$-string does not have to move which means that no additional dynamical phase because of the extrinsic nature of the $d$-string should come in to play. Furthermore, the three-string braiding phase $\omega_{c,f}^d$ can be related to the braiding phase $\omega_\text{c-str,f-str}=2\pi L/p$ between two vertical strings winding around the torus, which we discussed at the beginning of this subsection.
An $L$ layer system can be formally considered as a $0$ layer system (vacuum) with an edge dislocation in the $xy$ plane at infinity with Burgers vector $L\hat{z}$, {\it i.e.} $L$ $d$-strings at infinity
With periodic boundary condition, the $c$-strings and the $f$-string that wrap around the nontrivial cycle in the $z$ direction in fact have non-trivial linking number with all of the L $d$-strings, and therefore, have non-trivial string braiding statistics $\omega_\text{c-str,f-str}$ that is proportional to the number of layers $L$.
There is an alternative approach to compute the string-string braiding statistics, which will be helpful for our later discussion. To explain this approach, we consider the string fusion and splitting process in similar way as Ref. \onlinecite{Levin2014StrBrd}.
When we consider our system as a multi-layer 2D system, the finite-size $c$-string and $f$-string are special configurations of quasi-particles that live in some number of layers. Therefore, their deformation and fusion can be understood in term of the fusion of quasi-particles in the 2D theory. The specific type of deformation that we are interested in is the deformation that change the number of strings.
Here, we will show that the string excitations in the layer constructed system can also deform, especially split and fuse, without accumulating non-trivial Berry's phase. To demonstrate this, we can consider the example shown in Fig. \ref{PinchedLoop}. One $c$-string can be split into two independent $c$-strings by bringing two different points on the string very close to each other and pinching this point off. In this process, we use nothing more than the fusion rules of quasi-particles in each layer. In the current setting, $c_2$ in each layer is an Abelian particle with no spin. Therefore, the splitting process does not have any non-trivial Berry's phase associated with it. One can also consider the reverse process of splitting, which is the fusion of two $c$-strings. The same analysis is also applicable to the $f$-strings.
\begin{figure}[t]
\centerline{
\includegraphics[width=3
in]{PinchedLoop-01.pdf}
}
\caption{\label{PinchedLoop} One $c$-string can be deformed and split into two $c$-strings through the fusion of $c_2$'s within the layers. The reverse process can be viewed as the fusion between two $c$-strings.
}
\end{figure}
\begin{figure*}[t]
\centerline{
\includegraphics[width=7
in]{BraidDeformV2-01.pdf}
}
\caption{\label{BraidDeform} (a-d) An alternative string braiding process that is topologically equivalent to that shown in Fig. \ref{TwoLoopsDis}.
Upon fusion and splitting the strings the process in (b) can be deformed to the combined process shown in (e-h). (i) shows the local identification of string configuration under splitting and fusion which plays a central role in the re-interpretation of the process shown in (b). The total Berry's phase in the string braiding process equals the Berry's phase in the process induced by the braiding of the link $L_{c,f}$ around the $d$-string, as is shown in (j)
}
\end{figure*}
\begin{figure}[t]
\centerline{
\includegraphics[width=3.5
in]{LinkedMove-01.pdf}
}
\caption{\label{LinkedMove} (a) The configuration of the link $L_{c,f}$ with the quasi-particles associated with it in each layer. The layers are horizontal which are not drawn in this figure. In this link $L_{c,f}$, the $c$-string and the $f$-string have linking number $1$. Their orientation is indicated by the colored arrow (blue for the $c$-string and orange for the $f$-string). (b) The operation that brings the linked part of the two strings down by one layer along the $-z$ direction is a $\pi$ rotation of the part of the link indicated by the grey dashed box.
}
\end{figure}
Since for each layer of 2D topological state, the fusion and braiding of quasi-particles are operations that commute with each other, we argue that the string braiding also commutes with string deformation including splitting and fusion of strings. To be more precise, we can deform, split and fuse strings without changing the total Berry's phase in the braiding of string as long as the overall process remains topologically equivalent. Using this argument, we can obtain a new view to the braiding process (Fig. \ref{TwoLoopsDis} (b1)) that we discussed above. The braiding process is equivalent to the process shown in Fig. \ref{BraidDeform} (a-d). First, we pick an infinitesimal segment (the red shaded area in Fig. \ref{BraidDeform}(a)) of the $c$-string, and braid it around the $f$-string without moving the rest of the $c$-string. This action creates locally a link (yellow region) and an anti-link (grey region). The link and anti-link can annihilate each other if we bring them close together. Now, instead of annihilation them together directly, we bring the local link in the yellow region around the $d$-string (Fig. \ref{BraidDeform}(b)). We then annihilate the yellow and the grey regions by bringing them together afterwards (Fig. \ref{BraidDeform}(c)). The whole process shown in Fig. \ref{BraidDeform} (a-d) is topologically equivalent to the braiding of strings defined in Fig. \ref{TwoLoopsDis} (b1). We argue that in the combined process of Fig. \ref{BraidDeform} (a-d), only the process in Fig. \ref{BraidDeform} (b) contributes to the non-trivial Berry's phase . This is because 1) The process shown in Fig. \ref{BraidDeform}(a) only moves an arbitrarily small segment and should not accumulate any non-trivial Berry's phase; and 2) The annihilation between the yellow and grey regions also does not contribute to the Berry's phase since it can also be thought of as bringing an infinitesimal segment of the $c$-string around the $f$-string as we did in Fig. \ref{BraidDeform}(a). Given that all the non-trivial Berry's phase comes from the process Fig. \ref{BraidDeform} (b), we will re-interpret it using the fusion and splitting of strings. As is shown in Fig. \ref{BraidDeform} (i), the strings in the yellow region can be deformed, upon splitting and fusion, to unlinked $c$-string and $f$-string accompanied by a link $L_{c,f}$ of linking number 1 between the $c$-string and $f$-string. Therefore, the process shown in Fig. \ref{BraidDeform}(b) is equivalent to the combined process shown in Fig. \ref{BraidDeform}(e-h). Since the fusion and splitting do not produce non-trivial Berry's phase in this process, the total Berry's phase equals to the Berry's phase $\Omega^d_{L_{c,f}}$ of braiding the link $L_{c,f}$ around the the $d$-string. The analysis above can be summarized by the equation:
\begin{align}
\omega^d_{c,f}=\Omega^d_{L_{c,f}}.
\label{BraidEquiv}
\end{align}
In the analysis above, for the sake of generality, we've implicitly assumed that the $c$-strings and $f$-strings can be considered as continuous object in the space. We expect that its validity still holds even when the space is discrete. In fact, we can verify Eq. (\ref{BraidEquiv}) by explicitly calculating $\Omega^d_{L_{c,f}}$ in our layer constructed system. We start by studying the Berry's phase induced by the local motion of the link $L_{c,f}$. If we hold the relative position of the $c$-string and $f$-string in the link $L_{c,f}$ fixed, it is obvious that moving $L_{c,f}$ along the $x$ or the $y$ direction does not produce any non-trivial phase. In contrast, moving the $L_{c,f}$ in the $z$-direction while keeping the relative positions of the $c$-string and the $f$-string fixed will induce non-trivial Berry's phase. The link configuration of $L_{c,f}$ is drawn in Fig. \ref{LinkedMove} (a). The orientation of the strings are indicated by the arrows of the corresponding colored circles (blue and orange). The layers that are supposed to be represented by horizontal lines are omitted in this figure, and the $z$ direction is indicated by the black arrow. Let's consider moving this configuration by one layer down towards the $-z$ direction. For the parts outside grey shaded region in Fig. \ref{LinkedMove} (a), since there is no non-trivial linking between these segments of the $c$-string and $f$-string, they can be deformed into the configuration that is one layer below their original position without inducing any Berry's phase. In the grey shaded region, the linked parts of $c$-string and $f$-string need to be treated more carefully. To move the linked part one layer towards the $-z$ direction, we essentially have to rotation the part of the link enclosed by the grey dashed line in Fig. \ref{LinkedMove} (b) by $\pi$ about the $z$ axis. The net Berry's phase associated to this rotation is equivalent to a full braid between $c_2$ and $f_2$ in a single layer and is, therefore, $\theta_{c_2,f_2}=2\pi/p$. By similar analysis, the Berry's phase in moving the link $L_{c,f}$ one layer up toward the $z$ direction is the opposite. In process of moving the $L_{c,f}$ around the the $d$-string (Fig. \ref{BraidDeform}(g)), namely an edge dislocation, the net effect is to move the $L_{c,f}$ by the Burgers vector, {\it i.e.}, to move it downwards by one layer. Therefore, we can conclude that
\begin{align}
\Omega^d_{L_{c,f}}=\theta_{c_2,f_2},
\end{align}
which verifies the consistency of Eq. (\ref{BraidEquiv}). In this discussion, we always have to keep the $d$-string fixed because it is an extrinsic defect rather than a dynamical excitation. If we can promote the $d$-strings to dynamical excitations, we can study other types of string braiding between the three types of strings $c$, $f$ and $d$, such as $\omega^c_{f,d}$ and $\omega^f_{d,c}$\cite{Levin2014StrBrd}. The model that we considered so far is not capable of capturing the dynamics of the $d$-strings, but the analysis above relating string-string braiding and link-string braiding can be used to derive general identities on the string braiding, as will be discussed in Sec. \ref{GeneralA}.
\section{Topological Field Theory Description}
\label{FieldTheory}
In this section, we introduce a topological field theory description of the layer-constructed system, which provides a phenomenological description of the particle-string braiding and string-string braiding. The field theory can be considered as a continuum limit of the discrete system of coupled layers, although a generic derivation starting from coupled layers remain an open question that we will leave for future work. In this work, we will treat the field theory as a phenomenological low energy effective theory.
We consider the following Lagrangian density:
\begin{align}
\mathcal{L}_\text{LC}=
&\frac{Q_{IJ}}{2\pi} \epsilon^{\mu\nu\lambda\sigma} b^I_{\mu\nu}\partial_\lambda a^J_\sigma
+ \frac{\Theta}{8\pi^2} R_{IJ} \epsilon^{\mu\nu\lambda\sigma} \partial_\mu a^I_\nu \partial_\lambda a^J_\sigma \nonumber \\
& +j^I_\mu a^I_\mu + \mathcal{J}^I_{\mu\nu} b^I_{\mu\nu}
\label{FTlag}
\end{align}
with the summation of repeated indices implicitly assumed. Here $a^I_\mu$ for each $I$ is a 1-form $U(1)$ gauge field that couples to the current $j^I_\mu$ of non-trivial point excitations, and $b^I_{\mu\nu}$ for each $I$ is a 2-form $U(1)$ gauge field that couples to the 2-form current $\mathcal{J}^I_{\mu\nu}$ of non-trivial string excitations. ($b_{\mu\nu}^I$ and $\mathcal{J}_{\mu\nu}^I$ are antisymmetric in permutation of $\mu,\nu$.) Their gauge transformations are given by $a^I_\mu\rightarrow a^I_\mu+\partial_\mu \alpha^I$ and $b^I_{\mu\nu}\rightarrow b^I_{\mu\nu}+\partial_\mu \beta^I_\nu + \partial_\nu \beta^I_\mu$. $Q_{IJ}$ and $R_{IJ}$ are integer-valued non-singular matrices, and $R$ is symmetric. $\Theta$ is an ``order parameter field" which is extrinsically determined and does not have dynamics. When we consider the compactification radius of each $a_\mu^I$ as $2\pi$, the integral $\frac1{8\pi^2}\int d^4xR_{IJ} \epsilon^{\mu\nu\lambda\sigma} \partial_\mu a^I_\nu \partial_\lambda a^J_\sigma$ is quantized to integer values for all closed space-time manifolds. Therefore the partition function of the system is invariant under the transformation $\Theta\rightarrow \Theta+2\pi$. For a manifold with boundary, $\Theta\rightarrow \Theta+2\pi$ will induce a boundary Chern-Simons term $\int_{\rm boundary}d^3n_\mu\frac{1}{4\pi}R_{IJ}\epsilon^{\mu\nu\sigma\tau}a_\nu^I\partial_\sigma a_\tau^J$. Here $\int dn_\mu$ is the boundary volume integration with $dn_\mu$ along the normal direction.
The topological field theory (\ref{FTlag}) has the form of BF theory\cite{birmingham1991,hansson2004}. Similar theory has been proposed to describe fermionic and bosonic topological insulators\cite{cho2011,vishwanath2013}, and the Walker-Wang model\cite{walker2012,simon2013}. Compared with previous works, the essential new ingredient in our theory (\ref{FTlag}) is the string-string braiding statistics, which, as will be explained in the following, is enabled by the possibility of $\Theta$ vortex loops.
We start from a system with constant $\Theta$. The equation of motion is given by
\begin{align}
&j^I_\mu =-\frac{1}{2\pi}Q_{JI}\epsilon^{\mu\nu\lambda\sigma} \partial_\nu B^J_{\lambda\sigma}, \nonumber \\
&\mathcal{J}^I_{\mu\nu} =-\frac{1}{2\pi}Q_{IJ} \epsilon^{\mu\nu\lambda\sigma} \partial_\lambda a^J_\sigma.
\end{align}
Therefore, the string particle carries flux of $a_\mu^I$, so that braiding the $I^\text{th}$ particle $\Pi_I$ (whose current is given by $j^I_\mu$) and the $J^\text{th}$ string $\Sigma_J$ (whose current is given by $\mathcal{J}^J_{\mu\nu}$) produces the Berry's phase
\begin{align}
\theta_{\Pi_I,\Sigma_J}=-2\pi (Q^{-1})_{IJ}.
\end{align}
By properly choosing the matrix $Q$, the Lagrangian density $\mathcal{L}_\text{CL}$ can capture the particle-string statistics in the layer constructed model. For example, to capture the braiding statistics in Eq. (\ref{EMstat}), we need to take $Q=p$ (where $Q$ is a $1\times 1$ matrix). For the case studied in Sec. \ref{Coexist}, we should have $Q=3$. For these two cases, $R$'s are taken to be $0$. However, even with non-zero matrix $R$, the string-string braiding is trivial when $\Theta$ is a constant. To obtain nontrivial string-string braiding, it is essential to consider space-time dependent $\Theta$.
Now, we can consider the case with specially varying $\Theta$. The equation of motion is given by
\begin{align}
&j^I_\mu =-\frac{1}{2\pi}Q_{JI}\epsilon^{\mu\nu\lambda\sigma} \partial_\nu B^J_{\lambda\sigma}- \frac{1}{4\pi^2} R_{IJ} \epsilon^{\mu\nu\lambda\sigma} \partial_\nu\Theta \partial_\lambda a^J_\sigma , \nonumber \\
&\mathcal{J}^I_{\mu\nu} =-\frac{1}{2\pi}Q_{IJ} \epsilon^{\mu\nu\lambda\sigma} \partial_\lambda a^J_\sigma.
\end{align}
For the case without any point particles, i.e. $j^I_\mu=0$, we have
\begin{align}
(Q^{-1T} R Q^{-1})_{IJ} \mathcal{J}^J_{\mu\nu} \partial_\nu \Theta= \epsilon^{\mu\nu\lambda\sigma}\partial_\nu B^I_{\lambda\sigma}.
\end{align}
Since the bulk of system is invariant under the transformation $\Theta\rightarrow \Theta+2\pi$, we can consider a $2D$ surface in the system across which the the value of $\Theta$ jumps by $2\pi$. For example, when this $2D$ surface is the $xy$ plane, we have
\begin{align}
\partial_\mu \Theta= \delta_{\mu,z} 2\pi \delta(z).
\label{ThetaConfig}
\end{align}
When the vortex string $\Sigma_J$ goes through the $z=0$ plane, the point of penetration acts as a point source for $\epsilon^{\mu\nu\lambda\sigma}\partial_\nu B^I_{\lambda\sigma}$. If we consider braiding strings $\Sigma_I$ and $\Sigma_J$ as is shown in Fig. \ref{FT} (a) (with the $z=0$ plane where $\partial_z \Theta=2\pi$ the green plane shown in the figure), the Berry's phase of this braid is
\begin{align}
\omega_{\Sigma_I,\Sigma_J}=2\pi(Q^{-1T} R Q^{-1})_{IJ}.
\label{FTStrBrd}
\end{align}
In this discussion, the only non-trivial effects from the $\Theta$ configuration in Eq. (\ref{ThetaConfig}) is the braiding statistics between strings $\Sigma_I$ and $\Sigma_J$ that go through the $z=0$ only once and close at infinity. It is straightforward to see that if we consider strings of finite size, they have to cross the $z=0$ plane even number of times and their braiding is always trivial. Therefore, the system behaves the same as the original system without the jump of $\Theta$ across $z=0$. To obtain non-trivial braiding of finite-size strings, we need to consider 2D surface $\gamma$ with boundary $\partial \gamma \neq \emptyset$ crossing which $\Theta$ jumps by $2\pi$. Such a surface is a vortex loop of $\Theta$ at $\partial\gamma$. The strings that are linked with the vortex loop will have non-trivial string braiding statistics. For example, the braid statistics of the string $\Sigma_I$ and $\Sigma_J$ in Fig. \ref{FT} (b) is given by Eq. (\ref{FTStrBrd}). Compare to the result we present earlier in layer construction, we can easily identify the defected layer in Fig. \ref{TwoLoopsDis} (a) as the surface $\gamma$ and the edge dislocation as $\partial \gamma$. Since the jump of $\Theta$ is not locally observable in the bulk, we can think of the boundary $\partial \gamma$ as a line defect which is consistent with the edge dislocation interpretation.
For a specific layer constructed model, we can choose proper $Q_{IJ}$ and $R_{IJ}$ to capture the phenomenology of the model. For example, for the example we studied in Sec. \ref{GenericState}, the matrix $Q_{IJ}$ is determined by the statistics between point particles and string excitations in Eq. (\ref{pt_f-str}):
\begin{align}
Q=\left(
\begin{array}{cc}
p/2 & 0 \\ 0 & 2
\end{array}
\right).
\end{align}
With the choice
\begin{align}
R=\left(
\begin{array}{cc}
0 & 1 \\ 1 & 0
\end{array}
\right),
\end{align}
we obtain the string braiding statistics
\begin{align}
2\pi(Q^{-1T} R Q^{-1})_{IJ} =2\pi\left(
\begin{array}{cc}
0 & 1/p \\ 1/p & 0
\end{array}
\right),
\end{align}
which is consistent with braiding statistics $\omega^d_{c,f}$ between $c$-string and $f$-string calculated in Eq. (\ref{cfbraid}) in the presence of the $d$-string.
\begin{figure}[t]
\centerline{
\includegraphics[width=3
in]{FT-01.pdf}
}
\caption{\label{FT} (a) The configuration of two strings $\Sigma_I$ and $\Sigma_J$ that are aligned along the $z$ direction and penetrate the $z=0$ plane (green) where $\Theta$ jumps by $2\pi$. The strings $\Sigma_I$ and $\Sigma_J$ have non-trivial braiding. (b) The configuration of a finite open surface $\gamma$ across which $\Theta$ jumps by $2\pi$ together with strings $\Sigma_I$ and $\Sigma_J$ that linked with the vortex loop $\partial \gamma$ (purple).
}
\end{figure}
\section{General Discussion on String braiding}
\label{GeneralStrBrd}
\subsection{General Identities on Abelian String Braiding Statistics}
\label{GeneralA}
In the last part of previous section, we discuss the braiding statistics of strings that involves two dynamical string excitations and one extrinsic string defect. It is conceivable that similar string braiding can happen in the same configuration for three dynamical string excitations. As is recently discussed in Ref. \onlinecite{Levin2014StrBrd,Ran2014StrBrd,JuvenWenStrBrd2014}, in twisted lattice gauge theories defined by non-trivial group cohomology classes, different flux strings can braid non-trivially. Ref. \onlinecite{Levin2014StrBrd} proposed an identity on braiding statistics of these flux strings. For two strings $a$ and $b$ both linked with $c$ (Fig. \ref{StringBrd} (a)), the braiding phase of $a$ and $b$ is defined as $\omega_{a,b}^c$. Ref. \onlinecite{Levin2014StrBrd} proved the following identity (up to $2\pi$ times integers) for 3D $Z_N^k$ gauge theories:
\begin{eqnarray}
N(\omega_{a,b}^c+\omega_{b,c}^a+\omega_{c,a}^b)=0.\label{TriId2ml}
\end{eqnarray}
In the following, we will present a more general proof of a slightly modified version of this identity for Abelian string braiding, by making use of splitting and fusion of strings.
As a general setup of our discussion, we will consider the string excitations that has the following properties: (1) The strings are Abelian in sense that there are no non-trivial degenerate states associated to a local string configuration; (2) The strings have no point-like excitation attached to them. (3) Strings of the same type can split and fuse without inducing additional phase factor; (4) The strings can link with each other.
Since we have assumed that there is no non-trivial degeneracy associated with the string configuration, the Berry's phase $\omega_{a,b}^c$ is simply a $U(1)$ phase. This Berry's phase $\omega_{a,b}^c$ should be a topological invariant that does not depend on the shape of the strings in the process of braiding. In Sec. \ref{GenericState}, we have shown that by fusion and splitting of the same type of strings, one can topologically deform the string-string braiding process of $a$, $b$ strings linked to $c$ (Fig. \ref{StringBrd} (a)) to a particle-string braiding process in which the ``particle" $L_{a,b}$ is formed by two linked loops of types $a$ and $b$. Denoting the Berry's phase obtained by the braiding of $L_{ab}$ with string $c$ as $\Omega_{L_{a,b}}^c$, we obtain the equation
$\omega_{a,b}^c=\Omega_{L_{a,b}}^c$. Now we consider a different three-string configuration shown in Fig. \ref{StringBrd} (c). In this configuration, every pair of strings are mutually linked, with linking number 1. The ``braiding" process (which will be referred to as the linked braiding) in this configuration is defined to be a $2\pi$ rotation the two string $a$ and $b$ about the string $c$ (see Fig. \ref{StringBrd} (c)) while keeping the relative position between $a$ and $b$ fixed. The Berry's phase associated with this process is denoted as $\tilde{\omega}^c_{a,b}$. Following similar string fusion and splitting procedure as has been presented earlier in Fig. \ref{StringBrd} (c), we can deform the linked braiding process to the same particle-string braiding shown in Fig. \ref{StringBrd} (b). This is illustrated in detail in Fig. \ref{StringBrd} (d).
Therefore, we conclude,
\begin{align}
\omega_{a,b}^c=\Omega_{L_{a,b}}^c =\tilde{\omega}^c_{a,b}.
\label{BraidEquiv2}
\end{align}
Now we proceed to prove the a modified version of the identity (\ref{TriId2ml}) by proving it for $\tilde{\omega}^c_{a,b}$. The configuration of three mutually linked loops in the Fig. \ref{StringBrd} (c) can be topologically deformed to that in Fig. \ref{ThreeLoop}. This configuration is symmetric upon cyclic permutation of the three strings. The ``linked braiding" process defined in Fig. \ref{StringBrd} (c) can be rephrased as the operation that keeps the string $c$ fixed and rotates simultaneously the strings $a$ and $b$ about an axis, which goes through the reference point $o$ and is perpendicular to the 2D plane, clock-wisely by $2\pi$. This process is the linked braiding of the string $a$ and $b$ with respect to the string $c$. Due to the cyclic symmetry between the three strings in this configuration in Fig. \ref{StringBrd}, we can also consider similar linked braiding processes between the strings $b$ and $c$ with respect to the string $a$, and that between $c$ and $a$ with respect to $b$. If we perform these three different types of linked braiding sequentially, the net Berry's phase obtained is by definition $\tilde{\omega}_{a,b}^c+\tilde{\omega}_{b,c}^a+\tilde{\omega}_{c,a}^b$. However, one can easily see that the combined process is actually a global $4\pi$ rotation of the whole string configuration around axis $o$, since each string is rotated by $2\pi$ during two of the three processes, and is held static during one of them.
Since a $4\pi$ rotation is a topologically trivial action in three dimension, the total Berry's phase induced by the rotation must be trivial. This implies that the sum of three different linked braiding Berry's phases should be trivial, namely
\begin{align}
\tilde{\omega}_{a,b}^c+\tilde{\omega}_{b,c}^a+\tilde{\omega}_{c,a}^b=0.
\label{TriId1}
\end{align}
Therefore by using Eq. (\ref{BraidEquiv2}), we prove the following identities:
\begin{align}
&\omega_{a,b}^c+\omega_{b,c}^a+\omega_{c,a}^b=0,
\label{TriId3} \\
&\Omega_{L_{a,b}}^c+\Omega_{L_{b,c}}^a+\Omega_{L_{c,a}}^b=0.
\label{TriId4}
\end{align}
Eq. (\ref{TriId3}) directly implies the previously proposed Eq. (\ref{TriId2ml}). Eq. (\ref{TriId4}) constraints the braiding statistics of strings and ``link particles".
Our proof here only relies on adiabatical deformations and fusion-splitting processes of strings, which is thus applicable to generic 3D topologically ordered states that involves Abelian string-like excitations.
\begin{figure}[t]
\centerline{
\includegraphics[width=3.5
in]{StringBrd-01.pdf}
}
\caption{\label{StringBrd} Three different string braiding processes, named as the string-string braiding (a), link-string braiding (b) and the linked braiding (c). As is explained in the main text, the Berry's phase associated to these three processes are identical. (d) The decomposed steps that relate processes in (c) and (b).
}
\end{figure}
\begin{figure}[t]
\centerline{
\includegraphics[width=1.5
in]{ThreeLoop-01.pdf}
}
\caption{\label{ThreeLoop} The configuration of three mutually linked strings. This configuration is topologically equivalent to the configuration shown in Fig. \ref{StringBrd} (c), but is deformed to a shape with three-fold rotation symmetry according to an axis perpendicular to the plane at the point $o$
}
\end{figure}
\subsection{Non-Abelian String Braiding Statistics}
\label{GeneralB}
The discussion above only applies to the 3D topological states with string excitations that satisfies the assumptions given in the beginning of the previous subsection. It is worthy to point out that there are certainly possibility of other more generic string like excitations that violates some of these assumptions. For instance, the same string configurations can in principle carry non-Abelian degrees of freedom. The general framework of non-Abelian string-string braiding is beyond the scope of this paper, but we would like to present one explicit example of a system with non-Abelian string-string braiding.
Consider two Weyl fermions in $(3+1)$-dimensions with opposite chirality, which are equivalent to a Dirac fermion. Each Weyl fermion is time-reversal invariant, and a superconducting pairing can be introduced within each Weyl fermion. The Fermi level can be either at or away from the Weyl point, which does not affect our discussion. The BCS Hamiltonian of such a superconductor can be written as
\begin{align}
\hat{H}=\sum_{\bf{k}} \left[ \psi^\dag_{R\textbf{k}} (v_F\sigma\cdot \textbf{k}-\mu) \psi_{R \bf{k}}
+\psi^\dag_{L\bf{k}} (-v_F\sigma\cdot \textbf{k}-\mu) \psi_{L\bf{k}} \right]+ \nonumber \\
\frac{1}{2}\sum_{\bf{k}} \left[
|\Delta_R|e^{i\theta_R}\psi^\dag_{R\textbf{-k}} i\sigma_y \psi^\dag_{R \bf{k}}
+|\Delta_L|e^{i\theta_L}\psi^\dag_{L\textbf{-k}} i\sigma_y \psi^\dag_{L \bf{k}}+h.c.
\right],\label{TSCHamiltonian}
\end{align}
where $\psi_{R \bf{k}}$ and $\psi_{L \bf{k}} $ are two-component fermion operators, $v_F$ is the fermi velocity, $\sigma$'s are the $2\times2$ Pauli matrices, and $\Delta_R=|\Delta_R|e^{i\theta_R}$ and $\Delta_L=|\Delta_L|e^{i\theta_L}$ are the pairing order parameters on the fermi surface of $\psi_{R \bf{k}}$ and $\psi_{L \bf{k}}$. Without the pairing term and with $\mu>0$, the Fermi surface of $\psi_{R \bf{k}}$ carries Chern number 1 while the Fermi surface of $\psi_{L \bf{k}}$ carries Chern number -1. This model was studied before\cite{Qi2013TSC} as a minimal model of time-reversal invariant topological superconductivity (TSC)\cite{QiTSC2009,Roy2008,Schnyder2008}, which describes a TSC when $\theta_L=0,~\theta_R=\pi$, and a trivial superconductor when $\theta_L=\theta_R=0$\cite{qi2010b}.
As a probe to the TSC, Ref. \onlinecite{Qi2013TSC} studied the chiral vortex strings in this system, which are $2\pi$ vortex strings of only one of the pairing phases, such as $\theta_R$.
This chiral vortex string carries a $(1+1)$-d chiral Majorana fermion mode, which has a axial anomaly. For the purpose of this paper, we will focus on the zero energy mode in the spectrum of the chiral Majorana fermion, usually named as the Majorana zero mode\cite{kitaev2001}. A contractible vortex string does not carry a zero energy mode, since the Majorana fermion has anti-periodic boundary condition. (Without going into a detailed calculation, one can see that this must be true since a Majorana zero mode, if exists, cannot disappear when the loop shrinks to a point and disappear. Therefore it cannot occur in a contractible loop.) When a vortex string $v_1$ is linked with another chiral vortex string $v_2$ (see Fig. \ref{TSC} (a)), there will be a Majorana zero modes $\gamma_{1(2)}$ on each vortex string. Each Majorana zero mode is separated from other chiral modes by a finite size gap that is inversely proportional to the length of the vortex strings. There are two degenerate states with the opposite fermion parity $(-1)^F\equiv i \gamma_1 \gamma_2$ associated with the two zero modes on the two linked strings. If we braiding the link between $v_1$ and $v_2$ around a third chiral vortex string $v_3$, the Berry's phase is equal to
$(-1)^F$. Therefore, we see that the string braiding statistics relies on the state in the degenerate space defined by the string configuration, which is a three-dimensional generalization of ``fusion channels" of a non-Abelian anyon. Consider a different chiral vortex string configuration in Fig. \ref{TSC} (b) with two vortex strings $v_1$ and $v_2$ both linked with $v_3$. In this configuration both $v_1$ and $v_2$ carry Majorana zero modes $\gamma'_1$ and $\gamma'_2$, while $v_3$ does not have a zero mode. If we braid the vortex string $v_1$ and $v_2$ in the way defined in Fig. \ref{TwoLoopsDis} (a), the Majorana fermion operators $\gamma'_{1,2}$ will each obtain a $-1$ sign after a full braiding, due to the phase winding around the vortices. Therefore one expects that the braiding between these two strings with zero modes is the same as the braiding of two point-like vortices in $2+1$-d $p+ip$ superconductor\cite{read2000,ivanov2001}, up to an undetermined overall phase factor. The non-Abelian Berry's phase is a unitary operator acting on the two-dimensional Hilbert space, which can be written as $e^{i\frac{2\pi i \gamma'_1 \gamma'_2}{4}}$.
Since the chiral vortices are generically confined with the anti-chiral vortices energetically\cite{Qi2013TSC}, they have to be considered as external defects and the Abelian phase factor of the braiding process is not well-defined.
Since we can consider all the chiral vortices to be identical, we can also consider the ``half braid" between $v_1$ and $v_2$ that is similar to the string braiding defined Fig. \ref{TwoLoopsDis} (a) but only exchanges the position of two vortex strings instead of braiding one string around the other. In this half braid process, we can keep the sizes of all three vortex strings $v_1$, $v_2$ and $v_3$ finite. The degenerate space defined by the two zero modes $\gamma'_{1,2}$ are separated from other excited states by a finite size gap. Therefore the half braid should not change the total fermion parity $(-1)^F=i \gamma'_1 \gamma'_2$ of the three-string configuration. The only possible non-trivial operation to the Majorana zero modes is
\begin{align}
&\mathcal{U}^\dag \gamma'_1 \mathcal{U} = \gamma'_2, \nonumber \\
&\mathcal{U}^\dag \gamma'_2 \mathcal{U} = -\gamma'_1,
\end{align}
where $\mathcal{U}$ is the unitary transformation induced by the half braid. This determines $\mathcal{U}=e^{\frac{2\pi i \gamma'_1 \gamma'_2}{8}}$ up to a $U(1)$ phase
When more generic configurations of vortex strings are considered, one can see more qualitative difference from the Abelian case. For example, one can consider the configuration of three mutually linked chiral vortex strings shown in Fig. \ref{ThreeLoop}. In the Abelian case, a ``linked braiding" procedure is defined in this configuration (Fig. \ref{StringBrd} (c) and (d)), which is proved to induce the same Berry's phase as the three-string braiding process in Fig. \ref{StringBrd} (a). In the case of chiral vortex strings, the configuration of three mutually linked strings is very different from two strings linked with the third one: When each loop is linked with two other loops, there is no Majorana zero mode on any of the three loops. Consequently there cannot be any nontrivial non-Abelian Berry's phase in the linked braiding process. If the linked braiding and unlinked braiding (the procedures shown in Fig. \ref{StringBrd} (a) and (c)) are still related, the linked braiding will at most be related to a given fusion channel of the two strings in the unlinked configuration.
In general, the fusion and splitting of non-Abelian strings involve the change of topological degeneracy, which is more complicated than the case of Abelian strings with ``adiabatic" fusion and splitting within the same species. Fusion channels have to be defined when we consider the splitting and fusion of strings. We will leave this for future study.
One can consider a multi-component generalization of the model (\ref{TSCHamiltonian}), which has a matrix-valued pairing order parameter, and allows monopole-like non-Abelian point defects in addition to the chiral vortices\cite{teo2010,ran2011,freedman2011,freedman2011b}. The non-Abelian statistics of monopole defects are also related to Majorana zero modes, and it is interesting to investigate the interplay of the braiding statistics of monopoles and chiral vortex strings.
There are also other similar examples in the 3D quantum double model\cite{Wen2014StrBrd} and 3D Dijkgraaf-Witten lattice theory\cite{JuvenWenStrBrd2014} that strings can carry non-Abelian degrees of freedom and have non-Abelian braiding statistics. When we dimensionally reduced the $z$ direction to a finite compactified circle for these lattice gauge theory systems, the strings align along the $z$ direction which can be viewed as a quasi-particle in the 2D sense can be asigned quantum dimensions. Since the 3D lattice guage theory, when dimensionally reduced, becomes 2D lattice guage theory. Each quasi-particles (in the 2D theory) must have integer quantum dimension, and so are the strings (along the $z$ direction). In constrast, if we dimensionally reduce the 3D TSC to 2D with a $\pi$ chiral flux in the $z$ direction cycle, the chiral vortex strings (along the $z$ direction) will have quantum dimension $\sqrt{2}$ due to the Majorana zero mode. It would be interesting to find more examples of string excitations with non-integer quantum dimensions. Moreover, Ref. \onlinecite{Levin2014StrBrd,JuvenWenStrBrd2014,Ran2014StrBrd} mainly concerns the braiding process defined in Fig. \ref{StringBrd} (a) between flux strings in the twisted Dijkgraaf-Witten lattice gauge theory. It would also be interesting to study the braiding process defined in Fig. \ref{StringBrd} (b) and (c) for these models.
\begin{figure}[t]
\centerline{
\includegraphics[width=2.5
in]{TSC-01.pdf}
}
\caption{\label{TSC} (a) and (b) show two different configurations of chiral vortex strings.
}
\end{figure}
\section{Conclusion and Discussion}
\label{conclusion}
In this paper, we propose a general approach to obtain non-trivial 3D topological states through layer construction. Starting from stacked layers of 2D Abelian topological states, anyon-condensation induces inter-layer coupling and drives the system into a three-dimensional state which may have interesting bulk and/or surface topological orders. We illustrate several different possibilities with explicit examples. The first class of states are those with trivial bulk topological order and nontrivial surface two-dimensional topological order. We obtain general criteria for such states, and show that this state has the same topological order on surfaces with different orientation, even though it appears to be anisotropic.
This implies that the layer construction indeed captures the properties of homogeneous 3D states. The second class of states are those with nontrivial bulk topological order, which means nontrivial ground state degeneracy on a three-dimensional torus. The simplest examples realize the discrete gauge theories. More generically, the layer construction can describe broader class of topologically ordered states which have
coexisting surface and bulk topological order. One of the most interesting properties that can be realized in some layer-constructed topological states is the string-string braiding statistics, in addition to the more conventional particle-string braiding that exists in discrete gauge theories. We provide an explicit example of a state with non-trivial string-string braiding, and also discuss the relation between string-string braiding and particle-string braiding by making use of fusion and splitting procedures of strings. A topological field theory is proposed
to describe the topological order and string braiding statistics of the layer constructed systems. The topological field theory has the form of BF theory ``twisted" by an ``axionic" topological coupling of the gauge field with an axion field $\theta$. Based on the explicit examples and field theory, we further discuss the general properties of three-string braiding, and give a more general proof of the string braiding identities Eq. (\ref{TriId1}), (\ref{TriId3}) and (\ref{TriId4}) with Eq. (\ref{TriId3}) being a a modified version of the previously proposed identity (\ref{TriId2ml}) proved for $Z_N^k$ lattice gauge theories. Finally, we discussed the generalization of the 3D topological order and string-string braiding to non-Abelian systems by proposing one example system which is a superconductor with chiral vortex strings. Compared with Abelian strings, the non-Abelian strings have multiple fusion channels, non-Abelian braiding phase, and topological degeneracy depending on string linking.
There are a lot of interesting open questions along the direction of this work. The first question is how to write down lattice models which realizes the coupled layers and anyon condensation.
For layers of $Z_p$ toric code with condensation described in Fig. \ref{LGT3d} (a), the lattice model is simply the 3D toric code. This can be understood in the following way. Take $p=2$ as an example. Consider the cubic lattice with $Z_2$ degrees of freedoms defined on each link. The vertex term on the vertex $v$ is given by $\mathcal{O}_v=-\prod_{l \text{ connects to v} } \sigma^z_l$ with $l$ labeling the link, and the plaquette term at plaquette $p$ is given by $\mathcal{O}_p=-\prod_{l \in \text{ plaquette }p } \sigma^x_l$. Now consider adding a term $h \sigma^z_l$ for all vertical links with $h$ a coupling constant. In the $h\rightarrow \infty$ limit, all the vertical links are polarized, 3D toric code splits up into layers of 2D toric code. As we lower the coupling $h$, a transition that is described in Fig. \ref{LGT3d} (a) will happen and bring the system back to the topological ground state of 3D toric code. Its $Z_p$ and twisted version can also be constructed in a similar fashion. It would be interesting to construct lattice models that capture the physics of more generic 3D states obtained through layer construction.
Another interesting question is the relation of the layer constructed states with the Walker-Wang models. Some similarities are discussed in Sec. \ref{Coexist} between these two systems, and the Walker-Wang's work\cite{walker2012} set a general mathematical framework of defining 3D topological states. However, string-string braiding has not been discussed in Walker-Wang model, so it will be an interesting future direction.
If incorporated with symmetry, the layer construction for 3D state with purely surface topological order can be used to build models for symmetry protected topological (SPT) states with gapped surfaces, as has been discussed in Ref. \onlinecite{wang2013}. Our analysis on the side surface can serve as a Hamiltonian formalism for these topologically ordered surfaces of 3D SPTs which allow us to study more carefully the symmetry breaking transition on the surface.
Another future direction is a more systematic understanding of the 3D topological order. For 3D gapped state with purely surface topological order and trivial bulk topological order, we obtained some general criteria on when it occurs and what is the surface topological order (Sec. \ref{GCpSTO}). However, for more generic 3D states with bulk and/or surface topological order, we have only done a case-by-case study of the topological order for each given set of condensed anyons. More specifically, what is needed is a general formula which determines the topological properties of the 3D state (such as the $Q$ and $R$ matrices in the field theory approach) from the ``microscopic" data, {\it i.e.} the layer $K$ matrix before the anyon condensation, and the condensed anyons given by $p_i$ and $q_i$ vectors. It is also interesting to consider anyon condensation and layer construction in non-Abelian layers, using the technique developed in literature\cite{Bais2009, Bais2013,Kong2013}. A specific question will be how to obtain the topological superconductor state with chiral vortex from layer construction.
Moreover, we would like to make some more general comments on the string braiding statistics in 3D topologically ordered states. In Sec. \ref{GeneralStrBrd}, we have shown that, for Abelian strings with trivial fusion and splitting within the same species, the string braiding statistics can be defined in various of different ways with configurations of different linking numbers, which yield exactly the same Berry's phase. This implies that linked string configuration (Fig. \ref{ThreeLoop}) might the objects that play a central role in 3D topological order in general. For example, by applying our analysis on the identification between three-string braiding (Fig. \ref{StringBrd} (a)) and the link-string braiding (Fig. \ref{StringBrd} (b)) to the Dijkgraaf-Witten lattice $Z_N^k$ or $Z_{N_1}\times Z_{N_2}\times Z_{N_3}\times Z_{N_4}$ gauge theories studied in Ref. \onlinecite{Levin2014StrBrd,Ran2014StrBrd, JuvenWenStrBrd2014}, we immediately conclude that the link between two flux strings is in fact charged under the gauge group. Further, our example of chiral vortex in TSC and the result in Ref. \onlinecite{Wen2014StrBrd, JuvenWenStrBrd2014} show that the link of flux strings or themselves can even carry non-Abelian degrees of freedom. It would be very interesting to construct the mathematical framework that captures the braiding, link and fusion of the strings and study the various operations to the linked string configurations discussed in Sec. \ref{GeneralStrBrd} in generality.
At a final comment, we would like to memtion that Ref. \onlinecite{Kong2014BF} recently proposed a general theory for topological order in any dimension. It would be interesting to study its connection to the layer construced models and to the general properties of string statistics in 3D.
{\it Acknowledgements} We would like to thank Maissam Barkeshli, Ling-Yan Hung, Senthil Todadri, Chong Wang, Zhenghan Wang and Xiao-Gang Wen for help discussions. This work is supported by the BOCO fellowship (CMJ) and David and Lucile Packard Foundation (XLQ).
|
\section{Introduction}\label{intro-sec}
We consider the linear regression setup with high dimensional
covariates where the number of covariates $p$ can be large relative to
the sample size $n$. When $p > n$, the estimation problem is ill-posed
without performing variable selection. A natural assumption to limit
the number of parameters in high dimensional settings is that the
regression function (i.e., the conditional mean) is sparse in the sense
that only a small number of covariates (called active covariates) have
nonzero coefficients. We aim to develop a new Bayesian methodology for
selecting the active covariates that is asymptotically consistent and
computationally convenient. A~large number of methods have been
proposed for variable selection in the literature from both frequentist
and Bayesian viewpoints. Many frequentist methods based on penalization
have been proposed following the well-known least absolute shrinkage
and selection operator [LASSO, \citet{Tibs96}]. We mention the smoothly
clipped absolute deviation [SCAD, \citet{Fan01}], adaptive LASSO [\citet
{Zou06}], octagonal shrinkage and clustering algorithm for regression
[OSCAR, \citet{Bondell08}] and the Dantzig selector [\citet{Candes07};
\citet{James09}] just to name a few. \citet{Fan10} provided a selective
overview of high dimensional variable selection methods. Various
authors reported inconsistency of LASSO and its poor performance for
variable selection under high dimensional settings; see \citet{Zou06}
and \citet{Johnson12}. On the other hand, several penalization based
methods were shown to have the oracle property [\citet{Fan01}] under
some restrictions on~$p$. For example, \citet{Fan04} and \citet{Huang07}
showed the oracle property for some nonconcave penalized likelihood
methods when $p = O(n^{1/3})$ and $p = o(n)$, respectively. \citet
{Shen12} showed that $L_0$ penalized likelihood method has the oracle
property under exponentially large $p = e^{o(n)}$.
Many Bayesian methods have also been proposed for variable selection
including the stochastic search variable selection [\citet{George93}],
empirical Bayes variable selection [\citet{George00}], spike and slab
selection method [\citet{Ishwaran05}], penalized credible regions
[\citet
{Bondell12}], nonlocal prior method [\citet{Johnson12}], among others.
We shall describe the typical framework used for Bayesian variable
selection methods before discussing their theoretical properties.
We use the standard notation $Y_{n \times1} = X_{n \times p} \beta_{p
\times1} + \varepsilon_{n \times1}$ to represent the linear regression
model. Bayesian variable selection methods usually introduce latent
binary variables for each of the covariates to be denoted by $Z = (Z_1,
\ldots, Z_p)$. The idea is that each $Z_i$ would indicate whether the
$i$th covariate is active in the model or not. For this reason, the
prior distribution on the regression coefficient $\beta_i$ under $Z_i =
0$ is usually a point mass at zero, but a diffused (noninformative)
prior under $Z_i =1$. The concentrated prior of $\beta_i$ under $Z_i =
0$ is referred to as the spike prior, and the diffused prior under $Z_i
= 1$ is called the slab prior. Further, a prior distribution on the
binary random vector $Z$ is assumed, which can be interpreted as a
prior distribution on the space of models. A Bayesian variable
selection method then selects the model with the highest posterior
probability. Various selection procedures with this structure have been
proposed; they essentially differ in the form of the spike and slab
priors, or in the form of the prior on the model space.
\citet{Mitchell88} considered a uniform distribution for the slab prior.
\citet{George93} used the Gaussian distribution with a zero mean and a
small but fixed variance as the spike prior, and another Gaussian
distribution with a large variance as the slab prior. This allowed the
use of a Gibbs sampler to explore the posterior distribution of $Z$.
However, as we argue in Section~\ref{ortho-sec}, this prior
specification does not guarantee model selection consistency at any
fixed prior.
\citet{Ishwaran05} also used Gaussian spike and slab priors, but with
continuous bimodal priors for the variance of $\beta$ to alleviate the
difficulty of choosing specific prior parameters. More recently, \citet
{Ishwaran11} established the oracle property for the posterior mean as
$n$ converges to infinity (but $p$ is fixed) under certain conditions
on the prior variances. They noted that in the orthogonal design case,
a uniform complexity prior leads to correct complexity recovery (i.e.,
the expected size of the posterior model size converges to the true
model size) under weaker conditions on the prior variances.
In another development, \citet{Yang12} used shrinking priors to explore
commonality across quantiles in the context of Bayesian quantile
regression, but the use of such priors for achieving model selection
consistency has not been explored. In this paper, we continue to work
with the framework where both the spike and slab priors are Gaussian,
but our prior parameters depend explicitly on the sample size through
which appropriate shrinkage is achieved. We shall establish model
selection consistency properties for general design matrices while
allowing $p$ to grow with $n$ at a nearly exponential rate. In
particular, the strong selection consistency property we establish is a
stronger result for model selection than complexity recovery.
One of the most commonly used priors on the model space is the
independent prior given by $P[Z = z] = \prod_{i =1}^p w_i^{z_i} (1 -
w_i)^{z_i}$, where the marginal probabilities $w_i$ are usually taken
to be the same constant. However, when $p$ is diverging, this implies
that the prior probability on models with sizes of order less than $p$
goes to zero, which is against model sparsity. We consider marginal
probabilities $w_i$ in the order of $p^{-1}$, which will impose
vanishing prior probability on models of diverging size. \citet{Yuan05}
used a prior that depends on the Gram matrix to penalize models with
unnecessary covariates at the prior level. The vanishing prior
probability in our case achieves similar prior penalization.
A common notion of consistency for Bayesian variable selection is
defined in terms of pairwise Bayes factors, that is, the Bayes factor
of any under- or over-fitted model with respect to the true model goes
to zero. \citet{Moreno10} proved that intrinsic priors give pairwise
consistency when $p = O(n)$, and similar consistency of the Bayesian
information criterion [BIC, \citet{Schwarz78}] when $p = O(n^{\alpha}),
\alpha<1$. Another notion of consistency for both frequentist and
Bayesian methods is that the selected model equals the true model with
probability converging to one. We refer to this as selection
consistency. \citet{Bondell12} proposed a method based on penalized
credible regions that is shown to be selection consistent when $\log p
= O(n^c), c<1$. \citet{Johnson12} proposed a stronger consistency for
Bayesian methods under which the posterior probability of the true
model converges to one, which we shall refer to as strong selection
consistency. The authors used nonlocal distributions (distributions
with small probability mass close to zero) as slab priors, and proved
strong selection consistency when $p < n$. However, apart from the
limitation $p < n$, their method involves approximations of the
posterior distributions and an application of MCMC methods, which are
computationally intensive if at all feasible for modest size problems.
We make the following contributions to variable selection in this
article. We introduce shrinking and diffusing priors as spike and slab
priors, and establish strong selection consistency of the approach for
$p = e^{o(n)}$. This approach is computationally advantageous because a
standard Gibbs sampler can be used to sample from the posterior. In
addition, we find that the resultant selection on the model space is
closely related to the $L_0$ penalized likelihood function. The merits
of the $L_0$ penalty for variable selection have been discussed by many
authors including \citet{Schwarz78}, \citet{Liu07}, \citet{Dicker13},
\citet
{Kim12} and \citet{Shen12}.
We now outline the remaining sections of the paper as follows. The
first part of Section~\ref{model-sec} describes the model, conditions
on the prior parameters and motivation for these conditions. The latter
part describes our proposed methodology for variable selection based on
the proposed model. Section~\ref{ortho-sec} motivates the use of sample
size dependent prior parameters by considering orthogonal design
matrices, and provides insight into the variable selection mechanism
using those priors. Section~\ref{results-sec} presents our main results
on the convergence of the posterior distribution of the latent vector
$Z$, and the strong selection consistency of our model selection
methodology. Section~\ref{penalty-sec} provides an asymptotic
connection between the proposed method and the $L_0$ penalization.
Section~\ref{cond-sec} provides a discussion on the conditions assumed
for proving the results of Section~\ref{results-sec}. Some
computational aspects of the proposed method are noted in Section~\ref{compu-sec}. We present simulation studies in Section~\ref{simul-sec}
to illustrate how the proposed method compares with some existing
methods. Application to a gene expression data set is given in
Section~\ref{data-sec}, followed by a conclusion in Section~\ref{conclusion-sec}. Section~\ref{proofs-sec} provides proofs of some
results not given in the earlier sections.
\section{The model} \label{model-sec}
From now on, we use $p_n$ to denote the number of covariates to
indicate that it grows with $n$. Consider the $n \times1$ response
vector $Y$, and the $n\times p_n$ design matrix $X$ corresponding to
the $p_n$ covariates of interest. Let $\beta$ be the regression vector,
that is, the conditional mean of $Y$ given $X$ is given by $X\beta$. We
assume that $\beta$ is sparse in the sense that only a few components
of $\beta$ are nonzero; this sparsity assumption can be relaxed as in
Condition~\ref{model-cond}. Our goal is to identify the nonzero
coefficients to learn about the active covariates. We describe our
working model as follows:
\begin{eqnarray}\label{modeleq}
\label{model} &&\hspace*{12pt}Y \mid\bigl( X, \beta, \sigma^2 \bigr) \sim N\bigl(X
\beta, \sigma^2 I\bigr),
\nonumber
\\
&&\beta_i \mid \bigl( \sigma^2, Z_i = 0
\bigr) \sim N\bigl(0, \sigma^2 {\tau ^2_{0,n}}
\bigr),
\nonumber
\\
&&\beta_i \mid \bigl(\sigma^2, Z_i = 1
\bigr) \sim N\bigl(0,\sigma^2 {\tau ^2_{1,n}}
\bigr),
\\
&&P(Z_i = 1 ) =1 - P(Z_i = 0 ) = q_n,
\nonumber
\\
&&\sigma^2 \sim \operatorname{IG}(\alpha_1,\alpha_2),\nonumber
\end{eqnarray}
where $i$ runs from 1 to $p_n$, $q_n, \tau_{0,n}, \tau_{1,n}$ are
constants that depend on $n$, and $\operatorname{IG}(\alpha_1,\alpha_2)$ is the
Inverse Gamma distribution with shape parameter $\alpha_1$ and scale
parameter $\alpha_2$.
The intuition behind this set-up is that the covariates with zero or
very small coefficients will be identified with zero $Z$ values, and the
active covariates will be classified as $Z = 1$. We use the posterior
probabilities of the latent variables $Z$ to identify the active
covariates.
\textit{Notation}:
We now introduce the following notation to be used throughout the
paper.
\textit{Rates}: For sequences $a_n$ and $b_n$, $a_n \sim b_n$ means
$\frac{a_n}{b_n} \rightarrow c$ for some constant $c >0$, $a_n \succeq
b_n$ (or $b_n \preceq a_n$) means $b_n = O(a_n)$, and $a_n \succ b_n$
(or $b_n \prec a_n$) means $b_n = o(a_n)$.
\textit{Convergence}: Convergence in probability is denoted by $\cprob
$, and equivalence in distribution is denoted by $\eqdist$.
\textit{Models}: We use $k$ to index an arbitrary model which is viewed
as a $p_n \times1$ binary vector. The $i$th entry $k_i$ of $k$
indicates whether the $i$th covariate is active (1) or not (0). We
use $X_k$ as the design matrix corresponding to the model $k$, and
$\beta_k$ to denote the corresponding regression coefficients. In
addition, $t$ is used to represent the true model.
\textit{Model operations}: We use $|k|$ to represent the size of the
model $k$. For two models $k$ and $j$, the operations $k \vee j$ and $k
\wedge j$ denote entry-wise maximum and minimum, respectively.
Similarly, $k^c = \mathbf{1} - k$ is entrywise operation, where
$\mathbf
{1}$ is the vector of 1's. We also use the notation $k \supset j$ (or
$k \geq j$) to denote that the model $k$ includes all the covariates in
model $j$, and $k \not\supset j$ otherwise.
\textit{Eigenvalues}: We use $\phi_{\mathrm{min}}(A)$ and $\phi_{\mathrm{max}}(A)$ to
denote the minimum and maximum eigenvalues, respectively, and $\phi
_{\mathrm{min}}^{\#}(A)$ to denote the minimum nonzero eigenvalue (MNEV) of the
matrix $A$. Moreover, we use $\lambda_M^n$ to be the maximum eigenvalue
of the Gram matrix $X'X/n$, and for $\nu> 0$, we define
\[
m_n(\nu) = p_n \wedge
\frac{n}{(2 + \nu)\log p_n }\quad \mbox{and}\quad \lambda_m^n (\nu): = \inf
_ {|k| \leq m_n(\nu)} \phi_{\mathrm{min}}^{\#} \biggl(
\frac{X_k'
X_k}{n} \biggr).
\]
\textit{Matrix inequalities}: For square matrices $A$ and $B$ of the
same order, $A \geq B$ or $ (A - B) \geq0$ means that $(A-B)$ is
positive semidefinite.
\textit{Residual sum of squares}: We define $\tilde{R}_k = Y' (I -
X(D_k + X'X)^{-1}X') Y$, where $D_k = \operatorname{Diag} (k \tau_{1n}^{-2} +
(\mathbf
{1} -k) \tau_{0n}^{-2})$. $\tilde{R}_k$ approximates the usual residual
sum of squares $R_k^* = Y' ( I - P_k )Y$, where $P_k$ is the
projection matrix corresponding to the model~$k$.
\textit{Generic constants}: We use $c'$ and $w'$ to denote generic
positive constants that can take different values each time they appear.
\subsection{Prior parameters} \label{prior-sec} We consider ${\tau
^2_{0,n}} \rightarrow0$ and ${\tau^2_{1,n}} \rightarrow\infty$ as $n$
goes to $\infty$, where the rates of convergence depend on $n$ and
$p_n$. To be specific, we assume that for some $\nu>0$, and $\delta>0$,
\[
n \tau_{0n}^2
\lambda_M^n= o(1)\quad \mbox{and}\quad n \tau_{1n}^{2}
\lambda _m^n (\nu) \sim \bigl(n \vee
p_n^{2 + 2 \delta} \bigr).
\]
As will be seen later, these rates ensure desired model selection
consistency for any $\delta> 0$, where larger values of $\delta$ will
correspond to higher penalization and vice versa.
Note that the variance $\tau_{0n}^{2}$ depends on the sample size $n$
and the scale of the Gram matrix. Since the prior distribution of a
coefficient under $Z=0$ is mostly concentrated in
\[
\biggl(-\frac{3 \sigma}{\sqrt{n \lambda_M^n }}, \frac{3 \sigma
}{\sqrt{n
\lambda_M^n }}
\biggr),
\]
one can view this as the shrinking neighborhood around 0 that is being
treated as the region of inactive coefficients. The variance $\tau
_{1n}^{2}$ increases to $\infty$, where the rate depends on $p_n$.
However, when $p_n \prec\sqrt{n}$, $\tau_{1n}^{2}$ can be of constant
order [if $\lambda_m^n(\nu)$ is bounded away from zero].
Now consider the prior probability that a coefficient is nonzero
(denoted by $q_n$). The following calculation gives insight into the
choice of $q_n$. Let $K_n$ be a sequence going to $\infty$, then
\[
P \Biggl( \sum_{i =1}^{p_n}
Z_i> K_n \Biggr) \approx1 - \Phi \biggl(
\frac{K_n - p_n q_n}{\sqrt{p_n q_n (1-q_n)}} \biggr) \longrightarrow 0,
\]
if $p_n q_n$ is bounded. Therefore, we typically choose $q_n$ such that
$q_n \sim p^{-1}_n$. This can be viewed as a priori penalization of the
models with large size in the sense that the prior probability on
models with diverging number of covariates goes to zero. To this
respect, if $K$ is an initial upper bound for the size of the model
$t$, by choosing $q_n = c/p_n$ such that $\Phi ((K - c)/\sqrt{c}
) \approx1 - \alpha$, our prior probability on the models with
sizes greater than $K$ will be $\alpha$.
We would like to note that the hierarchical model considered by \citet
{George93} is similar to our model \eqref{modeleq}, but their prior
parameters are fixed and, therefore, do not satisfy our conditions. In
Section~\ref{ortho-sec}, we give an example illustrating model
selection inconsistency under fixed prior parameters.
\subsection{Methodology for variable selection}
We use the posterior distribution of the latent variables $Z_i$ to
select the active covariates. Note that the sample space of~$Z$,
denoted by $M$, has $2^{p_n}$ points, each of which corresponds to a
model. For this reason, we call $M$ the model space. To find the model
with the highest posterior probability is computationally challenging
for large $p_n$. In this paper, we use a simpler alternative, that is,
we use the $p_n$ marginal posterior probabilities $P(Z_i = 1|Y, X)$,
and select the covariates with the corresponding probability more than
a fixed threshold $\underbar{p} \in(0, 1)$. A threshold probability of
$0.5$ is a natural choice for $\underbar{p}$. This corresponds to what
\citet{Barbieri04} call the median probability model. In the orthogonal
design case, \citet{Barbieri04} showed that the median probability model
is an optimal predictive model. The median probability model may not be
the same as the maximum a posteriori (MAP) model in general, but the
two models are the same with probability converging to one under strong
selection consistency.
On the other hand, \citet{Dey08} argued that the median probability
model tends to underfit in finite samples. We also consider an
alternative by first ranking the variables based on the marginal
posterior probabilities and then using BIC to choose among different
model sizes. This option avoids the need to specify a threshold. In
either case, it is computationally advantageous to use the marginal
posterior probabilities, because we need fewer Gibbs iterations to
estimate only $p_n$ of them. The proposed methods based on marginal
posteriors achieve model selection consistency because the results in
Section~\ref{results-sec} assure that (i) the posterior probability of
the true model converges to~1, and (ii) the marginal posterior based
variable selection selects the true model with probability going to~1.
We now motivate these results and the necessity of sample size
dependent priors in a simple but illustrative case with orthogonal designs.
\section{Orthogonal design}\label{ortho-sec}
In this section, we consider the case where the number of covariates
$p_n < n$, and assume that the design matrix $X$ is orthogonal, that
is, $X'X = n I$. We also assume $\sigma^2$ to be known. Though this may
not be a realistic set-up, this simple case provides motivation for the
necessity of sample size dependent prior parameters as well as an
insight into the mechanism of model selection using these priors. At
this moment, we do not impose any assumptions on the prior parameters.
\textit{All the probabilities used in the rest of the paper are
conditional on~$X$}. Under this simple set-up, the joint posterior of
$\beta$ and $Z$ can be written as
\begin{eqnarray*}
&&P\bigl(\beta, Z \mid\sigma^2, Y
\bigr)
\\
&&\qquad\propto\exp \biggl\{{-\frac{1}{2 \sigma^2} \|Y - X\beta\| _2^2}
\biggr\} \prod_{i=1}^{p_n}
\bigl((1-q_n) \pi_0(\beta_i)
\bigr)^{1 - Z_i} \bigl(q_n \pi_1(
\beta_i) \bigr)^{Z_i}
\\
&&\qquad\propto\exp \biggl\{-\frac{1}{2 \sigma^2} \bigl(\beta'
X'X \beta- 2\beta 'X'Y\bigr) \biggr\}
\prod_{i=1}^{p_n} \bigl((1-q_n)
\pi_0(\beta_i) \bigr)^{1
- Z_i}
\bigl(q_n\pi_1(\beta_i)
\bigr)^{Z_i}
\\
&&\qquad\propto\exp \Biggl\{-\frac{n}{2 \sigma^2} \sum_{i=1}^p
(\beta_i - \hat{\beta}_i)^2 \Biggr\} \prod
_{i=1}^{p_n} \bigl((1-q_n)\pi
_0(\beta _i) \bigr)^{1 - Z_i}
\bigl(q_n\pi_1(\beta_i)
\bigr)^{Z_i},
\end{eqnarray*}
where for $k =0, 1$, $\pi_k(x) = \boldsymbol\phi(x, 0, \sigma^2
\tau
^2_{k,n})$ is the probability density function (p.d.f.) of the normal
distribution with mean zero and variance $\sigma^2 {\tau^2_{k,n}}$
evaluated at $x$, and $\hat{\beta}_i$ is the OLS estimator of $\beta
_i$, that is, $\hat{\beta}_i = X_i'Y/n$.
The product form of the joint posterior of $(Z_i, \beta_i)$ implies
that $(Z_i, \beta_i)$ and $\{(Z_j, \beta_j), j \neq i \}$ are
independent given data. Hence, the marginal posterior of $Z_i$ is given by
\[
P\bigl(Z_i \mid
\sigma^2, Y\bigr) \propto\int\exp \biggl\{-\frac{n}{2 \sigma
^2} (b -
\hat{\beta}_i)^2 \biggr\} \bigl((1-q_n)
\pi_0(b) \bigr)^{1 - Z_i} \bigl(q_n
\pi_1(b) \bigr)^{Z_i} \,db.
\]
Therefore,
\begin{equation}
\label{post-orth}
P\bigl(Z_i = 0 \mid
\sigma^2, Y\bigr) = \frac{(1-q_n) E_{\hat{\beta}_i} (\pi_0
(B))}{(1-q_n)E_{\hat{\beta}_i} (\pi_0 (B)) +q_n E_{\hat{\beta}_i}
(\pi
_1 (B))},
\end{equation}
where $E_{\hat{\beta}_i}$ is the expectation under $B$ following the
normal distribution with mean $\hat{\beta}_i$ and variance $ \sigma^2/n
$. These expectations can be calculated explicitly, that is, for $k$ =
0 and 1,
\begin{eqnarray*}
\label{exp-orth}
E_{\hat{\beta}_i} \bigl(
\pi_{k} (B) \bigr)& =& \frac{\sqrt {n}}{2\pi
\sigma\tau_{k,n}} \int\exp \biggl\{{-
\frac{n}{2 \sigma^2} (b- \hat {\beta}_i)^2} -
\frac{{b}^2}{2 {\tau^2_{k,n}}} \biggr\}\,db
\\
& =& \frac{1}{\sqrt{2 \pi}a_{k,n}} \exp \biggl\{-\frac{\hat{\beta
}^2_i}{2 a_{k,n}^2} \biggr\},
\end{eqnarray*}
where $a_{k,n} = \sqrt{\sigma^2/n + {\tau^2_{k,n}}}$.
This simple calculation gives much insight into the role of our priors
and the influence of the prior parameters on variable selection, which
we explain in some detail below. In the following subsections, we
assume that the $i$th covariate is identified as active if and only
if $P(Z_i = 1 \mid\sigma^2, Y) > 0.5$ for simplicity, and similar
arguments can be produced for threshold values other than 0.5.
\subsection{Fixed parameters}
Let us first consider the case of fixed parameters $\tau_{0n}^2 = \tau
_0^2 < \tau_{1n}^2 = \tau_1^2 $ and $q_n = q = 0.5$. We then have for
$k = 0, 1$,
\begin{equation}
E_{\hat{\beta}_i} \bigl(\pi_{k} (B)
\bigr) \cprob\frac{1}{{\tau_k}} \exp \biggl\{-\frac{\beta^2_i}{ 2 \tau^2_k}
\biggr\} \qquad\mbox{as } n \rightarrow\infty\mbox{ for } \beta_i \neq0.
\end{equation}
Now for $\beta_i = \tau_0 \neq0$, we have $ \exp \{- \beta
^2_i /
2 \tau^2_0 \}/\tau_0 > \exp \{-\beta^2_i/2 \tau^2_1
\}/\tau_1$ for any $\tau_1 \neq\tau_0$. Therefore, the limiting value
of $P(Z_i = 1 \mid\sigma^2, Y)$ will be less than 0.5 (with high
probability) as $n \rightarrow\infty$. This implies that even as $n
\rightarrow\infty$, we would not be able to identify the active
coefficient in this case.
\subsection{Shrinking \texorpdfstring{$\tau_{0,n}^2$}{tau 0,n 2}, fixed \texorpdfstring{$\tau_{1,n}^2$}{tau 1,n 2} and $q_n$}
Now consider the prior parameters such that $\tau_{1,n}^2$ and $q_n$ are
fixed, but $\tau_{0,n}^2$ goes to 0 with $n$. If $\beta_i = 0$,
$\sqrt {n} \hat{\beta}_i$ converges in distribution to the standard normal
distribution, and we have, for $k = 0, 1$,
\[
\exp \biggl\{-\frac{\hat{\beta}^2_i}{2 (\sigma^2/n) + 2 \tau^2_{k,n}} \biggr\}
=
O_P(1).
\]
In this case, \eqref{exp-orth} will imply that $E_{\hat{\beta}_i}
(\pi
_{1} (B)) = O_p(1)$, while $E_{\hat{\beta}_i} (\pi_{0} (B)) \cprob
\infty$. Therefore, from \eqref{post-orth}, we have $P(Z_i = 0 \mid
\sigma^2, Y) \cprob1$. For $\beta_i \neq0$, using $\hat{\beta}^2_i
\cprob{\beta}^2_i$ and the fact that $x e^{-rx^2}\rightarrow0$ as $x
\rightarrow\infty$ (for fixed $r >0$), we obtain $E_{\hat{\beta}_i}
(\pi_0 (B)) \rightarrow0$. As $E_{\hat{\beta}_i} (\pi_1 (B)) \sim c'$,
for some $c'>0$, we have $P(Z_i = 1 \mid\sigma^2, Y) \cprob1$.
To summarize, we have argued that $P(Z_i = 0 \mid\sigma^2, Y) \cprob
I{(\beta_i =0)}$, where $I(\cdot)$ is the indicator function.
That is, for orthogonal design matrices, the marginal posterior
probability of including an active covariate or excluding an inactive
covariate converges to one under shrinking prior parameter $\tau
_{0,n}^2$, with fixed parameters $\tau_{1,n}^2$ and $q_n$. However, it
should be noted that this statement is restricted to the convergence of
marginals of $Z$, and does not assure consistency of overall model
selection. To achieve this, we will need to allow $\tau_{1,n}^2$, $q_n$
to depend on the sample size, too.
\subsection{Shrinking and diffusing priors}
Note that the $i$th covariate is identified as active if and only if
\begin{eqnarray*}
&&P\bigl(Z_i = 1 \mid
\sigma^2, Y\bigr) > 0.5
\\
&&\qquad\Leftrightarrow q_n E_{\hat{\beta}_i} \bigl(\pi_1 (B)
\bigr) > (1-q_n) E_{\hat
{\beta}_i} \bigl(\pi_0 (B)\bigr)
\\
&&\qquad\Leftrightarrow\hat{\beta}^2_i \bigl(
a_{0,n}^{-2} - a_{1,n}^{-2} \bigr)> 2
\bigl(\log(1- q_n) a_{1,n} - \log q_n
a_{0,n} \bigr)
\\
&&\qquad\Leftrightarrow\hat{\beta}^2_i > 2 \bigl(\log(1-
q_n) a_{1,n} - \log q_n a_{0,n} \bigr)/
\bigl(a_{0,n}^{-2} - a_{1,n}^{-2}\bigr):=
\varphi_n.
\end{eqnarray*}
In particular, when $\tau^2_{0,n} = o(1/n)$, but the other parameters
$\tau^2_{1,n}$ and $q_n$ are fixed, we have $\varphi_n \sim\sigma^2
\log n /n $. Without loss of generality, assume that the first $|t|$
coefficients of $\beta$ are nonzero. For $i > |t|$, $\beta_i = 0$ which
implies that $n \hat{\beta}^2_i \eqdist\chi^2_1$. Therefore,
\begin{eqnarray*} P\biggl[\hat{\beta}^2_i
> \frac{\sigma^2 \log n }{n}\biggr] & = & P\bigl[\chi^2_1> \log
n\bigr]
\\
& \geq& \biggl(\frac{1}{\sqrt{\log n}} - \frac{1}{\sqrt{\log n}^3}\biggr)
e^{-{\log n}/{2}}
\\
& \geq& n^{-1/2 - \varepsilon},
\end{eqnarray*}
for $\varepsilon>0$ and sufficiently large $n$. Therefore, we have
\begin{eqnarray*} P\bigl[Z = t | \sigma^2, Y\bigr] &
\leq& P \biggl[\hat{\beta}^2_i \leq\frac
{\sigma
^2 \log n }{n},
\forall i > |t| \biggr]
\\
&\leq& \bigl(1 - n^{-1/2 - \varepsilon}\bigr)^{p_n - |t|}
\\
&\rightarrow& 0\qquad \mbox{if } p_n > n^{1/2 + 2\varepsilon}.
\end{eqnarray*}
The above argument shows that having $\tau^2_{1,n}$ and $q_n$ fixed
leads to inconsistency of selection if the number of covariates is much
greater than $\sqrt{n}$. In this case, the threshold $\varphi_n$ should
be larger to bound the magnitude of all the inactive covariates
simultaneously. By using the diffusing prior parameters Section~\ref{prior-sec}, the threshold will be $(2 + \delta) \sigma^2 \log p_n /n$
in place of $ \sigma^2 \log n/n $. Model selection consistency with
this threshold can be proved using similar arguments in the orthogonal
design case. We will defer the rigorous arguments to the next section.
\section{Main results}\label{results-sec}
In this section, we consider our model given by \eqref{modeleq} and
general design matrices. Because the model selection consistency holds
easily with $p_n = O(1)$, we assume throughout the paper that $p_n
\rightarrow\infty$ as $n \rightarrow\infty$.
\subsection{Conditions}
We first state the main conditions we use.
\begin{condition}[(On dimension $p_n$)] \label{dim-cond} $p_n = e^{n
d_n}$ for some $d_n\rightarrow0$ as $n \rightarrow\infty$, that is,
$\log p_n = o(n)$.
\end{condition}
\begin{condition}[(Prior parameters)]\label{prior-cond}
$n \tau_{0n}^2 = o(1)$, $ n \tau_{1n}^{2} \sim (n\vee p_n^{2 + 3
\delta} )$, for some $\delta>0$, and $q_n \sim p^{-1}_n$.
\end{condition}
\begin{condition}[(On true model)]\label{model-cond}
$Y|X \sim N (X_t \beta_t + X_{ t^c} \beta_{t^c}, \sigma^2 I)$ where the
size of the true model $|t|$ is fixed. The coefficients corresponding
to the inactive covariates can be nonzero but satisfy $b_0:= \|X_{ t^c}
\beta_{t^c}\|_2 = O(1)$.
\end{condition}
For any fixed $K$, define
\[
\Delta_n(K):= \inf_{\{k: |k| < K |t|, k \not\supset t\}} \bigl\|(I - P_{k})
X_t \beta_t\bigr\|_2^2,
\]
where $P_{k}$ is the projection matrix onto the column space of $X_k$.
\begin{condition}[(Identifiability)] \label{identify-cond}
There is $K > 1 + 8/\delta$ such that $\Delta_n(K) > \gamma_n:= 5
\sigma
^2 |t| (1 + \delta) \log ( \sqrt{n} \vee p_n )$.
\end{condition}
\begin{condition}[(Regularity of the design)]\label{eigen-cond}
For some $\nu< \delta$, $\kappa< (K-1) \delta/2$,
\[
\lambda_M^n \prec
\bigl( \bigl(n \tau_{0n}^2\bigr)^{-1} \wedge n
\tau_{1n}^2 \bigr)\quad \mbox{and}\quad \lambda_m^n
(\nu) \succeq \biggl( \frac{ n\vee p_n^{2
+ 2
\delta} }{n \tau_{1n}^{2}} \vee p_n^{- \kappa}
\biggr).
\]
\end{condition}
The moderateness of these conditions will be examined in some detail in
Section~\ref{cond-sec}.
\subsection{Results for fixed \texorpdfstring{$\sigma^2$}{sigma2}} We suppress $\nu$ and $K$
from the notation of $\lambda_m^n(\nu)$, $m_n(\nu)$ and $\Delta_n(K)$
for stating the results for convenience. In addition, we introduce the
following notation.
The posterior ratio of model $k$ with respect to the true model $t$ is
defined as
\[
\operatorname{PR}(k, t):={P\bigl(Z = k\mid Y,
\sigma^2\bigr)}/{P\bigl(Z = t\mid Y, \sigma^2\bigr)}.
\]
The following lemma gives an upper bound on the posterior ratio.
\begin{lem}\label{bf-lem}
Under Conditions \ref{prior-cond} and \ref{eigen-cond}, for any model
$k \neq t$, we have
\begin{eqnarray*}
\operatorname{PR}(k, t) &=& \frac{Q_k}{Q_t}
s_n^{|k| -|t|} \exp \biggl\{ -\frac{1}{2
\sigma^2} (
\tilde{R}_k - \tilde{R}_t) \biggr\}
\\
&\leq& w' \bigl(n \tau_{1n}^2
\lambda_m^n (1 - \phi_n)
\bigr)^ {-({1}/{2}) (r_{k}^* - r_t)} \bigl(\lambda_m^n
\bigr)^{- ({1}/{2}) |t \wedge
k^c|} s_n^{|k| -|t|}\\
&&{}\times\exp \biggl\{ -\frac{1}{2 \sigma^2} (\tilde{R}_k - \tilde
{R}_t) \biggr\},
\end{eqnarray*}
where $Q_k = | I + X D_k^{-1}X'|^{-1/2}$, $s_n = q_n/(1 - q_n) \sim
p_n^{-1}$, $w' > 0$ is a constant, $r_k = \operatorname{rank}( X_k)$, $r_k^* = r_k
\wedge m_n$, $\phi_n = o(1)$, $ \tilde{R}_k = Y' (I - X(D_k +
X'X)^{-1}X') Y $, and $D_k = \operatorname{Diag} (k \tau_{1n}^{-2} + (\mathbf{1} -k)
\tau_{0n}^{-2})$.
\end{lem}
The following arguments give some heuristics for the convergence of
pair-wise posterior ratio. Note that $\tilde{R}_k $ is the residual sum
of squares from a shrinkage estimator of $\beta$, and the term $LR_n:=
\exp\{ - (\tilde{R}_k - \tilde{R}_t)/2 \sigma^2 \}$ corresponds to the
usual likelihood ratio of the two models $k$ and $t$. Consider a model
$k$ that does not include one or more active covariates, then $ (\tilde
{R}_k - \tilde{R}_t)$ goes to $\infty$ at the same rate as $n$, because
it is (approximately) the difference in the residual sums of squares of
model $k$ and model $t$. We then have the posterior ratio converging to
zero since $LR_n \sim e^{-c n}$ for some $c>0$, and due to
Conditions \ref{dim-cond}--\ref{eigen-cond}, $P_n:= (n \tau_{1n}^2
\lambda_m^n (1 - \phi_n) )^ {(r_t - r^*_k)/2} ({\lambda
_m^n})^{-|t \wedge k^c|/2} s_n^{|k| -|t|} (1 - \phi_n)^{-|t|/2} =
o(e^{cn})$. On the other hand, if the model $k$ includes all the active
covariates and one or more inactive covariates, we have $|k| > |t|$,
but $(\tilde{R}_k - \tilde{R}_t)$ is probabilistically bounded. The
posterior ratio in this case also converges to zero because $P_n$ goes to
zero. Note that when $r_k > r_t $, larger values of $\tau_{1n}^2$ will
imply smaller $P_n$. That is, the posterior ratio for large sized models
go to zero faster for larger values of $\tau_{1n}^2$. A~similar
observation is made by \citet{Ishwaran11}. To state our main result, we
first consider the posterior distributions of the models $Z$, assuming
the variance parameter $\sigma^2$ to be known. We consider the case
with the prior on $\sigma^2$ in Theorem~\ref{sigprior-thmm}.
\begin{thmm}\label{sigfix-thmm}
Assume Conditions \ref{dim-cond}--\ref{eigen-cond}. Under
model \eqref
{modeleq}, we have $P(Z = t \mid Y, \sigma^2) \cprob1$ as $n
\rightarrow\infty$, that is, the posterior probability of the true
model goes to 1 as the sample size increases to $\infty$.
\end{thmm}
\begin{remark} \label{sum-bf}
The statement of Theorem~\ref{sigfix-thmm} is equivalent to
\begin{equation}
\label{sum-bfeq}
\frac{1 - P(Z = t\mid Y, \sigma^2)}{P(Z = t\mid Y, \sigma^2)} = \sum
_{k \neq t} \operatorname{PR}(k, t) \cprob0.
\end{equation}
\end{remark}
\begin{remark} It is worth noting that for Theorem~\ref{sigfix-thmm} to
hold, we do not actually need the true $\sigma^2$ to be known. Even for
a misspecified $\tilde{\sigma}^2 \neq\sigma^2$, $P(Z = t \mid Y,
\tilde
{\sigma}^2) \cprob1$ under the conditions $\Delta_n > \tilde{\sigma}^2
\gamma_n/ \sigma^2$ and $2 (1 +\delta) \tilde{\sigma}^2 > (2 +
\delta)
\sigma^2$. The same proof for Theorem~\ref{sigfix-thmm} works.
\end{remark}
To see why \eqref{sum-bfeq} holds, we provide specific rates of
convergence of individual posterior ratio summed over subsets of the
model space. We divide the set of models (excluding the model $t$) into
the following subsets:
\begin{longlist}[1.]
\item[1.]Unrealistically large models: $M_1 = \{k\dvtx r_k > m_n\}$, all the
models with dimension (i.e., the rank) greater than $ m_n$.
\item[2.]Over-fitted models: $M_2 = \{k\dvtx k \supset t, r_k \leq m_n\}$,
that is, the models of dimension smaller than $ m_n $ which include all
the active covariates plus one or more inactive covariates.
\item[3.]Large models: $M_3 = \{k\dvtx k \not\supset t, K |t| < r_k \leq
m_n \}$, the models which do not include one or more active covariates,
and dimension greater than $K |t|$ but smaller than $ m_n$.
\item[4.]Under-fitted models: $M_4 = \{k\dvtx k \not\supset t, r_k \leq K
|t| \}$, the models of moderate dimension which miss an active covariate.
\end{longlist}
The proof of Theorem~\ref{sigfix-thmm} shows the following results.
\begin{lem}[(Rates of convergence)]\label{rates-lem}
For some constants $c', w' >0 $ (which may depend on $\delta$), we have
\begin{longlist}[1.]
\item[1.] The sum of posterior ratio $\sum_{ k \in M_1}\operatorname{PR}(k, t) \preceq\exp
\{- w' n \}$, with probability at least $1 - 2 \exp\{ - c' n \}$.
\item[2.] The sum $\sum_{k \in M_2} \operatorname{PR}(k, t) \preceq v_n:=
(p_n^{-\delta/2}\wedge\frac{p_n^{1+\delta/2}}{ \sqrt{n}} )$,
with probability greater than $1 - \exp\{ - c' \log p_n\}$.
\item[3.] The sum $\sum_{k \in M_3} \operatorname{PR}(k, t) \preceq\nu_n^{ (K-1) |t|/2
+1}$, with probability greater than $1 - \exp\{ - c' K |t| \log p_n\}$.
\item[4.] For some $w'' <1$, we have $\sum_{k \in M_4} \operatorname{PR}(k, t) \preceq
\exp\{ - w' (\Delta_n - w''\gamma_n)\}$, with probability greater than
$1 - \exp\{ - c' \Delta_n \}$.
\end{longlist}
\end{lem}
\subsection{Results with prior on \texorpdfstring{$\sigma^2$}{sigma2}}
We now consider the case with the inverse Gamma prior on the variance
parameter $\sigma^2$. Define the constant $w$ as $w:= \delta/8 (
1+\delta)^2$ in the rest of the section.
\begin{thmm}\label{sigprior-thmm}
Under the same conditions as in Theorem~\ref{sigfix-thmm}, if we only
consider models of dimension at most $ |t| + w n/\log p_n$, we have
$P(Z = t \mid Y) \cprob1$ as $n \rightarrow\infty$.
\end{thmm}
\begin{remark}
Note that the dimension of the models that need to be excluded for
Theorem~\ref{sigprior-thmm} to hold is in the order of $n/\log p_n$.
These are unrealistically large models that are uninteresting to us.
From now on, we implicitly assume this restriction when a prior
distribution is used for $\sigma^2$.
\end{remark}
The following corollary ensures that the variable selection procedure
based on the marginal posterior probabilities finds the right model
with probability tending to 1. It is a direct consequence of
Theorems \ref{sigfix-thmm} and \ref{sigprior-thmm}, but is particularly
useful for computations because it ensures that the marginal posterior
probabilities can be used for selecting the active covariates.
\begin{cor}\label{marg-cor}
Under the conditions of Theorem~\ref{sigprior-thmm}, we have for any $0
< \underbar{\textup{p}} < 1$,
$P [ P(Z_i = t_i \mid Y) > \underbar{\textup{p}} \mbox{ for all } i = 1,
\ldots, p_n ] \rightarrow1$ as $n \rightarrow\infty$.
\end{cor}
\begin{pf}
Let $E_i$ be the event that the marginal posterior probability of
$i$th covariate $P(Z_i =t_i \mid Y) > \underbar{p}$. We shall show
that $P[ \bigcup_{i = 1}^{p_n} E_i^c] \rightarrow0$ as $n \rightarrow
\infty$. For each $i = 1, \ldots, p_n$, we have
\begin{eqnarray*}
P(Z_i \neq t_i \mid Y)
& = & \sum_{k:k_i \neq t_i} P(Z = k \mid Y)
\\
& \leq& \sum_{k \neq t} P(Z = k \mid Y)
\\
& = & 1 - P(Z = t|Y).
\end{eqnarray*}
Then $P[ \bigcup_{i = 1}^{p_n} E_i^c] = P [ P(Z_i = t_i \mid Y) \leq
\underbar{p} \mbox{ for some } i = 1, \ldots, p_n ] \leq P [P(Z
= t|Y) \leq\underbar{p} ] \rightarrow0$, due to Theorem~\ref
{sigprior-thmm}.
\end{pf}
\section{Connection with penalization methods}\label{penalty-sec}
Due to Lemma~\ref{bf-lem}, the maximum aposteriori (MAP) estimate of
the model using our Bayesian set-up is equivalent to minimizing the
objective function
\begin{eqnarray}
\label{obj-fn}
B(k) &:=& \tilde{R}_k
+ 2 \sigma^2 \bigl( - \bigl(|k| - |t|\bigr) \log s_n - \log
(Q_k /Q_t) \bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& = & \tilde{R}_k + \bigl(|k| - |t|\bigr) \psi_{n,k},
\end{eqnarray}
where
\[
\psi_{n,k} = 2 \sigma^2
\biggl( - \log s_n - \frac
{\log ( Q_k /Q_t )}{(|k| - |t|) } \biggr).
\]
Lemma~\ref{rates-lem} implies that with exponentially small
probability, the sum of posterior ratio of the models with dimension
greater than $ m_n $ goes to zero (exponentially) for the fixed $\sigma
$ case. We therefore focus on all the models with dimension less than
$m_n$ in this section. In addition, assume that the maximum and minimum
nonzero eigenvalues of models of size $2|t|$ are bounded away from
$\infty$ and 0, respectively. Then, due to Condition \ref{eigen-cond}
and the proof of Lemma~\ref{scale-lem}(iii), we have
\begin{equation}
\label{penal-bound}
c \log(n \vee p_n)
\leq- \frac{\log ( Q_k/Q_t )} {(r_k -
r_t) } \leq C \log(n \vee p_n),
\end{equation}
for some $0 < c \leq C < \infty$.
In particular, if the models with dimension less than $m_n$ are of full
rank, that is, $|k| = r_k$, then due to \eqref{penal-bound}, we have
\begin{equation}
\label{psi-bound}
2 \sigma^2
c' \log(n \vee p_n) \leq\psi_{n,k} \leq2
\sigma^2 C' \log (n \vee p_n),
\end{equation}
where $0 < c' \leq C' < \infty$. As $ n \tau_{0n}^2 \lambda_{M}^n
\rightarrow0$, and $n \tau_{1n}^2 \lambda_{m}^n \rightarrow\infty$,
\[
\tilde{R}_k \sim Y' \bigl(I - X \bigl(1/
\tau_{1n}^2 + X' X\bigr)^{-1}X'
\bigr) Y = \| Y - \hat{Y}_{k}\|^2 + O(1).
\]
Therefore, the MAP estimate can be (asymptotically) described as the
model corresponding to minimizing the following objective function:
\begin{equation}
\label{l0}
m (\beta):= \| Y - X \beta
\|_2^2 + \psi_{n,k} \bigl(\|\beta\|_0 -
|t|\bigr).
\end{equation}
Due to the bounds \eqref{psi-bound} on $\psi_{n,k}$, any inactive
covariate will be penalized in the order of $\log(n \vee p_n)$
irrespective of the size of the coeffecient. This is however not the
case with the $L_1$ penalty or SCAD penalty, which are directly
proportional to the magnitude of the coefficient in some interval
around zero.
The commonly used model selection criteria AIC and BIC are special
cases of $L_0$ penalization. The objective functions of AIC and BIC are
similar to $m(\beta)$, which have the quotient of penalty equal to $2$
and $\log n$ in place of $\psi_{n,k}$. Due to the results in Section~\ref{results-sec} and the above arguments, selection properties of our
proposed method are similar to those of the $L_0$ penalty. In
particular, it attempts to find the model with the least possible size
that could explain the conditional mean of the response variable.
A~salient feature of our approach is that the $L_0$-type penalization is
implied by the hierarchical model. The tuning parameters are more
transparent than those in penalization methods. Another feature to note
is that our model allows high (or even perfect) correlations among
inactive covatiates. This is practically very useful in high
dimensional problems because the number of inactive covariates is often
large and the singularity of the design matrix is a common occurrence.
Also, high correlations between active and inactive covariates is not
as harmful to the proposed method as they are to the $L_1$-type
penalties. This point is illustrated in Table~\ref{tab-highcor} of our
simulation studies in Section~\ref{simul-sec}.
\section{Discussion of the conditions}\label{cond-sec}
The purpose of this section is to demonstrate that Conditions~\ref
{dim-cond}--\ref{eigen-cond} that we use in Section~\ref{results-sec}
are quite mild. Condition~\ref{dim-cond} restricts the number of
covariates to be no greater than exponential in $n$, and Condition \ref
{prior-cond} provides the shrinking and diffusing rates for the spike
and slab priors, respectively. We note that Conditions \ref
{model-cond}--\ref{eigen-cond} allow $\beta$ to depend on $n$. For
instance, consider $p_n <n$ and the design matrix $X$ with $X'X/n
\rightarrow D$, where $D$ is a positive definite matrix. \citet
{Ishwaran05}, \citet{Zou06}, \citet{Bondell12} and \citet{Johnson12}
assumed this condition on the design under which Conditions \ref
{model-cond} and \ref{identify-cond} only require $\beta$ to be such that
\[
\|\beta_{t^c}\|_2^2
= O \biggl(\frac{1}{n} \biggr) \quad\mbox{and}\quad \| \beta _t
\|_2^2 > c' \frac{\log n}{n},
\]
for some $c' >0$. Condition \ref{eigen-cond} is also satisfied in this
case, so Conditions \ref{model-cond}--\ref{eigen-cond} allow a wider
class of design matrices.
In general, Condition \ref{identify-cond} is a mild regularity
condition that allows us to identify the true model. It serves to
restrict the magnitude of the correlation between active and inactive
covariates, and also to bound the signal to noise ratio from below. The
following two remarks provide some insight into the role of Condition
\ref{identify-cond} in these aspects.
\begin{remark}\label{identify-rem}
Consider the case where the active coefficients $\beta_t$ are fixed.
We then have some $w' >0$, such that
\begin{eqnarray*}
\Delta_n (K) &\geq& \|
\beta_t\|_2^2 \inf_{\{k: |k| < K |t|, k \not
\supset t\}}
\phi_{\mathrm{min}} \bigl( X_t' (I - P_{k})
X_t \bigr)
\\
& \geq& w'n \inf_{\{k: |k| < K |t|, k \not\supset t\}} \phi_{\mathrm{min}}
\biggl(\frac{ X_{k \vee t}' X_{k \vee t}}{n} \biggr),
\end{eqnarray*}
where we have used the fact that $\phi_{\mathrm{min}} ( X_{k \vee t}' X_{k
\vee t} ) \leq\phi_{\mathrm{min}} ( X_t' (I - P_{k}) X_t
)$. To
see this, we just need to consider the cases where $X_{k \vee t}$ is of
full rank. Then it follows from the observation that $ ( X_t' (I -
P_{k}) X_t )^{-1}$ is a submatrix of $ ( X_{k \vee t}' X_{k
\vee t} )^{-1}$.
Therefore, Condition \ref{identify-cond} is satisfied if the minimum
eigenvalues of the submatrices of $X'X/n$ with size smaller than $(K+1)
|t|$ are uniformly larger than $c' \log(n \vee p_n)/n$. In the other
end of the spectrum, where the inactive covariates can be perfectly
correlated, Condition \ref{identify-cond} could still hold.
\end{remark}
\begin{remark} If the infimum of $\phi_{\mathrm{min}} ( X_t' (I - P_{k})
X_t /n )$ is uniformly bounded away from zero, then $\Delta_n (K)
\geq w' n \|\beta_t\|_2^2$. Then Condition \ref{identify-cond} is
satisfied if
\[
\biggl\llVert \frac{\beta_t}{\sigma}\biggr\rrVert _2^2 \geq
\frac{c' \log(n
\vee p_n)}{n}.
\]
\end{remark}
Condition \ref{eigen-cond} provides conditions on the eigenvalues of
the Gram matrix in terms of the prior parameters. The condition is
weaker than the assumption that the maximum and minimum nonzero
eigenvalues of the Gram matrix are bounded away from infinity and zero,
respectively. In Condition \ref{eigen-cond}, $\lambda_M^n \prec(n
\tau
_{0n}^2)^{-1}$ will be satisfied if $\tau_{0n}^2$ is small enough.
However, the assumption on $\lambda_m^n (\nu) $ is nontrivial as it
needs to be greater than $p_n^{-\kappa}$. We now show that this
requirement is satisfied with high probability if the design matrix
consists of independent sub-Gaussian rows.
\begin{lem}[(MNEV for sub-Gaussian random matrices)] \label{mnev-lemma}
Suppose that the rows of $X_{n \times p_n}$ are independent isotropic
sub-Gaussian random vectors in $R^{p_n}$. Then there exists a $\nu>0$
such that, with probability greater than $1 - \exp(- w' n)$,
\[
\inf_{|k| \leq m_n(\nu)}
\phi_{\mathrm{min}} \biggl(\frac{X_k' X_k}{n} \biggr) >0.
\]
\end{lem}
A proof of Lemma~\ref{mnev-lemma} is provided in Section~\ref{proofs-sec}. Lemma~\ref{mnev-lemma} implies that the Gram matrix of a
sub-Gaussian design matrix has the minimum eigenvalues of all the
$m_n(\nu)$ dimensional\vadjust{\goodbreak} submatrices to be uniformly bounded away
from
zero. This clearly is stronger than Condition \ref{eigen-cond}, which
only requires the minimum nonzero eigenvalues to be uniformly greater
than $p_n ^{-\kappa}$. In particular, unlike the restricted isometry
conditions which control the minimum eigenvalue, Condition \ref
{eigen-cond} allows the minimum eigenvalue to be exactly zero to allow
even perfect correlation among inactive (or active) covariates.
\section{Computation} \label{compu-sec}
The implementation of our proposed method involves using the Gibbs
sampler to draw samples from the posterior of $Z$. The full
conditionals are standard distributions due to the use of conjugate
priors. The conditional distribution of $\beta$ is given by
\[
f\bigl(\beta\mid Z, \sigma^2, Y
\bigr) \propto\exp \biggl\{{-\frac{1}{2 \sigma^2
} \|Y - X\beta\|_2^2}
\biggr\} \prod_{i=1}^{p_n} \phi \bigl(
\beta_i, 0, \sigma^2 \tau^2_{Z_i,n}
\bigr),
\]
where $ \phi(x, 0, \tau^2)$ is the p.d.f. of the normal distribution with
mean zero, and variance $\tau^2$ evaluated at $x$.
This can be rewritten as
\[
f\bigl(\beta\mid Z = k, \sigma^2, Y
\bigr) \propto\exp \biggl\{-\frac{1}{2
\sigma
^2} \bigl(\beta'
X' X \beta- 2\beta'X'Y\bigr) \biggr\}
\exp \biggl\{-\frac{1}{2
\sigma^2} \beta' D_k \beta
\biggr\},
\]
where $D_k = \operatorname{Diag} (\tau_{k_i, n}^{-2})$. Hence, the conditional
distribution of $\beta$ is given by $\beta\sim N(m, \sigma^2 V)$,
where $V =(X' X + D_k)^{-1}$, and $ m = V X'Y$. Furthermore, the
conditional distribution of $Z_i$ is
\[
P\bigl(Z_i = 1\mid\beta, \sigma^2\bigr) =
\frac{q_n \phi(\beta_i, 0, \sigma^2
{\tau^2_{1,n}} )}{q_n \phi(\beta_i, 0, \sigma^2 {\tau^2_{1,n}} ) +
(1-q_n) \phi(\beta_i, 0, \sigma^2 {\tau^2_{0,n}} ) }.
\]
The conditional of $\sigma^2$ is the inverse Gamma distribution $\operatorname{IG}(a,
b)$ with $a = \alpha_1 + n/2 + p_n/2$, and $b = \alpha_2 + \beta'D_k
\beta/2 + (Y - X \beta)' (Y - X\beta)/2$.
The only possible computational difficulty in the Gibbs sampling
algorithm is the step of drawing from the conditional distribution of
$\beta$, which is a high dimensional normal distribution for large
values of $p_n$. However, due to the structure of the covariance matrix
$(X' X + D_k)^{-1}$, it can be efficiently sampled using block updating
that only requires drawing from smaller dimensional normal
distributions. Details of the block updating can be found in \citet{Ishwaran05}.
\section{Simulation study}\label{simul-sec}
In this section, we study performance of the proposed method in several
experimental settings, and compare it with some existing variable
selection methods. We will refer to the proposed method as BASAD for
BAyesian Shrinking And Diffusing priors.
The proposed BASAD method has three tuning parameters. In all our
empirical work, we use
\[
\tau_{0n}^2=
\frac{\hat\sigma^2 }{10n}, \qquad \tau_{1n}^2= \hat
\sigma^2 \max \biggl( \frac{p_n^{2.1}}{100 n}, \log n \biggr),
\]
where $\hat\sigma^2$ is the sample variance of $Y$, and we choose $q_n
= P[Z_i = 1]$ such that $P [\sum_{i=1}^{p_n} Z_i = 1 >K ] =
0.1$, for a prespecified value of $K$. Our default value is $K=\max(10
, \log(n))$, unless otherwise specified in anticipation of a less
sparse model. The purpose of using $\hat\sigma^2$ is to provide
appropriate scaling. If a preliminary model is available, it is better
to use as $\hat\sigma^2$ the residual variance from such a model. It is
clear that those choices are not optimized for any given problem, but
they provide a reasonable assessment on how well BASAD can do. In the
simulations, we use 1000 burn-in iterations for the Gibbs sampler
followed by 5000 updates for estimating the posterior probabilities. As
mentioned in Section~\ref{model-sec}, we consider both the median
probability model (denoted by BASAD) and the BIC-based model (denoted
by BASAD.BIC) where the threshold for marginal posterior probability is
chosen by the BIC.
The R function used for obtaining the results in this section is publicly
available on the authors' website.
In this paper, we report our simulation results for six cases under
several $(n, p)$ combinations, varied correlations, signal strengths
and sparsity levels.
\begin{itemize}
\item \textit{Case} 1: In the first case, we use the set-up of \citet{Johnson12}
with $p = n$. Two sample sizes, $n = 100$ and $n = 200$, are
considered, and the covariates are generated from the multivariate
normal distributions with zero mean and unit variance. The compound
symmetric covariance with pairwise covariance of $\rho= 0.25$ is used
to represent correlation between covariates. Five covariates are taken
active with coefficients $\beta_t = (0.6, 1.2, 1.8, 2.4, 3.0)$. This is
a simple setting with moderate correlation between covariates and
strong signal strength.
\item \textit{Case} 2: We consider the $p >n$ scenario with $(n, p) = (100,
500)$ and $(n, p) = (200, 1000)$, but the other parameters are same as
in case 1.
For the next three cases (cases 3--5), we keep $(n,p) = (100,500)$ but
vary model sparsity, signal strength and correlation among covariates.
\item \textit{Case} 3: We keep $\rho= 0.25 $ and $|t|=5$ but have low signals
$\beta_t = (0.6, 0.6, 0.6,\break 0.6, 0.6)$.
\item \textit{Case} 4: We consider a block covariance setting where the active
covariates have common correlation ($\rho_1$) equal to 0.25, the
inactive covariates have common correlation ($\rho_3$) equal to 0.75
and each pair of active and inactive covariate has correlation ($\rho
_2$) 0.50. The other aspects of the model are the same as in case 1.
\item \textit{Case} 5: We consider a less sparse true model with $|t| = 25$ and
$\beta_t$ is the vector containing 25 equally spaced values between 1
and 3 (inclusive of 1 and 3).
\item \textit{Case} 6: We consider the more classical case of $n > p$ with
$(n,p) = (100,50)$ and $(n,p) = (200,50)$. Following \citet{Bondell12},
the covariates are drawn from a normal distribution with the covariance
matrix distributed as the Wishart distribution centered at the identity
matrix with $p$ degrees of freedom. Three of the 50 covariates are
taken to be active with their coefficients drawn from the uniform
distribution $U(0, 3)$ to imply a mix of weak and strong signals.
\end{itemize}
\begin{table}
\caption{Performance of BASAD for case 1:
$n = p$. The methods under comparison are $\mathit{piMOM}$ with nonlocal priors
of
\citet{Johnson12}, $\mathit{BCR.Joint}$ of \citet{Bondell12}, $\mathit{SpikeSlab}$ of
\citet{Ishwaran05} and three penalization methods Lasso, elastic net
(EN), and SCAD tuned by the BIC. The other columns of the table are as
follows: $pp_0$ and $pp_1$ (when applicable) are the average posterior
probabilities of inactive and active variables, respectively; $Z=t$ is
the proportion that the exact models is selected; $Z \supset t$ is the
proportion that the selected model contains all the active covatiates;
FDR is the false discovery rate, and $\mathit{MSPE}$ is the mean squared
prediction error of the selected models}
\label{tab1}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccc@{}}
\hlin
& $\bolds{pp_0}$ & $\bolds{pp_1}$ & $ \bolds{Z = t} $& $ \bolds{Z \supset t} $
& \textbf{FDR}& \textbf{MSPE}\\
\hline
&\multicolumn{6}{c}{$(n, p) = (100, 100); \rho= 0.25; |t| = 5 $} \\
BASAD&0.016&0.985&0.866&0.954&0.015&1.092\\
BASAD.BIC&0.016&0.985&0.066&0.996&0.256&1.203\\
piMOM&0.012&0.991&0.836&0.982&0.030&1.083\\
BCR.Joint&&&0.442&0.940&0.157&1.165\\
SpikeSlab&&&0.005&0.216&0.502&1.660\\
Lasso.BIC&&&0.010&0.992&0.430&1.195\\
EN.BIC&&&0.398&0.982&0.154&1.134\\
SCAD.BIC&&&0.356&0.990&0.160&1.157\\[3pt]
&\multicolumn{6}{c}{$(n, p) = (200, 200); \rho= 0.25; |t| = 5 $} \\
BASAD&0.002&1.000&0.944&1.000&0.009&1.037\\
BASAD.BIC&0.002&1.000&0.090&1.000&0.187&1.087\\
piMOM&0.003&1.000&0.900&1.000&0.018&1.038\\
BCR.Joint&&&0.594&0.994&0.102&1.064\\
SpikeSlab&&&0.008&0.236&0.501&1.530\\
Lasso.BIC&&&0.014&1.000&0.422&1.101\\
EN.BIC&&&0.492&1.000&0.113&1.056\\
SCAD.BIC&&&0.844&1.000&0.029&1.040\\
\hline
\end{tabular*}
\end{table}
\begin{table}
\caption{Performance of BASAD for case 2: $p >
n$}\label{tab2}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccc@{}}
\hlin
& $\bolds{pp_0}$ & $\bolds{pp_1}$ & $ \bolds{Z = t} $& $ \bolds{Z \supset t} $
& \textbf{FDR}& \textbf{MSPE}\\
\hline
&\multicolumn{6}{c}{$(n, p) = (100, 500); \rho= 0.25; |t| = 5$} \\
BASAD&0.001&0.948&0.730&0.775&0.011&1.130\\
BASAD.BIC&0.001&0.948&0.190&0.915&0.146&1.168\\
BCR.Joint&&&0.070&0.305&0.268&1.592\\
SpikeSlab&&&0.000&0.040&0.626&3.351\\
Lasso.BIC&&&0.005&0.845&0.466&1.280\\
EN.BIC&&&0.135&0.835&0.283&1.223\\
SCAD.BIC&&&0.045&0.980&0.328&1.260\\[3pt]
&\multicolumn{6}{c}{$(n, p) = (200, 1000); \rho= 0.25; |t| = 5$} \\
BASAD&0.000&0.986&0.930&0.950&0.000&1.054\\
BASAD.BIC&0.000&0.986&0.720&0.990&0.046&1.060\\
BCR.Joint&&&0.090&0.250&0.176&1.324\\
SpikeSlab&&&0.000&0.050 &0.574 &1.933\\
Lasso.BIC&&&0.020&1.000&0.430&1.127\\
EN.BIC&&&0.325&1.000&0.177&1.077\\
SCAD.BIC&&&0.650&1.000&0.091&1.063\\
\hline
\end{tabular*}
\end{table}
\begin{table}[b]
\caption{Performance of BASAD for case 3: $(n, p) =
(100, 500)$}\label{tablowsignal}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccc@{}}
\hlin
& $\bolds{pp_0}$ & $\bolds{pp_1}$ & $ \bolds{Z = t} $& $ \bolds{Z \supset t} $
& \textbf{FDR}& \textbf{MSPE}\\
\hline
&\multicolumn{6}{c}{$ \rho= 0.25; |t| = 5; \beta_t = (0.6,
0.6,0.6,0.6,0.6)$} \\
BASAD&0.002&0.622&0.185&0.195&0.066&2.319\\
BASAD.BIC&0.002&0.622&0.160&0.375&0.193&1.521\\
BCR.Joint&&&0.030&0.315&0.447&1.501\\
SpikeSlab&&&0.000&0.000&0.857&2.466\\
Lasso.BIC&&&0.000&0.520&0.561&1.555\\
EN.BIC&&&0.040&0.345&0.478&1.552\\
SCAD.BIC&&&0.045&0.340&0.464&1.561\\
\hline
\end{tabular*}
\end{table}
\begin{table}
\caption{Performance of BASAD for case 4: $(n, p) =
(100, 500)$}\label{tab-highcor}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccc@{}}
\hlin
& $\bolds{pp_0}$ & $\bolds{pp_1}$ & $ \bolds{Z = t} $& $ \bolds{Z \supset t} $
& \textbf{FDR}& \textbf{MSPE}\\
\hline
&\multicolumn{6}{c}{$\rho_1 = 0.25, \rho_2 = 0.50, \rho_3 = 0.75$}
\\
BASAD&0.002&0.908&0.505&0.530&0.012&\phantom{0}1.199\\
BASAD.BIC&0.002&0.908&0.165&0.815&0.179&\phantom{0}1.210\\
BCR.Joint&&&0.000&0.000&0.515&\phantom{0}2.212\\
SpikeSlab&&&0.000&0.000&0.995&10.297\\
Lasso.BIC&&&0.000&0.015&0.869&\phantom{0}8.579\\
EN.BIC&&&0.000&0.000&0.898&\phantom{0}8.360\\
SCAD.BIC&&&0.000&0.000&0.899&\phantom{0}8.739\\
\hline
\end{tabular*}
\end{table}
\begin{table}[b]
\caption{Performance of BASAD for case 5: $(n, p) =
(100, 500)$. In this case, two versions of BASAD are included, where
BASAD.K10 uses our default value of $K=10$, and BASAD.K50 uses a less
sparse specification of $K=50$}\label{tabnz25}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccd{3.3}@{}}
\hlin
& $\bolds{pp_0}$ & $\bolds{pp_1}$ & $ \bolds{Z = t} $& $ \bolds{Z \supset t} $
& \textbf{FDR}& \multicolumn{1}{c@{}}{\textbf{MSPE}}\\
\hline
&\multicolumn{6}{c}{$ \rho= 0.25; |t| = 25 $} \\
BASAD.K50&0.020&0.988&0.650&0.950&0.036&3.397\\
BASAD.BIC.K50&0.020&0.988&0.005&0.960&0.283&4.019\\
BASAD.K10&0.003&0.548&0.405&0.420& 0.011&170.862\\
BASAD.BIC.K10& 0.003&0.548&0.035&0.430&0.076&88.881\\
BCR.Joint&&&0.000&0.000&0.622&49.299\\
SpikeSlab&&&0.000&0.000&0.816&111.911\\
Lasso.BIC&&&0.000&0.005&0.685&58.664\\
EN.BIC&&&0.000&0.000&0.693&59.058\\
SCAD.BIC&&&0.000&0.000&0.666&72.122\\[3pt]
&\multicolumn{6}{c}{$ \rho= 0.75; |t| = 25 $} \\ [0.5ex]
BASAD.K50&0.048&0.914&0.005&0.355&0.289&6.103\\
BASAD.BIC.K50&0.048&0.914&0.000&0.445&0.498&6.611\\
BASAD.K10&0.003&0.298&0.025&0.030&0.018&349.992\\
BASAD.BIC.K10&0.003&0.298&0.000&0.060&0.087&61.709\\
BCR.Joint&&&0.000&0.000&0.772&34.113\\
SpikeSlab&&&0.000&0.000&0.899&48.880\\
Lasso.BIC&&&0.000&0.000&0.734&24.310\\
EN.BIC&&&0.000&0.000&0.754&29.171\\
SCAD.BIC&&&0.000&0.000&0.736&27.236\\
\hline
\end{tabular*}
\end{table}
\begin{table}
\caption{Performance of BASAD for case 6: $n >
p$}\label{tabnmore
\centerin
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccc@{}}
\hlin
& $\bolds{pp_0}$ & $\bolds{pp_1}$ & $ \bolds{Z = t} $& $ \bolds{Z \supset t} $
& \textbf{FDR}& \textbf{MSPE}\\
\hline
&\multicolumn{6}{c}{$(n,p) = (100,50)$} \\
BASAD&0.037&0.899&0.654&0.714&0.026&1.086\\
BASAD.BIC&0.037&0.899&0.208&0.778&0.267&1.151\\
piMOM&0.011&0.892&0.656&0.708&0.021&1.066\\
SpikeSlab&&&0.064&0.846&0.567&1.226\\
BCR.Joint&&&0.336&0.650&0.216&1.124\\
Lasso.BIC&&&0.076&0.744&0.397&1.152\\
EN.BIC&&&0.378&0.742&0.194&1.110\\
SCAD.BIC&&&0.186&0.772&0.284&1.147\\[3pt]
&\multicolumn{6}{c}{$(n,p) = (200,50)$} \\
BASAD&0.026&0.926&0.738&0.784&0.017&1.029\\
BASAD.BIC&0.026&0.926&0.338&0.842&0.193&1.055\\
piMOM&0.005&0.908&0.694&0.740&0.020&1.036\\
BCR.Joint&&&0.484&0.770&0.133&1.045\\
SpikeSlab&&&0.038&0.900&0.629&1.121\\
Lasso.BIC&&&0.082&0.752&0.378&1.059\\
EN.BIC&&&0.428&0.748&0.165&1.039\\
SCAD.BIC&&&0.358&0.812&0.193&1.046\\
\hline
\end{tabular*}\vspace*{-5pt}
\end{table}
The summary of our results are presented in Tables \ref{tab1}--\ref
{tabnmore}. In those tables, BASAD denotes the median probability
model, BASAD.BIC denotes the model obtained by using the threshold
probability chosen by the BIC. Three competing Bayesian model selection
methods are: (1) piMOM, the nonlocal prior method proposed by \citet
{Johnson12} but only when $p \le n$; (2) BCR.Joint, the Bayesian joint
credible region method of \citet{Bondell12} (using the default priors
followed by an application of BIC); (3) SpikeSlab, the generalized
elastic net model obtained using the R package spikeslab [\citet
{Ishwaran10}] for the spike and slab method of \citet{Ishwaran05}. Three
penalization methods under consideration are: (1) LASSO; (2) Elastic
Net (EN); and (3) SCAD, all tuned by the BIC. Our simulation experiment
used 500 data sets from each model when $n \geq p$, but used 200 data
sets when $p > n$ to aggregate the results.
The columns of the tables show the average marginal posterior
probability assigned to inactive covariates and active covariates
($pp_0$ and $pp_1$, resp.), proportion of choosing the true
model ($ Z = t $), proportion of including the true model ($ Z \supset
t $) and false discovery rate (FDR). The last column (MSPE) gives the
average test mean squared prediction error based on $n$ new
observations as testing data. From our simulation experiment, we have
the following findings:
\begin{longlist}[(iii)]
\item[(i)] The Bayesian model selection methods BASAD and piMOM (whenever
available) tend to perform better then the other methods in terms of
selecting the true model and controlling the false discovery rate in
variable selection, and our proposed BASAD stands out in this regard.
The penalization methods often have higher probabilities of selecting
all the active covariates at the cost of overfitting and false
discoveries. In terms of the prediction error, however, BASAD does not
always outperform its competitors, but remains competitive.
\item[(ii)] When the signals are low (case 3), all the methods under
consideration have trouble finding the right model, and BASAD.BIC
results in lower prediction error than BASAD with 0.5 as the threshold
for posterior probabilities.
In most cases, BASAD.BIC leads to slightly higher false positive rates
than BASAD with similar prediction errors.
\item[(iii)] In case 4, there is a moderate level of correlation among
inactive covariates and some level of correlation between active and
inactive covariates. This is where BASAD outperforms the other methods
under consideration because BASAD is similar to the $L_0$ penalty and
is able to accommodate such correlations well. Please refer to our
discussion in Sections \ref{penalty-sec} and \ref{cond-sec}.
\item[(iv)] When the true model is not so sparse and has $|t|= 25$ active
covariates (case 5), our default choice of $K=10$ in BASAD did not
perform well, which is not surprising. In fact, no other methods under
consideration did well in this case, highlighting the difficulty of
finding a nonsparse model with a limited sample size. On the other
hand, there is some promising news. If we anticipate a less sparse
model with $K=50$, the proposed method BASAD improved the performance
considerably. Our empirical experience suggests that if we are
uncertain about the level of sparsity of our model, we may use a
generous choice of $K$ or use BIC to choose between different values of $K$.\vadjust{\goodbreak}
\end{longlist}
\section{Real data example}\label{data-sec}
In this section, we apply our variable selection method to a real data
set to examine how it works in practice. We consider the data from an
experiment conducted by \citet{Lan06} to study the genetics of two
inbred mouse populations (B6 and BTBR). The data include expression
levels of 22,575 genes of 31 female and 29 male mice resulting in a
total of 60 arrays. Some physiological phenotypes, including the
numbers of phosphoenopyruvate carboxykinase (PEPCK) and
glycerol-3-phosphate acyltransferase (GPAT) were also measured by
quantitative real-time PCR. The gene expression data and the phenotypic
data are available at GEO (\url{http://www.ncbi.nlm.nih.gov/geo}; accession
number GSE3330). \citet{Zhang09} used orthogonal components regression
to predict each phenotype based on the gene expression data. \citet
{Bondell12} used the Bayesian credible region method for variable
selection on the same data.
Because this is an ultra-high dimensional problem with $p_n = 22\mbox{,}575$,
we prefer to perform simple screenings of the genes first based on the
magnitude of marginal correlations with the response. The power of
marginal screening has been recognized by \citet{Fan08}. After the
screening, the dataset for each of the responses consisted of $p = 200$
and $400$ predictors (including the intercept and gender) by taking 198
and 398 genes based on marginal screening. We performed variable
selection with BASAD along with LASSO, SCAD and the BCR method.
Following \citet{Bondell12}, we randomly split the sample into a
training set of 55 observations and a test set with the remaining five
observations. The fitted models using the training set were used to
predict the response in the test set. This process was repeated 100
times to estimate the prediction error.
\begin{figure}
\includegraphics{1207f01.eps}
\caption{Mean squared prediction error (MSPE) versus
model size for analyzing PEPCK and GPAT in the upper and lower panel,
respectively, \textup{(a)} $p = 200$ and \textup{(b)} $p = 400$.}
\label{fig1}
\end{figure}
In Figure~\ref{fig1}, we plot the average mean square prediction error
(MSPE) for models of various sizes chosen by BASAD, BCR and SCAD
methods for the two responses PEPCK and GPAT. We find that the MSPE of
BASAD is mostly smaller than that for other methods across different
model sizes. In particular, BASAD chooses less correlated variables and
achieves low MSPE with fewer predictive genes than the other methods.
We also note that the 10-covariate models chosen by BASAD is very
different (with the overlap of just one covariate for PEPCK and three
covariates for GPAT) from those of SCAD which chose mostly the same
covariates as LASSO. There are four common covariates identified by
both BASAD and BCR methods. When we perform a linear regression by
including the covariates chosen by BASAD and SCAD, we noticed that
majority of the covariates chosen by BASAD are significant, which
indicates that those genes chosen by BASAD are significant in
explaining the response even in the presence of those chosen using
SCAD. Most of the genes selected by SCAD, however, are not significant
in the presence of those chosen by BASAD. Despite the evidence in favor
of the genes selected by BASAD in this example, we must add that the
ultimate assessment of a chosen model would need to be made by
additional information from the subject matter science and/or
additional experiment.
\section{Conclusion}\label{conclusion-sec}
In this paper, we consider a Bayesian variable selection method for
high dimensional data based on the spike and slab priors with shrinking
and diffusing priors. We show under mild conditions that this approach
achieves strong selection consistency in the sense that the posterior
probability of the true model converges to one. The tuning parameters
needed for the prior specifications are transparent, and a standard
Gibbs sampler can be used for posterior sampling. We also provide the
asymptotic relationship between the proposed approach and the $L_0$
penalty for model selection. Simulation studies in Section~\ref{simul-sec} and real data example in Section~\ref{data-sec} show
evidence that the method performs well in a variety of settings even
though we do not attempt to optimize the tuning parameters in the
proposed method.
The strong selection consistency of Bayesian methods has not been
established in the cases of $p >n$ until very recently. For higher
dimensional cases, we just became aware of \citet{Liang13}, which
provided the strong selection consistency for Bayesian subset selection
based on the theory developed by \citet{Jiang07} for posterior density
consistency. However, to translate density consistency into selection
consistency, \citet{Liang13} imposed a condition on the posterior
distribution itself, which is not verifiable directly. The techniques
we use in this paper might also be used to complete the development of
their theory on strong selection consistency.
Throughout the paper, we assume Gaussian errors in the regression
model, but this assumption is not necessary to obtain selection
consistency. For proving Lemma~\ref{bf-lem}, we did not need
assumptions on the error distribution, and to prove Theorem~\ref
{sigprior-thmm}, we just need deviation inequalities of the quadratic
forms $\varepsilon' P_k \varepsilon$, which follow the chi-squared
distribution for normal errors. Similar proofs with an application of
deviation inequalities for other error distributions would work. For
instance, \citet{Hsu12} provide deviation inequalities for quadratic
forms of sub-Gaussian random variables.
The primary focus of our paper is model selection consistency. The
model is selected by averaging over the latent indicator variables
drawn from the posterior distributions. The strengths of different
model selection methods need to be evaluated differently if prediction
accuracy is the goal. In our empirical work, we have included
comparisons of the mean squared prediction errors, and found that our
proposed method based on default tuning parameters is highly
competitive in terms of prediction. However, improvements are possible,
mainly in the cases of low signals, if the parameters are tuned by BIC
or cross-validation, or if model-averaging is used instead of the
predictions from a single model.
\section{Proofs}\label{proofs-sec}
In this section, we prove Lemmas \ref{bf-lem} and \ref{mnev-lemma}.
Please refer to \citet{Narisetty14} for proofs of the remaining results.
\begin{pf*}{Proof of Lemma~\ref{bf-lem}}
The joint posterior of $\beta, \sigma^2, Z$ under model \eqref{modeleq}
is given by
\begin{eqnarray}
\label{fullpost-eq}&& P\bigl(\beta, Z =k, \sigma^2 \mid Y\bigr)
\nonumber
\\
&&\qquad\propto\exp \biggl\{ -\frac{1}{2
\sigma^2} \bigl(\|Y - X\beta\|_2^2
- \beta' D_k \beta- 2\alpha_2 \bigr)
\biggr\} \sigma^{- 2 ({n}/{2} + {p_n}/{2} +
\alpha_1 +1 )}\\
&&\qquad\quad{}\times |D_k|^{{1}/{2}}
s_n^{|k|},\nonumber
\end{eqnarray}
where $D_k = \operatorname{Diag} (k \tau_{1n}^{-2} + (\mathbf{1} -k) \tau_{0n}^{-2})$,
$s_n =q_n/(1-q_n)$, $\alpha_1, \alpha_2$ are the parameters of {IG}
prior, and $|k|$ is the size of the model $k$. By a simple
rearrangement of terms in the above expression, we obtain
\begin{eqnarray*}
&&\hspace*{-6pt}P\bigl(\beta, Z=k \mid Y, \sigma^2\bigr)
\\
&&\hspace*{-6pt}\quad \propto\exp \biggl\{ -\frac{1}{2 \sigma^2} \bigl( (\beta- \tilde{\beta})'
\bigl(D_k + X'X\bigr) (\beta- \tilde{\beta}) - \tilde{
\beta}'\bigl(D_k + X'X\bigr)\tilde {
\beta} \bigr) \biggr\} |D_k|^{{1}/{2}} s_n^{|k|},
\end{eqnarray*}
where $\tilde{\beta} = (D_k + X'X)^{-1} X'Y$. Note that $\tilde
{\beta}$
is a shrinkage estimator of the regression vector $\beta$. Shrinkage of
$\tilde{\beta}$ depends on $D_k$, which is the precision matrix of
$\beta$ given $Z = k$. The components of $\tilde{\beta}_i$
corresponding to $k_i = 0$ are shrunk towards zero while the shrinkage
of coefficients corresponding to $k_i =1$ is negligible (as $\tau
_{1n}^{-2}$ is small).
\begin{eqnarray}
\label{post-final}
P\bigl(Z =k \mid Y,
\sigma^2\bigr) & \propto &Q_k s_n^{|k|}
\exp \biggl\{ -\frac{1}{2 \sigma^2} \bigl(Y'Y - \tilde {
\beta}'\bigl(D_k + X'X\bigr)\tilde{\beta}
\bigr) \biggr\}\nonumber
\\
& =& Q_k s_n^{|k|} \exp \biggl\{ -
\frac{1}{2 \sigma^2} \bigl(Y'Y - Y'X\bigl(D_k
+ X'X\bigr)^{-1}X'Y \bigr) \biggr\}
\\
& =& Q_k s_n^{|k|} \exp \biggl\{ -
\frac{1}{2 \sigma^2} \tilde{R}_k \biggr\}, \nonumber
\end{eqnarray}
where $Q_k = |D_k + X'X|^{-{1}/{2}} |D_k|^{{1}/{2}}$. Next, we
obtain bounds on $Q_k$.
\begin{lem}\label{scale-lem} Let $A$ be an invertible matrix, and $B$
be any matrix with appropriate dimension. Further, let $k$ and $j$ be
any pair of models. Then,
\begin{longlist}[(iii)]
\item[(i)
$ |(A + B'B)^{-1} A| = |I + B A^{-1}B'|^{-1}$,
\item[(ii)
$ ( I + \tau_{1n}^2 X_k X_k' + \tau_{0n}^2 X_j X_j' )^{-1}
\geq (I + \tau_{1n}^2 X_k X_k' )^{-1} (1 - \xi_n) $,
where $\xi_n = n \tau_{0n}^2 \lambda_M^n = o(1)$, and
\item[(iii)
$Q_k \leq w' (n \tau_{1n}^2 \lambda_m^n (1
- \phi_n) )^ {- ({1}/{2}) (r_{k}^* - r_t)} (\lambda_m^n )^{-
({1}/{2}) |t \wedge k^c|} Q_t$,
where $w' >0$, $r_k = \operatorname{rank}(X_k)$, $r_k^*= r_k \wedge m_n$,
and $\phi_n
= o(1)$.
\end{longlist}
\end{lem}
\begin{pf}
(i) We use the Sylvester's determinant theorem, and the
multiplicative property of the determinant to obtain
\begin{eqnarray*}
\bigl|\bigl(A + B'B\bigr)^{-1}
A\bigr| & = & \bigl|I + A^{-{1}/{2}} B' B A^{-{1}/{2}}\bigr|^{-1}
\\
& = &\bigl |I + B A^{-1} B'\bigr|^{-1}.
\end{eqnarray*}
(ii) By the Sherman--Morrison--Woodbury (SMW) identity,
assuming $A, C$ and $(C^{-1} + D A^{-1} B)$ to be nonsingular,
\begin{equation}
\label{smw-eq} (A + BCD)^{-1} = A^{-1} - A^{-1} B
\bigl(C^{-1} + D A^{-1} B\bigr)^{-1} D A^{-1},
\end{equation}
we have, for any vector $a$,
\[
a' \bigl(I + \tau_{1n}^2
X_k X_k' + \tau_{0n}^2
X_{j} X_{j}'\bigr)^{-1} a =
a' G^{-1} a - \tau_{0n}^2 H,
\]
where $G = I + \tau_{1n}^2 X_k X_k' $ and $H = a' G^{-1} X_{j}(I+
\tau
^2_{0n}X_{j}' G^{-1} X_{j})^{-1}X_{j}' G^{-1} a$. Note that
\begin{eqnarray}
0 \leq\tau_{0n}^2 H &\leq&
\tau_{0n}^2 a' G^{-1}
X_{j} X_{j}' G^{-1} a
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&\leq& n \tau_{0n}^2 \lambda_M^n
a' G^{-1} a,
\end{eqnarray}
where $\lambda_M^n$ is the maximum eigenvalue of the Gram matrix $X'
X/n$. Therefore,
\[
a' \bigl(I + \tau_{1n}^2
X_k X_k' \bigr)^{-1} a \bigl(1 -
n \tau_{0n}^2 \lambda_M^n\bigr)
\leq a' \bigl(I + \tau_{1n}^2 X_k
X_k' + \tau_{0n}^2
X_{j} X_{j}'\bigr)^{-1} a,
\]
and hence (ii) is proved.
(iii)
From part (i) of the lemma, we have
\begin{eqnarray}
Q_k &=&\bigl |I + XD_k^{-1}X'\bigr|^{-{1}/{2}}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& = & \bigl|I + \tau_{1n}^2 X_k
X_k' +\tau_{0n}^2
X_{k^c} X_{k^c}' \bigr|^{-
{1}/{2}}.
\end{eqnarray}
Define $A = I + \tau_{1n}^2 X_{k \wedge t} X_{k \wedge t}' +\tau_{0n}^2
X_{{k^c \vee t}^c} X_{{k^c \vee t}^c}' $. Then, by (ii) we have
\[
(1 - \xi_n) \bigl(I + \tau_{1n}^2
X_{k \wedge t} X_{k \wedge t}'^{-1}\bigr) \leq
A^{-1} \leq\bigl(I + \tau_{1n}^2 X_{k \wedge t}
X_{k \wedge t}'\bigr)^{-1}.
\]
This, along with Condition \ref{eigen-cond} implies
\begin{eqnarray*}
\label{ratio-scale1}
\frac{Q_k}{Q_{k \wedge t}} & = &
\bigl|I + \tau_{1n}^2 X_k X_k'
+\tau_{0n}^2 X_{k^c} X_{k^c}'
\bigr|^{-
{1}/{2}} |A|^{{1}/{2}}
\\
& = & \bigl|A + \bigl(\tau_{1n}^2 - \tau_{0n}^2
\bigr) X_{k \wedge t^c} X_{k \wedge
t^c}'\bigr|^{-{1}/{2}}
|A|^{{1}/{2}}
\\
& = & \bigl|I + \bigl(\tau_{1n}^2 - \tau_{0n}^2
\bigr) X_{k \wedge t^c} ' A^{-1} X_{k
\wedge t^c}
\bigr|^{- {1}/{2}}
\\
& \leq& \bigl|I + \bigl(\tau_{1n}^2 - \tau_{0n}^2
\bigr) (1 - \xi_n) X_{k \wedge t^c} ' \bigl(I +
\tau_{1n}^{2} X_{k \wedge t} X_{k \wedge t}'
\bigr)^{-1} X_{k \wedge
t^c}\bigr |^{- {1}/{2}}
\\
& = & \bigl|I + \tau_{1n}^{2} X_t
X_t' + \tau_{1n}^2 (1 -
\phi_n) X_{k
\wedge t^c} X_{k \wedge t^c}'\bigr|^{-{1}/{2}}
\bigl|I + \tau_{1n}^{2} X_{k
\wedge t} X_{k \wedge t}'
\bigr|^{{1}/{2}}
\\
& \leq&\bigl |I + \tau_{1n}^2 (1 - \phi_n)
X_k X_k'\bigr|^{-{1}/{2}}\bigl|I +
\tau_{1n}^2 X_{k \wedge t} X_{k \wedge t}'\bigr|^{{1}/{2}}
\\
&\leq& \bigl(n \tau_{1n}^2 \lambda_m^n
(1 - \phi_n) \bigr)^ {- (r_k^* - r_{t
\wedge k})/2} (1 - \phi_n)^{-|t \wedge k|/2},
\end{eqnarray*}
where $(1 - \phi_n) = (\tau_{1n}^2 - \tau_{0n}^2) (1 - \xi_n) /
\tau
_{1n}^2 \rightarrow1$.
Similarly, let $A = I + \tau_{1n}^2 X_{t} X_{t}' +\tau_{0n}^2 X_{t^c}
X_{t^c}'$ to obtain
\begin{eqnarray*}
\label{ratio-scale2}
\frac{Q_{k \wedge t}}{Q_t}
& = & \bigl|A - \bigl(\tau_{1n}^2 - \tau_{0n}^2
\bigr) X_{k \wedge t^c} X_{k \wedge
t^c}'\bigr|^{-{1}/{2}}
|A|^{{1}/{2}}
\\
& \leq& \bigl|I + \tau_{1n}^{2} X_{k \wedge t}
X_{k \wedge t}'\bigr|^{-{1}/{2}} \bigl|I + \tau_{1n}^{2}
X_t X_t' \bigr|^{{1}/{2}}
\\
&\leq&\bigl |I + \tau_{1n}^{2} X_{t \wedge k^c}
X_{t \wedge k^c}'\bigr|^{{1}/{2}}
\\
& \leq& \bigl(n \tau_{1n}^2 c'
\bigr)^{|t \wedge k^c|/2}.
\end{eqnarray*}
The above two inequalities give
\[
\label{ratio-scale}
\frac{Q_k}{Q_t}
\leq w' \bigl(n \tau_{1n}^2
\lambda_m^n (1 - \phi_n)
\bigr)^ {- (r_k^* -
r_t)/2} \bigl({\lambda_m^n}
\bigr)^{-|t \wedge k^c|/2}.
\]
\upqed\end{pf}
Due to \eqref{post-final}, we have
\[
\operatorname{PR}(k, t) = \frac{Q_k}{Q_t}
s_n^{|k| -|t|} \exp \biggl\{ -\frac{1}{2 \sigma
^2} (
\tilde{R}_k - \tilde{R}_t) \biggr\} .
\]
Therefore, Lemma~\ref{scale-lem}(iii) implies Lemma~\ref{bf-lem}.
\end{pf*}
\begin{pf*}{Proof of Lemma~\ref{mnev-lemma}}
The rows of $X_k$ are $n$ independent sub-Gaussian random isotropic
random vectors in $R^{|k|}$. Note that $|k| \leq m_n$ implies $|k| =
o(n)$. Due to Theorem~5.39 of \citet{Vershynin12}, with probability at
least $ 1 - 2 \exp(-c s)$, we have
\begin{equation}
\label{mineig-eq}
\phi_{\mathrm{min}} \biggl(
\frac{X_k' X_k}{n} \biggr) > \biggl(1 - C \sqrt {\frac
{|k|}{n}} - \sqrt{
\frac{s}{n}} \biggr)^2,
\end{equation}
where $c$ and $C$ are absolute constants that depend only on the
sub-Gaussian norms of the rows of the matrix $X_k$.
Let us fix $s= n(1 - \phi)$ for some $\phi>0$, and define the event
given by equation~\eqref{mineig-eq} as $A_k$. We then have $P[A_k^c] <
2 \exp(-c (1 - \phi) n)$ for all $k$. By taking an union bound over
$\{
k\dvtx |k| \leq m_n \}$, we obtain
\begin{eqnarray*}
P\biggl[\bigcup
_{|k| \leq m_n} A_k^c\biggr] &\leq&
p_n^{ m_n} \exp\bigl(-c (1 - \phi) n\bigr)
\\
&=& \exp{ \biggl\{ \frac{n}{2 + \nu} -c (1 - \phi) n \biggr\}} \rightarrow0,
\end{eqnarray*}
if $ \nu> (\frac{1}{c (1 - \phi)} - 2)$.
Therefore, in the event $\bigcap_{|k| \leq m_n} A_k$, whose probability
goes to~1, we have $\phi_{\mathrm{min}} (X_k' X_k/n ) \geq\phi
^2/4 -
O(\sqrt{m_n/n})>0$, for all $k$.
\end{pf*}
\section*{Acknowledgments}
The authors are grateful to anonymous
referees and an Associate Editor for their encouraging and helpful
comments on an earlier version of the paper. The authors would also
like to thank Professors Howard Bondell, Val Johnson and Faming Liang
for sharing with us their code to perform Bayesian model selection.
\begin{supplement}[id=suppA]
\stitle{Supplement to ``Bayesian variable selection with shrinking and
diffusing priors''}
\slink[doi]{10.1214/14-AOS1207SUPP}
\sdatatype{.pdf}
\sfilename{aos1207\_supp.pdf}
\sdescription{This supplement contains the proofs of Theorems \ref
{sigfix-thmm}, \ref{sigprior-thmm} and Lemma~\ref{rates-lem}.}
\end{supplement}
|
\section*{Acknowledgements}
\noindent We express our gratitude to our colleagues in the CERN
accelerator departments for the excellent performance of the LHC. We
thank the technical and administrative staff at the LHCb
institutes. We acknowledge support from CERN and from the national
agencies: CAPES, CNPq, FAPERJ and FINEP (Brazil); NSFC (China);
CNRS/IN2P3 and Region Auvergne (France); BMBF, DFG, HGF and MPG
(Germany); SFI (Ireland); INFN (Italy); FOM and NWO (The Netherlands);
SCSR (Poland); MEN/IFA (Romania); MinES, Rosatom, RFBR and NRC
``Kurchatov Institute'' (Russia); MinECo, XuntaGal and GENCAT (Spain);
SNSF and SER (Switzerland); NASU (Ukraine); STFC and the Royal Society (United
Kingdom); NSF (USA). We also acknowledge the support received from EPLANET,
Marie Curie Actions and the ERC under FP7.
The Tier1 computing centres are supported by IN2P3 (France), KIT and BMBF (Germany),
INFN (Italy), NWO and SURF (The Netherlands), PIC (Spain), GridPP (United Kingdom).
We are indebted to the communities behind the multiple open source software packages on which we depend.
We are also thankful for the computing resources and the access to software R\&D tools provided by Yandex LLC (Russia).
\section{Conclusions}
\label{sec:Conclusions}
The dependences of the production rate of \ensuremath{\L_\bquark^0}\xspace baryons with respect to \ensuremath{\B^0}\xspace mesons are measured as functions of
the transverse momentum \mbox{$p_{\rm T}$}\xspace and of the pseudorapidity $\eta$ of the $b$~hadron.
The \mbox{$p_{\rm T}$}\xspace dependence is accurately described by an exponential function.
The ratio of fragmentation fractions \ensuremath{f_{\Lb}/f_\dquark}\xspace decreases by a factor of three in the range $1.5 < \mbox{$p_{\rm T}$}\xspace < 40$~\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace.
The ratio of fragmentation fractions \ensuremath{f_{\Lb}/f_\dquark}\xspace versus $\eta$ is described by a linear dependence in the range $2 < \eta < 5$.
The absolute scale of \ensuremath{f_{\Lb}/f_\dquark}\xspace is fixed using the measurement of \ensuremath{f_{\Lb}/f_\dquark}\xspace from semileptonic $b$-hadron
decays~\cite{LHCb-PAPER-2011-018}.
The branching fraction of the decay \ensuremath{\Lb \to \Lc \pim}\xspace is determined with a total precision of 8\%,
\begin{equation*}
\label{eq:BFResult_conc}
{\ensuremath{\cal B}\xspace}(\ensuremath{\Lb \to \Lc \pim}\xspace) = \BRLbResultRound,
\end{equation*}
which is the most precise determination of a branching fraction of a beauty baryon to date.
\section{Detector and simulation}
\label{sec:Detector}
The \mbox{LHCb}\xspace detector~\cite{Alves:2008zz} is a single-arm forward
spectrometer covering the \mbox{pseudorapidity} range $2<\eta <5$,
designed for the study of particles containing \ensuremath{\Pb}\xspace or \ensuremath{\Pc}\xspace
quarks. The detector includes a high-precision tracking system
consisting of a silicon-strip vertex detector surrounding the $pp$
interaction region, a large-area silicon-strip detector located
upstream of a dipole magnet with a bending power of about
$4{\rm\,Tm}$, and three stations of silicon-strip detectors and straw
drift tubes placed downstream.
The combined tracking system provides a momentum measurement with
relative uncertainty that varies from 0.4\% at 5\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace to 0.6\% at 100\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace,
and impact parameter resolution of 20\ensuremath{{\,\upmu\rm m}}\xspace for
tracks with large \mbox{$p_{\rm T}$}\xspace.
Different types of charged hadrons are distinguished by information
from two ring-imaging Cherenkov detectors. Photon, electron and
hadron candidates are identified by a calorimeter system consisting of
scintillating-pad and preshower detectors, an electromagnetic
calorimeter and a hadronic calorimeter. Muons are identified by a system
composed of alternating layers of iron and multiwire proportional chambers.
The trigger consists of a hardware stage, based on information from the calorimeter and muon
systems, followed by a software stage, which applies a full event reconstruction.
The events used in this analysis are selected at the hardware stage by requiring a cluster in
the calorimeters with transverse energy greater than $3.6\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace$.
The software trigger requires a two-, three- or four-track secondary vertex~(SV) with a large sum
of the \mbox{$p_{\rm T}$}\xspace of the particles and a significant displacement from the primary $pp$ interaction
vertices~(PVs). At least one charged particle should have $\mbox{$p_{\rm T}$}\xspace > 1.7\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace$ and \ensuremath{\chi^2_{\rm IP}}\xspace
with respect to any PV greater than 16, where \ensuremath{\chi^2_{\rm IP}}\xspace is defined as the
difference in fit \ensuremath{\chi^2}\xspace of a given PV reconstructed with and without the considered track.
A multivariate algorithm is used for the identification of SVs consistent with
the decay of a \ensuremath{\Pb}\xspace hadron.
Simulated collision events are used to estimate the efficiency of the reconstruction and
selection steps for signal as well as background $b$-hadron decay modes.
In the simulation, $pp$ collisions are generated using \mbox{\textsc{Pythia}}\xspace~\cite{Sjostrand:2006za}
with a specific \mbox{LHCb}\xspace configuration~\cite{LHCb-PROC-2010-056}. Decays of hadronic particles
are described by \mbox{\textsc{EvtGen}}\xspace~\cite{Lange:2001uf}, in which final-state radiation is generated using
\mbox{\textsc{Photos}}\xspace~\cite{Golonka:2005pn}. The interaction of the generated particles with the detector and
its response are implemented using the \mbox{\textsc{Geant4}}\xspace toolkit~\cite{Allison:2006ve, *Agostinelli:2002hh}
as described in Ref.~\cite{LHCb-PROC-2011-006}.
\section{Event selection}
\label{sec:EvtSel}
Since the \ensuremath{\Lb \to \Lc (\to \Pp \Km \pip) \pim}\xspace and \ensuremath{\Bdb \to \Dp (\to \Km \pip \pip) \pim}\xspace decays have the same topology,
the criteria used to select them are chosen to be similar.
This minimises the systematic uncertainty on the ratio of the selection efficiencies.
Following the trigger selection, a preselection is applied using the reconstructed masses, decay times and
vertex qualities of the $b$-hadron and $c$-hadron candidates.
Further separation between signal and background is achieved using a boosted decision tree (BDT)~\cite{Breiman}.
The BDT is trained and tested on a sample of \ensuremath{\Bsb \to \Dsp \pim}\xspace events from the same data set as the signal events.
This sample of events is not used elsewhere in the analysis.
For the signal, a weighted data sample based on the \mbox{\em sPlot}\xspace technique~\cite{Pivk:2004ty} is used.
A training sample representative of combinatorial background is selected
from \ensuremath{\B^0_\squark}\xspace candidates with mass greater than 5445~\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace.
The variables with the most discriminating power are found to be the \ensuremath{\chi^2_{\rm IP}}\xspace of the $b$-hadron candidate
with respect to the PV, the \mbox{$p_{\rm T}$}\xspace of the final-state particles, and the angle between the $b$-hadron momentum vector
and the vector connecting its production and decay vertices.
In events with multiple PVs, the $b$ hadron is associated to the PV giving the smallest \ensuremath{\chi^2_{\rm IP}}\xspace.
The BDT requirement is chosen to maximise the signal yield divided by the square root of the sum of the signal and
background yields. It rejects approximately 84\% of the combinatorial background events while retaining
approximately 84\% of the signal events.
The $\ensuremath{\D^+}\xspace$ ($\ensuremath{\L_\cquark^+}\xspace$) candidates are identified by requiring the invariant mass under the \ensuremath{\Km \pip \pip}\xspace (\ensuremath{\Pp \Km \pip}\xspace) hypothesis
to fall within the range 1844--1890~(2265--2305)~\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace.
The mass resolution of the charm hadrons is approximately 6~\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace.
The ratio of selection efficiencies is evaluated using simulated events.
The \linebreak \ensuremath{\Dp \to \Km \pip \pip}\xspace decay is generated using the known Dalitz structure~\cite{Bonvicini:2008jw},
while the \ensuremath{\Lc \to \Pp \Km \pip}\xspace decay is generated using a combination of non-resonant and resonant decay modes with
proportions according to Ref.~\cite{Aitala:1999uq}.
Interference effects in the \ensuremath{\L_\cquark^+}\xspace decay are not taken into account.
Consistency checks, using a phase-space only model for the \ensuremath{\Lc \to \Pp \Km \pip}\xspace decay, show negligible differences in
the relative efficiencies.
The distributions of the input variables to the BDT are compared in data and simulation.
Good agreement is observed for most variables.
The largest deviation is seen for quantities related to the track quality.
The simulated events are reweighted so that the distributions of these quantities reproduce the distributions in data.
The final stage of the event selection applies particle identification (PID) criteria on all tracks,
based on the differences in the natural logarithm of the likelihood between the
pion, kaon and proton hypotheses~\cite{LHCb-DP-2012-003}.
The PID performance as a function of the \mbox{$p_{\rm T}$}\xspace and $\eta$ of the charged particle is estimated from data.
This is performed using calibration samples, selected using only kinematic criteria, and consisting of
approximately 27 million $\ensuremath{\D^{*-}}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{\Dbar^0}\xspace(\ensuremath{\kaon^+}\xspace\ensuremath{\pion^-}\xspace)\ensuremath{\pion^-}\xspace$ decays for kaons and pions,
and 13 million $\ensuremath{\PLambda}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{\Pp}\xspace \ensuremath{\pion^-}\xspace$ decays for protons.
The size of the proton calibration sample is small at high \mbox{$p_{\rm T}$}\xspace of the proton and does not allow
a reliable estimate of the efficiency of the proton PID requirement in this kinematic region.
Hence, proton PID criteria are only applied to candidates restricted to
a kinematic region in proton momentum and pseudorapidity corresponding to low-\mbox{$p_{\rm T}$}\xspace protons.
Outside of this region, no PID criteria are imposed on the proton.
The ratio of total selection efficiencies, $\varepsilon_{\ensuremath{\Bdb \to \Dp\pim}\xspace}/\varepsilon_{\ensuremath{\Lb \to \Lc \pim}\xspace}$,
is shown in Fig.~\ref{fig:effratio}.
Fluctuations are included in the calculation of the efficiency-corrected
yield ratio.
\begin{figure}[!b]
\begin{picture}(500,180)(0,0)
\put(-5,1){\includegraphics[scale=0.4]{Fig1a.pdf}}
\put(230,1){\includegraphics[scale=0.4]{Fig1b.pdf}}
\put(50,130){(a) LHCb}
\put(280,130){(b) LHCb}
\end{picture}
\caption{\label{fig:effratio} Ratio of total selection efficiencies in bins of the (a) \mbox{$p_{\rm T}$}\xspace and
(b) $\eta$ of the $b$ hadron. The horizontal error bars indicate the range of each bin in \mbox{$p_{\rm T}$}\xspace or $\eta$ respectively.}
\end{figure}
\section{Introduction}
\label{sec:Introduction}
Measurements of beauty hadron production in high-energy proton-proton ($pp$) collisions
provide valuable information on fragmentation and hadronisation
within the framework of quantum chromodynamics~\cite{Berezhnoy:2013cda}.
The study of beauty baryon decays also provides an additional channel for investigating \ensuremath{C\!P}\xspace violation~\cite{Dunietz:1992ti}.
While significant progress has been made in the understanding of the production and decay properties of beauty mesons,
knowledge of beauty baryons is limited.
The relative production rates of beauty hadrons are described by the fragmentation fractions $f_u$, $f_d$, $f_s$, $f_c$ and $f_{\ensuremath{\L_\bquark^0}\xspace}$,
which describe the probability that a $b$ quark fragments into a $B_q$~meson (where $q=u,d,s,c$) or a \ensuremath{\L_\bquark^0}\xspace~baryon, respectively,
and depend on the kinematic properties of the $b$ quark.
Strange $b$ baryons are less abundantly produced~\cite{LHCb-PAPER-2014-010} and are neglected here.
Measurements of ground state $b$~hadrons produced at the $pp$ interaction point also include decay products of excited $b$~hadrons.
In the case of $B$ mesons, such excited states include $B^{*}$ and $B^{**}$ mesons, while \ensuremath{\L_\bquark^0}\xspace baryons can be
produced via decays of \ensuremath{\Lz^{*0}_\bquark}\xspace or \ensuremath{\PSigma_\bquark^{(*)}}\xspace baryons.
Knowledge of the relative production rate of \ensuremath{\L_\bquark^0}\xspace~baryons is necessary to measure absolute \ensuremath{\L_\bquark^0}\xspace branching fractions.
The measurement of the branching fraction of the \ensuremath{\Lb \to \Lc \pim}\xspace decay reported in this paper improves the determination of any \ensuremath{\L_\bquark^0}\xspace branching
fraction measured relative to the \ensuremath{\Lb \to \Lc \pim}\xspace decay.
The inclusion of charge conjugate processes is implied throughout this paper.
The average branching fraction and production ratios are measured.
Previous measurements of \ensuremath{f_{\Lb}/f_\dquark}\xspace have been made in $e^+ e^-$ collisions at LEP~\cite{HFAG}, $p\overline{p}$
collisions at CDF~\cite{Aaltonen:2008zd,Aaltonen:2008eu} and $pp$ collisions at LHCb~\cite{LHCb-PAPER-2011-018}.
The value of \ensuremath{f_{\Lb}/f_\dquark}\xspace measured at LEP differs significantly from the values measured at the hadron colliders, indicating
a strong dependence of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the kinematic properties of the $b$ quark.
The LHCb analysis~\cite{LHCb-PAPER-2011-018} was based on semileptonic $\ensuremath{\L_\bquark^0}\xspace \rightarrow \ensuremath{\L_\cquark^+}\xspace\mu^{-}\bar{\nu} X$ and
$\ensuremath{\Bbar}\xspace \rightarrow \ensuremath{\PD}\xspace \mu^{-} \bar{\nu} X$ decays, where the $B$ meson is charged or neutral, and
$X$ represents possible additional decay products of the $b$ hadron that are not included in the candidate reconstruction.
Near equality of the inclusive semileptonic decay width of all $b$~hadrons was assumed.
The analysis measured \ensuremath{f_{\Lb}/(f_\uquark + f_\dquark)}\xspace, which can be converted into \ensuremath{f_{\Lb}/f_\dquark}\xspace under the assumption of isospin symmetry, \mbox{\itshape i.e.}\xspace $\ensuremath{f_{\uquark}}\xspace = \ensuremath{f_{\dquark}}\xspace$.
A clear dependence of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the transverse momentum \mbox{$p_{\rm T}$}\xspace of the $\ensuremath{\L_\cquark^+}\xspace\mu^-$ and $\ensuremath{\PD}\xspace\mu^-$ pairs was observed.
A CMS analysis~\cite{CMS_Lb} using $\ensuremath{\L_\bquark^0}\xspace \rightarrow \ensuremath{{\PJ\mskip -3mu/\mskip -2mu\Ppsi\mskip 2mu}}\xspace \ensuremath{\PLambda}\xspace$ decays also found that the cross-section for \ensuremath{\L_\bquark^0}\xspace~baryons fell faster
with \mbox{$p_{\rm T}$}\xspace than the $b$-meson cross-sections.
The present paper uses a data sample, corresponding to an integrated luminosity of 1\ensuremath{\mbox{\,fb}^{-1}}\xspace of $pp$ collision data at a
centre-of-mass energy of 7 TeV, collected with the LHCb detector.
This is a substantially increased data sample compared to that in Ref.~\cite{LHCb-PAPER-2011-018}.
The analysis aims to clarify the extent and characteristics of the \mbox{$p_{\rm T}$}\xspace dependence of \ensuremath{f_{\Lb}/f_\dquark}\xspace.
Moreover, the dependence of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the pseudorapidity $\eta$, defined in terms of the polar angle $\theta$ with respect to the beam
direction as $-\ln(\tan{\theta/2})$, is studied for the first time.
The analysis covers the fiducial region $1.5 < \mbox{$p_{\rm T}$}\xspace < 40$~\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace and $2 < \eta < 5$.
The hadronic decays \ensuremath{\Lb \to \Lc \pim}\xspace and \ensuremath{\Bdb \to \Dp\pim}\xspace are used, with the charm hadrons reconstructed using the decay modes \ensuremath{\Lc \to \Pp \Km \pip}\xspace and \ensuremath{\Dp \to \Km \pip \pip}\xspace, respectively.
The data sample and the selection of \ensuremath{\Bdb \to \Dp\pim}\xspace decays are identical to those used in Ref.~\cite{LHCb-PAPER-2012-037}.
Although a precise measurement of the absolute value of \ensuremath{f_{\Lb}/f_\dquark}\xspace is not possible with these decays, since the \ensuremath{\Lb \to \Lc \pim}\xspace branching fraction is
poorly known~\cite{PDG2012}, they can be used to measure the dependence of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the $b$-hadron kinematic properties to high precision.
This is achieved by measuring the efficiency-corrected yield ratio $\mathcal{R}$ in bins of \mbox{$p_{\rm T}$}\xspace or $\eta$ of the beauty hadron
\begin{equation}
\label{eq:DefnOfR}
\mathcal{R} (x) \equiv \frac{N_{\ensuremath{\Lb \to \Lc \pim}\xspace} (x) }{N_{\ensuremath{\Bdb \to \Dp\pim}\xspace} (x) } \times \frac{\varepsilon_{ \ensuremath{\Bdb \to \Dp\pim}\xspace} (x)}{\varepsilon_{\ensuremath{\Lb \to \Lc \pim}\xspace} (x)},
\end{equation}
where $N$ is the event yield, $\epsilon$ is the total reconstruction and selection efficiency, and $x$ denotes \mbox{$p_{\rm T}$}\xspace or $\eta$.
The quantity $\mathcal{R}$ is related to \ensuremath{f_{\Lb}/f_\dquark}\xspace through
\begin{eqnarray}
\label{eq:DefnOfS}
\frac{\ensuremath{f_{\Lb}}\xspace}{\ensuremath{f_{\dquark}}\xspace} (x) &=& \frac{\BF(\ensuremath{\Bdb \to \Dp\pim}\xspace)}{\BF(\ensuremath{\Lb \to \Lc \pim}\xspace)} \times \frac{\BF(\ensuremath{\Dp \to \Km \pip \pip}\xspace)}{\BF(\ensuremath{\Lc \to \Pp \Km \pip}\xspace)} \times
\mathcal{R} (x) \nonumber \\
& & \nonumber \\
&\equiv& \mathcal{S} \times \mathcal{R} (x),
\end{eqnarray}
where $\mathcal{S}$ is a constant scale factor.
Since the value of \ensuremath{f_{\Lb}/f_\dquark}\xspace in a given bin of \mbox{$p_{\rm T}$}\xspace or $\eta$ is independent of the decay mode of the $b$~hadron, the values of
$\ensuremath{f_{\Lb}/f_\dquark}\xspace(\mbox{$p_{\rm T}$}\xspace)$ from the semileptonic analysis~\cite{LHCb-PAPER-2011-018} can be compared to the measurement of $\mathcal{R}(\mbox{$p_{\rm T}$}\xspace)$,
which allows for the extraction of the value of $\mathcal{S}$.
The branching fraction $\BF(\ensuremath{\Lb \to \Lc \pim}\xspace)$ can then be readily obtained using Eq.~\eqref{eq:DefnOfS}.
Notably, the dependence on $\BF(\ensuremath{\Lc \to \Pp \Km \pip}\xspace)$ cancels when extracting $\BF(\ensuremath{\Lb \to \Lc \pim}\xspace)$ in this way, because this branching fraction
also enters in the semileptonic measurement of \ensuremath{f_{\Lb}/f_\dquark}\xspace. Furthermore, the branching fractions $\BF(\ensuremath{\Bdb \to \Dp\pim}\xspace)$~\cite{PDG2012} and
$\BF(\ensuremath{\Dp \to \Km \pip \pip}\xspace)$~\cite{Dobbs:2007ab} are well known, leading to a precise determination of $\BF(\ensuremath{\Lb \to \Lc \pim}\xspace)$.
The dependence of the semileptonic \ensuremath{f_{\Lb}/f_\dquark}\xspace measurement on $\BF(\ensuremath{\L_\bquark^0}\xspace \rightarrow \ensuremath{\L_\cquark^+}\xspace\mu^{-}\bar{\nu} X)$, and the assumption of near
equality of the inclusive semileptonic decay width of all $b$~hadrons, implies that the measurement of $\BF(\ensuremath{\Lb \to \Lc \pim}\xspace)$ from the current
paper cannot be used to normalise existing measurements of $\BF(\ensuremath{\L_\bquark^0}\xspace \rightarrow \ensuremath{\L_\cquark^+}\xspace\mu^{-}\bar{\nu} X)$~\cite{PDG2012}.
\section{Results}
\label{sec:Results}
The study of the dependences of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the \mbox{$p_{\rm T}$}\xspace and $\eta$ of the $b$ hadron and the
measurement of the branching fraction of \ensuremath{\Lb \to \Lc \pim}\xspace decays are performed using candidates restricted to
the fiducial region $1.5 < \mbox{$p_{\rm T}$}\xspace < 40$ \ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace and $ 2 < \eta < 5$. A discussion on the
systematic uncertainties related to these measurements can be found in the next section.
The ratio of efficiency-corrected event yields as a function of \mbox{$p_{\rm T}$}\xspace is shown in
Fig.~\ref{fig:flbfd-semi}(a), and is fitted with an exponential function,
\begin{equation}
\label{eq:ptRatio_syst}
\mathcal{R}(\mbox{$p_{\rm T}$}\xspace) = a + \exp\left(b + c \times \mbox{$p_{\rm T}$}\xspace [\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace] \right),
\end{equation}
with
\begin{eqnarray}
\label{eq:ptRatioNumbers_syst}
a & = & +0.181 \pm 0.018 \pm 0.026, \nonumber \\
b & = & -0.391 \pm 0.023 ^{\,\, + 0.069} _{\,\, - 0.067}, \nonumber \\
c & = & -0.095 \pm 0.007 \pm 0.014 \,\,\, [\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace]^{-1}, \nonumber
\end{eqnarray}
where the first uncertainties are statistical and the second systematic.
The correlation matrix of the parameters is
\begin{displaymath}
\rho(a,b,c) =
\left(
\begin{array}{ccc}
1 & -0.22 & -0.94 \\
-0.22 & 1 & -0.10 \\
-0.94 & -0.10 & 1
\end{array}
\right).
\end{displaymath}
The correlation between the parameters leads to a relatively large apparent uncertainty on the individual parameters.
Systematic uncertainties are not included in this matrix.
The \ensuremath{\chi^2/\mathrm{ndf}}\xspace value of the fit is 23.3/17, which corresponds to a p-value of
0.14.
The $\eta$ dependence is described by a linear function,
\begin{equation}
\label{eq:etaRatio_syst}
\mathcal{R}(\eta) = a + b \times \left( \eta-\overline{\eta} \right),
\end{equation}
with
\begin{eqnarray}
\label{eq:etaRatioNumbers_syst}
a & = & 0.464 \pm 0.003 ^{\,\, + 0.008} _{\,\, - 0.010}, \nonumber \\
b & = & 0.081 \pm 0.005 ^{\,\, + 0.013} _{\,\, - 0.009}, \nonumber \\
\overline{\eta} & = & 3.198, \nonumber
\end{eqnarray}
where the first uncertainties are statistical and the second systematic.
The offset $\overline{\eta}$ is fixed to the average value of the measured $\eta$ distribution.
The correlation between the two fit parameters is negligible for this choice of $\overline{\eta}$.
The \ensuremath{\chi^2/\mathrm{ndf}}\xspace value of the fit is 13.1/8, corresponding to a p-value of 0.11.
\begin{figure}[!h]
\begin{picture}(500,180)(0,0)
\put(-5,1){\includegraphics[scale=0.4]{Fig3a.pdf}}
\put(230,1){\includegraphics[scale=0.4]{Fig3b.pdf}}
\put(50,130){(a) LHCb}
\put(280,130){(b) LHCb}
\end{picture}
\caption{\label{fig:flbfd-semi} (a)
Dependence of the efficiency-corrected ratio of yields, $\mathcal{R}$,
between \ensuremath{\Lb \to \Lc \pim}\xspace and \ensuremath{\Bdb \to \Dp\pim}\xspace decays on the \mbox{$p_{\rm T}$}\xspace of the beauty hadron, fitted with an exponential function.
The error bars on the data show the statistical and systematic uncertainties added in quadrature.
(b) The resulting parametrisation is then fitted to the rescaled \ensuremath{f_{\Lb}/f_\dquark}\xspace measurements from the semileptonic
analysis~\cite{LHCb-PAPER-2011-018}, to obtain the scale factor $\mathcal{S}$. The error bars include
only the statistical uncertainty.}
\end{figure}
To extract the scale factor $\mathcal{S}$ given in Eq.~\eqref{eq:DefnOfS}, the normalisation of $\mathcal{R}(x)$,
with fixed parameters $a$, $b$ and $c$, is allowed to vary in a fit to the published \ensuremath{f_{\Lb}/f_\dquark}\xspace data~\cite{LHCb-PAPER-2011-018},
as shown in Fig.~\ref{fig:flbfd-semi}(b).
The result quoted in Ref.~\cite{LHCb-PAPER-2011-018} was measured as
a function of the \mbox{$p_{\rm T}$}\xspace of the $\ensuremath{\L_\cquark^+}\xspace\mu^{-}$ system.
A shift, estimated from simulation, is applied to the \mbox{$p_{\rm T}$}\xspace values
to obtain the corresponding average \mbox{$p_{\rm T}$}\xspace of the $b$~hadron for each bin.
Furthermore, the semileptonic results are updated using recent determinations of
$\BF(\ensuremath{\Lc \to \Pp \Km \pip}\xspace) = (6.84 \pm 0.24 ^{\,\,+0.21}_{\,\,-0.27})$\%~\cite{Zupanc:2013iki} and the ratio of lifetimes
$(\tau_{B^+}+\tau_{B^0})/2\tau_{\ensuremath{\L_\bquark^0}\xspace}=1.071 \pm 0.008$~\cite{LHCb-PAPER-2013-065,LHCb-PAPER-2014-003}.
The following value of the scale factor $\mathcal{S}$ is determined,
\begin{equation} \nonumber
\label{eq:ScaleFactorValue}
\mathcal{S} = 0.834
\overbrace{\pm 0.006 \, (\mathrm{stat}) ^{\,\,+0.023}_{\,\,-0.021} \, (\mathrm{syst})}^{\mathrm{hadronic}}
\overbrace{\pm 0.027 \, (\mathrm{stat}) ^{\,\,+0.058}_{\,\,-0.062} \, (\mathrm{syst})}^{\mathrm{semileptonic}},
\end{equation}
where the statistical and systematic uncertainties associated with the hadronic
and semileptonic measurement are shown separately.
The \ensuremath{\chi^2/\mathrm{ndf}}\xspace value of the fit is 8.68/3, which corresponds to a p-value of 0.03.
By multiplying the ratio of the efficiency-corrected yields $\mathcal{R}$ with the scale factor
$\mathcal{S}$, the dependences of \ensuremath{f_{\Lb}/f_\dquark}\xspace on \mbox{$p_{\rm T}$}\xspace and $\eta$ are
obtained.
The \mbox{$p_{\rm T}$}\xspace dependence is described with the exponential function
\begin{equation}
\label{eq:flfd_model_pt}
\ensuremath{f_{\Lb}/f_\dquark}\xspace(\mbox{$p_{\rm T}$}\xspace) = a' + \exp(b' + c' \times \mbox{$p_{\rm T}$}\xspace [\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace]),
\end{equation}
with
\begin{eqnarray}
\label{eq:flfd_model_pt_numbers}
a' & = & +0.151 \pm 0.016 ^{\,\,+ 0.024} _{\,\,- 0.025} , \nonumber \\
b' & = & -0.573 \pm 0.040 ^{\,\,+ 0.101} _{\,\,- 0.097}, \nonumber \\
c' & = & -0.095 \pm 0.007 \pm 0.014 \,\,\, [\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace]^{-1}, \nonumber
\end{eqnarray}
where the first uncertainty is the combined statistical and the second is the combined systematic from the hadronic and semileptonic measurements.
The correlations between the three fit parameters change due to the uncertainty on the scale factor $\mathcal{S}$. The correlation matrix of the parameters is
\begin{displaymath}
\rho(a',b',c') =
\left(
\begin{array}{ccc}
1 & 0.55 & -0.73 \\
0.55 & 1 & -0.03 \\
-0.73 & -0.03 & 1
\end{array}
\right).
\end{displaymath}
The $\eta$ dependence is described by the linear function
\begin{equation}
\label{eq:flfd_model_eta}
\ensuremath{f_{\Lb}/f_\dquark}\xspace(\eta) = a' + b' \times \left( \eta-\overline{\eta} \right),
\end{equation}
with
\begin{eqnarray}
\label{eq:flfd_model_eta_numbers}
a' & = & 0.387 \pm 0.013 ^{\,\,+ 0.028} _{\,\,- 0.030}, \nonumber \\
b' & = & 0.067 \pm 0.005 ^{\,\,+ 0.012} _{\,\,- 0.009}, \nonumber
\end{eqnarray}
where the first uncertainty is the combined statistical and the second is the combined systematic from the hadronic and semileptonic measurements.
The dependences of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the \mbox{$p_{\rm T}$}\xspace and $\eta$ of the
$b$ hadron are shown in Fig.~\ref{fig:flbfd}.
\begin{figure}[!t]
\begin{picture}(500,180)(0,0)
\put(-5,1){\includegraphics[scale=0.4]{Fig4a.pdf}}
\put(230,1){\includegraphics[scale=0.4]{Fig4b.pdf}}
\put(50,130){(a) LHCb}
\put(280,130){(b) LHCb}
\end{picture}
\caption{\label{fig:flbfd} Dependence of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the (a) \mbox{$p_{\rm T}$}\xspace and (b) $\eta$ of the beauty hadron.
To obtain this figure, the ratio of efficiency-corrected event yields is scaled to the absolute value
of \ensuremath{f_{\Lb}/f_\dquark}\xspace from the semileptonic analysis~\cite{LHCb-PAPER-2011-018}. The error bars include the
statistical and systematic uncertainties associated with the hadronic measurement. The dashed red lines indicate the uncertainty
on the scale of \ensuremath{f_{\Lb}/f_\dquark}\xspace from the semileptonic analysis.}
\end{figure}
The absolute value for \BF(\ensuremath{\Lb \to \Lc \pim}\xspace) is obtained by substituting the results for $\mathcal{S}$
and $\BF(\ensuremath{\Bdb \to \Dp\pim}\xspace)=(2.68 \pm 0.13) \times 10^{-3}$\cite{PDG2012} into Eq.~\eqref{eq:DefnOfS}.
The value for \BF(\ensuremath{\Lc \to \Pp \Km \pip}\xspace) is also used in the determination of \ensuremath{f_{\Lb}/f_\dquark}\xspace using
semileptonic decays and therefore cancels in the final result.
The branching fraction for \ensuremath{\Lb \to \Lc \pim}\xspace is measured to be
\begin{equation*}
\label{eq:BFResult}
{\ensuremath{\cal B}\xspace}(\ensuremath{\Lb \to \Lc \pim}\xspace) = \BRLbResultRound,
\end{equation*}
where the first uncertainty is statistical, the second is systematic,
the third is from the previous \mbox{LHCb}\xspace measurement
of \ensuremath{f_{\Lb}/f_\dquark}\xspace, and the fourth is due to the knowledge of
$\BF(\ensuremath{\Bbar^0}\xspace~\rightarrow~\ensuremath{\D^+}\xspace\pi^-)$.
This value is
in agreement with the current world average~\cite{PDG2012}. It also agrees within
2.4 standard deviations with the recent \mbox{LHCb}\xspace measurement using \ensuremath{\Lb \to \Lc (\to \Pp \KS) \pim}\xspace decays~\cite{LHCb-PAPER-2013-061},
taking into account the correlated uncertainty
from the semileptonic value for \ensuremath{f_{\Lb}/f_\dquark}\xspace (6.1\%). Combining the two \mbox{LHCb}\xspace
measurements, and using a consistent value for the lifetime ratio of
$(\tau_{B^+}+\tau_{B^0})/2\tau_{\ensuremath{\L_\bquark^0}\xspace}=1.071 \pm 0.008$, we obtain
${\ensuremath{\cal B}\xspace}(\ensuremath{\Lb \to \Lc \pim}\xspace) = (4.46 \pm 0.36)\times 10^{-3}$, where the uncertainty is the combined
statistical and systematic uncertainty of both measurements.
\section{Systematic uncertainties}
\label{sec:Systematics}
Systematic uncertainties on the measurement of the relative
efficiency-corrected event yields of the \ensuremath{\Lb \to \Lc \pim}\xspace and \ensuremath{\Bdb \to \Dp\pim}\xspace decay modes relate to the
fit models and to the efficiencies of the PID, BDT and trigger selections.
The effect of each systematic uncertainty on the
efficiency-corrected yield ratio is calculated separately for each bin of
$\mbox{$p_{\rm T}$}\xspace$ or $\eta$.
The systematic uncertainties are considered to be correlated across the bins,
unless mentioned otherwise.
The effect of the systematic uncertainties on
the model of the $\mathcal{R}(x)$ dependence and
the measurement of \BF(\ensuremath{\Lb \to \Lc \pim}\xspace)
are determined by refitting the data points when
the $\mathcal{R}$ value in each bin is
varied by its associated uncertainty.
The various sources of systematic uncertainty are discussed below and summarised in Table~\ref{t:syst}.
\begin{table}[b]
\begin{center}
\caption{Relative systematic uncertainties for the measurements of $\mathcal{R}(x)$ (first five columns) and \BF(\ensuremath{\Lb \to \Lc \pim}\xspace) (last column). The uncertainties from the various sources are uncorrelated and added in quadrature to obtain the total uncertainty. Sample size
refers to the size of the simulated events sample.}
\label{t:syst}
{\small
\begin{tabular}{l|rrr|rr|rrr}
& \multicolumn{3}{c|}{$\mbox{$p_{\rm T}$}\xspace$ bins} & \multicolumn{2}{c|}{$\eta$ bins} & \\
& \multicolumn{3}{c|}{$\mathcal{R} = a + \exp(b+c \times \mbox{$p_{\rm T}$}\xspace)$} & \multicolumn{2}{c|}{$\mathcal{R} = a + b \times (\eta - \overline{\eta})$} & \multicolumn{3}{c}{\BF(\ensuremath{\Lb \to \Lc \pim}\xspace)} \\
& \multicolumn{1}{c}{$a$} & \multicolumn{1}{c}{$b$} & \multicolumn{1}{c|}{$c$} & \multicolumn{1}{c}{$a$} & \multicolumn{1}{c|}{$b$} & & & \\
\hline
\multicolumn{1}{l}{Fit model } & \multicolumn{8}{c}{ } \\
\hline
Signal & $^{+0.7}_{-0.4}$\% & $^{+0.5}_{-0.2}$\% & $^{+0.2}_{-0.3}$\% & $^{+0.3}_{-0.1}$\% & $^{+1.1}_{-1.8}$\% & & $^{+0.2}_{-0.1}$\% & \\[0.3em]
Background & $^{+5.5}_{-1.7}$\% & $^{+2.8}_{-2.1}$\% & $^{+2.6}_{-1.1}$\% & $^{+0.6}_{-0.1}$\% & $^{+2.4}_{-4.7}$\% & & $^{+0.6}_{-0.0}$\% & \\
\hline
\multicolumn{1}{l}{Efficiencies} & \multicolumn{8}{c}{ } \\
\hline
PID & 0.0\% & 0.5\% & 2.5\% & $-1.3$\% & 12.7\% & & $-1.1$\% & \\
BDT & $^{+5.8}_{-7.6}$\% & $^{-15.1}_{+14.2}$\% & $^{+~9.6}_{-10.2}$\% & $^{+1.3}_{-1.3}$\% & $^{+4.7}_{-4.8}$\% & & $^{+2.3}_{-2.2}$\% & \\
Sample size & $\pm$12.1\% & $\pm$9.0\% & $\pm$10.8\% & $\pm$0.9\% & $\pm$9.3\% & & $\pm$1.2\% & \\
Trigger & 0.9\% & 1.0\% & 1.0\% & $-0.3$\% & $-0.1$\% & & $-0.3$\% & \\
\hline
\multicolumn{1}{l}{Other} & \multicolumn{8}{c}{ } \\
\hline
Bin centre & $\pm$0.3\% & $\pm$0.3\% & $\pm$0.1\% & $\pm$0.1\% & $\pm$1.3\% & & 0.0\% & \\
\specialrule{.1em}{.05em}{.05em}
\specialrule{.1em}{.05em}{.05em}
\bf{Total} \normalfont & $^{+14.6}_{-14.5}$\% & $^{+17.1}_{-17.7}$\% & $^{+14.9}_{-14.9}$\%& $^{+1.8}_{-2.1}$\% & $^{+16.6}_{-11.6}$\% & & $^{+2.6}_{-2.8}$\% & \\
\end{tabular}
}
\end{center}
\end{table}
The uncertainty due to the modelling of the signal shape is estimated by
replacing the modified Gaussian with two modified Gaussians, which share
the same mean but are allowed to have different widths.
In addition, the
parameters that describe the tails are varied by $\pm 10\%$ relative to their nominal
values, which is the maximum variation found for these parameters when leaving them
free in the fit. This affects the ratio of yields by a maximum of 0.3\%.
A possible variation of the slope of the combinatorial background shape
across the bins is observed in a data sample of \ensuremath{\Lc \pip}\xspace candidates.
To account for this, the slope is varied from $\pm50\%$ in
the lowest \mbox{$p_{\rm T}$}\xspace or $\eta$ bin to $\mp50\%$ in the highest \mbox{$p_{\rm T}$}\xspace or $\eta$ bin.
The signal yield ratio varies by less than 1\%, with
the exception of one \mbox{$p_{\rm T}$}\xspace bin which shows a variation of approximately 2\%.
The uncertainty on the shapes of partially reconstructed backgrounds is estimated by modelling
them with a non-parametric distribution~\cite{Cranmer:2000du} for \ensuremath{\Lb \to \PSigma_c^+ \pim}\xspace and \ensuremath{\Lb \to \Lc \rho^-}\xspace decays and with two
modified Gaussian distributions with tails on either side
for the \ensuremath{\Bdb \to \Dstarp \pim}\xspace shape. The effect on the signal yield ratio is below 0.5\% in most bins,
increasing to about 2\% for the highest \mbox{$p_{\rm T}$}\xspace bin.
The contribution of $b$-hadron decays without an intermediate $c$ hadron
is ignored in the fit. To evaluate the systematic uncertainty due to these decays, the $b$-hadron
mass spectra for candidates in the sidebands of the $c$-hadron mass distribution are examined.
A contribution of $0.4\%$ relative to the signal yield is found in the \ensuremath{\Bdb \to \Dp\pim}\xspace decay mode,
and its full size is taken as systematic uncertainty.
No contribution is seen in the \ensuremath{\Lb \to \Lc \pim}\xspace decay mode and no systematic uncertainty is assigned.
The uncertainty on the PID efficiency and misidentification rate is estimated
by comparing the PID performance measured using simulated $\ensuremath{\D^*}\xspace$ and $\PLambda$
calibration samples with that observed in simulated signal events.
The efficiency ratio varies by between 1\% and 4\% across the bins.
As discussed in Sec.~\ref{sec:EvtSel}, the simulated events are reweighted so that the distributions of
quantities related to the track quality match the distributions observed in data.
The systematic uncertainty on the selection efficiency is obtained by recalculating the efficiency without this reweighting.
The yield ratio varies by between 0.2\% and 6\%.
In addition, there is a 5\% statistical uncertainty per bin due to the simulated sample size,
which is uncorrelated across bins.
The uncertainty due to the trigger efficiency, caused by possible differences in
the response to a proton compared to a charged pion in the calorimeter, is
estimated to be about 0.4\%, taking into account that at most 10\% of the
events containing \ensuremath{\Lb \to \Lc \pim}\xspace candidates are triggered by the proton.
The systematic uncertainty due to the choice of bin centre is evaluated by redefining the bin centres
using the average \mbox{$p_{\rm T}$}\xspace or $\eta$ of the \ensuremath{\L_\bquark^0}\xspace or \ensuremath{\B^0}\xspace sample only, instead of the mean of the \ensuremath{\L_\bquark^0}\xspace and \ensuremath{\B^0}\xspace samples.
\section{Event yields}
\label{sec:Yields}
The dependences of \ensuremath{f_{\Lb}/f_\dquark}\xspace on the \mbox{$p_{\rm T}$}\xspace and
$\eta$ of the $b$~hadron are studied in the ranges $1.5 < \mbox{$p_{\rm T}$}\xspace < 40$ \ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace and
$2 < \eta < 5$. The event sample is sub-divided in 20 bins in \mbox{$p_{\rm T}$}\xspace and 10
bins in $\eta$, with bin boundaries chosen to obtain approximately equal numbers
of \ensuremath{\Bdb \to \Dp\pim}\xspace candidates per bin.
The bin centres are obtained from simulated events without any selection applied,
and are defined as the mean of the average \mbox{$p_{\rm T}$}\xspace or $\eta$ of the \ensuremath{\Lb \to \Lc \pim}\xspace and
\ensuremath{\Bdb \to \Dp\pim}\xspace samples in each bin.
The yields of the two decay modes are determined from extended maximum
likelihood fits to the unbinned mass distributions of the reconstructed $b$-hadron
candidates, in each bin of \mbox{$p_{\rm T}$}\xspace or $\eta$.
To improve the mass resolution, the value of the beauty hadron mass is refit
with the invariant mass of the charm hadron constrained to its known value~\cite{PDG2012}.
Example fits in the \mbox{$p_{\rm T}$}\xspace bin with the lowest fitted signal yield and in an arbitrarily chosen $\eta$ bin are shown in
Fig.~\ref{fig:masspeak} for $\ensuremath{\L_\cquark^+}\xspace\ensuremath{\pion^-}\xspace$ and $D^+\ensuremath{\pion^-}\xspace$ candidates. The total signal yields,
obtained from fits to the total event samples, are $44\,859 \pm 229$ for the \ensuremath{\Lb \to \Lc \pim}\xspace sample
and $106\,197 \pm 344$ for the \ensuremath{\Bdb \to \Dp\pim}\xspace sample.
\begin{figure}[!b]
\begin{picture}(500,300)(0,0)
\put(-5,151){\includegraphics[bb=0 0 500 310, scale=0.48, clip=]{Fig2a.pdf}}
\put(230,151){\includegraphics[bb=0 0 500 310, scale=0.48, clip=]{Fig2b.pdf}}
\put(50,250){(a)}
\put(60,280){$20.2<\mbox{$p_{\rm T}$}\xspace(\ensuremath{\L_\bquark^0}\xspace)<40\,\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace$}
\put(280,250){(b)}
\put(300,280){$20.2<\mbox{$p_{\rm T}$}\xspace(\ensuremath{\Bbar^0}\xspace)<40\,\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c}}\xspace$}
\put(-5,1){\includegraphics[bb=0 0 500 310, scale=0.48, clip=]{Fig2c.pdf}}
\put(230,1){\includegraphics[bb=0 0 500 310, scale=0.48, clip=]{Fig2d.pdf}}
\put(50,100){(c)}
\put(60,130){$3.05<\eta(\ensuremath{\L_\bquark^0}\xspace)<3.2$}
\put(280,100){(d)}
\put(300,130){$3.05<\eta(\ensuremath{\Bbar^0}\xspace)<3.2$}
\end{picture}
\caption{\label{fig:masspeak} Invariant mass distributions of (a,c) $\ensuremath{\L_\cquark^+}\xspace\ensuremath{\pion^-}\xspace$ candidates
and (b,d) $\ensuremath{\D^+}\xspace\ensuremath{\pion^-}\xspace$ candidates for specific ranges in \mbox{$p_{\rm T}$}\xspace and $\eta$ of the $b$~hadron,
with fit projections overlaid.
The different components are defined in the legend, where ``part reco" refers to the sum of partially reconstructed
decays.}
\end{figure}
The signal mass shape is described by a modified Gaussian distribution
with power-law tails on either side to model the radiative tail and non-Gaussian detector
effects.
The parameters of
the tails are obtained from simulated events and fixed in the fit. The mean and the width of the Gaussian distribution
are allowed to vary.
Three classes of background are considered: partially reconstructed decays
with or without misidentified tracks, fully reconstructed
decays where at least one track is misidentified, and combinatorial background. The shapes
of the invariant mass distributions for the partially reconstructed decays are
obtained using large samples of simulated events.
For the \ensuremath{\Bdb \to \Dp\pim}\xspace sample, the decays \ensuremath{\Bdb \to \Dp \rho^-}\xspace and
\ensuremath{\Bdb \to \Dstarp \pim}\xspace are modelled with non-parametric distributions~\cite{Cranmer:2000du}.
The main sources for the \ensuremath{\Lb \to \Lc \pim}\xspace sample are \ensuremath{\Lb \to \Lc \rho^-}\xspace and \ensuremath{\Lb \to \PSigma_c^+ \pim}\xspace decays, which are
modelled with a bifurcated Gaussian function.
All these processes involve a neutral pion that is not included
in the candidate's reconstruction.
The invariant mass distributions of the misidentified decays are affected by the
PID criteria. The shapes are obtained from simulated events,
reweighted
according to the momentum-dependent particle identification
efficiency, with the mass hypothesis of the signal applied.
The \ensuremath{\Bdb \to \Dp\pim}\xspace background in the \ensuremath{\Lb \to \Lc \pim}\xspace sample is most abundant in the
highest \mbox{$p_{\rm T}$}\xspace bins, since the proton PID criteria are least effective in this kinematic region.
The Cabibbo-suppressed decays \ensuremath{\Lb \to \Lc \Km}\xspace and \ensuremath{\Bdb \to \Dp \Km}\xspace contribute to the background in the
\ensuremath{\Lb \to \Lc \pim}\xspace and \ensuremath{\Bdb \to \Dp\pim}\xspace fits, respectively, when the kaon of the $b$-hadron decay is misidentified as
a pion. The yields of these backgrounds relative to the signal yield are
constrained in the fits, using \mbox{LHCb}\xspace measurements of the relevant ratios of
branching fractions~\cite{LHCb-PAPER-2012-037,LHCb-PAPER-2013-056} and the
misidentification probabilities with their associated uncertainties.
The combinatorial background consists of events with random pions, kaons and
protons forming a mis-reconstructed \ensuremath{\D^+}\xspace or \ensuremath{\L_\cquark^+}\xspace candidate, as well as genuine \ensuremath{\D^+}\xspace or \ensuremath{\L_\cquark^+}\xspace
hadrons, that combine with a random pion. The combinatorial background is
modelled with an exponential shape. The slope is fixed in
the fit in each kinematic bin to the value found from a fit to the total sample.
|
\section{Introduction}
\subsection{Background of open quantum systems}
Coupling between microscopic quantum mechanical systems and their environment is important in essentially every system where quantum mechanical behaviour is observed. No quantum system that has been measured in the laboratory is an ideal, perfectly isolated (or \emph{closed}) system. Rather, coherent dynamics (as described by a Schr\"odinger equation) typically last only over short timescales, before the dynamics become dominated by coupling of the \emph{open} system \cite{Breuer2007,Gardiner2005,Carmichael1993,Carmichael1999,Weiss2012,Muller2012} to its environment, leading to decoherence, and the onset of more classical behaviour. Over the last few decades, quantum mechanical behaviour has been observed and controlled in a diverse range of systems, to the point where many systems can be controlled on the level of a single atom, ion, molecule, photon or electron. As a result, the need to better understand dynamics in open quantum systems has increased. While philosophically it would be possible to extend the boundary of the system and include the environment in a larger system, this is typically impractical mathematically due to the enormous numbers of degrees of freedom involved in describing the environment. Hence, we usually look to find an effective description for dynamics in the approximately isolated system.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=12cm]{figures/systemenvironment}
\end{center}
\caption{General framework for an open quantum system: A small quantum system interacts with its environment, leading to a combination of coherent and dissipative dynamics for the small system. In quantum optics, and atomic, molecular, and optical (AMO) systems more generally, the resulting dynamics can both be microscopically well understood, as well as controlled to engineer dynamics or specific quantum states. This involves a separation of energy and frequency scales, where the energy and frequency scales of the small system that are directly coupled to the environment ($\omega_{\rm sys}$) and the relaxation rates for relevant correlation functions in the environment ($\omega_{\rm env}$) are much larger than the frequency scales of dynamics induced by the coupling ($\Gamma$).}\label{fig:openquantumsystem}
\end{figure}
With the recent focus on controlling quantum coherence, especially to store and manipulate quantum information \cite{Nielsen2000,Monroe2013,Awschalom2013,Devoret2013}, investigation of open quantum systems has become prominent across various fields of physics. Quantum coherence has been observed, and decoherence studies performed in atomic systems including photon modes in cavities \cite{Kimble2008,Deleglise2008,Gleyzes2007,Guerlin2007,Nogues1999}; single trapped atoms \cite{Bloch2008}, ions \cite{Monroe2013,Haffner2008,Blatt2008,Monz2011,Leibfried2003,Myatt2000}, and molecules \cite{Carr2009,Yan2013}; and in solid state systems \cite{Awschalom2013,Hanson2008} including superconducting junctions \cite{Clarke2008}, and spin systems including quantum dots, colour centres, Cooper pair boxes \cite{Devoret2013,Reed2012}, also in conjunction with microwave stripline cavities. In condensed matter physics, there have been many investigations of dissipative phenomena, a whole class of which began with the Caldeira-Leggett treatment of a spin coupled to a bosonic bath \cite{Caldeira1981,Leggett1987}. The study of open quantum systems also has importance in cosmology \cite{Hu1995}, and quantum-optical approaches have recently been applied to treat pion decay in high-energy physics \cite{Lello2013,Lello2013a}.
While many studies seek to characterise and reduce the destruction of quantum coherence in an open system, a key guiding aim in quantum optics over the last thirty years has been the use of controlled coupling to the environment to control and manipulate quantum coherence with high precision. This includes manipulating the environment in such a way as to drive the system into desired quantum states, increasing quantum coherence in the sample. In the laboratory, this philosophy began with optical pumping in atomic physics \cite{Brossel1952,Hawkins1953}, whereby using laser driving and spontaneous emission processes, atoms can be driven into a single atomic state with extraordinarily high fidelities approaching 100\%. This has lead on to techniques for laser cooling of trapped ions \cite{Leibfried2003,Wineland1978,Neuhauser1978} and of atomic and molecular samples \cite{Metcalf1999}, which has allowed the production of atomic gases with temperatures of the order of $1~\mu$K. This, in turn, set the stage for realising Bose-Einstein condensation \cite{Anderson1995a,Davis1995b,Bradley1995a} and degenerate Fermi gases \cite{DeMarco1999a,Truscott2001a}, as well as providing a level of high-precision control necessary for quantum computing with trapped ions \cite{Haffner2008,Leibfried2003,Cirac1995}. Such driving processes, making use of the coupling of the system to its environment, have been extended to algorithmic cooling in Nuclear Magnetic Resonance (NMR) systems \cite{Boykin2002,Baugh2005}, and are recently being applied to cool the motional modes of macroscopic oscillators to near their quantum ground state \cite{Chan2011}. A detailed understanding of the back-action on the quantum state of the system from coupling to the environment has also been useful in the context of high-fidelity state detection (e.g., in electron shelving \cite{Dehmelt1982,Nagourney1986,Bergquist1986,Sauter1986,Hegerfeldt1992b,Hegerfeldt1995}), or in quantum feedback \cite{Wiseman2010, Handel2005, Wiseman1993, Foster2000, Bushev2006}, and have also been exploited for control over atoms and photons in Cavity QED \cite{Haroche2013,Raimond2001}.
This high level of control is possible because in quantum optics several key approximations can often be made regarding the coupling of a system to its environment, which in combination make the control of this coupling both tractable theoretically and feasible experimentally. This includes the fact that the dynamics are often Markovian (i.e., the environment relaxes rapidly to an equilibrium state and on relevant timescales has no memory of previous interactions with the system), and that the system-environment coupling is typically weak compared with relevant system and environment energy scales. The possibility of making such approximations has resulted in the development of several complementary formalisms, including quantum Langevin equations \cite{Gardiner1985,Yurke1984,Caves1982,Ritsch1988,Parkins1988}, quantum master equations \cite{Gardiner2005,Lax1963,Collett1984,Gardiner1985}, and continuous measurement theory \cite{Wiseman2010,Gardiner2005}. These techniques and related ideas have been applied extensively over the last thirty years to a variety of single-particle and few-particle quantum systems, especially in the context of atomic, molecular, and optical (AMO) experiments. Techniques such as quantum master equations in Lindblad form \cite{Gardiner2005} are sometimes applied in other contexts, especially in a range of solid-state systems. However, many solid-state systems are dominated by non-Markovian aspects of their dissipative dynamics \cite{Breuer2007}, which are consequently not captured by this form. This is in strong contrast to the AMO/quantum optics context, where the approximations involved in deriving these equations are typically good to many orders of magnitude.
Many numerical techniques were developed for solving the equations produced within these formalisms. One simple but highly effective technique in this context is the quantum trajectory technique (also known as the Monte Carlo wavefunction method) \cite{Plenio1998,Carmichael1993,Dalibard1992,Molmer1993,Dum1992}, which allows for the numerical solution of a master equation without propagating a density matrix directly. Instead, pure states are propagated in time, with the dissipative process being described by a modification to the Hamiltonian, combined with quantum jumps - sudden changes in the state that take place at particular times. By taking an appropriate stochastic average over the times and type of quantum jumps, expectation values for the system propagated under the master equation can be faithfully and efficiently reconstructed within well-controlled statistical errors. This technique is particularly appealing, because when combined with continuous measurement theory, it can also help to give a physical intuition into the workings of the dissipative process.
\subsection{Open many-body quantum systems in a AMO context}
Over the last decade and a half, AMO systems (cold atoms and molecules, as well as trapped ions and photons) have been increasingly used to study strongly-interacting many-body systems. This has reached a level where these systems can be used to engineer microscopic Hamiltonians for the purposes of quantum simulation \cite{Cirac2012,Bloch2012,Blatt2012,Aspuru-Guzik2012,Muller2012,Bloch2008b,Lewenstein2007,Lewenstein2012}. As a result, it has become important to explore the application of existing experimental techniques and theoretical formalisms from quantum optics in this new many-body context. Particularly strong motivation in this respect has come from the desire to engineer particularly interesting many-body states, which often require very low temperatures (or entropies) in the system. Understanding dissipative dynamics in the many-body system is then necessary both to control heating processes, and to provide new means to drive the systems to lower temperatures.
Such studies are particularly facilitated by the fact that many of the same approximations that are made in describing open few-particle systems in quantum optics can also be made in these AMO many-body systems. This means both that the physics of open systems as they appear in quantum optics can be investigated in a wholly new context, and that mathematical techniques such as master equations and quantum Langevin equations, as well as numerical techniques such as quantum trajectory methods can be immediately applied to these systems in a way that faithfully represents the microscopic physics.
This has already lead to many important developments. Some of these have direct practical importance in the experiments - e.g., in order to cool many-body systems to the temperatures required to realise important many-body physics \cite{McKay2011}, it has become important to understand decoherence in a many-body context in order to characterise the related heating processes \cite{Pichler2010,Poletti2012,Poletti2013,Pichler2012,Pichler2013,Schachenmayer2014}. Other studies have involved fundamentally different approaches to state preparation, including studies of how to use controlled dissipation to produce new cooling methods \cite{Griessner2006,Griessner2007,Chen2014,Griessner2005,Daley2004a}, or to drive the system into desired many-body states - including Bose-Einstein condensates (BECs) or metastable states such as $\eta$ pairs \cite{Diehl2008,Kraus2008}, paired states \cite{Diehl2010,Yi2012,Han2009,Sandner2011,Ates2012}, or even states with topological order \cite{Diehl2011,Bardyn2013}. Dissipation can also play a key role in suppressing two-body \cite{Syassen2008,Garcia-Ripoll2009,Zhu2014,Yan2013} and three-body \cite{Daley2009,Mark2012,Schutzhold2010} loss processes via a continuous quantum Zeno effect, enabling the production of interesting many-body states requiring effective three-body interactions \cite{ChenYC2011,Privitera2011,Bonnes2011,Titvinidze2011,Diehl2010b,Diehl2010d,Lee2010,Diehl2010a,Roncaglia2010,Kantian2009,Capogrosso-Sansone2009,Daley2009,Ottenstein2008,Chen2008,Rapp2008,Rapp2007}, and similar effects are seen in locally induced single-particle losses in cold gases \cite{Barontini2013,Zezyulin2012,Barmettler2011}. Such effects can also be used to manipulate or protect states in quantum computations \cite{Verstraete2009,Kliesch2011} and quantum memories \cite{Pastawski2011}, as well as to prepare entangled states dissipatively \cite{Cho2011,McCutcheon2009,Vacanti2009}, or realise specific quantum gates for group-II atoms \cite{Daley2011,Daley2008b}. It is also possible to generate spin-squeezed states using collisional loss in fermions \cite{Foss-Feig2012} or collisions with background gases \cite{Caballar2014}, as well as to protect states during adiabatic state preparation \cite{Stannigel2013}.
Studies of dissipation have more recently included investigating the competition between dissipation and coherent dynamics, including dissipative phase transitions \cite{Diehl2010c,Sieberer2013a,Honing2012}, excitation dynamics in clouds of Rydberg atoms \cite{Olmos2012,Lesanovsky2013,Lesanovsky2014,Petrosyan2013,Honing2013,Olmos2013,Ates2012}, and collective spin dynamics on the clock transition in group II atoms \cite{Martin2013}. Other work has begun to characterise dissipative driving processes by identifying universal classes of states and phase transitions that can be realised in this way \cite{Tauber2013,Sieberer2013}. The observation of many-body dissipation in AMO systems extends further to interacting ultracold atoms in optical cavities \cite{Ritsch2013}, beginning with the observation of the Dicke phase transition \cite{Baumann2010,Dimer2007} and continuing to more complex dynamics of many moving atoms interacting with the cavity \cite{Brooks2012,Habibian2013,Brahms2012,Torre2013,Buchhold2013}, as well as to systems of strongly interacting photons \cite{Peyronel2012,Reinhard2012,Petrosyan2011,Sevincli2011,Gorshkov2011}. In addition, dissipative quantum simulation and state preparation \cite{Muller2012} has been observed in trapped ions \cite{Barreiro2011,Schindler2013} and proposed in optomechanical arrays \cite{Tomadin2012,Lechner2013}.
\subsection{Purpose and outline of this review}
As is clear from the previous section, this area of research, combining ideas and techniques from quantum optics with strongly interacting many-body systems, has developed dramatically in the last few years. Control over dissipation promises many new tools to prepare interesting and important many-body states, and to investigate models in quantum simulators that are inaccessible to current methods.
In this review, we set about facilitating discussion between the AMO and many-body communities by introducing open many-body systems as they are understood in the AMO context. We will focus on real-time dynamics, especially of open quantum system out of thermal equilibrium. We will also build our discussion around the use of quantum trajectory techniques, which provide a numerical method for detailing with dissipative dynamics, and also together with continuous measurement theory, provide a way to understand the dynamics for this class of open AMO quantum systems. We discuss the numerical implementation of quantum trajectories for many-body systems, and address a series of examples highlighting interesting physics and important tools that are already being explored in these dissipative many-body systems.
The review is structured as follows -- we begin with a brief introduction to open quantum systems as they are discussed in quantum optics, noting the key approximations that can be made for many open AMO systems, and introducing notation. We then give an introduction to the quantum trajectories method and its physical interpretation, as well as its integration with time-dependent methods for many-body systems, especially the time-dependent Density Matrix Renormalisation Group (t-DMRG). We then introduce several examples of open many-body AMO systems, including light scattering from strongly interacting atoms in an optical lattice, collisional two-body and three-body losses as well as loss processes with single atoms, and quantum state preparation by reservoir engineering -- including methods to drive systems dissipatively into states with important many-body character. We finally return to further details of the physical interpretation of quantum trajectories by briefly introducing continuous measurement theory, before finishing with an outlook and summary.
\section{Open Quantum Systems in Quantum Optics}
Throughout this review, we will work in the framework of a quantum optics description of open quantum systems. Open quantum systems have been regularly discussed in many areas of physics, but the AMO systems that are studied in quantum optics often enable a series of approximations that allow specific types of understanding and control over the dynamics. In this section we introduce the concept of an open system, and the key approximations that lead to simplified descriptions of system dynamics. We then introduce a key example of those descriptions, specifically the Lindblad form of the master equation.
\subsection{General framework}
\label{sec:generalqo}
The general framework for an open quantum system is sketched in Fig.~\ref{fig:openquantumsystem} \cite{Breuer2007,Gardiner2005}. We consider a small quantum system\footnote{In a many-body context, a ``small'' quantum system might be relatively large and complex -- it should just be very substantially smaller, e.g., in the scale of total energy, or total number of degrees of freedom, when compared with the environment to which it is coupling}, which is coupled to a large \emph{environment}, which can also be thought of as a \emph{reservoir}. This is a similar relationship to the heat bath and the system in the canonical ensemble of statistical mechanics, and for this reason the environment is also referred to as the ``bath'' in some literature.
In this setup, the Hamiltonian for the \emph{total system}, including both the system and its environment, consists of three parts,
\begin{equation}
H_{\rm total} = H_{\rm sys} + H_{\rm env} + H_{\rm int}.
\end{equation}
Here $H_{\rm sys}$ is the system Hamiltonian, describing the coherent dynamics of the system degrees of freedom alone, in the absence of any coupling to the reservoir. In quantum optics, this is often a two-level system (e.g., an atom, a spin, or two states of a Cooper pair box), in which case we typically might have $H_{\rm sys} = \hbar\omega_{\rm sys} \sigma_z$, where $\omega_{\rm sys}$ is a system frequency scale, and $\sigma_z$ is a Pauli operator for a two-level system. Another common system is to have a harmonic oscillator (e.g., a single mode of an optical or microwave cavity, or a single motional mode of a mechanical oscillator), with system Hamiltonian $H_{\rm sys}=\hbar \omega_{\rm sys} a_{ho}^\dagger a_{ho}$, where $a_{ho}$ is a lowering operator for the harmonic oscillator. The degrees of freedom for the environment and their dynamics are described by the hamiltonian $H_{\rm env}$, and the interaction between the system and the reservoir is described by the hamiltonian $H_{\rm int}$.
In quantum optical systems the reservoir usually consists of a set of bosonic modes, so that
\begin{equation}
H_{\rm res} = \sum_l \int_0^\infty d\omega\, \hbar \omega \, b_l^\dagger (\omega) b_l(\omega), \label{hres}
\end{equation}
where $b_l(\omega)$ is a bosonic annihilation operator for a mode of frequency $\omega$. The index $l$ is convenient for describing multiple discrete modes at a given frequency, e.g., in the case that these are the modes of an external radiation field, $l$ can play the role of a polarisation index. These operators obey the commutation relation $[b_l(\omega),b_{l}^\dagger (\omega')]=\delta(\omega-\omega') \delta_{ll'}$, where $\delta(\omega)$ denotes a Dirac delta function and $\delta_{l l'}$ a Kronecker delta. Note that in general, a reservoir can have a frequency-dependent density of modes $g(\omega)$, which we have assumed for notational convenience is constant in frequency. Below when we make a Markov approximation for the dynamics, we will assume that this density of modes is slowly varying over the frequencies at which the system couples to the reservoir.
The interaction Hamiltonian $H_{\rm int}$ analogously takes a typical form
\begin{eqnarray}
H_{\rm int} &=& - {\rm i} \hbar \sum_l \int_0^{\infty} d\omega \, \kappa_l (\omega)\left(x^+_l+ x^-_l\right) \left[ b_l(\omega) - b_l^\dagger(\omega) \right]\\
&\approx& - {\rm i} \hbar \sum_l \int_0^{\infty} d\omega \, \kappa_l (\omega) \left[x^+_l b_l(\omega) - x^-_l b_l^\dagger(\omega) \right],\label{hintrwa}
\end{eqnarray}
where $x_l^{\pm}$ is a system operator, and $\kappa(\omega)$ specifies the coupling strength. In connection with the two-level systems listed above, for a two-level system, we might have $x^-_l = \sigma^-$, the spin lowering operator, and $x^+_l=\sigma^+$, the spin raising operator. In the case that the system is a Harmonic oscillator, we would typically have $x_l^{-} = a_{ho}$. In the second line of this expression, we have explicitly applied a rotating wave approximation, which we will now discuss.
\subsection{Key approximations in AMO systems}
\label{sec:approxamo}
A key to the microscopic understanding and control that we have of open AMO quantum systems is the fact that we can usually make three approximations in describing the interaction between the system and the reservoir that are difficult to make in other systems \cite{Gardiner2005}:
\begin{enumerate}
\item{\emph{The rotating wave approximation} - In the approximate form eq.~\eqref{hintrwa} of the interaction Hamiltonian we gave above, we neglected energy non-conserving terms of the form $x^+_l b_l^\dagger (\omega)$ and $x^-_l b_l (\omega)$. In general, such terms will arise physically in the couplings, and will be of the same order as the energy conserving terms. However, if we transform these operators into a frame rotating with the system and bath frequencies (an interaction picture where the dynamics due to $H_{\rm sys}$ and $H_{\rm env}$ are incorporated in the time dependence of the operators), we see that the energy-conserving terms will be explicitly time-independent, whereas these energy non-conserving terms will be explicitly time-dependent, rotating at twice the typical frequency scale $\omega_{\rm sys}$. This is shown explicitly below in eq.~\eqref{inthamcmeq} in section \ref{sec:inthamcm}. In the case that this frequency scale is much larger than the important frequency scales for system dynamics (or the inverse of the timescales for which we wish to compute the dynamics), the effects of these energy non-conserving terms will average to zero over the relevant dynamical timescales for the system, and so we can neglect their effects in describing the dynamics.}
\item{\emph{The Born approximation} - We also make the approximation that the frequency scales associated with dynamics induced by the system-environment coupling is small in scale compared with the relevant system and environment dynamical frequency scales. That is, if the frequency scales $\omega_{\rm sys}$ at which the system couples are much larger than the frequency scales $\Gamma$ corresponding to the system dynamics induced by the environment, then we can make a Born approximation in time-dependent perturbation theory. }
\item{\emph{The Markov approximation} - We assume that the system-environment coupling is frequency/time-independent over short timescales, and that the environment returns rapidly to equilibrium in a manner essentially unaffected by its coupling to the system, so that the environment is unchanged in time, and the dynamics of the system are not affected by its coupling to the environment at earlier times, i.e., the time evolution of the system does not depend on the history of the system.}
\end{enumerate}
\begin{figure}[tb]
\begin{center}
\includegraphics[width=10cm]{figures/couplingandtls}
\end{center}
\caption{(a) Coupling constant as a function of frequency, $\kappa(\omega)$, illustrating a slowly varying coupling strength around the central system frequency $\omega_{\rm sys}$. This is a key to several of the approximations made for open quantum systems in quantum optics. (b) Comparison of the scales for a two-level system. The system frequency is the optical frequency, $\omega_{\rm sys}$, associated with the energy difference between the ground and excited levels. In order to make the standard approximations, we require the timescale for the decay of the excited state, $\Gamma^{-1}$, i.e., the timescale for the dynamics induced by coupling to the environment, to be much longer than the optical timescale associated with the transition, or in freqeuncy units, $\Gamma^{-1} \gg \omega_{\rm sys}^{-1}$. We also work in a limit where any remaining system dynamics are associated with much smaller frequency scales. In this case, the coupling strength $\Omega_0$, associated with the frequency of Rabi oscillations between the states when they are coupled strongly on resonance, as well as the frequency of detuning from resonance $\Delta$, satisfy $\Omega_0, \Delta \ll \omega_{\rm sys}$.}\label{fig:tls}\label{fig:largefrequency}
\end{figure}
Each of these three approximations are normally exceptionally well justified, with the neglected terms in equations of motion being many orders of magnitude smaller than those that are retained. This results from the existence of a large frequency/energy scale that dominates all other such scales in the system dynamics described by $H_{\rm sys}$, and in the system-environment interaction $H_{\rm int}$. Specifically, we usually have a situation where $[H_{\rm sys}, x^\pm_l ]\approx \pm \hbar \omega_{l} x_l^\pm$, and the frequencies $\omega_l$, which are all of the order of some system frequency $\omega_{\rm sys}$, (1) dominate any other processes present in the system dynamics, and (2) are much larger than the frequency scales of dynamics induced by coupling to the reservoir, $\Gamma$. The first of these two conditions allows us to make rotating wave approximations in the interaction Hamiltonian, without which we could not make the Markov approximation. The second of these allows us to make the Born approximation.
The Markov approximation can be physically interpreted in two parts: it implies firstly that the system-environment coupling should be independent of frequency, to remove any back-action of the environment on the system that is not local in time, and secondly, that any correlation functions in the reservoir should retain no long-term memory of the interaction with the system. The first of these two pieces is again justified by the large frequency scale for the energy being transferred between the system and the reservoir relative to the effective coupling strength. As depicted in Fig.~\ref{fig:largefrequency}, the system and environment couple in a range of frequencies that is of the order of the effective coupling strength, $\Gamma$, and if $\Gamma \ll \omega_{\rm sys}$, then the variation in coupling strength $\kappa(\omega)$ over the range of frequencies where the modes couple significantly to the system is very small. For a single index $l$, we can write $\kappa(\omega) \approx \sqrt{\Gamma/(2\pi)}$. Note that we implicitly made this assumption for the density of modes in the reservoir $g(\omega)$ when we didn't include such a density of modes in eqs.~\eqref{hres} and \eqref{hintrwa} above. The second requirement depends on the physical details of the reservoir, and justifying this part of the approximation usually requires comparing the typical coupling timescales $\Gamma^{-1}$ with environment relaxation timescales $\omega_{\rm env}^{-1}$, which are in general different to $\omega_{\rm sys}^{-1}$. However, for many typical cases, we also find that $\omega_{\rm env} \sim \omega_{\rm sys}$.
In the case of decaying two-level atoms, or loss of photons from an optical cavity, these timescales are both related to the relevant optical frequency. On a physical level, when a two-level atom decays and emits a photon of wavelength $\lambda_{\rm opt}$ and frequency $\omega_{\rm opt}$, then the relevant system timescale $\omega_{\rm sys}^{-1}=\omega_{\rm opt}^{-1}$, and we have a requirement that the decay time, which is the relevant system-environment coupling timescale $\Gamma^{-1}$, is much longer than $\omega_{\rm opt}^{-1}$. In terms of the reservoir relaxation time, physically the time it takes for the photon to propagate away from the atom within the photon wavelength $\lambda$ is also of the order $1/\omega_{\rm opt}$. In that sense, the single frequency $\omega_{\rm opt}$ is the large frequency scale for each of the three approximations. The ratio of this frequency scale to $\Gamma$ is, for typical optical systems $\sim 10^{15}/10^8=10^7$, so that the approximations made here are good to many orders of magnitude.
\subsection{Master Equation}
The approximations given above make it possible to derive an equation of motion for the behaviour of the system alone in the presence of the dissipation due to coupling to the environment. Because of the combination of dissipative and coherent dynamics, this is expressed in terms of the density operator $\rho$ describing the state of the \emph{system}, which can be defined based on the total density operator for the system and the environment $\rho_{\rm total}$ as $\rho={\rm Tr}_{\rm env} \{\rho_{\rm total}\}$, where the trace is taken over the environment degrees of freedom. There are many routes to derive such an equation of motion (see, e.g., \cite{Lax1963,Collett1984,Gardiner1985,Gardiner2005,Breuer2007}), one of which is outlined in appendix \ref{sec:continuousmeasurement}, via a continuous measurement formalism.
The resulting markovian \emph{master equation} takes a Lindblad form \cite{Lindblad1976} (see, e.g., \cite{Gardiner2005,Collett1984,Gardiner1985})
\begin{eqnarray}
\dot \rho &=&-\frac{\rmi}{\hbar} [H,\rho] -\frac{1}{2} \sum_m \gamma_m\left[\tilde c_m^\dag \tilde c_m \rho + \rho \tilde c_m^\dag \tilde c_m - 2\tilde c_m \rho \tilde c_m^\dag \right]. \label{mastereqversion1}
\end{eqnarray}
Here, $H$ is the remaining system Hamiltonian, after the frequency scale $\omega_{\rm sys}$ of the coupling between the system and the environment has been transformed away. For a two-level system, this is a Hamiltonian describing the coupling between levels and a detuning in the rotating frame (see section \ref{sec:twolevel} below), i.e., in an interaction picture we we transform the operators so that we make the dominant coupling terms in the Hamiltonian time-independent. The operators $\tilde c_m$ are sometimes called Lindblad operators (or jump operators, as described below), and describe dissipative dynamics (including decoherence and loss processes) that occur at characteristic rates $\gamma_m$.
Note that the $\tilde c_m$ are all system operators, and the time-dependence of the environment does not appear in this equation. For the interaction Hamiltonian in Eq.~\eqref{hintrwa}, if we assume that all of the reservoir modes that couple to the system are unoccupied, we will have $\tilde c_l = x^-_l$. One example of $\tilde c_m$ would be a transition operator $\sigma^-$ from an excited state to a ground state in the two-level system depicted in Fig.~\ref{fig:largefrequency}b, describing spontaneous emissions, and further examples are given below in sections \ref{sec:physicalinterp}, \ref{sec:illustrative} and \ref{amosys}. We also note at this point that in general, the $\tilde c_m$ are non-Hermitian, although they can always be interpreted as resulting from continuous measurement of Hermitian operators acting only on the environment, as discussed in appendix \ref{sec:continuousmeasurement}. In the case that all $\tilde c_m$ are Hermitian, we can think of this master equation as representing a continuous measurement of the system operators $c_m$.
The general form in eq.~\eqref{mastereqversion1} also demonstrates explicitly the trace-preserving property of the master equation (${\rm Tr}\{ \dot \rho\}=0$). This form of dissipative dynamics is sometimes used as a toy model for dissipation because of its relatively simple interpretation. In the AMO / quantum optics context, it is particularly important because it describes the microscopic behaviour under the well-controlled approximations described in section \ref{sec:approxamo}.
In order to reduce the notational complexity, we will combine the rate coefficients with the operators in the form $c_m=\sqrt{\gamma_m}\tilde c_m$, and we will also equate energy and frequency scales setting $\hbar\equiv 1$, to obtain
\begin{eqnarray*}
\dot \rho &=&-\rmi [H,\rho] -\frac{1}{2} \sum_m \left[c_m^\dag c_m \rho + \rho c_m^\dag c_m - 2c_m \rho c_m^\dag \right]\nonumber,
\end{eqnarray*}
Note that we can also express this in the convenient alternative form
\begin{eqnarray}
\dot \rho &=&-\rmi (H_{\rm eff}\rho - \rho H_{\rm eff}^\dag) + \sum_m c_m \rho c_m^\dag, \label{mastereq}
\end{eqnarray}
where we refer to
\begin{eqnarray*}
H_{\rm eff}=H - \frac{i}{2} \sum_m c_m^\dag c_m
\end{eqnarray*}
as the \emph{effective Hamiltonian} for the dissipative system. in this form, the term $\sum_m c_m \rho c_m^\dag$ is often called the \emph{recycling term}, as it recycles the population that is lost from certain states due to the non-Hermitian effective Hamiltonian, placing it in other states. For a two level system with decay from an excited state to the ground state, one may think of the non-Hermitian part as removing amplitude from the excited state, and the recycling term as reinstating this in the ground state (see the many-body examples for a more detailed discussion of this).
For some simple systems, master equations can be solved analytically, to obtain expectation values of a given physical observable $A$ at particular times, $\langle A(t) \rangle = {\rm Tr}\{ A \rho(t)\}$ or also two-time correlation functions of the form $\langle A(t+\tau) B(t) \rangle$. In the next section we will discuss how the dynamics described by the master equation can be both computed and physically interpreted using quantum trajectories techniques. We then give some specific examples of small physical systems described by master equations. In the following two sections, we then discuss how master equations of the form eq.~\eqref{mastereq} describing many-body systems can be solved by combining quantum trajectories techniques with many-body numerical methods, and then discuss a number of examples of recent work on such physical systems.
\section{Quantum Trajectories}
Quantum trajectory techniques were developed in Quantum Optics in the early 1990s \cite{Dalibard1992,Molmer1993,Dum1992,Dum1992a,Carmichael1993,Plenio1998} as a means of numerically simulating dissipative dynamics,
and can be applied to any system where the time evolution of the density operator is described via a master equation (in Lindblad
form). These techniques involve rewriting the master equation as a stochastic average over individual \emph{trajectories}, which can be
evolved in time numerically as pure states. These techniques avoid the need to
propagate a full density matrix in time (which is often numerically
prohibitive), and replace this complexity with stochastic sampling. The key advantage this gives is that if the Hilbert space has dimension $N_H$, then propagating the density matrix means propagating an object of size $N_H^2$, whereas stochastic sampling of states requires propagation of state vectors of size $N_H$ only. Naturally, the penalty that is paid is the need to collect many samples for small statistical errors, and it is important that the number of samples required remains smaller than the size of the Hilbert space in order to make this efficient.
In quantum optics, these techniques were developed in parallel by a number of groups, as they arose out of studies of different open-quantum system phenomena. The versions of these techniques that most closely resemble what we present here were developed by Dalibard, Castin and M{\o}lmer \cite{Dalibard1992,Molmer1993} as a Monte Carlo method for simulation of laser cooling; by Dum et al. \cite{Dum1992,Dum1992a}, arising from studies of continuous measurement; by Carmichael \cite{Carmichael1993}, in studying of the generation of non-classical states of light; and by Hegerfeldt and Wilser \cite{Hegerfeldt1992}, in modelling a single radiating atom. The term \emph{quantum trajectories} was coined by Carmichael, whereas the other approaches were referred to as either a \emph{quantum jump approach}, or the \emph{Monte Carlo wavefunction method}. This later term should not be confused with so-called \emph{quantum Monte Carlo} techniques - the Monte Carlo treatment performed here is classical, in the sense that there is no coherent sum of amplitudes involved in evaluating expectation values from the samples we obtain. Mathematically, all of these approaches are essentially equivalent, but there are differences in the numerical implementations.
In quantum optics these methods have been applied to a wide variety of problems, including laser cooling \cite{Molmer1993,Marte1993,Castin1995,Marte1993b} and coherent population trapping \cite{Bardou1994}, the behaviour of cascaded quantum systems \cite{Carmichael1993b,Gardiner1993,Kochan1994}, the continuous quantum Zeno effect \cite{Itano1990,Power1996,Beige1996}, two-photon processes \cite{Garraway1994}, decoherence in atom-field interactions \cite{Moya-Cessa1993}, description of quantum non-demolition measurements \cite{Imamoglu1993} and decoherence in the atom-optics kicked rotor \cite{Ammann1998,Doherty2000,Daley2002,Wimberger2003}. For further examples, see the review by Plenio and Knight \cite{Plenio1998} and references therein. Note that complementary stochastic approaches were developed at the same time, especially by Gisin and his collaborators \cite{Gisin1984,Gisin1989,Gisin1992}. These and other related quantum state diffusion approaches \cite{Goetsch1994, Garraway1994b} involve continuous stochastic processes, but can be directly related back to the quantum trajectories formalism by considering homodyne detection of the output of a quantum system (see, e.g.,\cite{Wiseman1993b,Carmichael1993}).
In the remainder of this section we introduce the quantum trajectories method and illustrate its use and physical interpretation. We begin by presenting the first-order Monte Carlo wavefunction method put forward by Dalibard et al. \cite{Dalibard1992} and Dum et al. \cite{Dum1992}, and discuss statistical errors and convergence in this method. We then discuss the physical interpretation of quantum trajectories, before continuing with generalisations to calculations that are higher-order in the timestep. We then give two illustrative examples. The first is a simple example involving the optical Bloch equations for a two-level atom, whereas the second example gives a preview of our discussion of open many-body systems by treating dephasing of a hard-core Bose gas on a lattice.
\subsection{First-order Monte Carlo wavefunction method}
\label{sec:firstorder}
The simplest form of the quantum trajectory method involves expanding the master equation to first order in a time step $\delta t$, and was first described in this form by Dalibard et al. \cite{Dalibard1992} and Dum et al. \cite{Dum1992}. It involves the evolution of individual trajectories $\ket{\phi(t)}$, over which we average values of observables.
For a single trajectory, the initial state $\ket{\phi(t=0)}$ should be sampled from the density operator at time $t=0$, $\rho(t=0)$. For some systems, $\rho(t=0)$ may be a pure state, and this is the case for a number of our many-body examples. In that case, the initial state is always the same state, $\rho(t=0)=|\phi(t=0)\rangle\langle \phi(t=0)|$. Once we have an initial state, we propagate forward in time, making use of the following procedure in each time step:
\begin{enumerate}
\item Taking the state at the beginning of the time step, $|\phi(t)\rangle$, we first compute the evolution under the effective Hamiltonian. This will give us one candidate for the new state at time $t+\delta t$,
\begin{eqnarray}
|\phi^{(1)}(t+\delta t)\rangle=\left(1-\rmi H_{\rm eff}\delta t \right)|\phi(t)\rangle,
\end{eqnarray}
and we compute the norm of the corresponding state, which will be less than one because $H_{\rm eff}$ is non-Hermitian:
\begin{eqnarray}
\langle \phi^{(1)}(t+\delta t)|\phi^{(1)}(t+\delta t)\rangle&=&\langle \phi(t) | \left(1+\rmi H^\dagger_{\rm eff}\delta t \right)\left(1-\rmi H_{\rm eff}\delta t \right)|\phi(t)\rangle\\
&=&1-\delta p .
\end{eqnarray}
Here, we can consider $\delta p$ as arising from different potential decay channels, corresponding to the action of different Lindblad operators $c_m$,
\begin{eqnarray}
\delta p&=&\delta t\langle \phi(t) | \rmi(H_{\rm eff}-H_{\rm eff}^\dag)|\phi(t)\rangle\\ &=&\delta t \sum_m \langle \phi(t) | c^\dag_m c_m |\phi(t) \rangle \equiv\sum_m \delta p_m.
\end{eqnarray}
We can effectively interpret $\delta p_m$ as the probability that the action described by the operator $c_m$ will occur during this particular time step.
\item Then, we choose the propagated state stochastically in the following manner. We would like to assign probabilities to different outcomes so that:
\begin{itemize}
\item With probability $1-\delta p$
\begin{eqnarray}
|\phi(t+\delta t) \rangle =\frac{ |\phi^{(1)}(t+\delta t)\rangle }{\sqrt{1-\delta p}}
\end{eqnarray}
\item With probability $\delta p$
\begin{eqnarray}
|\phi (t+\delta t) \rangle = \frac{ c_m |\phi(t) \rangle }{\sqrt{\delta p_m / \delta t}}
\end{eqnarray}
where we choose one \emph{particular} $m$, which is taken from all of the possible $m$ with probability
\begin{eqnarray}
\Pi_m=\delta p_m / \delta p
\end{eqnarray}
\end{itemize}
In a practical numerical calculation, this action requires drawing a uniform random number $r_1$ between 0 and 1, and comparing it with $\delta p$. If $r_1>\delta p$, then no jump occurs, and the first option arrising from propagation under $H_{\rm eff}$, i.e., $|\phi (t+\delta t) \rangle \propto |\phi^{(1)}(t+\delta t)\rangle$ is chosen. If $r_1<\delta p$, then a jump occurs, and we must choose the particular $c_m$ operator to apply. We therefore associate each $m$ with an interval of real numbers, with the size of the interval being proportional to $\delta p_m$. Normalising the total interval length to one so that every $m$ corresponds uniquely to a range between 0 and 1, we then choose a second random number $r_2$, also uniformly distributed between 0 and 1, and choose the associated $c_m$ for which the assigned interval contains $r_2$.
\end{enumerate}
In order to see that this stochastic propagation is equivalent to the master equation, we can form the density operator,
\begin{eqnarray}
\sigma(t)=|\phi(t)\rangle\langle \phi(t) | .
\end{eqnarray}
From the prescription above, the propagation of this density operator in a given step is:
\begin{eqnarray}
\overline{\sigma(t+\delta t)} = (1-\delta p) \frac{| \phi^{(1)} (t+\delta t) \rangle}{\sqrt{1-\delta p}} \frac{\langle \phi^{(1)} (t+\delta t) |}{{\sqrt{1-\delta p}}}
+ \delta p \sum_m \Pi_m \frac{c_m |\phi(t)\rangle}{\sqrt{\delta p_m / \delta t}}\frac{\langle\phi(t)| c_m^\dag}{\sqrt{\delta p_m / \delta t}},
\end{eqnarray}
where $\overline X$ for any $X$ denotes a statistical average over trajectories, as opposed to the quantum mechanical expectation values or mathematically exact averages, which we will denote $\langle X \rangle$. Rewriting the terms from the above definitions, we obtain
\begin{eqnarray}
\overline{\sigma(t+\delta t)}&=& \sigma(t) -\rmi \delta t (H_{\rm eff} \sigma(t) - \sigma(t) H_{\rm eff}^\dag) +\delta t \sum_m c_m \sigma(t) c_m^\dag,
\end{eqnarray}
which holds whether $\sigma(t)$ corresponds to a pure state or to a mixed state. In this way, we see that taking a stochastic average over trajectories is equivalent to the master equation
\begin{eqnarray}
\dot \rho =-\rmi (H_{\rm eff}\rho - \rho H_{\rm eff}^\dag) + \sum_m c_m \rho c_m^\dag.
\end{eqnarray}
Note that this equivalence doesn't require a particular choice of $\delta t$, and in particular $\delta t$ can be chosen to be small in evaluating this evolution. Naturally, choosing very large $\delta t$ would strongly compromise the accuracy of the method when propagating the state in time. In the subsection \ref{sec:higherorder}, we will deal with how to improve upon this first-order method numerically.
\subsubsection{Computing single-time expectation values}
In order to compute a particular quantity at time $t$, i.e., $\langle A \rangle_t={\rm Tr} \{A\rho(t) \}$, we simply compute the expectation value of $A$ for each of our stochastically propagated trajectories, $\langle \phi(t) | A |\phi(t) \rangle$, and take the average of this quantity over all of the trajectories,
\begin{equation}
\langle A \rangle_t \approx \overline{ \langle \phi(t) | A | \phi(t) \rangle}.
\end{equation}
Provided our random number generators are well behaved, the trajectories are statistically independent, allowing for simple estimate of statistical errors in the computation of $\langle A \rangle_t$, as discussed in section \ref{sec:staterror}.
\subsubsection{Computing two-time correlation functions}
Two-time correlation functions of the form $C(t,\tau)=\langle A(t+\tau) B(t)\rangle$ appear in many contexts, including spectral functions \cite{Cohen-Tannoudji1998}, and also current autocorrelation functions in many-body physics \cite{Mahan2000}. These functions are a little more complicated to compute than single-time expectation values. To compute such correlation functions from a master equation, we typically apply the quantum regression theorem \cite{Lax1968,Gardiner2005}, which can be applied generally to an operator $X_{ij}=|i\rangle\langle j|$, where $|i\rangle$ and $|j\rangle$ belong to an orthonormal set of basis states for the Hilbert space. If we write $C_{ij}(t,\tau)=\langle X_{ij}(t+\tau) B(t)\rangle$, then we note that $C_{ij}(t,0)$ are one-time averages that can be calculated directly from the density matrix at a single time, and the $\tau$ dependence can be computed as
\begin{equation}
\frac{\partial C_{ij}}{\partial \tau} (t,\tau) = \sum_{kl} M_{ijkl} C_{kl}(t,\tau),
\end{equation}
where the $M_{ijkl}$ are the same matrix elements that appear in the equation of motion for one-time averages \cite{Lax1968,Gardiner2005}
\begin{equation}
\frac{d \langle X_{ij} (t)\rangle}{dt}=\sum_{kl} M_{ijkl}\langle X_{kl} (t)\rangle.
\end{equation}
To reproduce these values from quantum trajectories, we follow a simple procedure \cite{Molmer1993,Plenio1998}. We propagate a sample trajectory to time $t$, and then generate four helper states:
\begin{eqnarray}
|\chi^R_\pm (0) \rangle&=&\frac{1}{\sqrt{\mu^R_\pm}} (1\pm B) |\phi(t)\rangle,\\
|\chi^I_\pm (0) \rangle&=&\frac{1}{\sqrt{\mu^I_\pm}} (1\pm \rmi B) |\phi(t)\rangle,
\end{eqnarray}
where $\mu^R_\pm$ and $\mu^I_\pm$ normalise the resulting helper states. We then evolve each of these four states using the quantum trajectories procedure, and compute the correlation functions
\begin{eqnarray}
c^R_\pm(\tau)= \langle \chi^R_\pm (\tau)|A|\chi^R_\pm (\tau)\rangle, \\
c^I_\pm(\tau)= \langle \chi^I_\pm (\tau)|A|\chi^I_\pm (\tau)\rangle.
\end{eqnarray}
We can then reconstruct a sample for
\begin{equation}
C(t,\tau)=\frac{1}{4} \left[\mu^R_+c^R_+(\tau)-\mu^R_-c^R_-(\tau)-\rmi \mu^I_+c_+^I(\tau)+\rmi \mu^I_-c_-^I(\tau) \right],
\end{equation}
and average this over both the evolution up to time $t$ and the propagation of helper states to time $\tau$.
\subsection{Statistical Errors and convergence}
\label{sec:staterror}
\subsubsection{Estimating statistical errors}
Using the above procedure or generalisations that are higher-order in the timestep $\delta t$ (see section \ref{sec:higherorder}), we generate $N$ sample trajectories. Under the assumption that the random numbers used in the numerical implementation are statistically random and uncorrelated\footnote{See, e.g., Ref.~\cite{Press1995} for a general discussion of random number generators. For typical calculations, only a few tens of thousands of random numbers must be generated, and good quality pseudorandom number generators suffice. See Ref.~\cite{Frauchiger2013} for a recent discussion of improvements to that are possible using quantum random number generators.}, these trajectories are statistically independent, and we can estimate the correct mean $\langle X \rangle$ of any operator of interest $\hat X$ as
\begin{equation}
\overline X (t)= \frac{1}{N} \sum_i X_i \equiv \frac{1}{N} \sum_i \langle \phi_i (t) | \hat X |\phi_i(t) \rangle.
\end{equation}
The central limit theorem implies that for sufficiently large $N$, the probability distribution for $\overline X$ will be well approximated by a Gaussian distribution with mean $\langle X \rangle$. The statistical error in this mean is the standard deviation of that distribution, which in turn can be estimated based on the variance of the values $X_i$ in the following way. We consider
\begin{eqnarray}
{\rm Var}[\overline X]&=& \left \langle \left( \overline X- \langle X \rangle \right)^2 \right \rangle = \left\langle \left( \frac{1}{N} \sum_{i=1}^N X_i \right)^2 - 2 \overline X \langle X \rangle + \langle X \rangle^2 \right\rangle\nonumber\\
&=& \frac{1}{N^2} \sum_{i=1}^N\sum_{j=1}^N \left\langle X_i X_j \right\rangle-\langle X \rangle^2
= \frac{1}{N^2} \sum_{i=1}^N \left\langle X_i^2 \right\rangle+ \frac{N-1}{N}\langle X \rangle^2-\langle X \rangle^2 \nonumber\\
&=&\frac{1}{N} (\langle X^2 \rangle - \langle X \rangle ^2)= \frac{1}{N} {\rm Var}[ X]. \nonumber\\
\end{eqnarray}
In deriving this standard result, we use both the independence of the trajectories $\langle X_i X_{j\neq i} \rangle =\langle X_i \rangle \langle X_{j\neq i} \rangle$, and the replacement that $\langle \overline X \rangle=\langle X \rangle$, i.e., the variance written in the last line is the true variance of the distribution for $X$. As a result, it can be shown (see, e.g., Ref.~\cite{Press1995}) that we should use the approximator
\begin{equation}
{\rm Var}[ X] \approx \frac{\sum_{i =1}^N(X_i - \overline X)^2 }{N-1},
\end{equation}
if we would like to estimate this based on the samples we obtain in the calculation.
In this sense, we can always calculate the statistical error $\sigma_A$ in our estimate of a quantity $\langle A \rangle$ by taking the \emph{estimate of the population standard deviation} $\Delta A$ from our $N$ samples, and \emph{dividing by $\sqrt{N}$},
\begin{equation}
\sigma_A = \frac{\Delta A}{\sqrt{N}}.
\end{equation}
How many trajectories we will require for good convergence will depend both on the details of the dynamics and on the quantity being calculated. For variables with non-zero mean, we would typically like to have $\sigma_A/\langle A \rangle \ll 1$, which implies that $\sqrt{N} \gg \Delta A / \langle A \rangle$. As the sample estimate overestimates the population standard deviation (see e.g., \cite{Molmer1993}), we can consider the sample estimate for the standard deviation $\Delta A$ here.
\subsubsection{Global quantities vs. local quantities}
\label{sec:globallocal}
In M{\o}lmer et al. \cite{Molmer1993}, there is a discussion of the different number of trajectories required for global quantities vs. local quantities to be straight-forwardly estimated. This is done for single-particle systems in a Hilbert space dimension $N_H$.
For \emph{global quantities} such as the total energy $A_G=E_{\rm tot}$, there is usually a fixed relationship between the estimate $\Delta A_G$ and $\langle A_G \rangle$ that is independent of the dimension of the Hilbert space $N_H$. For example, in the Brownian motion of a particle thermalised with a reservoir at temperature $T$, we would expect $\langle A_G \rangle\approx (3/2) k_B T$, where $k_B$ is the Boltzmann constant, and $\Delta A_G\approx \sqrt{3/2} k_B T$. The requirement on the number of trajectories is then simply $N\gg 1$, and we should expect that the relative error is well estimated by $1/\sqrt{N}$ (so that for 10\% relative statistical error, we would require $N\sim 100$ trajectories. These same arguments apply to many-particle systems, we also expect global averages involving all of the particles in a system (or all of the spins or lattice sites in a spin or lattice system) to be efficiently treatable with quantum trajectories methods.
For \emph{local quantities} in a single-particle calculation, the opposite is true. If we try to determine the population of a given eigenstate (of the Hamiltonian, or of the momentum operator), then we expect that in a Hilbert space of dimension $N_H$, our local quantities $A_L$ and their statistical variance will behave as $\langle A_L \rangle \sim 1/N_H$ and $\sigma_A^2 \sim 1/N_H$. As a result, we see that we require $N\gg N_H$ for good statistical convergence, and there is no advantage in using a quantum trajectories technique over direct integration of the master equation.
However, for \emph{local quantities in many-body systems}, the situation is not as clear-cut. If we attempt to compute the population of a given many-body eigenstate out of $N_H$ possible many-body eigenstates, then the previous conclusion for local quantities in a single-particle system still applies. However, single-particle or few-particle quantities in a finite-size system that are local in space or momentum will often scale simply as the size of the system, $L$, whereas for large systems with fixed particle density, $N_H\propto \exp(L)$. As a result, quantities such as the local density on one site in a lattice of $L$ sites, or individual elements of a single-particle density matrix on such a lattice will tend to scale as $\langle A_S \rangle \sim 1/L$ and $\sigma_A^2 \sim 1/L$. We therefore require $N\gg L$ for good statistical convergence, and not $N
\gg N_H$. Although this will require more trajectories for small relative statistical error than in the case of global quantities (by a factor of $\sqrt{L}$), there are often still large advantages in using quantum trajectory techniques to calculate these quantities when the size of the Hilbert space $N_H$ becomes large.
We will discuss this again below, when we give two illustrative examples for the use of quantum trajectories in section \ref{sec:illustrative}. Specifically, in section \ref{sec:hcb}, we treat the dephasing of hard-core bosons on a lattice, and consider calculation of the total energy, local single-particle correlations, and local densities using quantum trajectory techniques.
\subsection{Physical interpretation}
\label{sec:physicalinterp}
One of the greatest strengths of the quantum trajectories approach to dissipative systems, but also one of the most subtle points in its usage, is that it gives us a simple physical interpretation for the physics induced by the environment on the system. The method is also known as the method of \emph{quantum jumps}, and the Lindblad operators $c_m$ are also called \emph{jump operators}, inviting the picture of a system that evolves under the non-Hermitian effective Hamiltonian $H_{\rm eff}$ and then undergoes \emph{quantum jumps} at certain points in time. The master equation is then an appropriately weighted stochastic average over all of the different times at which the jumps could occur, and all of the different types of jumps that can occur.
To see this working in practice, consider a two-level system like that depicted in Fig.~\ref{fig:tls}b. If we drive the system, coupling the excited and ground states ($\ket{e}$ and $\ket{g}$) respectively with an effective Rabi frequency $\Omega_0$ and a detuning $\Delta$, then the system will undergo coherent dynamics, interrupted at particular times by spontaneous emission events, which return the atom to the ground state $\ket{g}$, due to the action of a jump operator $c=\sqrt{\Gamma}|g\rangle\langle e|$. If photons can be scattered in multiple directions, then the different directions constitute different channels $c_m$, and we can average over the probability distribution of emission directions. Note that this is presented carefully below in section \ref{sec:lightscattering}. The stochastically chosen times for the jumps are the times at which spontaneous emissions occur, and stochastically sampling different $m$ values amounts to a stochastic sampling of the direction in which the photon is emitted.
This is a very appealing intuitive picture of the dynamics, and connects strongly with an intuition of what would happen if we are actually able to measure the environment. Indeed, quantum trajectories techniques were originally developed by certain groups from studies of photon counting \cite{Hegerfeldt1992}, or continuous measurement \cite{Dum1992}. If we are able to make perfect measurements and we see a photon appear in the time window $\delta t$, then we know that a jump has occurred, and that the state of the atom should be projected on the ground state. On the other hand, if we know that no jump has occurred, then the corresponding evolution of the system is an evolution under the effective Hamiltonian $H_{\rm eff}$. Already here, we see a key piece of physics that will recur multiple times: \emph{knowing} that no jump has occurred means that we gain information about the system, just as knowing that a jump has occurred gives us information that the atom is projected into the ground state.
To see this in a simple example, consider preparing the system in a state $\ket{\psi(t)}=\alpha \ket{g}+ \beta \ket{e}$, where $\alpha$ and $\beta$ are complex coefficients, and set $\Omega_0=\Delta=0$. We can then ask what we expect to happen in a single step, depending on whether we observe a spontaneously emitted photon or not. As we are not concerned about the motion of the atom, or therefore about the direction of spontaneously emitted photons, we can consider a single jump operator $c=\sqrt{\Gamma}|g\rangle\langle e|$, and an effective Hamiltonian which is simply $H_{\rm eff} = -\rmi (\Gamma/2) |e\rangle\langle e|$. If a jump occurs in a time step $\delta t$, then the state after this jump is
\begin{equation}
\ket{\psi(t+\delta t)}=\frac{c\ket{\psi(t)}}{\Vert c\ket{\psi(t)}\Vert }=\ket{g}.
\end{equation}
So if there is a spontaneous emission, then the state is projected onto the ground state, as expected. Consider now what happens if no jump occurs: We then evolve the state as
\begin{equation}
\ket{\psi(t+\delta t)}=\frac{\exp(-\rmi H_{\rm eff} \delta t) \ket{\psi(t)}}{\Vert \exp(-\rmi H_{\rm eff} \delta t) \ket{\psi(t)}\Vert }=\frac{ \alpha \ket{g}+ \beta{\rm e}^{-\Gamma \delta t/2} \ket{e}}{\Vert \alpha \ket{g}+ \beta{\rm e}^{-\Gamma \delta t/2} \ket{e}\Vert }=\frac{ \alpha \ket{g}+ \beta{\rm e}^{-\Gamma \delta t/2} \ket{e}}{\sqrt{|\alpha|^2 + |\beta|^2{\rm e}^{-\Gamma \delta t} }}.
\end{equation}
In this way, the probability of finding the system in the excited state decreases relative to the probability of the system being in the ground state, provided $\alpha\neq 0$. Essentially, through the lack of a spontaneously emitted photon, we have gained the knowledge that the system is somewhat more likely to be in the ground state than in the excited state, and this gain in knowledge is reflected in the relative probabilities for occupation of these states at time $t+\delta t$ relative to time $t$. Thus, the dynamics are affected by coupling to the environment even in the absence of actual spontaneously emitted photons.
There are two important cautions to over-interpreting this intuitive physical picture. Firstly, the master equation can typically be expanded in different sets of jump operators.
If there exists a unitary transformation $T$ in the Hilbert space of the system such that for the dissipative part of the master equation
\begin{equation}
\mathcal{L}\rho = -\frac{1}{2} \sum_m \left[c_m^\dag c_m \rho + \rho c_m^\dag c_m - 2c_m \rho c_m^\dag \right],
\end{equation}
the operator $T$ satisfies the condition
\begin{equation}
T[\mathcal L \rho]T^\dagger = \mathcal L \left(T\rho T^\dagger\right),
\end{equation}
then we can rewrite each jump operator $c_m$ such that we obtain new operators $d_m=T^\dagger c_m T$, and show \cite{Molmer1993} that
\begin{equation}
\mathcal{L}\rho = -\frac{1}{2} \sum_m \left[d_m^\dag d_m \rho + \rho d_m^\dag d_m - 2d_m \rho d_m^\dag \right].
\end{equation}
In this sense, the operators $d_m$ are equally good choices for the jumps as the operators $c_m$. In general, for the application of quantum trajectories as a numerical method, we should simply choose the jump operators that provide most rapid convergence of sampling. This has been investigated in a number of contexts \cite{Molmer1993,Holland1998,Breslin1995,Wiseman2000,Atkins2005}, including quantum feedback \cite{Wiseman2005}. However, to ascribe physical meaning to an individual trajectory it is important to be able to associate the specific jumps with measurable properties of the environment after the jump. This detector dependence should even produce measurable differences under certain conditions \cite{Wiseman2012}.
In addition, ascribing a pure state to the system as a function of time implies knowledge of measurement results from the environment. Although each trajectory implies a particular physical picture of the dissipation, it is the average over trajectories that properly specifies our knowledge of the system if we do not measure the environment. At the same time, there are sometimes means to make a measurement on the system and infer the measurement result for the environment. A simple case of this type of \emph{postselection} appears when the dissipative process we deal with is dominated by particle loss, as we will treat in section \ref{sec:particleloss} below. There, if we know that no particles are lost by making a measurement at the end of a dynamical process, then we can project the state of the system on the state simply evolved under the effective Hamiltonian $H_{\rm eff}$, as this is an equivalent mechanism to measure that no jumps occurred during the evolution.
In order to address how the approximations that we can make in AMO systems give rise to the behaviour exhibited by the system, we will re-derive the quantum trajectory approach in appendix \ref{sec:continuousmeasurement}. There we will make explicit this connection between measurement of the environment and inference of the state of the system.
\subsection{Alternate formulation for higher-order integration in time}
\label{sec:higherorder}
The downside of the method presented in section \ref{sec:firstorder} is that it is only first order in the time step $\delta t$. Under some conditions, where the effect of dissipation is much slower than other dynamical timescales for the system, it might be possible to continue to apply this first order time step for the dissipative dynamics, but take a step towards higher-order expansions by propagating the state under the effective Hamiltonian more accurately, i.e., replacing the step $|\phi^{(1)}(t+\delta t)\rangle\propto\left(1-\rmi H_{\rm eff}\delta t \right)|\phi(t)\rangle$ with a higher-order version, or exact computation of $|\phi^{(1)}(t+\delta t)\rangle\propto\exp(-\rmi H_{\rm eff}\delta t)|\phi(t)\rangle$ (and corresponding normalisation of the resulting state).
However, this still doesn't alleviate the problem that the jump takes an entire time step $\delta t$, and effective results in an underestimation of the typical time between jumps of the order of $\delta t$. In this way, there is a systematic first-order overestimation of the rate at which jumps occur. Direct higher-order adjustments to the method presented in section \ref{sec:firstorder} were made by Steinbach et al., \cite{Steinbach1995}, allowing calculations of a Runge-Kutta type up to fourth order.
A general way to improve the method was originally proposed by Dum et al., \cite{Dum1992a}, in which they approach the problem from the point of view of continuous measurement, and essentially take the limit $\delta t \rightarrow 0$ in thinking about the occurrences of jumps within individual trajectories\footnote{We can't completely take the limit $\delta t \rightarrow 0$ physically, as we have made important approximations regarding the fact that we consider dynamics over long timescales compared with $\omega_{\rm sys}^{-1}$. However, we can take the limit of vanishing $\delta t$ from the point of view of numerical implementations after the other approximations have been implemented. See appendix \ref{sec:continuousmeasurement} for a more detailed discussion.}.
The revised version of the scheme takes the following form:
\begin{enumerate}
\item{Sample the initial state and begin the propagation under the effective Hamiltonian as in the scheme from section \ref{sec:firstorder}.}
\item{Sample a random number $r$, uniformly distributed between $0$ and $1$.}
\item Numerically solve the equation
\begin{equation}
\Vert\exp(-\rmi H_{\rm eff} t_1) \ket{\phi(t_0)}\Vert^2=r
\end{equation}
in order to find the time $t_1$ at which the next jump occurs, given that the previous jump or the start of the calculation was at time $t_0$. This can be solved using higher order integration methods, including Runge-Kutta.
\item $\ket{\phi(t)}$ is then computed numerically in the time interval $t\in [t_0,t_1]$ as
\begin{equation}
\ket{\phi(t)}=\frac{\exp[-\rmi H_{\rm eff} (t-t_0)] \ket{\phi(t_0)}}{\Vert \exp[-\rmi H_{\rm eff} (t-t_0)] \ket{\phi(t_0)}\Vert }.
\end{equation}
\item At time $t_1$, a quantum jump is applied, with probabilities for application of each $c_m$ determined as in step 2 of the method in section \ref{sec:firstorder}. That is, we choose a particular $m$ based on the probabilities
\begin{equation}
\delta p_m \propto \langle \phi(t_1)|c_m^\dagger c_m|\phi(t_1)\rangle,
\end{equation}
and apply the jump as
\begin{equation}
\ket{\phi(t_1^+)}=\frac{c_m \ket{\phi(t_1^-)}}{\Vert c_m \ket{\phi(t_1^-)}\Vert }.
\end{equation}
Here, $|\phi(t_1^-)\rangle$ is the state obtained in step 4 by propagating in time under $H_{\rm eff}$ up to time $t_1$, and $|\phi(t_1^+)\rangle$ is the state after the jump, i.e., the state we use to continue the time evolution.
\item We now continue the time evolution from step 2, choosing a new random number $r$.
\end{enumerate}
In this method, jumps occur at a particular point in time, rather than taking a fixed length of evolution time, and both the times of the jumps and the evolution under the effective Hamiltonian between the jumps can be solved numerically to arbitrary precision. In this way, we remove the reliance on a first-order Euler expansion.
\subsection{Illustrative examples}
\label{sec:illustrative}
We now give two example applications for quantum trajectories. The first is the simple case of the optical Bloch equations, which describe a driven two-level atom undergoing spontaneous emissions. While the underlying master equation is easily tractable, this gives a good basis for our intuition on the meaning of individual trajectories, and a good example for the calculation of statistical errors. The second example then provides a lead-in to thinking about many-body dissipative systems by considering the dephasing of a hard-core Bose gas on a lattice. This system is somewhat more complicated, though analytical experessions can be found for certain quantities, against which we can straight-forwardly benchmark the statistical error estimates discussed in section \ref{sec:staterror}.
\subsubsection{Optical Bloch Equations}
\label{sec:twolevel}
The optical Bloch equations describe a two-level atom that is driven by a classical laser field \cite{Cohen-Tannoudji1998}, with a detuning $\Delta$ between the frequency of the laser field and the atomic transition frequency, as depicted in Fig.~\ref{fig:largefrequency}b. In the absence of dissipation, this system undergoes Rabi oscillations at a frequency $\Omega$ that depends on the intensity and polarization of the classical laser field, and the dipole matrix elements between the two internal states $\{\ket{g}$ and $\ket{e}\}$. This gives rise to the well-known Rabi Hamiltonian for a spin-1/2 system. In the presence of damping, where an atom can undergo spontaneous emissions, decaying from the excited state to the ground state, we obtain the master equation
\begin{equation}
\frac{d}{dt} \rho = -{\rm i}[H_{\rm opt},\rho] - \frac{\Gamma}{2} \left( \sigma_+\sigma_- \rho + \rho \sigma_+\sigma_- -2 \sigma_- \rho \sigma_+ \right),
\end{equation}
where the Hamiltonian
\begin{equation}
H_{\rm opt} = - \frac{\Omega}{2} \sigma_x - \Delta \sigma_+\sigma_- .
\end{equation}
Here, in the basis $\{\ket{e},\ket{g}\}$, we denote the system density matrix $\rho$ and the Pauli matrices as
\begin{equation}
\rho=\left(\begin{array}{cc}
\rho_{ee} & {\rho}_{ge}\\
{\rho}_{eg} &\rho_{gg}
\end{array}\right), \,\,\,\sigma_x=\left(\begin{array}{cc}
0 & 1\\
1 & 0
\end{array}\right), \,\,\, \sigma_+=\left(\begin{array}{cc}
0 & 1\\
0 & 0
\end{array}\right), \,\,\,\sigma_- =[\sigma_+]^\dagger .
\end{equation}
We assume $\Omega$ is real for simplicity of notation.
This master equation (the so-called \emph{optical Bloch equations}) can be expressed in terms of the matrix elements of the system density operator as
\begin{equation}
\frac{d}{dt}\left[\begin{array}{c}
{\rho}_{eg}\\
{\rho}_{ge}\\
\rho_{ee}\\
\rho_{gg}
\end{array}\right]=\left[\begin{array}{cccc}
i\Delta-\Gamma/2 & 0 & -{\rm i}\Omega/2 & {\rm i}\Omega/2\\
0 & -{\rm i}\Delta-\Gamma/2 & {\rm i}\Omega/2 & -{\rm i}\Omega/2\\
-{\rm i}\Omega/2 & {\rm i}\Omega/2 & -\Gamma & 0\\
{\rm i}\Omega/2 & -{\rm i}\Omega/2 & \Gamma & 0
\end{array}\right]\left[\begin{array}{c}
{\rho}_{eg}\\
{\rho}_{ge}\\
\rho_{ee}\\
\rho_{gg}
\end{array}\right].
\end{equation}
These equations can be solved exactly, and describe damped oscillations of the system between the two internal states, damping to an excited state population $P_e \equiv \rho_{ee}$
\begin{equation}
\rho_{ee}=\frac{\frac{1}{4}|\Omega|^{2}}{\Delta^{2}+\frac{1}{4}\Gamma^{2}+\frac{1}{2}|\Omega|^{2}}.
\end{equation}
It is straight-forward to formulate the quantum trajectories approach for this master equation, which contains a single jump operator, $c\equiv \sigma_-$, and we can write the corresponding effective Hamiltonian as
\begin{equation}
H_{\rm eff} = H_{\rm opt} - {\rm i}\frac{\Gamma}{2} \sigma_+\sigma_-.
\end{equation}
The states can be propagated straight-forwardly in time numerically, and we show two example trajectories in the left panel of Fig.~\ref{fig:twolevelexample}. Each of these trajectories exhibits Rabi oscillations, which are interrupted by spontaneous emissions at points in time that are randomly chosen, and thus vary from trajectory to trajectory. After each spontaneous emission, the Rabi oscillations in an individual trajectory reappear with their original amplitude, implying that if we know exactly at which time(s) the atom was reset to its ground state, then we could predict the exact form of the system state at any point in time. In the absence of knowledge of these times, the excited state population damps to its steady-state value, which is marginally below $1/2$ for the case of $\Delta=0$, $\Gamma \ll \Omega$, which we have here. This damping comes from the incoherent averaging over different trajectories, and averaging over 1000 trajectories reliably predicts $P_e$ as a function of time up to statistical errors that are less than a few percent of $P_e$ throughout most of the evolution. This is shown in the right-hand panel of Fig.~\ref{fig:twolevelexample}, where the dotted line represents the exact solution for $P_e$ from the master equation, and the solid line represents the average over trajectories. The statistical error bars shown here are calculated as discussed in section \ref{sec:staterror}, i.e., from the sample of $N_{\rm traj}=1000$ trajectories we compute the population variance of $P_e$, $\Delta P_e^2$, and compute the error as $\sigma_{P_e}=\sqrt{\Delta P_e^2} / \sqrt{N_{\rm traj}}$.
\begin{figure}
\begin{center}
\includegraphics[width=6.8cm]{figures/twolevelexample}
\includegraphics[width=6.8cm]{figures/twolevelmeans}
\end{center}
\caption{Illustrative example of quantum trajectories averaging for a two-level system. (left) Probability of finding the atom in the excited state $P_e$ as a function of time $t\Omega$ for two example trajectories (with blue solid and red dashed lines showing different random samples). We see the effect of quantum jumps, where the atom is projected on the ground state. Here the detuning $\Delta=0$, and $\Gamma=\Omega/6$. (right) Values for $P_e$ averaged over 1000 sample trajectories (Solid line), compared with the exact result from direct integration of the master equation (dotted line). The quantum trajectories results agree with the exact results within the statistical errors, which are shown here as error bars calculated as described in section \ref{sec:staterror}. } \label{fig:twolevelexample}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=12cm]{figures/twolevelerror10000}
\end{center}
\caption{Statistical errors in the quantum trajectories computation for a two-level system. Here we compare the statistical error estimate (blue solid line) to the absolute error in the averaged value of $P_e$ when compared with the exactly computed value (red dashed line) at a time $t\Omega=40$. We show these values as a function of the number of trajectories over which we average, with all quantities plotted on a logarithmic scale. As expected, the behaviour of both the statistical error and the discrepancy is somewhat erratic for small numbers of trajectories, but decrease steadily $\propto 1/\sqrt{N}$ for larger numbers. The absolute error is mostly below the statistical error estimate, consistent with a gaussian distribution of possible errors with a standard deviation equal to our statistical error estimate. Note that these values depend on the particular random sample of trajectories obtained in the calculation. From the central limit theorem and the properties of Gaussian distributions that for any given computed values, we expect the absolute error to be smaller than the statistical error as it is quoted here for $68.2\%$ of all possible samples. } \label{fig:twolevelerror}
\end{figure}
In order to look into the statistical errors in more detail, we fix the time, and investigate how the statistical error and the discrepancy between the value of $P_e$ calculated from the trajectory average and the exact value obtained from the master equation vary with $N_{\rm traj}$. This comparison is shown in Fig.~\ref{fig:twolevelerror}, where the solid blue line represents the statistical error, and the dashed red line the discrepancy between the trajectory average and the exactly computed value as a function of the number of trajectories. Each of these lines is shown on a double logarithmic scale to make the scaling $\propto 1/\sqrt{N}$ clearer. The lines shown in the figure represent a particular sample of trajectories, and were chosen so that we simply grew the sample as we increased $N_{\rm traj}$. Naturally, we expect that different samples will produce different plots, but the results shown here constitute a typical example. From the central limit theorem, as described in \ref{sec:staterror}, we expect the means over different samples of trajectories to be approximately normally distributed, and the statistical error represents the standard deviation of that distribution. Thus we expect $68.2\%$ of all possible samples to have discrepancies to exact values that fall below the quoted statistical error. The small ranges of sample sizes where the discrepancy is marginally larger than the estimated statistical errors are perfectly consistent with this analysis.
\subsubsection{Dephasing for hard-core lattice bosons }
\label{sec:hardcorebosons}
\label{sec:hcb}
Our second example focusses on a gas of hard-core bosonic particles moving on a lattice, and introduces dissipation in a simple many-body system. We begin by considering a 1D Hamiltonian of the form \begin{equation}
H_{\rm bos} = - J \sum_{l} a^\dagger_{l+1} a_l + a^\dagger_{l} a_{l+1},\,\,\, (a_l^\dagger)^2\equiv 0,
\end{equation}
where $a_l$ is a bosonic destruction operator for a particle on lattice site $l$, $J$ is the tunnelling amplitude for a particle moving between sites in a 1D chain, and we have introduced a hard-core constraint, that is, we allow at most one particle on each lattice site. As will be discussed in more detail in sections \ref{sec:numerics} and \ref{sec:amosys} below, bosonic atoms confined to the lowest band of an optical lattice \cite{Jaksch1998,Jaksch2005,Lewenstein2007,Bloch2008b,Lewenstein2012}, the system are described under well-controlled approximations by the Bose-Hubbard model,
\begin{equation}
H_{BH}=-J\sum_{\langle i,j\rangle}{a}_{i}^{\dag}{a}_{j}+\frac{U}{2}\sum_{i}{a}_{i}^{\dag}{a}_{i}({a}_{i}^{\dag}{a}_{i}-1),\label{HBH}
\end{equation}
where $U$ is the on-site interaction energy shift, and $\langle i,j\rangle$ denotes a sum over nearest-neighbour sites. The model represented by $H_{\rm bos}$ is a limiting case of this when atoms are confined to move along one lattice direction in an optical lattice setup \cite{Jaksch2005,Lewenstein2007,Bloch2008b}, and where the particles are very strongly interacting $U/J\rightarrow \infty$ so that energy conservation strongly disfavours doubly-occupied lattice sites. At less than unit filling of particles in lattice sites, the ground state of $H_{\rm bos}$ is a bosonic superfluid, in which the tails of the momentum distribution decrease algebraically with increasing momentum. The properties of this gas can be derived straight-forwardly via a Jordan-Wigner transformation \cite{Sachdev2011}, which allows the solution to be expressed in terms of the properties of a gas of non-interacting Fermions.
Below, in section \ref{sec:lightscattering}, we will discuss how undergoing spontaneous emissions, i.e., incoherently scattering light from the lattice lasers or another source, tends to localise particles in the system, providing the environment with information about the location of the particles. In the typical limit for optical lattice experiments, this localisation essentially projects a particle onto a single lattice site (although we typically cannot or do not measure which site this was). This localisation in space delocalises the atom in quasi-momentum space across the first Brillouin zone, leading to a broadening of the quasi-momentum distribution and an increase in the kinetic energy beginning from the ground state. For hard-core bosons in 1D, these processes can be approximately described by the master equation
\begin{equation}
\frac{d}{dt} \rho = -{\rm i}[H_{\rm bos},\rho] - \frac{\Gamma}{2} \sum_l \left( a_l^\dagger a_l a_l^\dagger a_l \rho + \rho a_l^\dagger a_l a_l^\dagger a_l -2 a_l^\dagger a_l \rho a_l^\dagger a_l \right),
\end{equation}
where $\Gamma$ is the effective rate of spontaneous emission events.
\begin{figure}
\begin{center}
\includegraphics[width=6.8cm]{figures/hardcoreexample}
\includegraphics[width=6.8cm]{figures/hardcoremeans}
\end{center}
\caption{Illustrative example of quantum trajectories averaging for a gas of hard-core bosons on a lattice, analogous to the example for a two-level system in Fig.~\ref{fig:twolevelexample}. Here we show results from exact diagonalisation calculations with 5 particles on 10 lattice sites. (left) Kinetic energy of the system of hard-core bosons as a function of time $tJ$ for two example trajectories (with blue solid and red dashed lines showing different random samples). We see the effect of quantum jumps, where the kinetic energy increases as individual atoms are localised in space, and hence spread over the Brillouin zone in quasi-momentum. Here the scattering rate $\Gamma=0.1 J$. (right) Values for the kinetic energy averaged over 1000 sample trajectories, compared with the exact result from eq.~\eqref{hardcorebosonexactke}. As in the case of the two-level system, the quantum trajectories results agree with the exact results within the statistical errors, which are shown here as error bars calculated as described in section \ref{sec:staterror}. } \label{fig:hardcoreexample}
\end{figure}
As in the case of the optical Bloch equations, it is straight-forward to formulate the quantum trajectories approach for this master equation. Now we have a set of jump operators, with as many jump operators as we have lattice sites in the system and $c_m\equiv a_m^\dagger a_m$, i.e., the on-site number operator, and a corresponding effective Hamiltonian
\begin{equation}
H_{\rm eff} = H_{\rm bos} - {\rm i}\frac{\Gamma}{2} \sum_m a_m^\dagger a_m a_m^\dagger a_m.
\end{equation}
Intuitively, this form for the jump operators makes sense, as application of a number operator on a given site, $a_l^\dagger a_l |\psi\rangle/\Vert a_l^\dagger a_l |\psi\rangle \Vert $, leads to the localisation of a single particle on site $l$. The probability that this occurs on a particular site $l$ is proportional to the expectation value $\langle (a_l^\dagger a_l )^2 \rangle$, which reflects the fact that a site that is unoccupied will not give rise to jumps, and for particle numbers greater than one, the rate is enhanced by superradiance \cite{Lehmberg1970,Lehmberg1970a}. We will see below that this enhancement is related to the excitation of a collection of atoms in one site being symmetric.
The localisation of an atom on a single site corresponds to a spreading of the localised particle over all of the possible states in quasimomentum space, and hence to an increase in kinetic energy. This can be seen in Fig.~\ref{fig:hardcoreexample}, where we propagate states in time numerically using quantum trajectories. As for the previous case of the optical Bloch equations, we show two example trajectories as well as a trajectory average, here for the kinetic energy in the system as a function of time. By direct computation of expectation values from the master equation, we see that
\begin{equation}
\frac{d}{dt}E_{\rm bos} \equiv \frac{d}{dt}\langle H_{\rm bos} \rangle = - \Gamma \langle H_{\rm bos} \rangle , \label{hardcorebosonexactke}
\end{equation}
so that $E_{\rm bos}(t) = E_{\rm bos}(t=0) \exp (-\Gamma t)$. In the right hand panel of Fig.~\ref{fig:hardcoreexample} we compare this analytical result to the numerical value as a function of time, and observe very good agreement to within the estimated statistical error.
\begin{figure}
\begin{center}
\includegraphics[width=13cm]{figures/hardcorerelativeerror}
\end{center}
\caption{Statistical errors in the quantum trajectories computation for hard-core bosons, taken from exact diagonalisation calculations with 5 particles on 10 lattice sites. Analogously to the two-level system case, we compare the statistical error estimate for our estimate of the energy $E_{\rm bos}$ (dark blue solid line) to the absolute error in the averaged value of $E_{\rm bos}$ when compared with the exactly computed value (dark red dashed line) at a time $tJ=8$. We also show the estimate of the error in a local correlation function $\langle (a^\dagger_5 a_6+a^\dagger_6a_5) \rangle$ (light green solid line) compared with the discrepancy between this quantity and the exactly value (light green dashed line) at the same point in time. We show each of these these values as a function of the number of trajectories over which we average, with all quantities plotted on a logarithmic scale. The statistical errors again decrease steadily $\propto 1/\sqrt{N}$ for larger numbers, but the relative error in the local quantity remains larger than the global quantity. } \label{fig:hardcoreerror}
\end{figure}
We investigate this statistical error as a function of the number of trajectories in Fig.~\ref{fig:hardcoreerror}. We fix the time at $tJ=8$, and plot the statistical error and the discrepancy between the value of $E_{\rm bos}$ calculated from the trajectory average and the exact value obtained from the master equation as a function of $N_{\rm traj}$. The solid blue line represents the statistical error, and the dashed red line the discrepancy between the trajectory average and the exactly computed value. From the double logarithmic scale we see the scaling $\propto 1/\sqrt{N}$. As in the example of the optical Bloch equations, we expect the means over different samples of trajectories to be approximately normally distributed, and therefore $68.2\%$ of all possible samples should have discrepancies to exact values that fall below the estimated statistical error. The values shown here are again consistent with that analysis. In addition to this data, we also show using lighter (green) solid and dashed lines the same analysis for a local correlation function $\langle (a^\dagger_5 a_6+a^\dagger_6 a_5) \rangle$. We see that the relative error is higher than for the globally averaged value, but the behaviour of the local correlation function with the number of trajectories is very similar.
\begin{figure}
\begin{center}
\includegraphics[width=6.67cm]{figures/hardcoredens5}
\includegraphics[width=6.93cm]{figures/hardcoremeandens5}
\end{center}
\caption{Illustrative example of quantum trajectories averaging for a gas of hard-core bosons on a lattice, showing results from exact diagonalisation calculations with 5 particles on 10 lattice sites, as in Fig.~\ref{fig:hardcoreexample}. (left) Density on site 5 of the lattice as a function of time $tJ$ for two example trajectories (with blue solid and red dashed lines showing different random samples). We see the effect of quantum jumps, where nearby jumps give rise to fluctuations either up or down of the local mean density. (right) Values for $\langle n_5\rangle$ averaged over 1000 sample trajectories, and compared with the exact result, which is $\langle n_5 \rangle\equiv \langle a_5^\dagger a_5 \rangle=0.5$, because the equations of motion are homogeneous. We see that the deviation from this value is well approximated by the statistical errors, shown here as error bars calculated as described in section \ref{sec:staterror}. } \label{fig:hardcoredensity}
\end{figure}
In Fig.~\ref{fig:hardcoredensity} we show additional data for another local correlation function, namely the on-site density. This is quite instructive, because it shows how individual trajectories need not exactly enforce symmetries that are present in the master equation as a whole. Specifically, for a system with periodic boundary conditions and an initial density that is uniform across the system, we expect that
\begin{equation}
\frac{d}{dt}\langle a^\dagger_l a_l \rangle=0.
\end{equation}
However, because the individual jumps are local, the spatial invariance in the expectation value of the density is not followed by individual trajectories, as we see clearly in the left panel of Fig.~\ref{fig:hardcoredensity}. This reflects an analogous lack of spatial invariance that might occur in specific runs of an experiment if spontaneous emission events are taking place locally (see appendix \ref{sec:continuousmeasurement} for an interpretation of this in terms of physical processes occurring on individual lattice sites). In the right panel of Fig.~\ref{fig:hardcoredensity}, we see that the expectation value of the density remains constant to within the statistical errors, as is expected.
In Sec.~\ref{sec:lightscattering} we return to this problem of spontaneous emissions, outlining the derivation of this master equation for many bosons in an optical lattice, also away from the approximations that are made here. We also explain the additional dynamics (including transfer of particles to higher Bloch bands) that are exhibited in that case, and we explain the limits in which the simplified master equation treated here arises.
\section{Integration of quantum trajectories with many-body numerical methods}
\label{sec:numerics}
Over the last few years, several groups have begun applying these techniques to
describe dissipative dynamics of many-body systems as they arise in AMO systems. A number of examples of such master equations will be given in section \ref{sec:amosys}, beginning with a general version of the light scattering master equation discussed in section \ref{sec:hardcorebosons}. In order to facilitate the solution of the corresponding master equations, quantum trajectories techniques have been combined with many-body methods. In this section we give an overview of what has been done so far, beginning with exact diagonalisation and mean-field methods, and then describing in detail the integration of quantum trajectories with time-dependent density matrix renormalisation group (t-DMRG) techniques.
\subsection{Integration with exact diagonalisation}
\label{sec:ed}
For systems that are not too large, we can store the many-body states exactly, and apply quantum trajectories methods in much the same way as they would be applied to single-particle systems. We only need a method to propagate a state under the effective Hamiltonian $H_{\rm eff}$, and apply jump operators. The main difference in applying this method to many-body systems, as opposed to single-particle systems, is that quantum trajectories can be an efficient way to compute quantities that are local in space or momentum in these systems, because the Hilbert space is large compared with the system size. This was discussed in more detail in section \ref{sec:globallocal} above. In most quantum trajectories involving t-DMRG techniques, the dynamics are first studied for a small system that is tractable to exact diagonalisation (see, e.g., \cite{Daley2009,Kantian2009}). In the case of the Bose-Hubbard model eq.~\eqref{HBH}, such systems are of the order of 10 particles on 10 lattice sites for rapid computations, although recent calculations with the Bose-Hubbard model (without quantum trajectories) have extended to 14 particles on 14 sites \cite{Lux2013,Bauer2011}.
One important optimisation that quantum trajectories can allow involves conserved symmetries of the effective Hamiltonian. Specifically, if the Hamiltonian commutes with a unitary transformation $\hat T$, $[H_{\rm eff},\hat T]=0$, it is usually possible to optimise a calculation by working in one symmetry sector of $\hat T$, reducing the number of basis states that must be used for the Hilbert space. In the case that $[c_m,\hat T]\neq 0$, this is not typically the case for a full master equation simulation. However, in a trajectories calculation, if $[c_m , \hat T] \propto c_m$, it is still possible to make use of the symmetry in reducing the relevant basis size for the Hilbert space, because the application of a jump simply switches the state from one symmetry sector to another. A simple example is the case of particle loss in a many-body system \cite{Daley2009} (see section \ref{amosys} below). Often, the effective Hamiltonian commutes with the total particle number operator, and we can work in a basis with fixed particle number. Loss events reduce this number by a well controlled amount, allowing a different number conserving basis to be used for propagation after the jump \cite{Daley2009}.
\subsection{Integration with the time-dependent Gutzwiller ansatz}
\label{sec:gutzwiller}
This discussion applies equally to mean-field methods that make it possible to rewrite the dynamics so that it is exactly computable. An important example of this for bosons on a lattice is the possibility to perform time-dependent calculations with a Gutzwiller ansatz \cite{Snoek2007,Wernsdorfer2010,Zakrzewski2005,Jaksch1998}. In calculating ground states of the Bose-Hubbard Hamiltonian, this ansatz takes the form of a product state over different lattice sites,
\begin{equation}
\ket{\psi_{G}}=\prod_{l}\ket{\psi_G^{l}}_{[l]}=\prod_l\sum_{n}f_n^{(l)}\ket{n}_{[l]}. \label{eqgutzwiller}
\end{equation}
Here, $\ket{n}_{[l]}$ is a state of $n$ particles locally on site $l$, and $f_n^{(l)}$ can be used as variational parameters to minimise the energy. For the Bose-Hubbard model $H_{BH}$, this is equivalent to using a mean-field approximation on the operators $a_l$ in the tunnelling part of the Hamiltonian, $a_i^\dagger a_j \rightarrow a_i^\dagger \langle a_j \rangle + \langle a_i^\dagger \rangle a_j - \langle a_i^\dagger \rangle \langle a_j \rangle$, which leads to a Gutzwiller mean-field Hamiltonian that can be expressed as a sum over terms involving only local operators on site $l$,
\begin{equation}
H_{MF}=\sum_l\left[-J\left({\psi_l a_i^{\dag}+\psi_l^{\ast}a_l}\right)+\frac{1}{2} U(a_l^{\dag})^2a_l^2\right].
\end{equation}
Here, $\psi_l\equiv \sum_{j|l}\langle \psi_G^j |a_j|\psi_G^j\rangle$ involves a sum over the sites $j$ that are immediate neighbours of site $l$, and in variational calculations these parameters must be found self-consistently. As a method of determining the ground state, this method faithfully reproduces states for weak interactions \cite{Rokhsar1991}, and is exact in the limit of infinite dimensions\cite{Byczuk2008,Hubener2009,Anders2011}. In this form, the superfluid phase of the Bose-Hubbard model, which is characterised by off-diagonal long-range order \cite{Sachdev2011}, is represented by a nonzero value of $\psi_l$, whereas for the Mott Insulator state, $\psi_l=0$. For unit filling on the lattice, this transition from the superfluid phase to the Mott Insulator phase in this ansatz occurs at $U/(zJ)\approx5.8$ \cite{Sachdev2011}, where $z$ is the number of neighbouring sites. This value is accurate for a 3D cubic systems to about 15\%, when compared with results from Quantum Monte Carlo calculations \cite{Pollet2013}.
These methods can be directly generalised to time-dependent calculations \cite{Jaksch1998}, and have been seen to give reasonable qualitative results in a number of contexts \cite{Snoek2007,Wernsdorfer2010,Zakrzewski2005}. Because these calculations involve direct evolution of on-site wavefunctions under the mean-field Hamiltonian, they are efficient to compute numerically. Care must be taken, as there are some cases where the method clearly fails - for example, no dynamics can be generated when the initial state is a Mott Insulator state, as in this case the terms in $H_{MF}$ that couple sites are exactly zero. This also leads to artefacts and unphysical slowing down in studying time-dependent transitions from the superfluid to Mott Insulator phase. However, these methods can be very useful in developing understanding of systems in a tractable manner.
Because dynamics are efficiently computable, these methods can be used directly together with quantum trajectory techniques. The averages are then performed over an ensemble of stochastically evolved states, where each state of the ensemble has the Gutzwiller ansatz form of eq.~\eqref{eqgutzwiller}.
An alternative Gutzwiller ansatz form for solving a master equation was used by Diehl et al. in Ref.~\cite{Diehl2010c}. This involved a factorisation of the system density operator,
\begin{equation}
\rho_G=\bigotimes_i\rho_i\;\; \rho_l={\rm Tr}_{i\neq l}\{\rho\}
\end{equation}
which gives rise to a nonlinear set of coupled master equations, each for one site $l$. This ansatz is particularly useful in certain cases where it can be treated analytically \cite{Diehl2010c}, and can also be useful as a numerical method to obtain certain qualitative effects (see, e.g., \cite{Pichler2010}, where this method was used to study transfer of particles to higher bands of an optical lattice in spontaneous emission events).
It is important to note that where sampling an ensemble of trajectories in the Gutzwiller ansatz form $|\psi_G\rangle$ can actually contain somewhat more information than $\rho_G$. While neither ansatz can represent quantum entanglement between different sites, by making a product state of local density operators, $\rho_G$ also does not capture the development of non-trivial classical correlations between different parts of the system. However, classical correlations can be captured by an ensemble of trajectories in the form $|\psi_G\rangle$ (each trajectory has no classical correlations, but the ensemble of trajectories can represent these). This can be important not only in treating dissipative dynamics with quantum mechanical reservoirs, but also when sampling over classical noise within such an ansatz, as was observed in Ref.~\cite{Pichler2013}.
\subsection{Time-dependent density matrix renormalization group methods}
An important development in treating dissipative dynamics of many-body systems has been the advent of time-dependent methods for dealing with 1D many-body systems. We will briefly summarise the key features of these methods in this section, and then discuss their integration with quantum trajectories methods\footnote{For a more detailed review of these methods, see, e.g., Ref.~\cite{Schollwock2011}.}.
Over the last ten years, a range of methods based on matrix product states \cite{Schollwock2011} have been developed, which effectively build on the success of density matrix renormalization group (DMRG) methods \cite{Schollwock2005,White1992} as the most powerful numerical method for generic lattice models in one dimension. Early development of time-dependent methods has included the time-evolving block decimation (TEBD) algorithm \cite{Vidal2003,Vidal2004}, and its integration with DMRG to produce an adaptive time-dependent DMRG (t-DMRG) algorithm \cite{White2004,Daley2004}. Similar methods based on matrix product operators have been developed \cite{McCulloch2007,Verstraete2008,Crosswhite2008,Pirvu2010,Verstraete2004}, which allow for both the study of time-dependent dynamics with long-range interactions \cite{Crosswhite2008}, and the direct study of dissipative dynamics described by a master equation, with matrix product states used to represent a density matrix \cite{Verstraete2004,Zwolak2004}. Such density matrix representations can be used to represent finite-temperature states \cite{Verstraete2004,Verstraete2008}, though finite temperature dynamics alone (without dissipative processes) can also be represented by use of ancilla states \cite{Feiguin2005}, or by sampling minimally entangled states \cite{Stoudenmire2010}. Generalisations have been proposed to higher dimensions \cite{Verstraete2004b,Vidal2008}, although these are typically very computationally intensive.
\begin{figure}
\begin{center}
\includegraphics[width=\textwidth]{figures/mps}
\end{center}
\caption{Diagram comparing a product state and a matrix product state. Each square represents a state $\ket{\psi_l}_{[l]}$ in a local Hilbert space $\mathcal{H}_l$, e.g., a single spin or lattice site, and the vertical leg on each square box indicates an index for the states of the local Hilbert space. The Hilbert space representing the whole system is the tensor product $\mathcal{H}=\mathcal{H}_1 \otimes \mathcal{H}_2 \otimes \ldots \otimes \mathcal{H}_L$. A product state can be written as $\ket{\psi_1}_{[1]}\otimes \ket{\psi_2}_{[2]} \otimes \ldots \otimes \ket{\psi_L}_{[L]}$, and represents a quantum state with no entanglement between any two local Hilbert spaces. By multiplying out $D\times D$ matrices, a matrix product state makes it possible to represent superpositions of product states, and this superposition contains entanglement between the local Hilbert spaces. For sufficiently large $D$, any state can be represented in this way, and the ground states of large classes of local 1D many-body Hamiltonians can be represented faithfully for small $D$ in this fashion. This is the basis of the time-dependent density matrix renormalization group method, as discussed in the main text. The maximum amount of entanglement we can represent across any bipartite splitting in the system is bounded in terms of the von Neumann entropy by $S_{\rm vN}\leq \log_2 D$. Note that while each square box here is a similar notation to that used in Ref.~\cite{Schollwock2011} and similar references, in the notation of those references, contractions between matrices are represented by connecting boxes with horizontal lines.} \label{fig:mps}
\end{figure}
Matrix product state methods can be applied to any system for which we can write the Hilbert space as a product of local Hilbert spaces, $\mathcal{H}=\mathcal{H}_1\otimes \mathcal{H}_2\otimes \cdots \otimes \mathcal{H}_L$. This works well for spin chains, where each local Hilbert space $\mathcal{H}_i$ corresponds to the states of a single spin, as well as for bosons or fermions on a lattice, where each local Hilbert space corresponds to the different possible occupations of particles on a single lattice site. If we write the $d$ basis states of the local space $\mathcal{H}_l$ as $|i_l\rangle_{[l]}$, $i_l=1\ldots d$, then any state $\ket{\psi}$ of the complete system can be expressed in the form
\begin{equation}
|\psi\rangle=\sum_{i_1,i_2,\ldots,i_L = 1}^d c_{i_1i_2\ldots i_L} |i_1\rangle_{[1]} \otimes |i_2\rangle_{[2]} \otimes \ldots.
\otimes |i_L\rangle_{[L]}. \label{mpssystemstate}
\end{equation}
The key to matrix product state methods is a decomposition of the state in which we write $c_{i_1i_2\ldots i_L}$ from this equation as the multiplication of a series of matrices $A_l [i_l]$,
\begin{equation}
c_{i_1i_2\ldots i_L}=A_1 [i_1] \, A_2[i_2] \ldots A_L [i_L] \label{mpseq},
\end{equation}
where the boundary matrices are assumed to be one-dimensional in order to produce coefficients from the matrix multiplication.
This state is represented in the diagram of Fig.~\ref{fig:mps}, where we visualise the basic concept of a matrix product state. If the matrices were just complex coefficients (i.e., matrices of dimension $D=1$), then this state would be precisely the same as the Gutzwiller ansatz state eq.~\eqref{eqgutzwiller} described in the previous section. It would be a product state, with a simple physical interpretation, and no entanglement between any of the local Hilbert spaces. However, by using matrices of dimension $D\times D$, where $D>1$, we can produce a superposition of different product states, in so doing, allowing entanglement between different local Hilbert spaces. For sufficiently large $D$, any state can be represented in this form, but in general $D$ must grow exponentially with the system size in order to represent an arbitrary state of the Hilbert space $D\propto \exp(L)$. In fact, to represent an arbitrary state of the form given in eq.~\eqref{mpssystemstate}, and $L$ is even, then we must have matrices at the centre of the system of dimension $D=d^{L/2}$.
The key to the success of these methods comes in the understanding of entanglement in many-body systems, which has developed extensively in the past few years \cite{Eisert2010,Amico2008,Calabrese2009,Calabrese2009b,Peschel2009,Cardy2010}. For a system where the total state of the system is a pure state $\psi$, the entanglement between two parts $A$ and $B$ can be quantified by computing the von Neumann entropy $S_{\rm vN}[\rho_A]=-{\rm Tr}[\rho_A \log_2 \rho_A]$ of the reduced density operator for one part of the system $\rho_A={\rm Tr}_B [|\psi\rangle\langle \psi |]$. If the two parts are in a product state, then the reduced density operator $\rho_A$ will represent a pure state, and we have $S_{\rm vN}[\rho_A]=0$. However, if the parts of the system are entangled, $\rho_A$ will represent a mixed state, and its entropy gives the quantification of the amount of entanglement between $A$ and $B$.
In a matrix product state with matrix dimension $D$, the maximum entanglement across any possible bipartite splitting in the system that can be represented by this state is $S_{\rm vN}=\log_2 (D)$. However, if we take a bipartite splitting of a gapped lattice model or spin chain in 1D, then we expect $S_{\rm vN}$ to be bounded as a function of the size of each part of the system $A$ and $B$, even as we extrapolate to unbounded block size in an infinite-sized system \cite{Eisert2010,Amico2008,Calabrese2009}. If, on the other hand, we have a 1D system that is critical, then we expect $S_{\rm vN}\propto \log(L_s)$, where $L_s$ is the size of the smaller block, $A$ or $B$ \cite{Eisert2010,Calabrese2009}. In either case, the entanglement is in some sense weak, and for finite-size systems it is reasonable to expect that ``slightly-entangled'' matrix product states will represent the states well. While the scaling of the von Neumann entropy is a practical guide, and in fact the scaling of R\'{e}nyi entropies $S_R\equiv (1-R)^{-1} {\rm Tr} [\rho^R]$ of order $R<1$ should be considered in order to prove that a state can be represented in this form \cite{Schuch2008,Schuch2008b}, it is found that these methods work very well under a wide range of circumstances. In fact, it has been shown \cite{Verstraete2006} that ground states of 1D spin systems are expected to be represented faithfully in this form, even at criticality, and the success of DMRG as an essentially exact method for the computation of ground states of 1D systems is practical confirmation of this \cite{Schollwock2011,Schollwock2005}.
Time-dependent methods can be readily developed, as it is straight-forward to apply either local operators or even long-range operators that are representable in matrix operator form to these states. In particular, applying an operator to two neighbouring sites can be done while only updating the matrices associated with those sites. Then, a simple method to compute time evolution for a local Hamiltonian $\hat H=\sum_{l} \hat H_{l,l+1}$ is to take a Trotter decomposition of the time evolution operator,
\begin{equation}
{\rm e}^{-{\rm i} \hat H (2\delta t) /\hbar}={\rm e}^{-{\rm i} \sum_l \hat H_{l, l+1} \delta t /\hbar}\approx \prod_{l=1}^{L-1} {\rm e}^{-{\rm i} \hat H_{l,l+1} \delta t /\hbar} \prod_{l=L-1}^{1}{\rm e}^{-{\rm i} \hat H_{l,l+1} \delta t /\hbar}+\mathcal{O}(\delta t^3),\label{trotter}
\end{equation}
where each of these operators can be applied to the state by updating only the matrices associated with those sites. This can be straight-forwardly generalised to higher-order Trotter decompositions \cite{Sornborger1999}, and a variety of other methods for efficiently propagating the state, especially Arnoldi methods \cite{Garcia-Ripoll2006,McCulloch2007} are also available. In each case, the numerical cost of propagating the state scales as $D^3$.
\begin{figure}
\begin{center}
\includegraphics[width=7cm]{figures/quasiparticlediagram}
\end{center}
\caption{After a local or global quench in a 1D system with short-range interactions, quasiparticle excitations are produced that propagate through the system. The rate of information propagation in the system is restricted by the Lieb-Robinson velocity. When we consider a bipartite splitting of the system (here into parts A and B), propagation of quasiparticles across the boundary leads to an increase in spatial entanglement between parts A and B of the system.} \label{fig:quasiparticlediagram}
\end{figure}
For how long a state can be accurately propagated in time using these methods depends on the rate of growth of entanglement during the time-evolution. For cases where the evolution is somewhat close to equilibrium, dynamics on very long timescales can be computed. For example, near-adiabatic state preparation has been studied in Hubbard models on timescales longer than $2000 J^{-1}$ \cite{Kantian2010}. However, general time evolution under a Hamiltonian for which the matrix product state is not a near-equilibrium state can only be computed over short timescales \cite{Gobert2005,Prosen2007,Schuch2008,Schuch2008b,Trotzky2012}. This is directly related to the growth of entanglement after a quantum quench \cite{Calabrese2005,Prosen2007,Lauchli2008,Unanyan2010,Gobert2005}. A simple picture for how this works, which can be clearly seen in the case of nearest-neighbour transverse Ising model \cite{Calabrese2005} is depicted in Fig.~\ref{fig:quasiparticlediagram}. There, we show the propagation in time of quasiparticle excitations, which carry information at a rate that is limited by the Lieb-Robinson velocity \cite{Lieb1972}. Propagation of quasiparticles across a boundary between two parts of the system leads to an increase in spatial entanglement between those parts of the system. In the generic case \cite{Calabrese2005,Schuch2008,Schuch2008b}, this leads to a linear growth in the entanglement entropy as a function of time. This has been investigated in a variety of different models, including the Bose-Hubbard model \cite{Lauchli2008,Daley2012}.
\begin{figure}
\begin{center}
\includegraphics[width=8cm]{figures/vnegrowth}
\end{center}
\caption{Increase in entanglement as a function of time after a quantum quench in the Bose-Hubbard model. Here we consider a quench in the Bose-Hubbard model eq.~\eqref{HBH} from $U=10J$ to $U=J$ in a system with 30 particles on 30 lattice sites, initially in the ground state with $U=10J$. Making a bipartite splitting in the centre of the system (with 15 sites on each side), we compute the von Neumann entropy increase as a function of time using t-DMRG methods with varying values of $D$. We see clearly the bounds in $S_{\rm vN}$ imposed by different values of $D$, which are indicated with dotted horizontal lines, as well as clear convergence of the results at short times to the exact value as we increase $D$. Adapted from A.~J.~Daley, H.~Pichler, J.~Schachenmayer, and P.~Zoller, Phys. Rev. Lett. {\bf 109}, 020505 (2012).} \label{fig:vnegrowth}
\end{figure}
As an example of this, in Fig.~\ref{fig:vnegrowth}, we plot the entanglement across a bipartite splitting the middle of the system as a function of time after a quench in the Bose-Hubbard model. The quench is from a Mott Insulator state in 1D at unit filling, with an initial ground state at $U/J=10$, and a quench to $U/J=1$. The linear growth of the entanglement entropy is clearly seen, as is the ability of t-DMRG methods to capture the quench dynamics only so long as the entanglement remains less than $\log_2 D$. Because we require a matrix size $D> 2^{S_{\rm vN}}$ to represent the entanglement in the system, and because the numerical cost scales as $D^3$, we see that the cost of the algorithm grows exponentially in order to simulate this further in time. For reference, in this computation, a computation on a single CPU core required several days for $D=512$. It has been recently discussed that measurement of such entanglement growth in experiments could be a good way to characterise the behaviour of a quantum simulator in a regime that goes beyond what is currently computable on classical computers \cite{Daley2012,Cardy2011,Moura-Alves2004,Palmer2005,Guhne2009,Abanin2012,Schachenmayer2013}.
In practice, the practical procedure employed when using t-DMRG methods is to fix an allowed truncation error value $\epsilon_T$ for the propagation of the state in each time step. We then ensure that $D$ is chosen to be large enough so that this error is small in each time step, in an analogous manner to the control of other such errors in numerical computations. The error $\epsilon_T$ we define by considering the exact state $|\Psi\rangle$, that would be produced via application of the time evolution operator exactly on the matrix product state we had at the beginning of the time step, and comparing this with the nearest matrix product state representation with dimension $D$ to this state, $|\Psi_{{\rm MPS},D}\rangle$. Specifically, we define $|\langle \Psi |\Psi_{{\rm MPS},D}\rangle |^2\equiv1-\epsilon_T$, where we compute the square of the inner product between $|\Psi\rangle$ and $|\Psi_{{\rm MPS},D}\rangle$. We then require that $1-\epsilon_T\approx 1$, and we can perform systematic convergence tests in $\epsilon_T$, or in some cases where it is well controlled, even extrapolation to $\epsilon_T \rightarrow 0$. This naturally involves performing calculations with increasing matrix size $D$.
On a practical level, the size of the the matrix $D$ depends on the particular system being investigated, and the length of time over which the dynamics are computed. Typical calculations often involve several hundred to several thousand time steps, with $D\sim$ 200--2000. A calculation with a small local Hilbert space of 2--5 states and $D\sim500$ takes of the order of a day on a single Intel Ivy Bridge processor core. A significant computational advantage of quantum trajectory techniques is that with access to a large computational cluster, the methods can be immediately parallelised as each trajectory can be run independently.
\subsection{Integration of quantum trajectories and t-DMRG methods}
Integration of quantum trajectories with t-DMRG methods is somewhat straight-forward in so far as t-DMRG provides a direct means to propagate the state under the effective Hamiltonian $H_{\rm eff}$, or to apply jump operators $c_m$ to the state. Initial states can be directly constructed or computed if they are pure states, or known mixtures of pure states, and finite temperature states could also potentially be sampled as minimally entangled thermal states \cite{Stoudenmire2010}. We can then implement the protocol from section \ref{sec:higherorder} directly within this formalism \cite{Daley2009}. In computing the norm of the state as a function of time, it is usually best to ensure that the state stored in a matrix product state is always normalised. If a Trotter decomposition is used to compute the time evolution, for example, it is often useful to renormalize the state stored in memory after every application of local evolution operators \eqref{trotter} to the state. The required normalisation factors can be stored separately in memory and used to compute the norm of the state for the purposes of the protocol in \ref{sec:higherorder}.
A key test for these out of equilibrium calculations is to perform convergence tests in the truncation error $\epsilon_{T}$ \emph{for each individual trajectory}. It is typically sufficient to perform convergence tests for a small representative sample of the trajectories, because the behaviour of the entanglement growth in the system and the related growth in $\epsilon_T$ generally depends on the type of dynamics being induced by the quantum jumps. Care should be taken with this, because jumps effectively produce quantum quenches (often local quantum quenches \cite{Schachenmayer2014}), which can lead to an increase in entanglement.
\begin{figure}
\begin{center}
\includegraphics[width=10cm]{figures/egrowthtraj}\end{center}
\caption{Six example trajectories showing time-dependent DMRG simulations of the dynamics of 48 bosons on 48 lattice sites in a master equation with $H=H_{BH}$ eq.~\eqref{HBH}, and jump operators $c_m=a_m^\dagger a_m$, and $\gamma=0.02J$. Here we investigate the increase in the difficulty of the simulation as a function of time by plotting the von Neumann entropy across a splitting in the centre of the system (with 24 sites on each side). We see that there are different delays to the increase in $S_{\rm vN}$, which depend on the timing of quantum jumps as well as their location with respect to the point of the bipartite splitting. All of these trajectories are exact calculations (t-DMRG truncation errors $\epsilon_T$ lead to errors in $S_{\rm vN}$ that are smaller than the line thickness), with $D=256$.} \label{fig:spementropy}
\end{figure}
As an example of the entanglement growth after quantum jumps, in Fig.~\ref{fig:spementropy}, we plot example values of the von Neumann entropy in the centre of a system as a function of time for bosons on an optical lattice with light scattering in the form discussed in section \ref{sec:hardcorebosons}. We show six trajectories that are representative of relatively typical behaviour, and we can see how the difficulty of the simulation increases with time. Note that there are different delays to the increase in $S_{\rm vN}$, which depend on both the timing of quantum jumps and also their location with respect to the point of the bipartite splitting. It should be noted that for local jumps, it can be favourable numerically to use different matrix dimensions $D$ for different matrices along the chain, in order to keep a fixed value of $\epsilon_T$. This is especially straight-forward when using a Trotter decomposition eq.~\eqref{trotter} and applying local time evolution operators. In general, jump operators that are local in space, and so provide a \emph{local quench} for the system should lead to a slower entanglement growth than global quenches \cite{Calabrese2007}, which should favour the use of quantum trajectories methods to describe the resulting dynamics.
As with the case of exact diagonalisation methods, matrix product state methods can be substantially optimised by restricting the values that are stored to particular symmetry sectors. This was long used as a means of optimising states in DMRG \cite{Schollwock2005}. Because of this, the added numerical advantages of quantum trajectories over density matrix propagation arising from additional symmetries present in the evolution of individual trajectories apply to t-DMRG methods in the same was as was discussed in section \ref{sec:ed} for exact diagonalisation methods. In Ref.~\cite{Daley2009}, calculations were performed for master equations describing three-body loss using quantum trajectories methods. In that case, the calculation of each trajectory was optimised by making use of a U(1) symmetry corresponding to total particle number conservation, because at each point in the evolution, each trajectory can have a fixed total particle number, even though this can change in time. Note that optimisation to such symmetries also provides a good choice when there is a freedom in choosing the jump operators in the way described in section \ref{sec:physicalinterp}.
As mentioned above, there is also an alternative way to compute dynamics of open quantum systems (as described, e.g., by a master equation for the density matrix), by using matrix product state to encode a density matrix directly \cite{Verstraete2004,Zwolak2004}. We can, e.g., replace states by density matrices and operators by super-operators. The dimension of the local Hilbert space then increases from $d$ to $d^2$. Whether such representations or quantum trajectories methods using stochastically propagated pure states is more efficient depends strongly on the particular problem being solved, and is still an area of ongoing research. In comparing the two methods it is important to make several general remarks.
Firstly, the most clear trade-off for the density matrix approach vs. quantum trajectories is that in the density matrix approach the local dimension of the Hilbert space must be increased (giving rise to a computational cost increase that is usually of the order of $d^3$) whereas in the quantum trajectories case it is important to sample sufficient numbers of trajectories in order to obtain accurate statistics for the quantities that are being computed. Given that the number of trajectories required for computation of global few-body correlation functions with a relative accuracy of a few percent tends to be of the order of several hundred, the density matrix approach could have advantages, especially in the case that the local Hilbert space dimension is small (e.g., $d=2$ for a spin-1/2 chain). This advantage is much less clear for lattice models, especially Bose-Hubbard models with potentially large on-site occupation, or ladder systems. For the Fermi Hubbard model on a two-leg ladder, for example, it is typical to take $d=16$.
Secondly, and perhaps more importantly, the matrix size $D$ required to solve the problem will depend strongly on the details of the problem. For a density matrix representation, the matrix dimension $D$ has to be increased in order to allow for classical correlations to be represented as well as quantum entanglement. This is directly analogous to the mathematics leading to a lack of classical correlations in the Gutzwiller ansatz-type density matrix representation presented in section \ref{sec:gutzwiller}. Often, this will mean that $D$ will have to be much larger for an accurate representation of the state in a density matrix form directly than would be required for states propagated in a quantum trajectories formalism. There is one important caveat for this: If the state becomes substantially mixed, then by propagating pure states, we can end up having more quantum entanglement represented than would be required for a mixed state representation of the same dynamics. One avenue of research for the future is the question of how to optimise the choice of jump operators (as discussed in sec.~{sec:physicalinterp}) in order to produce pure states in a quantum trajectories formulation that would minimise the amount of entanglement generated in time.
As a general rule, short-time dissipative dynamics that begin with a pure state are straight-forwardly described by integrating t-DMRG with quantum trajectories \cite{Daley2009,Kantian2009,Barmettler2011}. Several examples of the use of quantum trajectories integrated with t-DMRG to describe open many-body systems in an AMO context are described in the next section.
\section{Open many-body AMO systems}
\label{sec:amosys}\label{amosys}
As we discussed above, the study of strongly interacting AMO systems has made it possible to realise open many-body quantum systems in which the usual quantum optics approximations apply. This makes it possible to have microscopically well-controlled models for many-body systems in an analogous form to the single-particle models of quantum optics. In this section, we give several examples, focusing initially on cold quantum gases in optical lattices \cite{Jaksch1998,Jaksch2005,Lewenstein2007,Bloch2008b,Lewenstein2012}, where we begin with a discussion of the corresponding microscopic models before presenting an overview of dissipation due to light scattering and due to particle losses. We then discuss dissipative state preparation in this system, as well as corresponding experiments with other experimental systems, including trapped ions, Bose-Einstein condensates in an optical cavity, and Rydberg atoms.
\subsection{Microscopic description for cold quantum gases and quantum simulation}
\label{sec:qsim}
A key starting point for a lot of recent progress with ultracold atomic gases is the possibility to realise microscopically well controlled models for strongly interacting systems, and to manipulate the parameters of those models via external fields. This comes about because the atomic physics associated with magnetic and optical trapping potentials is very well understood \cite{Pethick2002,Metcalf1999}, and for the experimentally relevant limit of dilute gases, inter-particle interactions are dominated by two-particle interactions. Moreover, because the collision energies of atoms are very small, the scattering (which typically involves complicated interatomic potentials with many bound states) reduces to a universal regime where the behaviour can be described by a single parameter, the s-wave scattering length $a_{\rm scatt}$ \cite{Dalibard1998}\cite{Castin2001}. Moreover, when $a_{\rm scatt}$ is small compared with typical scattering momentum scales, we can make a Born approximation for the scattering. For cold bosonic atoms, we then write the well-controlled microscopic Hamiltonian in terms of bosonic field operators\footnote{Although atoms consist of protons, neutrons, and electrons, a tightly bound atom at low energy can be assumed to maintain bosonic or fermionic character provided short-distance physics can be eliminated, and the typical separation of atoms is much larger than their size.} $\hat{\psi}(\mathbf{r})$ that obey $\left[ \hat{\psi}(\mathbf{x}), \hat{\psi}^\dagger(\mathbf{y})\right]=\delta(\mathbf{x}-\mathbf{y})$ as:
\begin{equation}
\hat{H}_{\rm atoms}\approx\int d^{3}r\,\hat{\psi}^{\dag}(\mathbf{r})\,\left[-\frac{\hbar^{2}}{2m}\nabla^{2}+V_{0}(\mathbf{r})\right]\,\hat{\psi}(\mathbf{r})+\frac{g}{2}\int d^{3}r\,\,\,\hat{\psi}^{\dag}(\mathbf{r})\,\hat{\psi}^{\dag}(\mathbf{r})\,\,\hat{\psi}(\mathbf{r})\,\hat{\psi}(\mathbf{r}), \label{hatoms}
\end{equation}
where $m$ is the atomic mass, $g=4\pi\hbar^2a_{\rm scatt}/m$, $V_{0}(\mathbf{r})$ is the effective single-particle potential (generated by external magnetic fields and optical traps).
Not only is this model well-understood microscopically, but experiments offer precise time-dependent control, both over $V_{0}$, through external magnetic fields and optical traps based on the AC-Stark Shift \cite{Pethick2002,Metcalf1999}, and over $a_{\rm scatt}$, through optical and magnetic Feshbach resonances \cite{Chin2010}. While the gas described here is in principle weakly interacting, Feshbach resonances can be used to increase the interaction strength towards a strongly interacting regime. Alternatively, even while maintaining values of $a_{\rm scatt} \sim 100 a_0$, where $a_0$ is the Bohr radius, we can achieve a strongly interacting system by loading atoms into an optical lattice potential, formed via the AC-Stark shift with standing waves of laser light. Because the temperature and energy scales can be straight-forwardly made much smaller than the bandgap in an optical lattice potential \cite{Jaksch1998}, this makes it possible to directly realise simple lattice models for atoms in a single band of an optical lattice, such as the Bose-Hubbard model eq.~\eqref{HBH}, or extended versions of this that allow for additional confinement and/or superlattice potentials in the form
\[
{H_{\rm BH+}}=-J\sum_{\langle j,l\rangle}{a}_{j}^{\dag}{a}_{l}+\frac{U}{2}\sum_{l}{a}_{l}^{\dag}{a}_{l}({a}_{l}^{\dag}{a}_{l}-1)+\sum_{l}\epsilon_{l}{a}_{l}^{\dag}{a}_{l},\]
where as before $\langle j,l\rangle$ denotes a sum over all combinations of neighbouring
sites, and $\epsilon_{i}$ is the local energy offset of each site. In order to obtain the Bose-Hubbard model, we expand the field operator $\hat{\psi}(\mathbf{r})$ in terms of Wannier function modes for the lowest Bloch Band centred at each site location $\mathbf{r_l}$, $w_0(\mathbf{r}-\mathbf{r}_l)$ \cite{Jaksch1998,Kohn1959} as $\hat{\psi}(\mathbf{r})=\sum_l w_0(\mathbf{r}-\mathbf{r}_l) a_l$. We then see that terms involving off-site interactions and tunnelling to more than the nearest neighbour site can be neglected in lattices deeper than about $5 E_R$, where $E_R=\hbar^2 k_L^2/(2m)$, and $k_L$ is the wavenumber of the laser generating a cubic lattice through forming standing waves in three dimensions \cite{Jaksch1998,Jaksch2005}.
It is possible to realise a large array of lattice models in this way, and to obtain an array of spin models as limiting cases \cite{Lewenstein2007}. Lattice geometries are specified by the optical trapping potentials, and it is also possible to confine atoms to low-dimensional systems by making the lattice sufficiently deep along particular axis directions that the atoms do not tunnel along those directions on typical experimental timescales \cite{Jaksch2005}. Because $J$ depends exponentially on the depth of the lattice, it is relatively straight-forward to make tunnelling timescales very long indeed \cite{Jaksch2005}. A particularly important model that can also be realised in this way is the (fermionic) Hubbard model \cite{Jordens2008,Esslinger2010},
\[
{H_{\rm FH}}=-J\sum_{\langle l,j\rangle,\sigma}\hat{a}_{l\sigma}^{\dag}\hat{a}_{j\sigma}+U\sum_{l}\hat{a}_{l\uparrow}^\dag\hat{a}_{l\uparrow}\hat{a}_{l\downarrow}^\dag\hat{a}_{l\downarrow}+\sum_{l,\sigma}\epsilon_{l}\hat{a}_{l\sigma}^\dag\hat{a}_{l\sigma},\]
which is well known as a description for strongly interacting electrons in the solid state. Here, $\hat{a}_{l\sigma}$ are fermionic operators obeying standard
anti-commutator relations, and $\sigma\in\{\uparrow,\downarrow\}$.
This is a simple example of the many two-species models that can be
engineered with atoms in optical lattices. It is similarly possible to create mixtures of bosons and fermions in a lattice, as well as multi-band Hubbard models \cite{Lewenstein2007, Pichler2010}.
This level of control over ultracold atomic gases offers enormous new opportunities for the field of analogue quantum simulation \cite{Cirac2012,Bloch2012}. In particular, by realising lattice models that are analytically and computationally intractable, there is great potential to use these systems as a special-purpose analogue quantum computer to determine the properties of these models. A simple but important example of this is the simulation of the 2D Hubbard model, the ground state of which under certain conditions is still debated. Moreover, the ability to construct an enormous range of models opens possibilities for realising theoretical ideas that have been predicted, but never observed, in a real physical system. In each of these ways, there are many possibilities for cold atoms in optical lattices to contribute strongly to our understanding of many-body quantum physics in the next few decades.
At the same time, the use of these systems as quantum simulators also offers up key challenges. Firstly, there are always imperfections in an experimental implementation that go beyond the Hamiltonian models written down here. Technically, this can include things such as noise on lattice potentials \cite{Pichler2012,Pichler2013}, and also spontaneous emissions, or incoherent scattering of light from the lasers generating the optical lattice potential \cite{Gerbier2010,Pichler2010, Dalibard1985,Gordon1980,Ellinger1994}. It is important to characterise and control the effects that these imperfections have on the many-body states in the lattice, in order to restrict heating and ensure that the most interesting many-body states can be reached \cite{McKay2011}. Secondly, many of the most interesting many-body states specifically require low temperatures for their realisation, especially those that arise in perturbation theory, e.g., as spin superexchange terms $\propto J^2/U$ in strongly interacting regimes where $U\ll J$ \cite{Lewenstein2007}. While current experimental temperatures have reached regimes of hundreds of picokelvin \cite{McKay2011} -- which is impressive on an absolute scale -- these temperatures are often higher than scales of the order of $J^2/U$ in optical lattices \cite{Jordens2010,Fuchs2011,Greif2013}.
However, the same understanding of the atomic physics that allows for design and engineering of lattice and spin models in these systems also allows us to microscopically model dissipative processes, including both the imperfections that can lead to heating, and also engineered dissipative processes that could provide a new route to cooling and preparation of many-body states. Below we investigate several examples of these processes, beginning with light scattering in an optical lattice and particle loss, and continuing on to engineered dissipation, and state preparation in driven, dissipative systems.
\subsection{Light scattering}
\label{sec:lightscattering}
Incoherent light scattering from trapped atoms has been studied in several different contexts over the past few decades. This has included the study of heating of single particles in optical dipole traps \cite{Dalibard1985,Gordon1980}, and more recently directly in the context of optical lattice potentials \cite{Gerbier2010,Pichler2010}. Where early studies with single atoms focussed on the rate of increase of energy, this is typically not sufficient to completely characterise how the many-body state changes in an optical lattice. Instead, we have to look more closely at the full out-of-equilibrium dynamics of the many-body system. This is more reminiscent of studies of the measurement of Bose-Einstein condensates in the late 1990s, which sought to understand the origins of interference patterns in experiments produced by multiple trapped condensates in a way that went beyond the assumption of each condensate having a fixed phase \cite{Cirac1996,Wong1996,Jack1996,Yoo1997,Castin1997,Ruostekoski1997,Ruostekoski1998,Dunningham1999,Dalvit2002}. These studies included relative phase measurement directly via light scattering \cite{Ruostekoski1997,Ruostekoski1999,Saba2005}, as well as decoherence of a Bose-Einstein condensate due to such scattering processes \cite{Ruostekoski1998}.
For atoms in an optical lattice, it is possible to derive a many-body master equation to describe incoherent scattering of light in the system. This treatment involves the generalisation of the optical Bloch equations from section \ref{sec:twolevel} to many atoms, as was done by Lehmberg \cite{Lehmberg1970,Lehmberg1970a}, and the inclusion of the atomic motion \cite{Ellinger1994}, in order to properly account for the mechanical effects of light on atoms \cite{Kazantsev1989,Cohen-Tannoudji1989,Cohen-Tannoudji1998}.
For a single particle, the resulting master equation for the system density operator $\rho$ was derived in several places \cite{Dalibard1985,Gordon1980,Gerbier2010,Pichler2010}, and takes the form
\begin{eqnarray}
\frac{d}{dt}{\rho}&=&-{\rm i}\left[H_{\rm atom},\rho\right] - \frac{\Gamma}{2} \int d^2\mathbf{u}N(\mathbf{u})\left( c^{\dag}_{\mathbf{u}}c_{\mathbf{u}}\rho+\rho\,c^{\dag}_{\mathbf{u}}c_{\mathbf{u}}-2c_{\mathbf{u}}\rho\,c^{\dag}_{\mathbf{u}}\right),\nonumber
\end{eqnarray}
where the Hamiltonian for the single atom with ground state $|g\rangle$ and excited state $|e\rangle$, as depicted in Fig.~\ref{fig:largefrequency}b, is given by
\begin{equation}\label{eq:single_article_Hamiltonian}
H_{\rm atom}=\frac{\hat{\mathbf{p}}^{\,2}}{2m}-\Delta \ket{e}\bra{e}-\left(
\frac{\Omega_0(\hat{\mathbf{x}})}{2} \ket{e}\bra{g}
+{\rm H.c.}\right).
\end{equation}
The decay rate is given by $\Gamma$, and the jump operators in the master equation are $c_{\mathbf{u}}=e^{-{\rm i}k_{A}\mathbf{u}\cdot
\hat{\mathbf{x}}} \ket{g}\bra{e}$, corresponding to a decay $\ket{e}\rightarrow\ket{g}$ and a momentum recoil. For the recoil, $k_{A}$ is the wavenumber of a photon at the atomic transition frequency, and $k_A \approx k_L$, where $k_L$ is the wavenumber of the laser driving the classical transition at a spatially-dependent Rabi frequency $\Omega_0(\hat{\mathbf{x}})$ and with detuning $\Delta$. The distribution of scatted photons, $N(\mathbf{u})$, is determined by the distribution of dipole radiation about the dipole $\hat{\mathbf{d}}$ between the states $|g\rangle$ and $|e\rangle$ as
\begin{equation}
N(\mathbf{u})=\frac{3}{8\pi}\left[1-\left(\hat{\mathbf{d}}\cdot\mathbf{u}\right)^2\right].
\end{equation}
Here, the equations of motion have been written in a frame rotating with the laser frequency, so that the optical frequencies have been eliminated, and we have made the standard set of approximations outlined in section \ref{sec:approxamo}. In the limit where $\Delta\gg \Omega_0$, we can adiabatically eliminate the excited state \cite{Gardiner2005}, giving an effective equation of motion for the system density operator for the ground state $\rho_g$,
\begin{align}
\frac{d}{dt}\rho_g&=-{\rm i}(H_{\textrm{eff}} \rho_g-\rho_g H_{\textrm{eff}}^{\dag})+{\mathcal{J}}{\rho_g}.
\end{align}
Here,
\begin{eqnarray} H_{\textrm{eff}}&=&\frac{\hat{\mathbf{p}}^{\,2}}{2m}+\frac{|\Omega_0(\hat{\mathbf{x}})|^2}{4\Delta}-{\rm i}\frac{1}{2}\frac{\Gamma|\Omega_0(\hat{\mathbf{x}})|^2}{4\Delta^2}\nonumber\\
&\equiv&\frac{\hat{\mathbf{p}}^{\,2}}{2m}+V_{\rm opt}(\hat{ \mathbf{x}}) -{\rm i}\frac{\gamma_{\rm eff}(\hat{\mathbf{x}})}{2}, \label{voptgeff}
\end{eqnarray}
showing how the optical potential $V_{\rm opt}(\mathbf{x})$ arises from the AC-Stark shift in this two-level system, as well as how the effective spontaneous emission rate $\gamma_{\rm eff}(\mathbf{x})$ appears. Note that we have also taken the limit $\Delta \gg \Gamma$ to simplify eq.~\eqref{voptgeff}. The recycling term in the master equation is given by
\begin{equation}
{\mathcal{J}}{\rho_g}=\Gamma\int d^2\mathbf{u}\,N(\mathbf{u})\left[
e^{-{\rm i}k_{A}\mathbf{u}\cdot\hat{\mathbf{x}}}\frac{\Omega_0(\hat{\mathbf{x}})}{2\Delta}\right]\rho_g \left[ e^{ik_{A}\mathbf{u}\cdot\hat{\mathbf{x}}}\frac{\Omega_0^{\ast}(\hat{\mathbf{x}})}{2\Delta}\right].
\end{equation}
From this, we clearly see the effect of the spontaneous emission on the motion of the atom in the ground state, based on the effective jump operator $\tilde c_{\mathbf{u}}(\hat{\mathbf{x}})\equiv e^{-{\rm i}k_{A}\mathbf{u}\cdot\hat{\mathbf{x}}}\Omega_0(\hat{\mathbf{x}})/(2\Delta)$, which contains the momentum recoils associated with the absorption of a laser photon from the classical field [$\Omega_0(\mathbf{x})$] followed by spontaneous scattering of a photon in the direction $\mathbf{u}$.
In Ref.~\cite{Pichler2010}, this derivation was generalised to $N$ bosonic atoms in a laser field, where the dynamics of the reduced density operator $\rho$, now for the many-body system, was derived. Because the dissipative dynamics are again dominated by the large optical frequency, all of the usual quantum optics approximations from section \ref{sec:approxamo} can be made here. As in the single-particle case, the excited state is adiabatically eliminated, making it possible to write a master equation in second quantisation using the field operators $\hat{\psi}(\mathbf{x})$,
\begin{eqnarray}\label{eq:master_equation}
\dot \rho&=&-{\rm i} \left( H_{\textrm{eff}}\rho-\rho H_{\textrm{eff}}^\dag \right) +\mathcal{J}\rho,
\end{eqnarray}
with effective Hamiltonian
\begin{eqnarray}
H_{\rm eff} &=&H_{\rm atoms} + H_{\rm eff, rad},
\end{eqnarray}
where $H_{\rm atoms}$ was given in eq.~\eqref{hatoms} up to the additional note that the effective potential $V_0(\mathbf{r})=V_{\rm opt}(\mathbf{r})$ as given in eq.~\eqref{voptgeff} above, and the effects of radiative coupling to the field are given by the effective Hamiltonian term
\begin{flalign}
H_{\rm eff,rad}= &-{\rm i}\frac{1}{2}\int d^3x \frac{\Gamma| \Omega_0(\mathbf{x})|^2}{4\Delta^2} \hat{\psi}^{\dag}(\mathbf{x})\hat{\psi}(\mathbf{x})\nonumber\\
& -{\rm i}\frac{1}{2}\iint d^3xd^3y \frac{\Gamma\Omega_0(\mathbf{y})\Omega_0^{\ast}(\mathbf{x})}{4\Delta^2}F(k_{A}(\mathbf{x}-\mathbf{y})) \fd{x}{}\fd{y}{}\f{y}{}\f{x}{}\nonumber\\
&+\iint d^3xd^3y \frac{\Gamma\Omega_0(\mathbf{y})\Omega_0^{\ast}(\mathbf{x})}{4\Delta^2}G(k_{A}(\mathbf{x}-\mathbf{y})) \fd{x}{}\fd{y}{} \f{y}{}\f{x}{}
\end{flalign}
and the recycling term \cite{Pichler2010}
\begin{flalign}
\mathcal{J}\rho&=\iint d^3x d^3y \frac{\Gamma\Omega_0(\mathbf{x})\Omega_0(\mathbf{y})}{4\Delta^2} F(k_{A}(\mathbf{x}-\mathbf{y}))
\hat{\psi}^{\dag}(\mathbf{x})\hat{\psi}(\mathbf{x})\rho
\hat{\psi}^{\dag}(\mathbf{y})\hat{\psi}(\mathbf{y}),&&\label{eq:dissipation}
\end{flalign}
where the fuctions $F$ and $G$ are defined as
\begin{flalign}
F(\mathbf{z})&=\int d^2{u}\,N(\mathbf{u})e^{-{\rm i}\mathbf{u}\cdot\mathbf{z}}&&\nonumber\\
&=\frac{3}{2}\left\{\frac{\sin z}{z}\left(1-(\hat{\mathbf{d}}\cdot\hat{\mathbf{z}})^2\right)+\right.
\left.\left(1-3(\hat{\mathbf{d}}\cdot\hat{\mathbf{z}})^2\right)\left(\frac{\cos z}{z^2}-\frac{\sin z}{z^3}\right)\right\},\\
G(\mathbf{z})&=-\frac{1}{z^3}\mathcal{P}\int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}\frac{\zeta^3}{\zeta-z}F(\zeta\mathbf{z}/z)&&\nonumber\\
&=\frac{3}{4}\left\{\left((\hat{\mathbf{d}}\cdot\hat{\mathbf{z}})^2-1\right)\frac{\cos z}{z}+\right.
\left.\left(1-3(\hat{\mathbf{d}}\cdot\hat{\mathbf{z}})^2\right)\left(\frac{\sin z}{z^2}+\frac{\cos z}{z^3}\right)\right\},
\end{flalign}
and $\mathcal{P}$ is used to denote the Cauchy principal value integral.
If the short-range collisional physics Hamiltonian accounting for short range collision physics in the presence of laser fields gives rise to additional dissipative processes, than this can also be accounted for by an additional two-body loss term in the effective Hamiltonian \cite{Pichler2010}.
\begin{figure}[tbh]
\begin{center}
\includegraphics[width=12cm]{figures/spemschematic}
\end{center}
\caption{Schematic plot of the effects of spontaneous emissions in a deep optical lattice. For an atom beginning in the lowest band of the lattice, a spontaneous emission (a) can transfer it to the first excited band with a probability $\eta^2$, were $\eta$ is the Lamb-Dicke parameter (see the text for details), and (b) will always lead to localisation of the atom, in the sense that the spontaneous emission provides information to the environment about the location of the atom.}\label{fig:spemschematic}
\end{figure}
In eq.~\eqref{eq:dissipation}, the function $F(k_{A}(\mathbf{x}-\mathbf{y}))$ is localised in $(\mathbf{x}-\mathbf{y})$ on a length scale given by $1/k_{A}$, i.e., $\lambda_A/(2\pi)$. As a result, particles undergoing spontaneous emission can be reinterpreted as a continuous measurement of the position of the atoms up to a position resolution of length scale $\lambda_A/(2\pi)$. Phrased another way, as the atoms scatter light incoherently in random directions, they provide their environment with information about where they are. This scattering does two things then, as depicted in Fig.~\ref{fig:spemschematic}. Firstly, with a probability given by $\eta^2$, where $\eta=2 \pi a_{\rm trap}/\lambda\sim 0.3$ for optical lattices with depth around 10$E_R$, an atom beginning in the lowest band will be transferred to a higher band via the momentum recoil. This depends on the size of the trap wavefunction $a_{\rm trap}$ in each lattice site\footnote{Note that this transfer can also be interpreted as measurement of the position of the atom within the lattice site.}. Then, whether or not the atom is transferred to a higher band, a delocalised atom will be localised within the lattice, in the sense that a coherent superposition of an atom delocalised over different lattice sites will be transferred to a mixed state of the atom on different lattice sites. Note that this happens because the localisation scale is $\lambda_A/(2\pi)$, whereas the lattice spacing for a lattice formed by counter-propagating laser beams is $\lambda_A/2$. This is the opposite limit to what might occur in solid state experiments, where the lattice spacing is often much shorter than typical optical wavelengths.
\begin{figure}[tbh]
\begin{center}
\includegraphics[width=8cm]{figures/pichler1}
\end{center}
\caption{Comparison between the total process rate $\Gamma_{\rm scatt}
/\gamma_0$ (solid lines) and the rate for processes back to the lowest band (dashed lines) for $\Delta<0$ (upper lines) and a $\Delta>0$ (lower lines) in the ground state of a 1D lattice (Using DMRG ground states for Bosons in a 1D lattice with lattice depth in the $x$, $y$ and $z$ directions $V_x=V_y=30 \,E_R, V_z=10 \,E_R$). The rates include scattering from all three lattice generating beams. Reprinted figure from H.~Pichler, A.~J.~Daley, and P.~Zoller, Phys. Rev. A {\bf 82}, 063605 (2010) \cite{Pichler2010}. Copyright 2010 by the American Physical Society.}\label{fig:pichler1}
\end{figure}
The rate with which spontaneous emission events occur is then dependent on the sign of the detuning for the atoms. For red detuning $\Delta<0$, the AC-Stark shift potential derived in eq.~\eqref{voptgeff} shows that atoms are located where the light is brightest, whereas for $\Delta>0$, the atoms are located where the light is darkest. This results in a difference in scattering rates between blue and red light $\sim \eta^2$ \cite{Gordon1980,Dalibard1985}, with the dominant process for blue-detuned fields being the transfer of atoms to the first excited band. This can be seen in the results from Ref.~\cite{Pichler2010} for scattering rates from many bosons in an optical lattice, which are reproduced in Fig.~\ref{fig:pichler1}. There the expected rate of scattering per particle $\Gamma_{\rm scatt}$ is plotted for atoms in red and blue detuned optical lattices, normalised to $\gamma_0$, which is the value of $\gamma_{\rm eff}$ (from eq.~\ref{voptgeff}) at the maximum depth of the lattice. The initial state involves bosons in a 1D optical lattice at unit filling, and we see that in the red detuned case, for small interactions $U/J$, multiple occupancy of particles on lattice sites gives rise to a super-radiant enhancement of the scattering rate.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=12cm]{figures/spemhigherband}
\end{center}
\caption{Schematic diagram indicating the energy scales involved in preventing thermalisation of the bandgap energy after a spontaneous emission. Atoms transferred to a higher band cannot thermalise the energy $\omega_{\rm gap}$ on typical experimental timescales, because this energy scale is much larger than the energy scales $J$ associated with tunnelling and $U$ associated with interactions in the lowest band.}\label{fig:spemhigherband}
\end{figure}
Although the scattering rates are different by almost an order of magnitude between red and blue detuning, the rate of energy increase is identical for the two cases, and also equal to $N$ times the single particle result \cite{Pichler2010,Gerbier2010,Dalibard1985}. This might seem to imply that red and blue detunings have the same effect on a many body state, but this is not the case. The reason for this comes back to the question of whether energy added to the system is thermalised in an out of equilibrium many-particle system. Here, the largest contribution towards the increase in energy in the system comes from transfer of particles to higher Bloch bands. But, as depicted in Fig.~\ref{fig:spemhigherband}, the bandgap energy $\omega_{\rm gap}\gg U,J$, so that the energy scales of dynamics in the lowest band are much smaller than the bandgap energy. This means that only a very high-order process in $J/\omega_{\rm gap}$ and $U/\omega_{\rm gap}$ will allow the sharing of the bandgap energy amongst atoms in the lowest band. These processes will not take place on typical experimental timescales, and so the system will not thermalise the energy input from the scattering on the relevant timescales.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=12cm]{figures/pichler2}
\end{center}
\caption{Relative rate of change of the (integrated) single particle density matrix $S_{1D}(z_1,z_2)=\iint dxdy\langle\hat{\psi}^{\dag}(x,y,z_1)\hat{\psi}(x,y,z_2)\rangle$ in an effectively 1D lattice with depth in recoil energies given by ($V_x=V_y=30\,E_R$; $V_z=10\,E_R$). In the tightly bound transversal directions the atoms are assumed to be in the lowest band. Scattering from all three lattice generating laser beams is taken into account (with weights corresponding to the lattice depths). (The lattice constant is denoted by $a$.). Reprinted figure from H.~Pichler, A.~J.~Daley, and P.~Zoller, Phys. Rev. A {\bf 82}, 063605 (2010) \cite{Pichler2010}. Copyright 2010 by the American Physical Society. } \label{fig:pichler2}
\end{figure}
For this reason, it is important to study the complete out-of-equilibrium many-body physics of many atoms as described by the master equation. As a first step in this direction, it is possible to calculate the rate of change of the single-particle density matrix $\langle \hat \psi^\dagger(\mathbf{x})\psi(\mathbf{y})\rangle$ at short times using perturbation theory. This correlation function is important, because it characterises the off-diagonal long-range order in a superfluid phase, as well as the expontential localisation of particles in a Mott Insulator. This calculation was performed in Ref.~\cite{Pichler2010}, and the results are shown in Fig.~\ref{fig:pichler2}. This shows that the key effect of spontaneous emissions is to lead to localisation of the atoms, and therefore a direct decrease in the off-diagonal elements. This rate of decrease is proportional to the scattering rate, and therefore is much larger for red-detuned lattices than for blue-detuned lattices. It is also more important in the superfluid state, where the off-diagonal correlations are initial strong, as opposed to the Mott Insulator state, where the atoms are anyway initially exponentially localised.
\begin{figure}[tb]
\centering
\includegraphics[width=12cm]{figures/pichler3}
\caption{Comparison of the decay of off-diagonal correlations in the single particle density matrix $S(i,j)=\langle a_i^\dag a_j \rangle$ as a function of time. These are computed from the master equation for the lowest band Eq.~\eqref{eq:QT_master_eq} with $\gamma/J=0.01$ for 24 particles on a 24 site lattice by combining quantum trajectories methods with t-DMRG. (a) Comparison of the decay of off-diagonal elements $\sum_i S(i,i+x)$ when $U/J=2$ (superfluid regime, upper lines) and $U/J=10$ (Mott Insulator regime, lower lines). In each case we show correlations at separation distances $x=1$ (solid lines) and $x=2$ (dashed lines), each of which are averaged over $i$. We see that the correlations for the Mott Insulator state at $U/J=10$ are almost constant, whereas for $U/J=2$, the decay becomes more rapid than at initial times. (b) Comparison of $S(i,i+2)$ averaged over $i$ and normalized to the value at time $t=0$, $|S(i,i+2)|[t=0]$ for $U/J=2,4,6,8,10$ (seen here from bottom to top). The dashed line shows the corresponding result from perturbation theory. Each computation was averaged over 1000 trajectories, and error bars are shown in (b). For (a), the statistical errors fit inside the line thickness. Adapted from H. Pichler et al., Phys. Rev. A {\bf 82}, 063605 (2010) \cite{Pichler2010}. Copyright 2010 by the American Physical Society.} \label{fig:pichler3}
\end{figure}
To go beyond this perturbation theory result, in Ref.~\cite{Pichler2010}, a combination of t-DMRG with quantum trajectories is used to solve the master equation for atoms in the lowest band. Under the approximation that atoms do not get transferred to higher bands, and making the approximation that atoms are localised exactly on a single site, eq.~\eqref{eq:master_equation} reduces to
\begin{equation}
\label{eq:QT_master_eq}
\dot{\rho}=-{\rm i}[H_{BH},\rho]-\sum_{i}\frac{\gamma}{2}\left( n_i n_i\rho+\rho n_i n_i-2n_i\rho n_i \right),
\end{equation}
where $n_i=a^\dagger_i a_i$, $\gamma$ is the effective scattering rate on a single site, and $H_{BH}$ is the Bose-Hubbard Hamiltonian for the lowest band from eq.~\eqref{HBH}.
In Fig.~\ref{fig:pichler3} we reproduce the results for the time dependence of the single-particle density matrix in the lowest band, $\langle a_i^\dagger a_j\rangle$. It can be seen that in the Mott Insulator limit, studying the proper interplay between coherent and dissipative dynamics in this system shows that the system is even more robust towards spontaneous emissions than might have been expected from perturbation theory, as local correlations are reestablished by coherent dynamics. The superfluid state, on the other hand, exhibits a faster rate of decay for off-diagonal elements of the single-particle density matrix at medium times than had been predicted using perturbation theory. This appears to result from better thermalisation of the energy being added to the system from spontaneous emission events.
\begin{figure}[tb]
\centering
\includegraphics[width=11cm]{figures/kollath2}
\caption{Diagonal terms of the density matrix $\rho_{n,n}$, with $ {\rho}=\sum \rho_{n,m} \ket{n}\bra{m}$ versus rescaled time $\tau$ in
log-log form: the exact solutions of Eq.~(\ref{eq:QT_master_eq}) for $n=41$(top)$...80$(bottom)
are represented by blue continuous lines; the element $n=40$ (thick red line) is proportional to the probability of finding
a balanced configuration $P_b$; the corresponding diffusion density $p(x=n/N,\tau)$ Eq.~(5) of Ref.~\cite{Poletti2012} up
to $n/N=60$ are shown in dashed green lines. Inset: 3D plot of the evolution.
Parameters: $U/J=20$, $\hbar\gamma/J=1$, $N=80$ and $L=2$. Reprinted figure with permission from D.~Poletti, J.-S.~Bernier, A.~Georges, and C.~Kollath, Phys. Rev. Lett {\bf 109}, 045302 (2012). Copyright 2012 by the American Physical Society. } \label{fig:kollath2}
\end{figure}
\begin{figure}[tb]
\centering
\includegraphics[width=9cm]{figures/kollath1}
\caption{Numerical evolution of the density matrix elements $\rho(n,\tau)$ (solid lines), where the density operator is given by the ansatz $\bigotimes_j\left[\sum_{n}\rho(n,t)\ket{n}\bra{n}\right]$, in a semi-logarithmic plot, versus $n$ for large rescaled times $\tau$ between 0.1 and 50 (not equidistant) in the direction of the (red) arrow. The inset shows the same evolution at shorter times $\tau$ between 0.0002 and 0.1 (not equidistant) in the direction of the (red) arrow in a linear plot. Parameters: $N/L=3$, $U/\hbar\gamma=10$. The (red) dashed lines show the (analytical) asymptotic limit. Reprinted figure with permission from D.~Poletti, P.~Barmettler, A.~Georges, and C.~Kollath, Phys. Rev. Lett {\bf 111}, 195301 (2013) \cite{Poletti2013}. Copyright 2013 by the American Physical Society. } \label{fig:kollath1}
\end{figure}
The interplay between interactions and dissipation also leads to interesting dynamics in on-site particle occupation numbers for bosons. In particular, dynamics resulting from spontaneous emissions generate significant probability that states with multiple occupancies on lattice sites -- which are disfavoured due to interaction energies -- will appear, and these probabilities continue to grow over time. In Refs.~\cite{Poletti2012,Poletti2013}, Poletti et al.~investigate these dynamics beginning from eq.~(\ref{eq:QT_master_eq}). They observe a number of striking effects, including the tendency of interactions to slow down decoherence, because spontaneous emissions distinguish only between states with different on-site particle numbers. Strong interactions lead to exponential localisation of particles at given lattice sites, with small amplitudes in perturbation theory for particles to be found at neighbouring sites. This results in anomalous diffusion in a particle number basis, even for a two-site model with large occupation \cite{Poletti2012}. For a lattice with many sites, it further gives rise to glass-like behaviour, exhibiting an anomalous diffusive evolution in configuration space at short times and dynamics dominated by rare events at long times \cite{Poletti2013}.
An example of these dynamics for a two-site system is reprinted from Ref.~\cite{Poletti2012} in Fig.~\ref{fig:kollath2}, and shows how the probability distribution for different occupation numbers on two sites diffuses as a function of time. Beginning with equal number of particles on each site, the initial behaviour is an exponential change away from this configuration, which is replaced by power-law decay with an exponent of 1/4, as the dynamics become dominated by processes populating states with high local occupancy. Related behaviour can be seen for a lattice with many sites, as demonstrated by a figure reprinted here from Ref.~\cite{Poletti2013} in Fig.~\ref{fig:kollath1}. The calculation shown involves an approximate solution to the master equation eq.~(\ref{eq:QT_master_eq}), obtained using a separable and translationally invariant ansatz for the density operator as $\bigotimes_j\left[\sum_{n}\rho(n,t)\ket{n}\bra{n}\right]$. Here the sum over $j$ includes all of the lattice sites and that over $n$ includes all of the possible occupations of each site. Again, it is clear that the long-time behaviour is dominated by anomalous diffusion that populates states with high onsite occupation numbers. Such dynamics should be directly observable in current experiments, especially in setups with local site addressing in a quantum gas microscope and control over light scattering rates. Initial experimental studies of dynamics in the presence of light scattering and density measurements were recently performed by Patil et al.~\cite{Patil2014}.
For fermionic species this interplay between interactions and dissipation can similarly lead to important many-body effects. In the case of multiple fermionic species, where the dissipation distinguishes between the two species, this can lead to a build-up of long-range correlations beginning from a fermionic Mott Insulator \cite{Bernier2013}. The same types of effects can also protect states of spin-ordered fermions, especially where spontaneous emission does not distinguish between different spin-species \cite{Sarkar2014}.
\begin{figure}[tb]
\centering
\includegraphics[width=10cm]{figures/schachenmayer1}
\caption{ (a) Absorption and spontaneous emission of a lattice photon
leads effectively to localization of single atoms. Tunnelling and interactions between
atoms then redistribute the energy added to the system. (b) Localization of an atom
in space corresponds initially to a distribution of the atom over the whole Brillouin
zone (the tails of the quasi-momentum distribution are lifted). Subsequent unitary
evolution leads to a broadened quasi-momentum distribution, {\it i.e.}, the $n_{q=0}$ peak and the tails
decrease, while the small quasi-momentum components increase (t-DMRG
results, $U=2J$, $N=48$ particles on $L=48$ sites, $d_l=6$,
$D=512$. Reprinted figure from J.~Schachenmayer, L.~Pollet, M.~Troyer, and A.~J.~Daley, Phys. Rev. A {\bf 89}, 011601(R) (2014) \cite{Schachenmayer2014}. Copyright 2014 by the American Physical Society.} \label{fig:schachenmayer1}
\end{figure}
Another important question related to the interplay between dissipative and coherent dynamics is that of whether spontaneous emission events that leave atoms in the lowest band of an optical lattice lead to energy increases that can be thermalised on typical experimental timescales. Though atoms being placed in higher bands cannot thermalise the bandgap energy, as we have already discussed, there is experimental evidence that suggests as a system heats due to spontaneous emissions that the momentum distributions can be quantitatively compared with thermal momentum distributions calculated via Quantum Monte Carlo methods \cite{Trotzky2010}.
In Ref.~\cite{Schachenmayer2014}, this is studied using quantum trajectory methods combined with t-DMRG, initially considering the thermalisation after a single spontaneous emission event (averaged over the different lattice sites on which the event can occur), and then generalising this to continuous light scattering. As shown in Fig.~\ref{fig:schachenmayer1}, after a spontaneous emission occurs, initially the momentum distribution develops strong tails as a result of a single atom being localised spatially on a lattice site, and therefore completely delocalised over the Brillouin zone in quasimomentum. As a result of interactions with other particles, the quasimomentum distribution then gradually relaxes towards a steady-state value. As a particular way of determining to what extent simple single-particle quantities have thermalised after a particular length of time, the time dependence of the kinetic energy and the quasimomentum distribution is then calculated, and compared with the thermal values from a canonical ensemble in which the temperature is chosen to match the \emph{total energy} in the system.
\begin{figure}[tb]
\centering
\includegraphics[width=10 cm]{figures/schachenmayer2}
\caption{Time evolution after a single spontaneous
emission (averaged over jumps on all possible sites). (a - b) For
a superfluid initial state ($U=2J$), the kinetic energy relaxes
to the equilibrium value obtained from a Quantum Monte Carlo (QMC) calculation
$E_{\rm eq}^{\rm kin}$. For MI states ($U=4J$), the energy
relaxes, but not to $E_{\rm eq}^{\rm kin}$. The zero value of
kinetic energy for this plot is the ground state kinetic energy
$E_{\rm gs}^{\rm kin}$. (c) The difference of the infinite time
value of the kinetic energy (obtained from an extrapolation of an
exponential fit) to the Monte Carlo equilibrium energy. For MI
states with $U/J \gtrsim 3.37$, the difference increases rapidly
for system sizes $M=24,48,96$ sites (d) The decay rate extracted
from the exponential fit as a function of $U$. (e) Comparison of
the time-evolved quasi-momentum distribution at $t=10/J$ (dots)
to the equilibrium distribution from a QMC calculation. (f)
Differences between the two distributions as a function of time
for the $qa=0$ peak and for a large quasi-momentum of $qa=(40/48) \pi$. In
the superfluid ($U=2J$), the components for large momenta relax rapidly to thermal values, for $qa\sim 0$, the relaxation timescale is much longer. In the MI with $U=5J$, the same is true, but for large momenta there is a discrepancy to the thermal value.
(t-DMRG results, $d_l=6$, $D=256,512$; error bars represent
fitting errors and statistical errors from QMC. Reprinted figure from J.~Schachenmayer, L.~Pollet, M.~Troyer, and A.~J.~Daley, Phys. Rev. A {\bf 89}, 011601(R) (2014) \cite{Schachenmayer2014}. Copyright 2014 by the American Physical Society.} \label{fig:schachenmayer2}
\end{figure}
It is found that both in the Mott Insulator and the superfluid regime, the system relaxes to a steady state value on a short timescale. As can be seen in Fig.~\ref{fig:schachenmayer2}a, in the superfluid regime the value to which the kinetic energy relaxes is equal to the expected thermal value $E_{\rm eq}^{\rm kin}$ in a canonical ensemble with the same total energy. Such a relaxation has also been recently explored for Luttinger liquids heated in a related manner \cite{Buchhold2014}. However, in the Mott Insulator regime (Fig.~\ref{fig:schachenmayer2}b), while the system relaxes rapidly to a steady-state value, this is not the same value as expected from a canonical ensemble with the same total energy. As shown in Fig.~\ref{fig:schachenmayer2}c, such discrepancies begin at the phase transition point, and become larger for larger $U/J$. This sudden transition is in contrast to the variation of the relaxation rate as a function of $U/J$, which is smoothly and slowly varying (Fig.~\ref{fig:schachenmayer2}d). Similar results and timescales are seen for the high-quasimomentum components of the quasimomentum distribution, as shown in Figs.~\ref{fig:schachenmayer2}e,f. For low-quasimomentum values also on the superfluid side the relaxation time is much longer, because these momentum values depend heavily on correlation functions over long distances. In Ref.~\cite{Schachenmayer2014}, it is shown that this behaviour, as a local quench, is strongly dependent on the low-energy spectrum of the Hamiltonian, which leads to the strong changes from the superfluid regime to the Mott Insulator regime.
Microscopic understanding and control over spontaneous emissions in many-body systems also has other applications, including the potential to observe Pauli-blocking of spontaneous emissions \cite{Sandner2011,Busch1998} for Fermi gases. This can be used as an ingredient in state preparation schemes that produce pairing of fermions dissipatively \cite{Diehl2010,Yi2012}
With this recent progress, it has been possible to understand various basic features of heating and decoherence of many-particle states with cold atoms in optical lattices. In the near future it should be possible to investigate such effects for a much broader range of many-body states and Hamiltonians. This will provide both a better fundamental understanding of decoherence in many-body states, as well as important practical information for controlling heating in experiments. In the near future it will be important to understand better how the dynamical processes that atoms undergo as they are loaded into a lattice or many-body states are prepared interacts with the dissipation to produce heating during this phase of the experiment. It will also be important to go further into other dissipative mechanisms and their effects, including classical noise \cite{Pichler2012,Pichler2013} and particle loss (see the next section).
\subsection{Particle loss}
\label{sec:particleloss}
Another dissipative process with strongly interacting cold quantum gases that can be treated in experimentally relevant regimes using open systems techniques from quantum optics is particle loss. In collisional loss events, the binding energy released in the collision is typically much larger than the depth of the corresponding trap. In this regime, this energy acts as a large scale, exactly in the sense of section \ref{sec:approxamo}. In particular in the case that atoms lost in such events will leave the lattice without additional collisions, this allows Born and Markov approximations to be made, and the derivation of a corresponding master equation. This has been investigated theoretically for two-body \cite{Syassen2008,Garcia-Ripoll2009}, three-body \cite{Daley2009}, and single-body \cite{Barmettler2011} loss processes for cold atoms in optical lattices.
One particularly useful and counter-intuitive result that emerges is that these systems are often found in a regime where losses can be suppressed by a continuous quantum Zeno effect \cite{Itano1990,Misra1977}. This has been observed in experiments for two-body losses of molecules \cite{Syassen2008,Yan2013}, and initial indications have also been seen for three-body loss in caesium atoms \cite{Mark2012}. In this section we introduce the continuous quantum Zeno effect, before reviewing these observations and the consequences for many-body physics in these systems.
\subsubsection{Continuous quantum Zeno effect}
\label{sec:zeno}
The quantum Zeno effect was first discussed by Misra and Sudarshan in Ref.~\cite{Misra1977}, and involves the tendency of repeated measurements on a quantum system to inhibit coherent processes. In an idealised scenario, this can be formulated with projective measurements as follows: if we consider, e.g., a two level system that begins in the state $\ket g$ and is coupled resonantly to state $\ket e$ with effective Rabi frequency $\Omega$, then after a short time $\Delta \tau$, the probability a measurement finding the system to be in state $\ket e$, $P_e \approx \Omega^2 \Delta \tau^2 /4$. Because this result $\propto \Delta \tau^2$, if measurements are repeated in equally spaced intervals up to a time $T$, then as we increase the number of measurements $T/\Delta \tau$ by taking $\Delta \tau\rightarrow 0$, we see that the probability of finding the system in the excited state at time $T$ also goes monotonically to zero. This effect was first observed in trapped ions by Itano et al.~\cite{Itano1990}, and an anti-Zeno effect that arises under more general conditions has also been measured \cite{Fischer2001}.
The \emph{continuous} quantum Zeno effect is a generalization of these ideas to continuous measurements \cite{Itano1990,Gagen1993}, in the sense that when a system is coupled to its environment, then the environment can be seen to be continuously performing measurements on the system. In essence it amounts to the fact that a strong dissipative process in a quantum system can suppress a related coherent process. Let us consider the specific case of a two-level atom described by the optical Bloch equations from section \ref{sec:twolevel}. In the presence of spontaneous decay of the excited state, we can write the quantum trajectories formulation of the evolution with an effective Hamiltonian given by
\begin{equation}
H_{\rm eff} = - \frac{\Omega}{2} \sigma_x - \Delta \sigma_+\sigma_- - {\rm i}\frac{\Gamma}{2} \sigma_+\sigma_-.
\end{equation}
If the system begins in the ground state $\ket g$, then for small $\Gamma$ it will undergo damped Rabi oscillations as computed in section \ref{sec:twolevel}. Let us now consider the overdamped regime $\Gamma \gg \Delta, \Omega$. For the general case where $\Omega \ll \Delta$ and/or $\Omega \ll \Gamma$, we can apply perturbation theory to the effective Hamiltonian, taking $H_0=-(\Delta + {\rm i}\Gamma/2)\sigma_+\sigma_-$ and $H_1=- \frac{\Omega}{2} \sigma_x$, and we see that we can write the effective energy shift for the ground state, $E_g$ as:
\begin{equation}
E_g=\frac{\bra g H_1 \ket e \bra e H_1 \ket g}{\Delta + {\rm i}\Gamma/2}= \frac{\Omega^2}{4\Delta^2 + \Gamma^2} (\Delta - {\rm i} \Gamma/2)
\end{equation}
Based on the physical interpretation of quantum trajectory methods, we know that the probability of spontaneous emission occurring after time $t$, $P_s(t)$ is given by $1-\Vert \ket \Psi \Vert$, where $\Psi$ is the corresponding wavefunction evolving under the effective Hamiltonian. In the perturbation theory limit, we therefore have
\begin{equation}
P_s(t)=\left\Vert \exp\left[ -{\rm i} \frac{\Omega^2}{4\Delta^2 + \Gamma^2} (\Delta - {\rm i} \Gamma/2) t \right] \ket g \right\Vert^2 =\exp\left[- \frac{\Omega^2}{4\Delta^2 + \Gamma^2} \Gamma t\right],
\end{equation}
so that the effective photon scattering rate is given by $\Gamma_{\rm scatt}\approx \Omega^2 \Gamma / (4\Delta^2+\Gamma^2)$. In the limit where $\Gamma \gg \Delta$, we see that this reduces to $\Gamma_{\rm scatt}\approx \Omega^2 /\Gamma$, i.e., as we continue to increase $\Gamma$, the scattering rate decreases.
This is the essence of the continuous quantum Zeno effect - for sufficiently strong dissipative processes, the dissipation will suppress coherent processes (in this case the classical drive with Rabi frequency $\Omega$) that occupy the states coupled to that dissipation. This leads to the counterintuitive situation that increasing the rate of dissipation for specific states actually leads to less actual dissipation in total (at least for a system beginning in a state that doesn't directly undergo dissipative processes -- if we begin in the state $\ket e$, it always decays at a rate $\Gamma$). This can also be rephrased in two other ways - one is that the linewidth of the excited state is so broad that coupling into it is suppressed. The other is that increasing $\Gamma$ increases the effective coupling to the environment, and therefore the rate at which the environment obtains information -- or \emph{measures} the system, analogously to the quantum Zeno effect with projective measurements.
\subsubsection{Two-body loss}
Evidence for the quantum Zeno effect resulting from two-body loss in a many-particle system was first obtained by Syassen et al.~\cite{Syassen2008}. Their setup involved weakly bound Feshbach molecules of Rb$_2$ \cite{Chin2010} held either in 3D optical lattice potentials, or confined to move in 1D tubes, in a regime where collisions between molecules resulted in the transfer of one molecule to a much more deeply bound state, and the separation of the atoms forming the second molecule. The binding energy released in this collision is typically much larger than the depth of the optical trap confining the molecules to begin with, resulting in loss of the particles involved in the collision. Because of the large energy scale involved in the inelastic collision, and because the particles are ejected immediately from the trap, the approximations made in section \ref{sec:approxamo} are justified, and a master equation can be derived to describe this process. Detailed analysis of these master equations are presented in Ref.~\cite{Garcia-Ripoll2009} for the 3D lattice case, and in \cite{Durr2009} for the case of particles in 1D tubes, based on a Lieb-Liniger model. In each case, a master equation can be derived for the density operator $\rho$ of the system, tracing over the states associated with the products of the collision. Assuming an effective zero-range potential to describe two-body loss in an equivalent form to the description of elastic interactions between particles, this can be written in terms of a field operator for bosonic Feshbach molecules in the initial state, $\hat \psi_m (\mathbf{r})$ as \cite{Garcia-Ripoll2009}
\begin{eqnarray}
\frac{d\rho}{dt}&=&-{\rm i} [H_{\rm mol},\rho]\nonumber\\ & & -\frac{L_2}{2} \int d^3 r \left\{ 2 [\hat\psi_m(\mathbf{r})]^2\rho [\hat\psi_m^\dagger(\mathbf{r})]^2 - [\hat\psi_m^\dagger(\mathbf{r})]^2 [\hat\psi_m(\mathbf{r})]^2\rho - \rho [\hat\psi_m^\dagger(\mathbf{r})]^2 [\hat\psi_m(\mathbf{r})]^2 \right\},\nonumber \\
\end{eqnarray}
where $L_2$ is the two-particle loss coefficient, and the coherent hamiltonian $H_{\rm mol}$ takes the same form as the atomic hamiltonian eq.~\eqref{hatoms} in section \ref{sec:qsim},
\begin{equation}
{H}_{\rm mol}\approx\int d^{3}r\,\hat{\psi}_m^{\dag}(\mathbf{r})\,\left[-\frac{\hbar^{2}}{2m}\nabla^{2}+V_{0}(\mathbf{r})\right]\,\hat{\psi}_m(\mathbf{r})+\frac{g}{2}\int d^{3}r\,\,[\hat{\psi}_m^{\dag}(\mathbf{r})]^2\,[\hat{\psi}_m(\mathbf{r})]^2. \label{hmols}
\end{equation}
Here, $m$ is again the particle mass (now the mass of a Feshbach molecule), and $g=4\pi \hbar^2 a_{\rm scatt} / m$, where $a_{\rm scatt}$ is the real part of the scattering length for two Feshbach molecules ($L_2$ may be interpreted as the imaginary part of an effective two-particle scattering length including inelastic processes \cite{Garcia-Ripoll2009}).
In the case of a 3D optical lattice, this can be rewritten as an equation of motion for molecules in the lowest band of the lattice, analogously to the derivation of the Bose-Hubbard model that was discussed in section \ref{sec:qsim}. The resulting master equation is given by \cite{Garcia-Ripoll2009}
\begin{equation}
\frac{d \rho}{dt} =- {\rm i}\left(H_{\rm eff,m} \rho - \rho H_{\rm eff,m}^\dag \right) +{\Gamma_0} \sum_l \hat m_l^2 \rho (\hat m_l^\dag)^2,
\label{3bodylossmastereq}
\end{equation}
with bosonic annihilation operators for Feshbach molecules in Wannier function modes denoted $m_l$, and the effective Hamiltonian
\begin{equation}
H_{\rm eff,m}=H_{\rm BH,m}-{\rm i} \frac{\Gamma_0}{2} \sum_l (\hat m_l^\dag)^2 \hat m_l^2.
\end{equation}
Here, $H_{\rm BH,m}$ is the Bose-Hubbard Hamiltonian for molecules on the lattice, taking an identical form to $H_{\rm BH}$ from eq.~\eqref{HBH}, but with bosonic molecule operators $m_l$, and $\Gamma_0$ denotes the effective on-site loss rate when two molecules occupy the same site of the lattice. As for the previous discussion of the Bose-Hubbard model in section \ref{sec:qsim}, we require all parameters to be small compared with the band gap energy, and assume that terms involving off-site processes other than nearest neighbour tunnelling are negligibly small (see below for further comments on this).
In order to interpret more directly the dynamics described by this master equation, we can expand the density operator in terms of the contributions with fixed particle number $N$, $\rho_{N}$. Beginning with $N_0$ particles in the system, $\rho=\sum_{N=N_0,N_0-2,N_0-4,...} \rho_{N}$, and
\begin{equation}
\dot \rho_{N} = - {\rm i}\left(H_{\rm eff} \rho_N- \rho_N H_{\rm eff}^\dag \right) +\Gamma_0 \sum_i \hat m_i^2 \rho_{N+2} (\hat m_i^\dag)^2.
\end{equation}
We can then consider the loss as having two effects: it produces states with two less particles, with the corresponding coupling between density matrices given by the recycling term, and for a fixed particle number, it provides an effective two-body on-site interaction term with an imaginary coefficient. This term generates the quantum Zeno effect analogously to the case of a two-level system discussed in section \ref{sec:zeno}.
\begin{figure}[tb]
\centering
\includegraphics[width=10cm]{figures/ye1}
\caption{Quantum Zeno effect for polar molecules in a 3D lattice.
\textbf{~a},~The lattice depths along $x$ and $z$ are kept at $40$~$E_R$, while the lattice depth along $y$ is reduced to allow tunneling along the $y$ direction at a rate $J_t/\hbar$. Once two molecules in different spin states tunnel to the same site, they are lost due to chemical reactions at a rate $\Gamma_0$.
\textbf{b},~Number loss of $|\downarrow \rangle$ state molecules versus time is shown for lattice depths along $y$ of $8.1$~$E_R$ and $15.1$~$E_R$.
\textbf{c},~The number loss rate $\kappa$ versus $\Gamma_0$ fits to a $1/\Gamma_0$ dependence, which is consistent with the quantum Zeno effect.
\textbf{d},~The number loss rate $\kappa$ versus $J_t$ fits to a $(J_t)^2$ dependence, as predicted from the quantum Zeno effect. Reprinted by permission from Macmillan Publishers Ltd: Nature, B.~Yan, S.~A.~Moses, B.~Gadway, J.~P.~Covey, K.~R.~A.~Hazzard, A.~M.~Rey, D.~S.~Jin, and J.~Ye, Nature {\bf 501}, 521 (2013), copyright 2013.} \label{fig:ye1}
\end{figure}
\begin{figure}[tb]
\centering
\includegraphics[width=\textwidth]{figures/rey1}
\caption{(a) Measured number loss of $\ket{\downarrow}$ molecules for an axial (transverse) lattice depth of $V_y=5\,E_R$ ($V_\perp=25\,E_R$) (circles) and best fit using a rate equation, $dN_{\downarrow}/dt=-\kappa N_{\downarrow}(0)/ [N_{\downarrow}(t)]^2$ (black dashed line).
(b) Number loss rate, $\kappa$, as a function of $\Gamma_0$ (fixing $J\approx570~$Hz and varying the bare on-site rate via $V_\perp$). (c) Number loss rate, $\kappa$, versus $J$ for fixed $\Gamma_0\approx87$~kHz (varying $V_y$ and adjusting $V_{\perp}$ accordingly). $V_y$ ($V_{\perp}$) was varied from 5 to $16\,E_R$ ($20$ to $40\,E_R$). Black circles are experimental measurements (error bars represent one standard error). Green short-dashed lines show solutions of the rate equation using an effective loss rate $\Gamma_{\rm{eff}}\propto J^2/\Gamma_0$ (single-band approximation). The blue long-dashed line shows the multiband rate equation using $\tilde{\Gamma}_{\rm{eff}}\propto \tilde{J}^2/\tilde{\Gamma}_0$, in which $\tilde J$ and $\tilde \Gamma_0$ are renormalized coefficents based on the effective multi-band theory. The multiband and single-band rate equation results were obtained by fixing the filling fraction to be $ 6\%$, and $25\%$ respectively. Panels (b) and (c) directly manifest the continuous quantum Zeno effect: in (b) the measured loss rate $\kappa$ decreases with increasing on-site $\Gamma_0$; in (c) a fit to the experimental data supports $\kappa\propto J^2$, with a $\chi^2$ (sum of the squared fitting errors) several times smaller than for a linear fit. Reprinted figure with permission from B.~Zhu et al., Phys. Rev. Lett. {\bf 112}, 070404 (2014). Copyright 2014 by the American Physical Society.} \label{fig:rey1}
\end{figure}
An identical treatment can also be performed to describe two-particle losses in ground-state molecules due to chemical reactions, as was recently done by Yan et al.~\cite{Yan2013} for the case of KRb molecules on a lattice, which can reach to form Rb$_2$ and K$_2$ molecules, again with a release of binding energy that is much larger than the lattice depth \cite{Yan2013,Zhu2014}. In Fig.~\ref{fig:ye1}, we reprint their results demonstrating the suppression of loss due to the quantum Zeno effect. In particular, by allowing tunnelling only along one direction in a 3D optical lattice, they make it possible to independently control the tunnelling $J$ and the on-site loss rate $\Gamma_0$ by controlling the lattice depth along the 1D axis and transverse to it \cite{Yan2013,Zhu2014}. The setup, initially involving KRb molecules in separate lattice sites, is depicted in Fig.~\ref{fig:ye1}a, and as a function of time leads to losses of molecules as depicted in Fig.~\ref{fig:ye1}b. By fitting a loss rate coefficient $\kappa$ based on the time-dependence of the molecule number $N_{\downarrow}(t)$,
\begin{equation}
\frac{dN_{\downarrow}}{dt}=-\kappa \frac{N_{\downarrow}(0)}{ [N_{\downarrow}(t)]^2},
\end{equation}
they can investigate the change in the effective loss rate as $J$ and $\Gamma_0$ are varied. The results shown in Figs.~\ref{fig:ye1}c,d demonstrate very clearly the suppression of loss as $\Gamma_0$ is increased, by showing that $\kappa\propto J^2/\Gamma_0$. The underlying mechanism for this is the quantum Zeno effect in the form discussed in section \ref{sec:zeno}, i.e., a strong dissipative loss process on a given site will suppress coherent tunnelling processes that populate sites with two particles, forming states that can potentially undergo loss.
As was noted above, the form we wrote based on the Bose-Hubbard model for particle loss in the lowest band of an optical lattice requires a series of assumptions, especially neglecting coupling to higher bands and off-site loss processes. In Ref.~\cite{Zhu2014}, a further theoretical analysis of the results of Yan et al.~\cite{Yan2013} was performed, taking into account additional terms involving higher bands and off-site loss processes. In Fig.~\ref{fig:rey1}, we reprint part of their theoretical comparison to experimental data, involving both the simple single-band loss model discussed above, and a more general multi-band model. They find that both models give quantitatively very similar results, and confirm that the quantum Zeno effect suppression of the losses survives these extensions of the model.
These effects naturally generalise to other systems undergoing two-body loss in which the energy released is larger than the trapping potential, and the resulting product particles are ejected from the system. For example, group-II atoms in metastable excited states would undergo similar loss processes. In Ref.~\cite{Foss-Feig2012}, Foss-Feig et al.~investigated this process for short-range s-wave collision processes. It was shown that this can lead to an interesting effect, because many-body states that are completely anti-symmetric under exchange of fermions in space do not undergo loss. If the Hamiltonian fulfils the requirement that spatially symmetric states can be eigenstates of the Hamiltonian (which is true, e.g., for two-species fermions in a harmonic trapping potential), then such states can be long lived. Then, if we begin with $N$ particles in the system in randomly chosen initial single-particle states, then of the order of $\sqrt{N}$ particles are expected to remain in a steady state. Moreover, because the resulting state is completely anti-symmetric in space, it will be completely symmetric in spin, and could be used for Heisenberg-limited spectroscopy. This results in a result for quantum-enhanced spectroscopy where the precision is not degraded by the loss of atoms, but collisional shifts are substantially reduced, potentially affording substantial improvements in accuracy \cite{Foss-Feig2012}.
Such collisional loss in group-II atoms has also been proposed as a means to implement dissipative blockade gates for quantum computing \cite{Daley2011,Daley2008b}, in which the strong loss plays the role of an effective interaction for the gate. Two-particle loss was also shown to generate pairing where Rydberg states are designed to be distance-selective \cite{Ates2012}.
Collective excitation dynamics including two-body loss have already been measured for group-II atoms in optical dipole traps \cite{Martin2013}. In these experiments, high-stability lasers make it possible to probe a system with relatively weak interactions using a Ramsey sequence on the clock transition, in such a way that the interactions become dominant. These experiments are also remarkable in that collective quantum dynamics emerge in the spin degree of freedom, despite relatively high motional temperatures for the atoms.
\subsubsection{Three-body loss}
While the suppression of two-body loss processes for molecules can be very important in stabilising quantum gases in experiments with molecules, and is of substantial fundamental interest, for many atomic species similar effects can be achieved by increasing the elastic two-body interactions (e.g., using Feshbach resonances \cite{Chin2010}). However, the potential to suppress three-body loss processes via a quantum Zeno effect opens very different opportunities, because it implies the possibility to produce effective three-body interactions, which are otherwise difficult to engineer in dilute quantum gases. Three-body loss for atoms involves collisional processes in which two particles form a molecule, and given that the resulting binding energy is large compared with the trap depth, the resulting molecule and atom will be ejected from the system. Such suppression of three-body loss events was first discussed for bosons in optical lattices in Ref.~\cite{Daley2009}, and the first evidence for such a suppression was recently obtained in experiments with Caesium atoms in an optical lattice in Ref.~\cite{Mark2012}.
\begin{figure}[tbh]
\begin{center}
\includegraphics[width=6cm]{figures/3bodyloss}
\end{center}
\caption{Schematic showing the continuous quantum Zeno effect in the case of three-body loss in an optical lattice. If we have a large on-site rate of loss $\gamma_3$, such that the tunnelling in the lowest Bloch band $J \ll \gamma_3$, then tunnelling processes that populate a single site with three particles (i.e., where a particle on one site tunnels onto an already doubly-occupied site) will be suppressed. In this form, large on-site loss rates can counterintuitively lead to a reduction in the overall rate of loss, and can act as an effective dynamical interaction, which suppresses triple occupation.}\label{fig:3bodyloss}
\end{figure}
The basic process, which is depicted in Fig.~\ref{fig:3bodyloss}, takes the same form as that for two-body loss. In a 3D optical lattice potential, if the on-site loss rate $\gamma_3$ for three atoms is very large, then this strong dissipative process can suppress coherent tunnelling that would otherwise populate triply occupied sites. In this form, both the effective resulting rate of three-body loss, and three-body occupation on a lattice site can be strongly suppressed. As with the case of two-body losses, in the limit that the binding energy released in the collision is much larger than the trap depth, and the products of the collision are immediately ejected from the lattice, the loss process can be described by a master equation. Restricting again to the lowest band of an optical lattice, the equation of motion for the system density operator $\rho$ is given by \cite{Daley2009}
\begin{equation}
\frac{d \rho}{dt} =- {\rm i} \left(H_{\rm eff} \rho - \rho H_{\rm eff}^\dag \right) +\frac{\gamma_3}{6} \sum_i a_i^3 \rho ( a_i^\dag)^3,
\label{meq3body}
\end{equation}
with the effective Hamiltonian given by
\begin{equation}
H_{\rm eff}=H_{\rm BH}-{\rm i} \frac{\gamma_3}{12} \sum_i ( a_i^\dag)^3 a_i^3,
\end{equation}
where $H_{\rm BH}$ is the Bose-Hubbard hamiltonian [eq.\eqref{HBH}], and $\gamma_3$ is three-body loss rate for three particles occupying a single lattice site.
\begin{figure}[t]
\par
\begin{center}
\includegraphics[width=8cm]{figures/diehl.pdf}
\end{center}
\par\vspace{-0.5cm}
\caption{Phase diagram for the Bose-Hubbard model with a three-body hard-core constraint, and $U<0$. The black curve
represents the mean-field phase border, while red (light gray) and blue (dark gray) curves include
shifts due to quantum fluctuations in $d=2,3$.
An Ising quantum critical point
is predicted in the vicinity of unit filling. The continuous supersolid is
reached asymptotically at unit filling. Reprinted figure from S.~Diehl, M.~Baranov, A.~J.~Daley, and P.~Zoller, Phys. Rev. Lett {\bf 104}, 165301 (2010). Copyright 2010 by the American Physical Society.
}
\label{fig:diehl}%
\end{figure}
In Ref.~\cite{Daley2009}, the dynamics is considered beginning in a subspace where no lattice sites are triply occupied. Analogously to the case of the two-level system in section \ref{sec:zeno}, in the limit $\gamma_3\gg J,U$ it is possible to perform perturbation theory on the effective Hamiltonian projected with an operator $P$ onto the subspace of states with at most two atoms per site. In second order-perturbation theory the effective model is given by the projected effective Hamiltonian \cite{Daley2009}
\begin{equation}
H_{\rm eff}^P \approx P H_{\rm BH} P - {\rm i} \frac{6J^2}{\gamma_3} P\sum_j c_j^{\dag} c_jP. \label{eq:model}
\end{equation}
Here the operator $ c_j = ( a_j^2/\sqrt{2}) \sum_{k\in {\mathcal N}_j} a_k$, with ${\mathcal N}_j$ denoting the set of nearest neighbours of site $j$. In this way, the term $PH_{\rm BH}P$ describes dynamics in a Bose-Hubbard Hamiltonian in the presence of the constraint $(a_i^\dag)^3\equiv 0$ \cite{Paredes2007}, and corrections to this description describe three-body recombination, which in the projected effective Hamiltonian begins from two particles present on a site neighbouring a single particle. These three-body loss events occur at a rate $\propto J^2/\gamma_3$, so that we see the continuous quantum Zeno effect again arising naturally in the limit $\gamma_3/J \gg 1$. For timescales that are shorter than the effective loss timescale, we are left with dynamics described by a constrained Bose-Hubbard model.
This constrained Bose-Hubbard model exhibits particularly interesting many-body properties \cite{Daley2009,Diehl2010a,Paredes2007}, beginning with the fact that the three-body constraint stabilises the system for attractive two-body interactions $U<0$. In the absence of the constraint, such attractive interactions would favour build-up of all bosons on one site, but with the maximum on-site occupancy limited to two, the system undergoes a second-order quantum phase transition. Using Gutzwiller mean-field techniques (Sec.~\ref{sec:gutzwiller}), this phase transition can be described as a transition from a usual atomic superfluid with mean-field parameter $\langle a_l \rangle\neq 0 $ (and also $\langle a_l^2\rangle\neq 0$) to a dimer superfluid, where $\langle a_l \rangle=0 $, but $\langle a_l^2\rangle\neq 0$. In Fig.~\ref{fig:diehl}, we reprint the phase diagram from Ref.~\cite{Diehl2010a}, showing the phase boundary at zero temperature in mean-field and including quantum fluctuations, as a function of $U/(Jz)$, where $z$ is the number of nearest neighbour sites, and $n$, the filling fraction in the system. This phase diagram has been investigated in further detail numerically and analytically \cite{Diehl2010b,Diehl2010d,Bonnes2011,Lee2010}, and exhibits an Ising critical point between the atomic superfluid and dimer superfluid phases near unit filling $n=1$. For stronger attractive interactions, bound dimers can form a phase with a coexistence of dimer superfluid and charge density wave order, in a type of continuous supersolid \cite{Diehl2010a,Diehl2010d}. Similar related three-body physics for bosons has been studied in a variety of contexts, including nearest-neighbour interactions that are realisable with polar molecules \cite{Capogrosso-Sansone2009,Buchler2007b, ChenYC2011,Ng2011}, and finite on-site three-body interactions \cite{Chen2008,Mazza2010,Silva-Valencia2011,Safavi-Naini2012,singh2012,Sowinski2012,Daley2013}.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=10cm]{figures/daley3body}
\end{center}
\caption{Dynamics of adiabatic ramps into a dimer superfluid regime. (a) We begin with (i) a Mott insulator state (ramping $U/J$), and (ii) a state with pre-prepared dimers in a superlattice (removing the superlattice). (b)-(c) The sum of kinetic ($E_K$) and interaction ($E_I$) energy and (inset) particle number as a function of time for two example trajectories, one with no loss events (dashed lines) and one with several loss events (solid lines). Here, (b) shows a ramp from $U/J=30$ to $U/J=-8$, with $U(t)=\alpha J/(100 + 3tJ)+\gamma$, with $\alpha$ and $\gamma$ ramp parameters, and (c) shows a ramp with a superlattice potential, $\varepsilon_l=V_0 \cos(2\pi l /3)$, where $V_0\approx30 J\exp(-0.1 tJ)$, adjusted so that $V_0(tJ=100)=0$, with fixed $U/J=-8$. In each case, $\gamma_3=250J$. For (b), we use 20 atoms on 20 lattice sites, for (c), 14 atoms on 23 lattice sites. (d) Plot showing the probability that no loss event has occurred after time $t$ for the ramps in (b) (dashed line) and (c) (solid). Reprinted figure from A.~J.~Daley, J.~M.~Taylor, S.~Diehl, M.~Baranov, and P.~Zoller, Phys. Rev. Lett {\bf 102}, 040402 (2009). Copyright 2009 by the American Physical Society.}\label{fig:daley3body}
\end{figure}
In an experiment real losses will also occur, making it important to model the complete dissipative many-body dynamics. For this purpose, in Ref.~\cite{Daley2009}, quantum trajectories techniques were first combined with t-DMRG to solve the master equation \eqref{meq3body}. In this paper, quantum trajectories were used to study the adiabatic preparation of a dimer superfluid state in the presence of three-body loss. Example results are shown in Fig.~\ref{fig:daley3body}, including ramps from an initial Mott Insulator state with a single atom per site (Fig.~\ref{fig:daley3body}b), and ramping in a superlattice as depicted in Fig.~\ref{fig:daley3body}a, using a site-dependent energy offset $\epsilon_l$, as was defined in $H_{\rm BH+}$ in section \ref{sec:qsim}. Figs~\ref{fig:daley3body}b,c show the total energy $\langle H_{\rm BH} \rangle$ and total particle number for example trajectories in a ramp that would adiabatically realise the ground state in the absence of losses. The dashed lines show trajectories without particle losses, whereas the solid lines show examples including losses. By propagating a state under the effective Hamiltonian, it is possible to determine for each ramp type what the probability of having no losses will be, which is depicted as a function of time in Fig.~\ref{fig:daley3body}d. These results clearly demonstrate advantages for the superlattice ramp, because the superlattice holds particles apart for a long time, and by beginning with a finite attractive interaction always includes elastic interactions to help suppress three-body occupation. The ramp from the Mott insulator, in contrast, passes through a regime of small coherent on-site interactions, where the loss is increased. In Ref.~\cite{Daley2009}, it is further shown that dimer superfluid order can survive some loss events, and that preparation of this state via the superlattice ramp should be realistic for experimentally realisable parameters, especially in Caesium.
These effects are also especially interesting for fermions in an optical lattice, where large three-body loss rates have been measured for three-component Lithium gasses \cite{Ottenstein2008}. There it is found that the three-body constraint induced for short times by the quantum Zeno effect will suppress formation of trimers for attractive interactions \cite{Kantian2009, Privitera2011,Titvinidze2011}, stabilising an atomic colour superfluid state with two-component BCS pairing in a three-component Fermi gas \cite{Rapp2008,Rapp2007}. Quantum Zeno suppression generating effective three-body interations has also been discussed for trapped gasses not in an optical lattice, including weakly interacting Bose-Einstein condensates \cite{Schutzhold2010}, and strongly interacting systems where Pfaffian-like states can be formed via the induced effective interactions \cite{Roncaglia2010}. Such effective interactions can also stabilise p-wave pairing in strongly interacting fermions \cite{Han2009}.
\subsubsection{Single-particle loss}
Single-particle atomic losses can be generated and controlled in a number of forms in an optical lattice, including photon scattering that drives atoms to momentum states with energy greater than the lattice depth. In certain regimes, this can lead to quantum optical analogues, in which the trapped atoms represent few-level systems, and atoms in untrapped states play the role of a radiation field \cite{Vega2008,Navarrete-Benlloch2011}. By tuning the laser field that drives atoms out of the lattice, super-radience and non-Markovian emission of atoms was predicted in Refs.~\cite{Vega2008,Navarrete-Benlloch2011}.
Local single-particle losses can be generated in experiments either by collisions with a background gas \cite{McKay2011}, or deliberately using individual addressing with an electron beam \cite{Gericke2008} or with light in quantum gas microscopes \cite{Weitenberg2011} that have single-site resolution in an optical lattice \cite{Bakr2009,Sherson2010}. As in the previous cases, the energy transferred to atoms that are removed from the lattice in this manner is much larger than the trap depth, and for these processes, a master equation analogous to eq.~\eqref{3bodylossmastereq}, but for single-particle loss can be derived \cite{Barmettler2011,Witthaut2011,Kordas2013,Vidanovic2014}. Theoretically, this master equation has been solved using quantum trajectory techniques with t-DMRG \cite{Barmettler2011} and a time-dependent Gutzwiller ansatz \cite{Vidanovic2014}. The dissipative dynamics produces complex effects, including the generation of entanglement under appropriate conditions \cite{Barmettler2011,Kordas2013}, and the ability to detect and control supersolid phases \cite{Vidanovic2014}.
\begin{figure}[tb]
\centering
\includegraphics[width=12cm]{figures/ott1}
\caption{a) The electrons locally collide with the atoms constantly dissipating the BEC. b) Temporal resolved signal from the ion detector. The bin size is 1 $\mu$s. Points are experimental data averaged over 1800 experimental repetitions, while the solid curve is the numerical simulation (see text). After 5 ms we typically collect $\simeq450$ ions. Reprinted figure with permission from G.~Barontini, R.~Labouvie, F.~Stubenrauch, A.~Vogler, V.~Guarrera, and H.~Ott, Phys. Rev. Lett. {\bf 110}, 035302 (2013). Copyright 2013 by the American Physical Society.} \label{fig:ott1}
\end{figure}
\begin{figure}[tb]
\centering
\includegraphics[width=12cm]{figures/ott3}
\caption{a) Theoretical curves of the number of ions produced in 5 ms as a function of $U/\hbar=\gamma_1(\mathbf{0})$ solving equation (\ref{dGPE}) for different values of the beam width $w$ \cite{Barontini2013}. The values of $U_M$ obtained using the approximate expression given in the text are shown as open diamonds over the corresponding curves. b) Number of ions measured after 5 ms of dissipation for a thermal cloud as a function of the EB current, with beam width $w=170 (7)$ nm. The solid line is the result of the corresponding numerical simulation using the molecular dynamics method. c) Comparison between the theoretical curves of the number of produced ions as a function of $I$ for the BEC and the \emph{corresponding} classical analogue (beam width $w=170 (7)$ nm) \cite{Barontini2013}. Reprinted figure with permission from G.~Barontini, R.~Labouvie, F.~Stubenrauch, A.~Vogler, V.~Guarrera, and H.~Ott, Phys. Rev. Lett. {\bf 110}, 035302 (2013). Copyright 2013 by the American Physical Society.} \label{fig:ott3}
\end{figure}
Recently, the quantum Zeno effect has been observed in experiments with Bose-Einstein condensates subjected to local losses due to an electron beam \cite{Barontini2013,Zezyulin2012}, in which atoms from the cloud are ionised and detected as depicted in Fig.~\ref{fig:ott1}a. In the weakly interacting case, the dynamics under the effective Hamiltonian can be modelled by a time-dependent Gross-Pittaevskii equation with an additional imaginary term \cite{Witthaut2008,Barontini2013},
\begin{equation}
i\hbar\frac{\partial\psi(\textbf{x},t)}{\partial t}=\left(-\frac{\hbar^2\nabla^2}{2m}+V_{ext}(\textbf{x})+g|\psi(\textbf{x},t)|^2-{\rm i}\hbar\frac{\gamma_1(\textbf{x})}{2}\right)\psi(\textbf{x},t).
\label{dGPE}
\end{equation}
Here $\psi(\textbf{x},t)$ is the condensate wavefunction, obeying the constraint $\int|\psi(\textbf{x},t)|^2d\textbf{x}=N(t)$, $V_{ext}(\textbf{x})$ is an external trapping potential, and $\gamma_1(\textbf{x})$ describes the profile of the single-particle loss in space. In Figs.~\ref{fig:ott1}b and \ref{fig:ott3}, we reprint experimental and theoretical results from Ref.~\cite{Barontini2013}. In Fig.~\ref{fig:ott3}a, the dependence of the number of ions detected as a function of the strength of the electron beam is plotted, and it can be seen that this clearly saturates in time, as a direct result of the quantum Zeno effect. This saturation is observed clearly in Ref.~\cite{Barontini2013}, and is contrasted to the case of a thermal cloud for which the dynamics are classical, and do not exhibit this effect is shown in Figs.~\ref{fig:ott3}b,c. These effects in weakly interacting systems have also been discussed theoretically for the case of a two-mode Bose-Einstein condensate in a double-well trap \cite{Witthaut2008}, and in the context of decoherence due to atom loss in bosonic Josephson junctions \cite{Pawlowski2013,Spehner2014}.
\subsection{State preparation in driven, dissipative many-body systems}
A key aspect of open systems as they appear in quantum optics involves the new tools to control quantum systems that are provided by dissipation. In particular, optical pumping \cite{Brossel1952,Hawkins1953} and laser cooling \cite{Metcalf1999} are dissipative processes that have become a key foundation for the majority of modern atomic physics experiments. The general idea behind these procedures is always to engineer a dissipative process in such a way that an arbitrary initial state of the system $\rho_0$ is driven via dissipation [e.g., as described by a master equation \eqref{mastereq}] into a steady state that is a chosen pure state $\ket{\Psi_s}$, $\rho(t\rightarrow \infty) \rightarrow \ket{\Psi_{s}}\bra{\Psi_s}$. In this process, entropy is removed from the system to the environment.
The idea of dissipative processes for state preparation and driving in quantum simulators has seen rapid recent progress. In this section, we will give a brief summary of some key directions, beginning with engineered dissipation in the presence of a reservoir gas, and going on to discuss implications of this for dissipative driving towards important many-body states. For several of these topics, further details can be found in the recent review on engineered dissipation and quantum simulation by M\"uller et al., Ref.~\cite{Muller2012}. The drive towards finding new means to cool many-body systems and prepare important many-body states is particularly important in light of current experimental challenges to reach lower temperatures in order to realise certain particularly interesting aspects of the many-body physics predicted for cold atoms in optical lattices \cite{McKay2011}.
\begin{figure}[tb]
\centering
\includegraphics[width=8cm]{figures/immersion}
\caption{Schematic picture of atoms in an optical lattice immersed in a Bose-Einstein condensate of another species that is not trapped by the lattice. Processes in which an atom in an excited band of the lattice decays to the lowest band are generated by spontaneous emission of Bogoliubov excitations in the condensate. This process is analogous to spontaneous emissions of photons, but takes place on much smaller energy and momentum scales, allowing for tools such as laser cooling and optical pumping to be reworked in a new context. Figure adapted from A.~Griessner, A.~J.~Daley, S.~R.~Clark, D.~Jaksch, and P.~Zoller, New J. Phys {\bf 9}, 44 (2007).} \label{fig:immersion}
\end{figure}
The example of engineered dissipation that is closest to the systems we have discussed up to now is the case in which dissipation is produced and controlled for cold atoms in an optical lattice by immersing the system in a reservoir gas of a second species of atoms \cite{Daley2004a}. This second species is not trapped by the lattice, and is usually chosen to be in a Bose-Einstein condensate, in a broad trapping potential so that the gas is almost uniform. In Refs.~\cite{Daley2004a,Griessner2006,Griessner2007} a master equation is derived for this system, making the assumption that the typical Bloch band separation in the lattice, $\omega_g$ is much larger than the corresponding energy scales in the reservoir gas, including its chemical potential, $\mu_R$. In this scenario, the quantum optics assumptions from section \ref{sec:approxamo} are satisfied because the Bloch band separation $\omega_g$ plays the role of the large energy and frequency scale, and the master equation is well-justified for realistic experimental parameters.
The resulting dissipative process is drawn schematically in Fig.~\ref{fig:immersion}. Atoms in higher bands of the lattice are sympathetically cooled via density-density interactions with atoms in the reservoir gas, with the atoms in the lattice decaying to the lowest band, while a Bogoliubov excitation is produced in the reservoir gas. As $\omega_g \gg \mu_R$, this excitation is typically a particle branch excitation of high energy \cite{Pitaevskii2003}, which can be allowed to rapidly leave the trap, ensuring that the Markov approximation remains valid \cite{Daley2004a}. Also, because the temperature of the reservoir gas, $T_R$ satisfies $k_B T_R\ll \mu_R \ll \omega_g$, where $k_B$ is the Boltzmann constant, excitations are not present to reheat lattice atoms to higher bands after they have decayed.
Moreover, while atoms in higher bands couple strongly to the reservoir, decaying on rapid timescales $\sim 1$kHz for typical parameters in an optical lattice and Bose-Einstein condensate of $^{87}$Rb, atoms in the lowest band couple very weakly, preventing re-heating of atoms within in the lowest band (of width $4J^0$). This occurs because of a mismatch in the dispersion relation of a Bose-Einstein condensate and atoms in the lowest band of the lattice, which in analogy to the Landau criterion for superfluidity \cite{Pitaevskii2003,Pethick2002}, prevents energy and momentum conservation when a single atom in the lattice interacts with a single excitation in the reservoir. Even in the case that these matched, the coupling constant would be small \cite{Griessner2006,Griessner2007}, due to the small structure factor of a Bose-Einstein condensate at low energies \cite{Pitaevskii2003,Pethick2002}. Both the decay of atoms in higher bands and the decoupling for atoms in the lowest band have recently been observed in experiments with Rb atoms in state-dependent traps \cite{Chen2014}. Similar effects have also been studied when motional superposition states are prepared in experiments using Li atoms immersed in a Bose-Einstein condensate of sodium \cite{Scelle2013}. In each of these experiments, the observed decay rates were in agreement with the assumptions made in the derivation of the master equation \cite{Daley2009}.
This combination of behaviours for the atoms in the lattice is strongly mathematically and physically analogous to spontaneous emissions into the vacuum modes of a radiation field. It immediately prompts the question as to whether ideas from quantum optics can be re-used in this context, but now with two-level atoms replaced by motional states or Bloch bands in the lattice, and photons being replaced by Bogoliubov excitations in the reservoir gas. By repeating such ideas on the smaller momentum and energy scales that are associated with excitations in a reservoir gas rather than photons, new possibilities for control and cooling to lower temperatures might then be realised.
\begin{figure}[tb]
\centering
\includegraphics[width=10cm]{figures/griessner}
\caption{Schematic picture of one excitation and decay step in Raman cooling in an optical lattice. (a) Atoms are transferred from a region with high quasi- momentum $|q| > 0$ in the lowest Bloch band to the first excited band. (b) The collisional interaction with the Bose-Einstein condensate atoms is switched on, the resulting decay of the excited lattice atoms leads to a randomization of the quasi-momentum. Sequences of pulses, each one followed by a decay time $\tau_c$, efficiently excite all atoms outside a narrow region around $q = 0$. Repeating the sequence leads to accumulation of atoms in the dark state region around $q = 0$, i.e., to cooling. Reprinted figure from A.~Griessner, A.~J.~Daley, S.~R..~Clark, D.~Jaksch, and P.~Zoller, New J. Phys {\bf 9}, 44 (2007). Copyright 2007 by IOP Publishing Ltd and Deutsche Physikalische Gesellschaft.} \label{fig:griessner}
\end{figure}
The first example of such a technique is a dark-state laser-cooling scheme based directly on an analogy with Raman dark-state cooling, as proposed by Kasevich and Chu \cite{Kasevich1992}. The key idea behind such schemes such as that of Ref.~\cite{Kasevich1992} and also velocity-selective coherent population trapping \cite{Aspect1988}, is that the desired state is engineered to be a \emph{dark state}, which does not couple to laser pulses that would transfer the atoms to a different internal state. A schematic diagram showing an analogous scheme for cooling non-interacting atoms within the lowest Bloch band of an optical lattice \cite{Griessner2006,Griessner2007} is shown in Fig.~\ref{fig:griessner}.
The cycle proceeds in two steps: (i) Raman pulses are applied to transfer atoms from the lowest Bloch band of the lattice to an excited Bloch band. Making use of the energy dependence on quasimomentum, together with control over the detuning and the time-dependence of the pulses, these are engineered to transfer only particles with large quasimomentum to the excited band -- in particular, a narrow dark state region near quasimomentum $q=0$ is deliberately not excited. (ii) Coupling to the reservoir gas (which can be left on if it is slower than the Raman pulse, but which can also be switched using Feshbach resonances for appropriate atomic species \cite{Chin2010}) is used to cause the atoms to decay to the lowest band. The quasimomentum will be redistributed in this process, and depending on the relative mass of atoms in the lattice and atoms in the reservoir gas, may be randomly distributed over the whole Brillouin zone \cite{Griessner2007}.
After repeating this cycle, atoms will begin to collect in the dark state region. Fundamentally, the cooling power of this method is determined only by the selectivity with which the dark state region is not excited, and the narrowness of this region. The final temperature can be much lower than that of the reservoir gas, because exactly as in regular laser-cooling, we are upconverting the energy to be removed from the system in such a way that it falls in an energy regime where there are no thermal excitations in the reservoir. This also clarifies that the decoupling of atoms in the lowest band from the reservoir gas is essential in this scheme, to prevent reheating of atoms already in the dark state.
There are many possible extensions that can be built upon this type of technique. In Ref.~\cite{Griessner2005}, a variant of this method for loading fermions in a state-dependent lattice was discussed. In this proposal, fault-tolerant loading of fermions in an optical lattice was investigated, in which particles are initially coupled to a higher Bloch band, and then cooled to the lowest band by creation of particle-hole pairs. In Refs.~\cite{Diehl2008,Kraus2008}, this was further extended to certain classes of many-body systems. By using additional internal atomic levels, schemes were devised to use a reservoir gas to implement a master equation with jump operators
\begin{equation}
c_m\equiv c_{\langle l,j\rangle} = (a_l^\dagger + a_j^\dagger)(a_l-a_j), \label{becjump}
\end{equation}
which results in a coherent driving of the system into a Bose-Einstein condensate on the lattice, i.e., is a designed master equation to produce a steady state $\propto (\sum_l a_l^\dagger)^N |{\rm vac}\rangle$. These ideas were also applied to implement a master equation for which the $\eta$-pairing states \cite{Yang1989} for two-species fermions are the steady state \cite{Diehl2008,Kraus2008}.
In Refs.~\cite{Diehl2010,Yi2012}, a master equation was found that -- with no direct or induced interactions between particles -- could produce either antiferromagnetic states, or pairing states for two-component fermions. In that case, the dissipative dynamics could be induced with the help of spontaneous emission processes in group-II atoms, which can exhibit Pauli blocking for fermions \cite{Sandner2011}. Quantum trajectory methods were employed together with exact diagonalisation methods (see section \ref{sec:numerics}) in order to demonstrate the time-dependent formation of the corresponding ordered states. Diehl, Bardyn, and their collaborators \cite{Diehl2011,Bardyn2013} also constructed master equations that would drive a system into topologically protected states. Specifically, they investigated atomic quantum wires coupled to a reservoir gas, and showed that with a dissipative drive, dissipative pairing gaps would enforce the isolation of Majorana edge modes \cite{Diehl2011}.
In this context, it is also both interesting and important to understand the competition between dissipation and coherent dynamics. Diehl et al.~\cite{Diehl2010c} treated the effect of interactions in the Hamiltonian on dissipative driving towards a Bose-Einstein condensate based on the jump operators in eq.~\eqref{becjump}, and found that competition between coherent and dissipative dynamics produced a mixed state in which the temperature was effectively controlled by the size of the interactions. Such competition can result further in dissipative phase transitions \cite{Diehl2010c,Sieberer2013a,Honing2012}, and some of the resulting transitions can form new types of universality classes for dissipative phase transitions, which can be studied with functional renormalization group methods \cite{Tauber2013,Sieberer2013}.
\subsection{Connections with other dissipative many-body systems}
\label{sec:othersys}
In addition to cold atoms and molecules in optical traps, there are a wide variety of AMO and solid state systems to which the ideas discussed in this review may be applied, or are already being applied. In this section we briefly summarise these corresponding developments.
A prime example of a system in which dissipative quantum simulation has already been implemented is the case of trapped ions \cite{Blatt2012,Muller2012}, which are also one of the most advanced systems in terms of control over single particles and few-particle coupling for quantum computing \cite{Haffner2008,Monroe2013}. Recent experiments have made use of this extreme control to demonstrate combinations of coherent and dissipative maps in digital quantum simulation \cite{Barreiro2011,Schindler2013}. In these experiments, qubits are prepared and then dynamical elements are applied as quantum gates, demonstrating -- amongst other attributes -- the dissipative preparation of entangled states, and simulations corresponding to many-body spin interactions. These experiments constitute a wonderful starting point for further high-precision studies of dissipative processes in many-particle systems. This is further enhanced by recent developments in realising analogue quantum simulation of spin chains, including transverse Ising and XY models with variable-range interactions, in these systems \cite{Porras2004,Friedenauer2008,Kim2009}. Dynamics in such setups were recently demonstrated in 1D chains \cite{Islam2013}, and also in 2D arrays \cite{Britton2012}. The possibility of combining these advances with control over dissipation is an exciting future direction. Controlled dissipative processes are also useful in a broader quantum computing context, where states can be protected in quantum computations \cite{Verstraete2009,Kliesch2011} or in quantum memories \cite{Pastawski2011} by carefully engineered dissipation.
Another system where dissipation is inherent in the most fundamental properties involves Bose-Einstein condensates in optical cavities \cite{Ritsch2013,Mekhov2012}. Recent observation of the Dicke phase transition \cite{Baumann2010,Dimer2007} and related optomechanical effects, including the detection of quantised motion and non-classical output light \cite{Brooks2012,Brahms2012} has opened new possibilities for studying many-body dissipative dynamics. This could include the realisation of more complicated collective dissipative spin models \cite{Morrison2008}, and quantum phases associated with the motion of atoms within the cavity \cite{Habibian2013,Torre2013,Buchhold2013}. Atoms in optical lattices within the cavity have also been discussed, including the potential use of cavity fields as a probe to detect detect atom number statistics \cite{Mekhov2007b,Mekhov2007}, or the use of cavity fields for state preparation \cite{Mekhov2009}. Bose-Hubbard dynamics in the presence of quantised light fields have also been discussed in Refs.~\cite{Maschler2008,Niedenzu2010}. Generalisations of this to other photon-atom systems have included studies of strongly interacting photons \cite{Peyronel2012,Reinhard2012,Petrosyan2011,Sevincli2011,Gorshkov2011}, dissipative state preparation and quantum simulation in optomechanical arrays \cite{Tomadin2012,Lechner2013}, and also the potential for photon condensation generated via engineered dissipation in a circuit QED system \cite{Marcos2012}.
There has similarly been a lot of recent interest in systems of atoms in excited Rydberg states, including possibilities to form crystaline states, and other long-range correlations. In the context of these excitations, several groups have become interested in the effect of decay of the Rydberg levels on the spatial patterns being realised. Their investigations have included the build-up of spatial correlations in the presence of dissipation \cite{Petrosyan2013,Honing2013}, the effects of non-local dissipation \cite{Olmos2013,Ates2012}, and studies of resulting glass-like dynamics and corresponding structures that emerge \cite{Olmos2012,Lesanovsky2013,Lesanovsky2014}.
Also going beyond these AMO systems, there have been several recent treatments of transport systems and dynamics of spin models with open quantum systems \cite{Prosen2014,Karevski2013,Prosen2011,Prosen2010,Znidaric2010,Prosen2008,Cai2013}, including applications of quantum trajectories to solve a master equation for spin dynamics. These studies have also employed matrix product state methods for such calculations \cite{Benenti2009,Prosen2009}. Quantum trajectory methods have also been used to investigate open system dynamics for ions traversing amorphous solids \cite{Minami2003,Seliger2007}. This is based on an approach with a master equation that allows probability flux for particles to be exchanged with the environment \cite{Seliger2005}.
\section{Summary and Outlook}
\label{sec:summary}
In this review we have seen how approaches from quantum optics to treating dynamics in open quantum systems has become important for our understanding of strongly interacting AMO systems. This opens many new opportunities to look at fundamental properties of open many-body systems, especially dynamics far from equilibrium, decoherence of states, and the back-action of continuous measurements. Not only is an understanding of dissipative dynamics necessary to treat processes such as light scattering and loss that are naturally present in quantum simulators, but the possibility to understand and engineer dissipative coupling provides an interesting route towards realisation of lower temperatures and important strongly-interacting many-body states, ranging from paired states of fermions to topological order.
At the same time, this is also a relatively new subfield with many open questions, ranging from specific questions such as the effect of spontaneous emissions on dynamical processes used to prepare states with cold atoms in optical lattices, to broader questions such as to what extent and under which conditions it is possible to identify universality classes for the steady-state of driven, dissipative many-body systems. While this review has focussed on markovian dissipative systems, which are prevalent in quantum optics and AMO implementations, one important area for future development will be the further development of these ideas in non-markovian regimes. Already there has been a great deal of progress on treating non-markovian effects in open systems \cite{Breuer2007,Breuer2012}, including through the extension of master equation techniques \cite{Breuer2004b,Breuer2007b} and stochastic wavefunction methods \cite{Breuer1999,Breuer2004,Breuer2004c} on expanded Hilbert spaces. In terms of quantum trajectories methods, there have been various extensions to the method to include memory effects of the reservoir for single-particle systems, beginning with work by Imamoglu in Ref.~\cite{Imamoglu1994}, which involved the coupling of a small quantum system to a damped cavity with a finite lifetime, and leading up to recent direct extensions to quantum trajectories methods on the Hilbert space of the reduced system \cite{Piilo2008,Piilo2009}. Recent studies have started combining this with Bose-Einstein condensates, including work where atoms are used as a probe immersed in the condensate \cite{Haikka2013,McEndoo2013}. Combined with strongly interacting systems, such ideas may open new pathways for investigations of fundamental dynamics, or provide practical tools for out-of-equilibrium preparation of many-body states as discussed in section \ref{amosys}. Extensions to non-markovian systems may also bring new opportunities for the study of solid-state systems via these techniques \cite{Breuer2007}.
In each of the ways described in this review, recent advances in AMO experiments and in theoretical tools are opening new opportunities to combine ideas from quantum optics and many-body physics. As discussed at the end of section \ref{sec:othersys}, numerical and analytical techniques with open quantum systems and quantum trajectories methods are also beginning to find applications not only in quantum simulators, but also outside the traditional AMO context. Understanding and even tailoring the dynamics of open systems will in this way lead to better control over a wide variety of systems, and to new possibilities to explore many-body physics in and out of equilibrium.
\section*{Acknowledgements}
I would also like to thank my colleagues and collaborators, Lars Bonnes, Sebastian Diehl, Adrian Kantian, Stephan Langer, Andreas L\"auchli, Hannes Pichler, Saubhik Sarkar, Johannes Schachenmayer, and Ulrich Schollw\"ock for insightful discussions on the application of these methods to many-body systems, and particularly Scott Parkins, who first introduced me to quantum trajectories methods and open quantum systems. I would also like to thank Anton Buyskikh, Sebastian Diehl, Stephan Langer, Saubhik Sarkar, Alexandre Tacla, and Guanglei Xu for providing comments on the manuscript, and Johannes Schachenmayer for providing the data from our project in Ref.~\cite{Schachenmayer2014} for the example in Fig.~\ref{fig:spementropy}. I would especially like to thank Peter Zoller for discussions and collaborations on open quantum systems the last 10 years -- the way of thinking about these systems that I learned from him and that I developed while working with him is strongly reflected in many parts of this review.
Work on dissipative systems in Pittsburgh is supported by AFOSR grant
FA9550-13-1-0093, and computational resources were provided by the Center for Simulation and Modeling at the University of Pittsburgh.
|
\section{Introduction}\label{intro}
In recent years, Heisenberg's uncertainty principle has received renewed
attention and scrutiny. The principle is often loosely associated with three sets
of ideas -- preparation uncertainty, joint measurement error trade-offs, and error-disturbance trade-offs.
While the first of these is uncontroversial, the latter two are subjects of an ongoing controversy.
For many decades, the only formally and operationally well-defined form of uncertainty relation known
in the physics literature was the familiar {\em preparation uncertainty relation} for
standard deviations of, say, position and momentum,
\begin{equation}\label{eqn:state-ur}
\Delta(Q,\rho)\,\Delta(P,\rho)\ge \tfrac{1}{2}\hbar.
\end{equation}
This relation is a statement about the widths of the probability distributions $\rho^Q,\rho^P$ of
the position $Q$ and momentum $P$ in a state $\rho$, and it can be
tested by measuring position and momentum in separate runs of experiments
on particles prepared in the same state $\rho$.
Notwithstanding this clear-cut interpretation, the relation (\ref{eqn:state-ur}) is often
paraphrased as constituting a limitation of the accuracies of any attempted joint
measurements of position and momentum. This unjustified conflation has equally often
been criticised, but then it happened not seldom that the critics (or their readers) jumped
to the conclusion that the uncertainty principle has nothing to do with the possibility or
impossibility of joint measurements of position and momentum.
The joint measurement uncertainty question was brought into focus with these conflicting views
but until recently a rigorous investigation of the problem has remained outstanding. The first seemingly plausible
attempt offered at quantifying measurement errors in quantum mechanics is based on the concept of
{\em noise operator} that was introduced into quantum optical amplifier theory in the 1960s and soon
after applied in the measurement context. For a brief history of the development of the notion of
{\em noise-(operator) based error}, we refer the reader to \cite{BLW2013a}. On the basis of this
{\em state-dependent} error measure it appeared that joint measurement error relations are much weaker than
suggested by the Heisenberg form (\ref{eqn:state-ur}); this has led to claims of a violation or circumvention
of Heisenberg-type measurement uncertainty relations, both theoretically (e.g., \cite{Ozawa03,Hall04})
and experimentally (e.g., \cite{Erh12,Roz12}). As shown in \cite{BLW2013a}, however, the noise-based error
measures do not purely quantify errors but also contain contributions of preparation uncertainty; moreover,
they are of limited operational significance as state-specific error measures.
In the meantime, different measures of measurement error were introduced that quantify the performance of
measuring devices and as such are state-independent. For these measures, joint measurement trade-off
relations have been formulated and proven. This development, which is reviewed in \cite{BuHeLa07}, was
enabled by making full use of the operational possibilities of quantum mechanics, notably by the generalised
representation of observables as positive operator valued measures
(POVMs).
A key concept for this solution to the joint-measurement problem is that of an
{\em approximate} measurement of a given observable (represented by a POVM) $\mathsf{E}$, which is any
measurement whose associated POVM $\mathsf{F}$ is close to $\mathsf{E}$ in a suitable operationally
relevant sense. This has made it possible to overcome the obstacle of the
noncommutativity of $\qhat$ and $\phat$, which precludes any {\em sharp}
joint measurement of these observables: there are (generally noncommuting)
{\em unsharp} observables $\mathsf{M}_1$, $\mathsf{M}_2$ that are jointly measurable and
still constitute reasonable approximations of $Q$ and $P$, respectively.
Two observables $\mathsf{M}_1,\mathsf{M}_2$ on $\bre$ are said to be {\em jointly measurable} if there
is a third, {\em joint observable} $\mathsf{M}$ on $\brr$ such that $\mathsf{M}_1,\mathsf{M}_2$ are the
Cartesian marginals of $\mathsf{M}$, $\mathsf{M}_1(X)=\mathsf{M}(X\times\R)$, $\mathsf{M}_2(Y)=\mathsf{M}(\R\times Y)$.
Three proposed measures of error for approximate measurements of position and momentum and their associated uncertainty relations were briefly reviewed in \cite{BuHeLa07}:
these were referred to as {\em standard error}, {\em (Monge) metric error}, and
{\em error bar width}. The first of these is what we called {\em noise-based error (measure)} above (due to the limitations of this concept it seems inappropriate to refer to it as ``standard''). In the meantime, measurement uncertainty relations have been proven for a wider class of metric error measures \cite{BLW2013b,BLW2013c}; these are based on the so-called Wasserstein distance of order $\alpha$ between probability measures on a metric space; here $\alpha$ is a parameter whose values range from 1 to $\infty$. The Monge metric corresponds to the value $\alpha=1$, while $\alpha=2$ is found to provide a natural operational quantum generalisation of the notion of root-mean-square (rms) error.
It is the purpose of the present paper to analyze further the concept of error bar width introduced in \cite{BuPe07} and to study its connections with the metric error measures. Some aspects of the noise operator based error will also be considered to the extent that they are useful as estimates of the other measures.
In addition to measurement errors, we also review other quantities describing the intrinsic unsharpness of the
approximators of $\qhat,\phat$.
The proofs of measurement uncertainty relations given in \cite{Werner04,BuHeLa07,BLW2013c} make it evident that the measurement error relations follow mathematically from related preparation uncertainty relations. Versions of these latter relations will be the starting points for the presentations of error and unsharpness measures to be given below.
We begin with a brief introduction of the requisite mathematical tools.
\section{Preliminaries}\label{sec:prelim}
Our considerations will be based on the usual description of a physical
system in a separable complex Hilbert space $\hi$, with states being represented
as positive operators $\rho$ of trace 1 (also called {\em density operators}). The convex set of all states will denoted
$S$. Pure states correspond to unit vectors $\varphi\in\hi$ or rather the associated rank-1
projections $\kb{\varphi}{\varphi}\equiv P_\varphi$. Observables are represented as positive operator measures (POVMs)
$\mathsf{E}$ on a measurable
space $(\Omega,\Sigma)$ that are normalised, i.e., $\mathsf{E}(\Omega)=I$. In this paper $(\Omega,\Sigma)$
will be one of the Borel spaces $(\R,\bre)$ or $(\R^2,\brr)$. An observable $\mathsf{E}$ is called
\emph{sharp} if it is projection valued; otherwise $\mathsf{E}$ is an \emph{unsharp}
observable. We write $\rho^\mathsf{E}$
for the probability measure induced by a state $\rho$ and an observable
$\mathsf{E}$ via the formula $\rho^\mathsf{E}(X):=\tra{\rho \mathsf{E}(X)}$, $X\in \Sigma$.
We use the notation $\mathsf{E}[x^k]$, $k\in\N$, for the $k^{\mathrm{th}}$ moment operators
$\int x^k\mathsf{E}(dx)$ of an observable $\mathsf{E}$ on $\bre$. These operators are defined
on their natural domains \cite{lahti1998moment}
$D(\mathsf{E}[x^k])$ of all $\varphi\in\hi$ for which the function $x\mapsto x^k$
is integrable with respect to the complex measures $\ip{\psi}{\mathsf{E}(dx)\varphi}$ for all
$\psi\in\hi$; this contains the square-integrability domain
$\{\varphi\in\hi\,:\, \int x^{2k}\ip{\varphi}{\mathsf{E}(dx)\varphi}<\infty\}$.
The moments
of a probability measure $\prob$ on $\R$ will be denoted $\prob[x^k]$, $k\in\N$.
All uncertainty relations to be studied here will be formulated for the case of
a quantum particle in one spatial dimension, with Hilbert space $\hi=L^2(\R)$
and canonical position and momentum operators $\qhat,\phat$, defined
via $(\qhat\psi)(x)=x\psi(x)$, $(\phat\psi)(x)=-i\hbar (d\psi/dx)(x)$ on the usual
maximal domains ensuring selfadjointness. Generalizations to more degrees of
freedom are straightforward. The spectral measures of $\qhat$ and $\phat$ will
be denoted $\mathsf{Q}$ and $\mathsf{P}$, respectively.
An important class of POVMs representing approximations of position and momentum
are given by {\em smeared} position and momentum observables $\mathsf{Q}^\mu,\mathsf{P}^\nu$,
defined as convolutions of $\mathsf{Q},\mathsf{P}$ with probability measures $\mu,\nu$ on $\bre$:
\begin{eqnarray}
\mathsf{Q}^\mu(X)&=\mathsf{Q}*\mu(X)=\int_\R\mu(X-q)\,\mathsf{Q}(dq),\nonumber\\
\mathsf{P}^\nu(Y)&=\mathsf{P}*\nu(Y)=\int_\R\nu(Y-p)\,\mathsf{P}(dp).\nonumber
\end{eqnarray}
The integrals are defined in the weak operator topology.
We will make use of the important class of covariant phase space observables which
is defined as follows.
Let $W(q,p)=\exp(iqp/2\hbar)\exp(-iq\phat/\hbar)\exp(ip\qhat/\hbar )$ be the Weyl operators
comprising an irreducible unitary projective representation of
the translations on phase space $\R^2$.
An observable $\mathsf{G}$ on $\R^2$ is called a \emph{covariant phase space observable} if it satisfies the covariance condition
\begin{equation*
W(q,p)\mathsf{G}(Z)W(q,p)^*=\mathsf{G}(Z-(q,p)),\quad Z\in\brr.\nonumber
\end{equation*}
This is satisfied by the following family of observables $\mathsf{G}=\mathsf{G}^\tau}%{{\mathbf m}$
on $\R^2$, which are thus covariant phase space observables:
\begin{equation}\label{gen-cov-obs}
\mathcal{B}(\R^2)\ni Z\mapsto \mathsf{G}^\tau}%{{\mathbf m}(Z)= \frac 1{2\pi\hbar}\int_Z W(q,p)\tau}%{{\mathbf m} W(q,p)^*dq\,dp;
\end{equation}
the integral is defined in the weak operator topology and the operator density is
generated by an arbitrary fixed positive operator $\tau}%{{\mathbf m}$ of trace 1
(for details of the proof of these properties, see, e.g., \cite{CRQM}).
Moreover, \emph{every} covariant phase space
observable is of the form (\ref{gen-cov-obs}) for some positive operator $\tau$ of trace 1. This fundamental fact
is implied by results of \cite{PSAQT} and \cite{Werner1984} and has
been made explicit in \cite{CaDeTo03}
using the theory of induced representations and in
\cite{KiLaYl06a} using the theory of integration with respect to
operator measures.
The marginal observables of $\mathsf{G}^\tau}%{{\mathbf m}$ are smeared position and momentum
observables $\mathsf{Q}^{\mu_\tau}%{{\mathbf m}},\mathsf{P}^{\nu_\tau}%{{\mathbf m}}$, where
$\mu_\tau}%{{\mathbf m}:=\tau}%{{\mathbf m}_\Pi^\mathsf{Q}$ and $\nu_\tau}%{{\mathbf m}:=\tau}%{{\mathbf m}_\Pi^\mathsf{P}$ are the probability distributions of $Q$ and $P$
in the state described by $\tau}%{{\mathbf m}_\Pi^{\phantom{P}}$, that is,
\begin{equation*}
\mathsf{G}^\tau}%{{\mathbf m}_1=\mathsf{Q}*\mu_\tau}%{{\mathbf m},\quad \mathsf{G}^\tau}%{{\mathbf m}_2=\mathsf{P}*\nu_\tau}%{{\mathbf m}.
\end{equation*}
Here $\tau}%{{\mathbf m}_\Pi^{\phantom{P}}=\Pi\tau}%{{\mathbf m}\Pi^*$ is the operator obtained from $\tau}%{{\mathbf m}$
under the action of the parity transformation $\Pi$
($\Pi\varphi(x)=\varphi(-x)$).
There is a simple but fundamental characterization of all pairs of smeared position
and momentum observables admitting a joint measurement.
\begin{theorem}\label{thm:jm-smeared-qp}
A pair of smeared position and momentum observables $\mathsf{Q}^\mu,\mathsf{P}^\nu$ are
jointly measurable exactly when there exists a covariant phase space observable
$\mathsf{G}^\tau}%{{\mathbf m}$ of which they are marginals. In that case, $\mu=\mu_\tau}%{{\mathbf m}$, $\nu=\nu_\tau}%{{\mathbf m}$.
\end{theorem}
This result has been obtained in a long series of investigations
by various authors, culminating and summarised in \cite{CaHeTo05}.
\section{Uncertainty: $\boldmath\alpha$-deviation and overall width}\label{sec:uncertainty}
We will make use of the following measures of the widths of a probability distribution
$\prob:\bre\to[0,1]$ on $\R$.
The \emph{standard deviation} $\Delta(\prob)$ is given by
\begin{equation*}
\Delta(\prob):=\left(\int \left(x-\int x\prob(dx)\right)^2\prob(dx)\right)^{1/2}=
\left(\prob[x^2]-\prob[x]^2\right)^{1/2}.
\end{equation*}
The standard deviation of an observable $\mathsf{E}$ on $\bre$ in a state $\rho$ is
$\Delta(\mathsf{E},\rho):=\Delta(\rho^\mathsf{E})$. For vector states $\varphi$ we write
$\Delta(\mathsf{E},\varphi):=\Delta(\prob^\mathsf{E}_\varphi)$.
The standard deviation is a special case of the so-called (Wasserstein) $\alpha$-deviation:
\begin{equation*}
\Delta_\alpha(\prob):=\inf_y\left(\int \left|x-y\right|^\alpha\prob(dx)\right)^{1/\alpha},\quad 1\le\alpha<\infty.
\end{equation*}
The uncertainty relation for the standard deviations of position and momentum has recently been generalised to $\alpha$-deviation \cite{BLW2013b}.
\begin{theorem}[Preparation Uncertainty]\label{thm:prepURpq}
Let $\mathsf{Q}$ and $\mathsf{P}$ be canonically conjugate position and momentum observables, and $\rho$ a density operator. Then, for any $1\leq \alpha,\beta<\infty$,
\begin{equation*
\Delta_\alpha(\rho^\mathsf{Q})\Delta_\beta(\rho^\mathsf{P})\geq c_{\alpha\mkern1mu\beta}\hbar,
\end{equation*}
The constant $c_{\alpha\mkern1mu\beta}$ is connected to the ground state energy $g_{\alpha\mkern1mu\beta}$ of the Hamiltonian $H_{\alpha\mkern1mu\beta}=|Q|^\alpha+|P|^\beta$ by the equation
\begin{equation*
c_{\alpha\mkern1mu\beta} = \alpha^{\frac 1\beta}\beta^{\frac 1\alpha}\left(\frac{g_{\alpha\mkern1mu\beta}}{\alpha+\beta}\right)^{\frac 1\alpha+\frac 1\beta}.
\end{equation*}
The lower bound is attained exactly when $\rho$ arises from the ground state of the operator $H_{\alpha\mkern1mu\beta}$ by phase space translation and dilatation.
For $\alpha=\beta=2$, $H_{\alpha\mkern1mu\beta}$ is twice the harmonic oscillator Hamiltonian with ground state energy $g_{22}=1$, and $c_{22}=1/2$.
\end{theorem}
The \emph{overall width} of $\prob$ (at confidence level $1-\varepsilon$) is defined for $\varepsilon\in[0,1)$ as
\begin{equation*}
W_\varepsilon(\prob):=\inf\{w>0\,|\,\exists x\in\R: \,\prob([x-\tfrac w2,x+\tfrac w2])\ge 1-\varepsilon\}.
\end{equation*}
This quantity is finite for any $\varepsilon>0$. For the overall width of the distribution of an observable
$\mathsf{E}$ on $\bre$ in a state $\rho$ we will write $\Delta_\varepsilon(\mathsf{E},\rho):=W_\varepsilon(\rho^\mathsf{E})$. This describes
the extent to which the quantity described by $\mathsf{E}$ can be approximately localised. As shown in
\cite{UH1985}, the overall width is generally a more stringent measure of the spread of a distribution
than the standard deviation.
In analogy to the uncertainty relation (\ref{eqn:state-ur}) for standard deviations,
the overall widths of the position and momentum distributions in a state $\rho$
also satisfy a trade-off relation: for positive
$\varepsilon_1,\varepsilon_2$ satisfying $\varepsilon_1+\varepsilon_2<1$ there is a constant $K(\varepsilon_1,\varepsilon_2)>0$ such that
\begin{equation}\label{eqn:ow-ur}
W_{\varepsilon_1}(\mathsf{Q},\rho)\cdot W_{\varepsilon_2}(\mathsf{P},\rho)\ge 2\pi\hbar\,K(\varepsilon_1,\varepsilon_2).
\end{equation}
An uncertainty relations of this form was first presented in a somewhat implicit way
in the context of signal analysis by Landau and Pollak in 1961 \cite{LaPo61}.
Its explicit form was given by Uffink in 1990 \cite{Uffink90}:
\begin{equation}\label{eqn:ow-U}
K(\varepsilon_1,\varepsilon_2)=\left(\sqrt{(1-\varepsilon_1)(1-\varepsilon_2)}-\sqrt{\varepsilon_1\varepsilon_2}\right)^2.
\end{equation}
A somewhat simpler (but weaker) bound was given in \cite{BuHeLa07}
using elementary arguments:
\begin{equation*
\widetilde{K}(\varepsilon_1,\varepsilon_2)=\left(1-(\varepsilon_1+\varepsilon_2)\right)^2 \le K(\varepsilon_1,\varepsilon_2).
\end{equation*}
Note that the two expressions coincide on the `diagonal':
\begin{equation*}
K(\varepsilon,\varepsilon)=\widetilde{K}(\varepsilon,\varepsilon)=(1-2\varepsilon)^2.
\end{equation*}
\section{Intrinsic unsharpness: resolution width}\label{sec:unsharpness}
For two noncommuting observables to be jointly measurable, it is necessary that they are unsharp.
One expects intuitively that the required degree of their unsharpness depends on the extent of
their noncommutativity.
We will see that two unsharp observables which approximate
position and momentum, respectively, cannot have arbitrarily small degrees of unsharpness if
they are to be jointly measurable. We will use the following measures as indicators of the
unsharpness of an observable $\mathsf{E}$ on $\R$.
For an observable $\mathsf{E}$ with support ${\rm supp}(\mathsf{E})$ given by $\R$ or a closed interval, the \emph{resolution width} (at confidence level
$1-\varepsilon$) is defined as \cite{CaHeTo06}:
\begin{equation*}
\gamma_\varepsilon(\mathsf{E}):=\inf\{ w>0\,| \forall x\in \R\,\exists\rho\in S:
\rho^\mathsf{E}([x-\tfrac w 2,x+\tfrac w 2])\ge 1-\varepsilon \}.\
\end{equation*}
For a sharp observable $\mathsf{E}$ on $\bre$
the resolution width
is $\gamma_\varepsilon(\mathsf{E})=0$ for all $\varepsilon\in(0,1)$. It is worth noting that vanishing resolution width does not
require the observable to be sharp: in fact, any observable whose nonzero effects have norm 1
has zero resolution width; an example is given by the so-called canonical phase observable \cite{Heinonen-etal2003}.
For the resolution width of $\mathsf{Q}^\mu,\mathsf{P}^\nu$ we obtain (see also \cite{CaHeTo06}):
\begin{equation*}
\gamma_{\varepsilon_1}(\mathsf{Q}^\mu)=W_{\varepsilon_1}(\mu),\qquad
\gamma_{\varepsilon_2}(\mathsf{P}^\nu)=W_{\varepsilon_2}(\nu).
\end{equation*}
If a pair of observables $\mathsf{Q}^\mu$, $\mathsf{P}^\nu$ is jointly measurable,
their resolution widths are determined
by the probability measures $\mu=\mu_\tau}%{{\mathbf m},\nu=\nu_\tau}%{{\mathbf m}$ which
obey the uncertainty relations (\ref{eqn:state-ur}) and (\ref{eqn:ow-ur}); we thus obtain:
\begin{equation*}
\gamma_{\varepsilon_1}(\mathsf{Q}^{\mu_\tau}%{{\mathbf m}})\,\gamma_{\varepsilon_2}(\mathsf{P}^{\nu_\tau}%{{\mathbf m}})=
W_{\varepsilon_1}(\mathsf{Q},\tau}%{{\mathbf m})\, W_{\varepsilon_2}(\mathsf{P},\tau}%{{\mathbf m})\ge 2\pi\hbar\,
K(\varepsilon_1,\varepsilon_2).\label{eqn:gam-gam-cov-ur}
\end{equation*}
The last inequality holds for any $\varepsilon_1,\varepsilon_2>0$ with $\varepsilon_1+\varepsilon_2<1$.
\section{Error measures I: Distance between observables}\label{sec:dist}
We review three distinct measures of error which quantify the difference
between an observable $\mathsf{E}$ on $\bre$ to be approximated and the approximator $\mathsf{F}$,
which is also a POVM on $\bre$. Any error measure should be \emph{operationally significant} in the sense that it quantifies the difference between the distributions $\rho^{\mathsf{F}}$ and $\rho^{\mathsf{E}}$. We begin with a family of metric error measures.
\subsection{Wasserstein $\alpha$-distance: definition.}
Next we briefly review a family of distances on the set of observables on $\R$ that was used in \cite{BLW2013b}
to formulate measurement uncertainty relations for canonically conjugate pairs of observables such as position and momentum. We adapt the presentation given there for general metric spaces to the case of $\R$.
For any two probability measures $\mu,\nu$ on $\R$ a {\it coupling} is defined to be a probability measure $\gamma$ on $\R\times \R$ with $\mu$ and $\nu$ as the Cartesian marginals.
The set of couplings between $\mu$ and $\nu$ will be denoted $\Gamma(\mu,\nu)$. Then, for any $\alpha$, $1\leq \alpha<\infty$ the $\alpha$-distance (also Wasserstein $\alpha$-distance \cite{Villani}) of $\mu$ and $\nu$ is defined as
\begin{equation}\label{Wasserstein}
\mathcal{D}_\alpha(\mu,\nu)= \inf_{\gamma\in\Gamma(\mu,\nu)} \mathcal{D}^\gamma_\alpha(\mu,\nu)=
\inf_{\gamma\in\Gamma(\mu,\nu)}\left(\int |x-y|^\alpha\,d\gamma(x,y) \right)^{\frac1\alpha}
\end{equation}
For $\alpha=\infty$, one defines $\mathcal{D}^\gamma_\infty(\mu,\nu)=\gamma-{\rm ess}\ \sup\{|x-y|\,|\, (x,y)\in\R\times\R\}$ and thus
\begin{equation}\label{Wassersteininfty}
\mathcal{D}_\infty(\mu,\nu) = \inf_{\gamma\in\Gamma(\mu,\nu)} \mathcal{D}^\gamma_\infty(\mu,\nu).
\end{equation}
It turns out that $\mathcal{D}^\gamma_\infty(\mu,\nu)$ actually depends only on the support of $\gamma$, that is, $\mathcal{D}^\gamma_\infty(\mu,\nu)=\sup\{|x-y|\,|\, (x,y)\in{\mathop{\rm supp}\nolimits}\,(\gamma)\}$.
The existence of an optimal coupling is known, for $1\leq\alpha<\infty$, see \cite[Theorem 4.1]{Villani}, the case $\alpha=\infty$ is shown in \cite[Theorem 2.6]{Jylha2014}, but it does not imply that $D_\alpha(\mu,\nu)$ is finite.
When $\nu=\delta_y$ is a point measure, there is only one coupling between $\mu$ and $\nu$, namely the product measure $\gamma=\mu\times\delta_y$. In that case (\ref{Wasserstein}) and (\ref{Wassersteininfty}) describe the deviation of the measure $\mu$ from a point $y$. In particular, the Wasserstein distances between point measures are seen to be extensions of the given metric for points, interpreted as point measures.
The metric can become infinite, but the triangle inequality still holds \cite[after Example~6.3]{Villani}. The proof relies on Minkowski's inequality and the use of a ``Gluing Lemma'' \cite{Villani}, which builds a coupling from $\mu$ to $\zeta$ out of couplings from $\mu$ to $\nu$ and from $\nu$ to $\zeta$. It also covers the case $\alpha=\infty$, which is not otherwise treated in \cite{Villani}.
We can now define the \emph{(Wasserstein) $\alpha$-distance between observables} $\mathsf{E},\mathsf{F}$ on $\R$:
\begin{equation*}
\Delta_\alpha(\mathsf{E},\mathsf{F}):=\sup_{\rho\in S}\mathcal{D}_\alpha(\rho^\mathsf{E},\rho^\mathsf{F}).
\end{equation*}
These distances are operationally significant and global error measures,
taking into account the largest possible deviations between corresponding probability measures of the
observables being compared.
\subsection{Working with Wasserstein distances: Kantorovich duality.}
A powerful tool for working with the distance functions is a dual expression of the infimum over couplings as a supremum over certain other functions obtained by the Kantorovich duality.
In this context we exclude the case $\alpha=\infty$.
First we note that the ``gap inequality''
\begin{equation}\label{gap}
\int\Phi(y)\,d\nu(y) \ -\ \int\Psi(x)\,d\mu(x) \leq \int |x-y|^\alpha\,d\gamma(x,y)
\end{equation}
holds for any pair of functions $(\Psi,\Phi)$ and any coupling $\gamma$ whenever the constraint
\begin{equation}\label{compete}
\Phi(y)-\Psi(x)\leq |x-y|^\alpha
\end{equation}
is satisfied.
The Kantorovich Duality Theorem asserts that the gap is actually closed:
\begin{equation}\label{DKanto}
\mathcal{D}_\alpha(\mu,\nu)^\alpha=\sup_{\Phi,\Psi} \left\{ \int\Phi(y)\,d\nu(y) - \int\Psi(x)\,d\mu(x)\right\}
\end{equation}
where functions $\Phi$ and $\Psi$ satisfy (\ref{compete}).
When maximizing the left hand side of (\ref{gap}), one can naturally choose $\Phi$ as large as possible under the constraint (\ref{compete}), i.e.,
$\Phi(y)=\inf_x\{\Psi(x)+|x-y|^\alpha \}$, and similarly for $\Psi$. Hence one can choose just one variable $\Phi$ or $\Psi$ and determine the other by this
formula. In the case $\alpha=1$ the triangle inequality for the metric on $\R$ entails that one can take $\Phi=\Psi$. In this case (\ref{compete}) just asserts that
this function is Lipshitz continuous with respect to the metric on $\R$, with constant $1$. The left hand side of (\ref{gap}) is thus a difference of
expectation values of the given measures $\mu$, $\nu$.
It is of interest to note that the duality gap still closes if the set of functions $\Phi,\Psi$ is further restricted.
The natural condition is, first of all, that $\Psi\in L^1(\mu)$. The statement of Kantorovich Duality in \cite[Thm.~5.10]{Villani}
includes that the supremum (\ref{DKanto}) is attained also when one restricts the set of functions to bounded continuous functions.
In \cite{BLW2013b} it is shown that this set can be further restricted
to positive continuous functions of compact support without changing the value of the supremum.
\subsection{Properties of the 1-distance.}
We now specialise to the case of the Wasserstein 1-distance ($\alpha=1$), also known as the Monge metric. This was the choice of metric for the first formulation of a rigorous measurement uncertainty relation for position and momentum in \cite{Werner04}.
Denoting by $\Lambda$ the set of Lipshitz functions, that is, the bounded measurable functions $h:\R\to\R$ for which $|h(x)-h(y)|\leq |x-y|$,
the Wasserstein-1 distance $\mathcal{D}_1$ then becomes \cite{Werner04}
\[
\mathcal{D}_1(\mu,\nu)=\sup_{h\in\Lambda}\,\left|\int hd\mu-\int hd\nu\right|.
\]
This gives rise to a metric on the set of observables on $\bre$ as follows.
We first recall that for any bounded measurable
function $h:\R\to\R$, the integral $\int_\R h\,d\mathsf{E}$ defines (in the weak
sense) a bounded selfadjoint operator, which we denote by $\mathsf{E}[h]$.
Thus, for any vector state $\varphi$ the number $\ip{\varphi}{\mathsf{E}[h]\varphi}=\int_\R h\,d\ip\varphi{\mathsf{E}(x)\varphi}$ is well-defined.
The Wasserstein 1-distance between observables $\mathsf{E}$ and $\mathsf{F}$ can then be expressed as
\begin{equation*
\Delta_1(\mathsf{E},\mathsf{F}) := \sup_{\rho\in S}\, \sup_{h\in\Lambda}\,
\biggl| \tr\left[\rho\,( \mathsf{E}[h]-\mathsf{F}[h]) \right]\biggr|=\sup_{h\in\Lambda}\,\biggl\|{\mathsf{E}[h]-\mathsf{F}[h]}\biggr\|.
\end{equation*}
An observable $\mathsf{C}$ will be called a \emph{metric approximation} to
$\mathsf{A}$ if $\Delta_1(\mathsf{A},\mathsf{C})<\infty$.
\begin{example}\label{ex:triv-dist}
Any trivial observables $\mathsf{E}=\mu\,I$ on $\R$ has infinite
distance from sharp position $\mathsf{Q}$: $d(\mu\,I,\mathsf{Q})=\infty$.
Take the family of functions $h_n(x)=n-|x-c_n-n|$ if $|x-c_n-n|\le
n$, and $h_n(x)=0$ otherwise; here $(c_n)_{n\in\N}$ is an increasing
sequence of positive numbers still to be determined. Note that
$h_n\in\Lambda$.
We have $||\mathsf{Q}[h_n]||=h_n(c_n+n)=n$, so this approaches infinity as
$n\to\infty$.
For a trivial observable $\mathsf{E}$ we get
$\mathsf{E}[h_n]=\int h_nd\mu\,I=:\mu(h_n)\,I$.
We show that for a suitable choice of the sequence $c_n$, one obtains
$\|\mathsf{E}[h_n]\|=\mu(h_n) \to 0$ as $n\to\infty$.
Let $c_n$ be such that the set $K_n=(-\infty,c_n]$ has measure
$\mu(K_n)> 1-1/n^2$, so that $\mu(\R\setminus
K_n)=\mu((c_n,\infty))<1/n^2$. Then
\begin{equation*
\|\mathsf{E}[h_n]\|=\mu(h_n)=\int_{c_n}^{c_n+2n} h_n(x)\mu(dx)
\le n\mu((c_n,\infty))<1/n.
\end{equation*}
By the triangle inequality for norms we get
\[
\|\mathsf{Q}(h_n)-\mathsf{E}(h_n)\|\ge\|\mathsf{Q}(h_n)\|-\|\mathsf{E}(h_n)\|> n-1/n.
\]
It follows that the distance $\Delta_1(\mathsf{E},\mathsf{Q})=\infty$. \qed
\end{example}
Our next example exhibits functions of position $\mathsf{Q}$ that may or may not
be good metric approximations to $\mathsf{Q}$.
\begin{example}\label{ex:bdd-dist}
Let $g$ be a bounded measurable function on $\R$. Then the distance
of $\mathsf{Q}$ and $\mathsf{Q}\circ g^{-1}$ is infinite, $\Delta_1(\mathsf{Q},\mathsf{Q}\circ g^{-1})=\infty$. For a function $f(x)=x+g(x)$ where
$g$ is bounded, the distance is finite: $\Delta_1(\mathsf{Q},\mathsf{Q}\circ f^{-1})=\sup(|g|)$.
\emph{Proof.} We will show that $||\mathsf{Q}[h_n]-\mathsf{Q}\circ g^{-1}[h_n]||\to\infty$ as $n\to\infty$
for a suitable sequence of functions $h_n\in\Lambda$. To this end we
use the inequality
$$
||\mathsf{Q}[h_n]-\mathsf{Q}\circ g^{-1}[h_n]||\ge \bigg| ||\mathsf{Q}[h_n]|| -
||\mathsf{Q}\circ g^{-1}[h_n]|| \bigg|,
$$
and choose $h_n$ such that
$||\mathsf{Q}[h_n]||\to\infty$ as $n\to\infty$, while $||\mathsf{Q}\circ g^{-1}[h_n]||$ will
remain bounded.
Let $|g(x)|\le g_0$.
Choose $h_n(x)=n-|x-n|$ if $|x-n|\le n$ and $h_n(x)=0$ otherwise.
Then we have $h_n\in\Lambda$. Further,
$\mathsf{Q}[h_n]=\int h_n(x) \mathsf{Q}(dx)$, so $\no{\mathsf{Q}[h_n]}=n\to\infty$ as $n\to\infty$.
Next, we see that for $n > g_0$
\begin{equation*}
\mathsf{Q}\circ g^{-1}[h_n]=\int h_n(t)\mathsf{Q}\circ g^{-1}(dt)=\int h_n(g(x))\mathsf{Q}(dx)
\end{equation*}
is a bounded operator since then $|h_n(g(x))|\le g_0$, and so
$\no{\mathsf{Q}\circ g^{-1}[h_n]}\le g_0$.
To verify the second claim, we note that ${\mathsf{Q}[h]-\mathsf{Q}\circ f^{-1}[h]}=h(Q)-h(f(Q))$, and so for $h\in\Lambda$ and unit vector $\varphi$,
\begin{eqnarray*}
\ip\varphi{\bigl[h(Q)-h(f(Q))\bigr]^2\varphi}&=&\int|\varphi(x)|^2\left[h(x)-h(f(x))\right]^2dx\\
&\le&\int|\varphi(x)|^2\left(x-f(x)\right)^2dx\\
&=&
\int|\varphi(x)|^2g(x)^2dx\le\no{g(Q)}^2=(\sup|g|)^2,
\end{eqnarray*}
from which the claim follows. \qed
\end{example}
\begin{example}\label{ex:approx-pos-dist}
For smeared position and momentum observables $\mathsf{Q}^\mu$, $\mathsf{P}^\nu$,
the distances from $\mathsf{Q}$ and $\mathsf{P}$ are
\begin{equation*}
\Delta_1(\mathsf{Q}^\mu,\mathsf{Q})=\int |q|\,\mu(dq),\quad \Delta_1(\mathsf{P}^\nu,\mathsf{P})=\int |p|\,\nu(dp) .
\end{equation*}
(See \cite{Werner04}.) Thus, $\mathsf{Q}^\mu$ and $\mathsf{P}^\nu$ are metric approximations
of $\mathsf{Q}$ and $\mathsf{P}$
exactly when these integrals are finite.
\end{example}
\subsection{Measurement uncertainty relations for metric errors}\label{sec:distance}
The measurement uncertainty relation for the metric error associated with the 1-deviation (or Monge metric)
proven in \cite{Werner04} has recently been generalised to all Wasserstein distances \cite{BLW2013b}.
\begin{theorem}
\label{thm:main}
Let $M$ be a phase space observable and $1\leq\alpha,\beta\leq\infty$. Then
\[
\Delta_\alpha(\mathsf{M}_1,\mathsf{Q})\, \Delta_\beta(\mathsf{M}_2,\mathsf{P})\ \geq\ c_{\alpha\mkern1mu\beta}\hbar
\]
provided that the quantities on the left hand side are finite. The constants $c_{\alpha\mkern1mu\beta}$ are the same as in
Theorem~\ref{thm:prepURpq}.
\end{theorem}
The proof is analogous to that of Werner's original theorem for $1$-deviations: it proceeds by reduction
to the covariant case, and the latter is immediately obtained by application of the preparation uncertainty
relation of Theorem \ref{thm:prepURpq}.
The case where one of the distances is zero is in fact covered by the Theorem: the other
distance must then be infinite. In fact if (in the case of the 1-deviation) one has
$\Delta_1(\mathsf{M}_1,\mathsf{Q})=0$, then $\mathsf{M}_1=\mathsf{Q}$ and $\Delta_1(\mathsf{M}_2,\mathsf{P})$ cannot be finite; otherwise
the associated covariant phase space observable would have to have $\mathsf{Q}$ as its first marginal,
which is impossible. Hence Theorem \ref{thm:main} implies that whenever $\Delta_1(\mathsf{M}_1,\mathsf{Q})=0$
then $\Delta_1(\mathsf{M}_2,\mathsf{P})=\infty$. It is an instructive exercise to verify this explicitly.
\begin{example}\label{ex3}
Let $\mathsf{M}$ be an observable on phase space $\R^2$ whose first marginal
$\mathsf{M}_1$ is sharp position $\mathsf{Q}$. Then the second
marginal $\mathsf{M}_2$ has infinite 1-distance from sharp momentum
$\mathsf{P}$.
\emph{Proof.} We note first that all positive operators (effects)
$\mathsf{M}_2(X)$ in the range of $\mathsf{M}_2$ commute with $\qhat$ (see, e.g., \cite{CRQM}) and are thus
functions of $\qhat$. Thus one can write
$\mathsf{M}_2(X)=\int \mathsf{Q}(dq)m(q,X)$, where the functions $m(\cdot,X)$ are
defined almost everywhere for all (Borel) subsets $X$ of $\R$, and
$X\mapsto m(q,X)$ is then a probability measure. We consider states
$\rho$ with the same fixed position distribution, $\rho^\mathsf{Q}=\prob$,
and compute
\[
\tra{\rho \mathsf{M}_2(h)}=\int h(x)m_\prob(dx),\ m_\prob(X):=\int \prob(dq)m(q,X).
\]
We will let $h$ run through a family $h_n\in\Lambda$ and $\rho$ through a family
$\mathsf{p}_{\rho_n}\in S_q$ such that $\mathsf{p}_{\rho_n}^\mathsf{Q}=\prob$ and $\tra{\rho_n \mathsf{M}_2(h_n)}\to 0$, while $\tra{\rho_n\mathsf{P}(h_n)}\to\infty$.
This shows that $\Delta_1(\mathsf{M}_2,\mathsf{P})=\infty$.
Choose $h_n$ as in Example \ref{ex:triv-dist}, where we have now $\mu=m_\prob$. This gives
$\tra{\rho_n \mathsf{M}_2(h_n)}\to 0$ for any $\rho_n$ (yet to be specified) with $\mathsf{p}_{\rho_n}^\mathsf{Q}=\prob$.
Let $\rho_n=W(0,c_n+n-(c_1+1))\rho_1W(0,c_n+n-(c_1+1))^*$, with $\rho_1$ a state whose
momentum distribution is centered symmetrically at $c_1+1$, the peak location of $h_1$. Then
the momentum distribution of $\rho_n$ is centered at the peak location $c_n+n$ of $h_n$.
Also note that $\mathsf{p}_{\rho_n}^\mathsf{Q}=\mathsf{p}_{\rho_1}^\mathsf{Q}=:\prob$.
Specifically we take $\rho_n$ such that the densities $\mathsf{p}_{\rho_n}^\mathsf{P}(p)=\chi_{J_n}(p)$,
$J_n=[c_n+n-1/2,c_n+n+ 1/2]$. Then we have $\tra{\rho_n\mathsf{P}(h_n)}= n-1/4\to\infty$ as $n\to\infty$.
\qed
\end{example}
\section{Error measures II: error bar width}
\subsection{Gross error bar.}
We now present a definition of measurement error in terms of likely error intervals
that follows most closely the usual practice of calibrating measuring instruments. In the
process of {\em calibration} of a measurement scheme, one seeks to obtain estimates of the
likely error and perhaps also the degree of disturbance that the scheme contains. In order to
estimate the error, one tests the device by applying it to a sufficiently large family of input
states in which the observable one wishes to measure with this setup has fairly sharp values.
The error is then characterised as an overall measure of the bias and
the width of the output distribution across a range of input values. Error bars give
the minimal average interval lengths that one has to allow to contain all output values with a
given confidence level.
For simplicity, we give the following definitions only for approximations of a sharp observable $\mathsf{E}$,
so that the assumption of localised input states $\rho$ can be described as $\rho^\mathsf{E}(J_{x;\delta})=1$, for
intervals $J_{x;\delta}:=[x-\delta/2,x+\delta/2]$, $x\in\R,\delta>0$.
Let $\mathsf{E}_1,\mathsf{E}$ be observables on $\R$ and $\mathsf{E}$ be sharp.
For each $\varepsilon\in(0,1)$, $\delta>0$, we define the \emph{error} of $\mathsf{E}_1$ relative to $\mathsf{E}$
\begin{eqnarray
\dele{\mathsf{E}_1}:=&\inf\{w>0\,|\ \forall\ x\in\R\ \forall\rho\in S:\nonumber\\
&\qquad\qquad\qquad \rho^\mathsf{E}(J_{x;\delta})=1\Rightarrow \rho^{\mathsf{E}_1}(J_{x,w})\ge 1-\varepsilon\}.\nonumber
\end{eqnarray}
The error describes the range within which the input values can be inferred
from the output distributions, with confidence level $1-\varepsilon$, given initial localizations within
$\delta$.
$\mathsf{E}_1$ is called an $\varepsilon$-\emph{approximation} to $\mathsf{E}$ if
$\dele{\mathsf{E}_1}<\infty$ for all $\delta>0$.
Note that the error is an increasing function of $\delta$, so that one can
define the \emph{(gross) error bar width} of $\mathsf{E}_1$ relative to $\mathsf{E}$:
\begin{equation*}
\dee{\mathsf{E}_1}:=\inf_\delta\dele{\mathsf{E}_1}=\lim_{\delta\to 0}\dele{\mathsf{E}_1}.
\end{equation*}
In the case $\dele{\mathsf{E}_1}=\infty$ for all $\delta>0$, we write $\dee{\mathsf{E}_1}=\infty$.
$\mathsf{E}_1$ will be called an \emph{approximation (in the sense of finite error bar width)}
to $\mathsf{E}$ if $\dee{\mathsf{E}_1}<\infty$ for all $\varepsilon\in(0,\frac 12)$. The restriction to $\varepsilon<\frac 12$
reflects the idea that a ``good" approximation should have confidence levels greater than $\frac 12$.
We note that if $\mathsf{E}_1$ is an approximation to $\mathsf{E}$, the map
$\varepsilon\mapsto \dele{\mathsf{E}_1}$ is a decreasing function of $\varepsilon\in(0,\frac 12)$
for every $\delta>0$.
The following result shows that our definition is not empty.
\begin{proposition}\label{prop:smeared-pos-approx}
The smeared position and momentum observables $\mathsf{Q}^\mu$, $\mathsf{P}^\nu$ are
approximations (in the sense of finite error bar widths) to $\mathsf{Q}$ and $\mathsf{P}$, respectively, for
any probability measures $\mu$, $\nu$.
\end{proposition}
\emph{Proof.} It is sufficient to consider the case of the position observable.
Let $\varepsilon\in(0,1), \delta>0$ be given. We have to show that there is a finite number $w>0$
such that for all $q\in\R$ one has $\rho^{\mathsf{Q}^\mu}(J_{q;w})\ge1-\varepsilon$ whenever
$\rho^\mathsf{Q}(J_{q;\delta})=1$.
Let $q_0,w_0$ be such that $\mu(J_{q_0;w_0})\ge 1-\varepsilon$. Then, if
$w\ge 2|q_0|+w_0+\delta$, it follows that $J_{q;\delta}\subseteq x+J_{q;w}$ for all $x\in J_{q_0;w_0}$,
that is, $\rho^\mathsf{Q}(x+J_{q;w})=1$ for all such $x$. Then:
\begin{eqnarray*}
\rho^{\mathsf{Q}^\mu}(J_{q;w})&=\int\mu(dx)\rho^\mathsf{Q}(x+J_{q;w})\ge
\int_{J_{q_0;w_0}}\mu(dx)\rho^\mathsf{Q}(x+J_{q;w})\\
&=\mu(J_{q_0;w_0})\ge1-\varepsilon.\qed
\end{eqnarray*}
\subsection{Properties of the error bar width}
It is not hard to construct approximations of $\mathsf{Q}$ that do not share the translation covariance
of $\mathsf{Q}$.
\begin{example}\label{ex:noncov-approx}
Let $f$ be a continuous function on $\R$ which is one-to-one and such that $f(q)-q$ is not constant but
$|f(q)-q|\le \alpha$ for
all $q\in \R$ and some fixed $\alpha>0$. An example is $f(q)=q+\frac 12\cos(q)$. Let $\mathsf{Q}^\mu$ be a smeared position
observable. Then $\mathsf{Q}^\mu\circ f^{-1}$ is a non-covariant approximation to $\mathsf{Q}$ in
the sense of finite error bars.
\emph{Proof.}
Let $\varepsilon\in(0,1)$and $\delta>0$ be given. We have to show that there is a finite positive $w$ such
that for all $q\in\R$ and all $\rho$ with $\rho^\mathsf{Q}(J_{q;\delta})=1$, then
$\rho^{\mathsf{Q}^\mu}(f^{-1}(J_{q;w}))\ge1-\varepsilon$.
We know that $\mathsf{Q}^\mu$ is an approximation to $\mathsf{Q}$. Hence there is $w'>0$ such that
for all $q\in\R$ and all $\rho$ with $\rho^\mathsf{Q}(J_{q;\delta})=1$, we have
$\rho^{\mathsf{Q}^\mu}((\intv{q}{w'}))\ge1-\varepsilon$.
Now take $w=w'+2\alpha$. This entails that $f^{-1}(J_{q;w})\supseteq\intv{q}{w'}$ for all $q\in\R$. Then, for
$q\in\R$ and $\rho$ such that $\rho^\mathsf{Q}(J_{q;\delta})=1$ we obtain
\begin{equation*}
\rho^{\mathsf{Q}^{\mu}\circ f^{-1}}(J_{q;w})=\rho^{\mathsf{Q}^\mu}(f^{-1}(J_{q;w}))\ge \rho^{\mathsf{Q}^\mu}(\intv{q}{w'})\ge 1-\varepsilon.
\end{equation*}
Noting that $f^{-1}(J_{q;w}+q')\ne f^{-1}(J_{q;w})+q' $ (since $f(q)-q$ is not constant) one concludes readily that
$\mathsf{Q}^{\mu}\circ f^{-1}$ is not covariant. \qed
\end{example}
\begin{example}\label{ex:const-inf-err}
For any bounded Borel function $f$ on $\R$, the observable $\mathsf{Q}\circ f^{-1}$ has infinite error
bars with respect to $\mathsf{Q}$.
\emph{Proof.}
Let $J$ be a bounded interval which contains the range of $f$. Then for any finite $w>0$, one can find
$q$ such that $J_{q;w}\cap J=\emptyset$. Then $f^{-1}(J_{q;w})=\emptyset$ and so
$\rho^{\mathsf{Q}\circ f^{-1}}(J_{q;w})=\rho^\mathsf{Q}(f^{-1}(J_{q;w}))=0$ for all $\rho$.
It follows that $\Delta_{\varepsilon,\delta}(\mathsf{Q}\circ f^{-1},\mathsf{Q})=\infty$ for all $\varepsilon\in(0,1)$ and all $\delta>0$.
\qed
\end{example}
It is possible to characterise the case of an accurate
measurement of the sharp observable $\mathsf{E}$.
\begin{proposition}\label{prop:zero-err}
Let $\mathsf{E}_1$ be an approximation of the sharp observable $\mathsf{E}$. Then the following are equivalent:
\begin{enumerate}
\item[(a)] $\dele{\mathsf{E}_1}\le\delta$ for all $\varepsilon\in(0,\frac12),\delta>0$;
\item[(b)] $\mathsf{E}_1=\mathsf{E}$.
\end{enumerate}
If either of these condition is fulfilled then $\dee{\mathsf{E}_1}=0$ for all $\varepsilon\in(0,\frac12)$.
\end{proposition}
\emph{Proof.} Assume (b) holds. Let $\varepsilon\in(0,\frac12)$, $\delta>0$.
Choose $w=\delta$; then for any $q\in\R$ and
any state $\rho$ with $\rho^\mathsf{E}(J_{x;\delta})=1$, we also have
$\rho^{\mathsf{E}_1}(J_{x;\delta})\ge 1-\varepsilon$. This shows that $\dele{\mathsf{E}_1}\le\delta$.
Conversely, assume that (a) holds. Consider any
$\varepsilon\in(0,\frac12),\delta>0$. For $w=\dele{\mathsf{E}_1}\le\delta$, we have, for
all $x\in\R$ and all $\rho$ with $\rho^\mathsf{E}(J_{x;\delta})=1$, that
$\rho^{\mathsf{E}_1}(J_{x;\delta})\ge\rho^{\mathsf{E}_1}(\intv{x}{w})\ge 1-\varepsilon$. This entails for any vector state
$\varphi$ for which $\mathsf{Q}(J_{x;\delta})\varphi=\varphi$ that
$\ip{\varphi}{\mathsf{E}_1(J_{x;\delta})\varphi}\ge 1-\varepsilon$. As this holds for any
$\varepsilon\in(0,\frac12)$, it follows that $\ip{\varphi}{\mathsf{E}_1(J_{x;\delta})\varphi}=1$. This
entails that $\mathsf{E}(J_{x;\delta})\le \mathsf{E}_1(J_{x;\delta})$. Since $x\in\R$ and
$\delta>0$ are arbitrary, this operator inequality holds for any closed
interval $J=[a,b]$.
We show that then also $\mathsf{E}((a,b))\le E_1((a,b))$ for any open
interval. Let $J_n$ be an increasing sequence of closed sets which
converges to a given open interval $(a,b)$. Put
$D_n:=\mathsf{E}_1(J_n)-\mathsf{E}(J_n)\ge O$. For any POVM $\mathsf{N}$ on $\bre$ we have
$\mathsf{N}(J_n)\to \mathsf{N}((a,b))$ (ultraweakly).
(This is a consequence of the regularity of Borel measures
on $\bre$, see, e.g., \cite{NOST}.)
So we obtain $D_n\to \mathsf{E}_1((a,b))-\mathsf{E}((a,b))$ (ultraweakly),
and since $D_n\ge O$, this limit operator is also nonnegative. In
this way we conclude that $\mathsf{E}(K)\le \mathsf{E}_1(K)$ for all open
intervals $K$. Similarly we can show that $\mathsf{E}((a,b])\le \mathsf{E}_1((a,b])$.
Due to the normalization of both POVMs $\mathsf{E},\mathsf{E}_1$, it follows that they must
coincide on all intervals and finally, since the intervals generate $\bre$, that they are identical. \qed
We remark that it is not known whether the condition $\dee{\mathsf{E}_1}=0$ for all $\varepsilon\in(0,\frac12)$
is sufficient to conclude that $\mathsf{E}_1=\mathsf{E}$.
\begin{proposition}\label{prop:errbar-res}
Let $\mathsf{E}_1,\mathsf{E}$ be observables with support $\R$, and $\mathsf{E}$ be a sharp observable.
The error bar width of $\mathsf{E}_1$ relative to $\mathsf{E}$ is never smaller than the intrinsic
resolution width of $\mathsf{E}_1$:
\begin{equation*
\dee{\mathsf{E}_1}\ge \gamma_\varepsilon(\mathsf{E}_1).
\end{equation*}
\end{proposition}
The proof is given in \cite[Prop. 1]{BuPe07}.
\begin{corollary}
Let $\mathsf{E}$ be a sharp observable on $\bre$ with support $\R$. Any $\varepsilon$-approximation $\mathsf{E}_1$
(supported on $\R$) of
$\mathsf{E}$ has finite resolution width, $\gamma_\varepsilon(\mathsf{E}_1)<\infty$.
\end{corollary}
\subsection{Bias-free error and bias}
We show next how the gross error can be decomposed into a (positive) bias term and
a random error.
Let $\varepsilon\in(0,1)$ and $\delta>0$ be given. Let $\mathsf{E}_1,\mathsf{E}$ be observables on $\R$ and $\mathsf{E}$ be sharp.
Note that the condition $\rho^\mathsf{E}(J_{x;\delta})=1$ (for some $x\in\R$) can be expressed as
$W_0(\rho^\mathsf{E})\le\delta$.
We define the \emph{bias-free}, or {\em random error} $\w_{\varepsilon,\delta}({\sfe_1},\sfe)$ as follows:
\begin{equation*}
\w_{\varepsilon,\delta}({\sfe_1},\sfe):=\sup\big\{W_\varepsilon(\rho^{\mathsf{E}_1})\,|\, W_0(\rho^\mathsf{E})\le\delta\big\}.
\end{equation*}
This is a measure of the overall minimal error, determined by the overall widths of all
output distributions, given input distributions supported in intervals $J_{q;\delta}$.
If this quantity is finite for some $\delta_0$, it is an increasing function for all $\delta\le\delta_0$.
In that case we can define the \emph{bias-free error bar width},
\begin{equation*}
\w_{\varepsilon}({\sfe_1},\sfe):=\lim_{\delta\to 0}\w_{\varepsilon,\delta}({\sfe_1},\sfe).
\end{equation*}
The following is obvious:
\begin{equation*}
\dele{\mathsf{E}_1}\ge\w_{\varepsilon,\delta}({\sfe_1},\sfe).
\end{equation*}
If these quantities are finite, one then has in the limit $\delta\to 0$:
\begin{equation}\label{unbiased-biased}
\dee{\mathsf{E}_1}\ge\w_{\varepsilon}({\sfe_1},\sfe).
\end{equation}
The difference between $\dele{\mathsf{E}_1}$ and $\w_{\varepsilon,\delta}({\sfe_1},\sfe)$ disappears when the output
distributions are concentrated at the locations of the input distributions, that is, around
the intervals $J_{x;\delta}$. This is to say that the difference is a measure of the overall magnitude
of the \emph{bias} $\bede{\mathsf{E}_1}$ inherent in $\mathsf{E}_1$ relative to $\mathsf{E}$:
\begin{equation*}
\bede{\mathsf{E}_1}:=\dele{\mathsf{E}_1}-\w_{\varepsilon,\delta}({\sfe_1},\sfe)\ge 0.
\end{equation*}
Rephrasing this as
\begin{equation*}
\dele{\mathsf{E}_1}=\w_{\varepsilon,\delta}({\sfe_1},\sfe)+\bede{\mathsf{E}_1},
\end{equation*}
we see that the gross error is decomposed into the bias-free error and the magnitude of the
bias. Note that one can take the limit of $\delta\to 0$:
\begin{equation*}
\bee{\mathsf{E}_1}:=\dee{\mathsf{E}_1}-\w_{\varepsilon}({\sfe_1},\sfe).
\end{equation*}
As an immediate consequence of these definitions, we can say that $\mathsf{E}_1$ is an
$\varepsilon$-approximation to $\mathsf{E}$ if and only if the bias and random errors are finite
for all $\delta>0$.
\begin{proposition}\label{prop:sm-bf-err}
Let $\mathsf{Q}^\mu$, $\mathsf{P}^\nu$ be smeared position and momentum observables. Then
\begin{equation*}
{\mathcal W}^0_{\varepsilon_1}(\mathsf{Q}^\mu,\mathsf{Q})=W_{\varepsilon_1}(\mu),
\quad {\mathcal W}^0_{\varepsilon_2}(\mathsf{P}^\nu,\mathsf{P})=W_{\varepsilon_2}(\nu).
\end{equation*}
\end{proposition}
\emph{Proof.} It suffices to consider the case of position.
We show first that $W_{\varepsilon_1}(\rho^{\mathsf{Q}^\mu})\ge W_{\varepsilon_1}(\mu)$. This is equivalent to the following:
$w\ge W_{\varepsilon_1}(\rho^{\mathsf{Q}^\mu})$ implies $w\ge W_{\varepsilon_1}(\mu)$.
Thus, let $w$ be such that $\rho^{\mathsf{Q}^\mu}(J_{q;w})\ge 1-\varepsilon_1$ for some $q\in\R$.
Assume $w<W_{\varepsilon_1}(\mu)$; this means that for all $q'\in\R$ one has $\mu(\intv{q'}{w})<1-\varepsilon_1$.
But then
\begin{equation*}
\rho^{\mathsf{Q}^\mu}(\intv{q'}{w})=\int \rho^{\mathsf{Q}}(dx)\mu(x-\intv{q'}{w})<1-\varepsilon_1,
\end{equation*}
which contradicts the premise.
Next we show that whenever $W_0(\rho^{\mathsf{Q}})\le\delta$, then
$W_{\varepsilon_1}(\rho^{\mathsf{Q}^\mu})\le W_{\varepsilon_1}(\mu)+\delta$.
We are given that $\rho^{\mathsf{Q}}(\intv{q_0}{\delta})=1$ for some $q_0\in\R$. Assume
$w\ge W_{\varepsilon_1}(\mu)$, that is, $\mu(\intv{q_1}{w})\ge1-{\varepsilon_1}$ for some $q_1$.We have to show
that $w+\delta\ge W_{\varepsilon_1}(\rho^{\mathsf{Q}^\mu})$, that is,
$\rho^{\mathsf{Q}^\mu}(\intv{q_2}{w+\delta})\ge1-\varepsilon_1$
for some $q_2\in\R$.
Let $q_2=q_0-q_1$. Then it follows that $q+\intv{q_2}{w+\delta}\supseteq \intv{q_0}{\delta}$ for
all $q\in\intv{q_1}{w}$. Then
\begin{eqnarray*}
\rho^{\mathsf{Q}^\mu}(\intv{q_2}{w+\delta})&=\int\mu(dq)\rho^{\mathsf{Q}}(q+\intv{q_2}{w+\delta})\\
&\ge
\int_{J_{q;w}}\mu(dq)=\mu(J_{q;w})\ge1-\varepsilon_1.
\end{eqnarray*}
This shows that $w\ge W_{\varepsilon_1}(\mu)$ implies $w+\delta\ge W_{\varepsilon_1}(\rho^{\mathsf{Q}^\mu})$
whenever $W_0(\rho^\mathsf{Q})\le\delta$. Thus, under this assumption we let $w$ approach
$W_{\varepsilon_1}(\mu)$ to obtain $W_{\varepsilon_1}(\mu)+\delta\ge W_{\varepsilon_1}(\rho^{\mathsf{Q}^\mu})$.
To summarise, we have shown:
$W_\varepsilon(\mu)\le W_\varepsilon(\rho^{\mathsf{Q}^\mu})\le W_\varepsilon(\mu)+\delta$, where the latter inequality
holds if $W_0(\rho^\mathsf{Q})\le\delta$. This entails that also
$W_\varepsilon(\mu)\le {\mathcal W}^0_{\varepsilon,\delta}(\mathsf{Q}^\mu,\mathsf{Q})\le W_\varepsilon(\mu)+\delta$. Now we can
take the limit $\delta\to 0$ to obtain the result. \qed
\subsection{Measurement uncertainty relations for error bar widths}\label{sec:errorbar}
The following error relations for covariant approximations are special cases of the general result quoted below.
Their proofs are straightforward consequences of the considerations of this paper, hence we present them here
as separate statements.
\begin{proposition}
Let $\mathsf{G}^\tau}%{{\mathbf m}$ be a covariant phase space observable. Then the bias-free error bar widths
of the marginals relative to $\mathsf{Q}$ and $\mathsf{P}$ obey the trade-off relation:
\begin{equation}\label{unbiased-ur}
\deq{\mathsf{Q}^{\mu_\tau}%{{\mathbf m}}}\,\dep{\mathsf{P}^{\nu_\tau}%{{\mathbf m}}} \ge {\mathcal W}^0_{\varepsilon_1}(\mathsf{Q}^{\mu_\tau}%{{\mathbf m}},\mathsf{Q})\, {\mathcal W}^0_{\varepsilon_2}(\mathsf{P}^{\nu_\tau}%{{\mathbf m}},\mathsf{P})
\ge 2\pi\hbar\,K(\varepsilon_1,\varepsilon_2),
\end{equation}
where $K(\varepsilon_1,\varepsilon_2)$ is given by Eq.~(\ref{eqn:ow-U}).
\end{proposition}
{\em Proof.} The first inequality follows from (\ref{unbiased-biased}) and the second
is a direct consequence of Proposition \ref{prop:sm-bf-err} and Eq.~(\ref{eqn:gam-gam-cov-ur})
The corresponding inequality for general phase space observables was proven in \cite{BuPe07}.
\begin{theorem}\label{thm}
Let $\mathsf{M}$ be an approximate joint observable for $\mathsf{Q},\mathsf{P}$, in the sense that its marginals have finite error bar
widths as approximations of position and momentum, respectively. Then,
for $\varepsilon_1,\varepsilon_2\in(0,\frac 12)$,
the error bar widths of $\mathsf{M}_1$ and $\mathsf{M}_2$ satisfy the
uncertainty relation
\begin{equation*
\deq{\mathsf{M}_1}\cdot\dep{\mathsf{M}_2}\ge
2\pi\hbar\,K(\varepsilon_1,\varepsilon_2),
\end{equation*}
where $K(\varepsilon_1,\varepsilon_2)$ is given by Eq.~(\ref{eqn:ow-U}).
\end{theorem}
This result entails the following statement: an
approximate joint observable for $\mathsf{Q},\mathsf{P}$ cannot have one of
these sharp observables as its marginal. It is instructive to show this explicitly by
considering the case $\mathsf{M}_1=\mathsf{Q}$.
\begin{proposition}\label{sharp-marg}
Let $\mathsf{M}$ be an observable on phase space whose first marginal
coincides with sharp position, $\mathsf{M}_1=\mathsf{Q}$ (so that $\deq{\mathsf{M}_1}=0$).
Then the second marginal $\mathsf{M}_2$ cannot satisfy the condition of an
$\varepsilon_2$-approximation to $\mathsf{P}$ for any $\varepsilon_2\in(0,\frac 12)$, that is,
$\dep{\mathsf{M}_2}=\infty$. Hence $\mathsf{M}$ cannot be an
$(\varepsilon_1,\varepsilon_2)$-approximate joint observable to $\mathsf{Q},\mathsf{P}$ for
any $\varepsilon_1,\varepsilon_2\in(0,\frac 12)$.
\end{proposition}
\emph{Proof.} Let $\varepsilon_2\in(0,\frac 12)$ be given and let $\delta>0$
and $w'>0$ be arbitrary. We have to show that there is an interval
$J_{p;\delta}$ and a state $\rho$ localised in $J_{p;\delta}$ so that $\tra{\rho
M_2(J_{p;w'})}<1-\varepsilon_2$.
As noted in Example \ref{ex3}, all positive operators (effects)
$\mathsf{M}_2(X)$ in the range of $\mathsf{M}_2$ commute with $\qhat$ and are thus
functions of $\qhat$.
Thus we can write: $\mathsf{M}_2(X)=\int m(q,X)\,\mathsf{Q}(dx)$. Consider the
sequence of intervals $J_{n;w'}$, $n=0,1,2,\dots$. Since
$I=\mathsf{M}_2(\R)$, then $\mathsf{M}_2((-\infty,n-w'/2))\to1$ as $n\to\infty$
(ultraweakly), and it follows that for every state $\rho$,
$\tra{\rho \mathsf{M}_2(J_{n;w'})}\le \tra{\rho \mathsf{M}_2([n-w'/2,\infty))}\to 0$,
hence:
\[
\tra{\rho \mathsf{M}_2(J_{n;w'})}=\int\rho^\mathsf{Q}(dq)m(q,J_{n;w'})\to 0 \quad {\rm as}\ n\to\infty.
\]
Let $\rho_0$ be such that $\rho_0^\mathsf{P}(J_{0;\delta})=1$, that is, the
distribution $\rho_0^\mathsf{P}$ vanishes outside that interval. Then
$\rho_n:=W(0,n)\rho_0$ is localised in $J_{n;\delta}$, while the
position distribution is unchanged, $\rho_n^\mathsf{Q}=\rho_0^\mathsf{Q}$.
For the given $\varepsilon_2\in(0,\frac 12)$, there is an $n\in\N$ such that for
the fixed state $\rho_0$, $\tra{\rho_0 \mathsf{M}_2(J_{n;w'})}<1-\varepsilon_2$.
Then, since $\rho_0^\mathsf{Q}=\rho_n^\mathsf{Q}$, we also have $\tra{\rho_n
\mathsf{M}_2(J_{n;w'})}<1-\varepsilon_2$, whereas $\rho_n$ is localised in
$J_{n;\delta}$. \qed
\vspace{12pt}
This result reproduces, in particular, the well-known fact that
there is no observable on phase space whose marginals are sharp
position and sharp momentum.
\begin{example}
Example \ref{ex:noncov-approx} can be used to construct an observable $M$ on phase space which is not
covariant but is still an approximate joint observable for $\mathsf{Q},\mathsf{P}$. Let $\mathsf{G}^\tau}%{{\mathbf m}$ be a
covariant phase space observable and define $\mathsf{M}:=\mathsf{G}^\tau}%{{\mathbf m}\circ \gamma^{-1}$, where $\gamma(q,p):=
(\gamma_1(q),\gamma_2(p))$. We assume that $\gamma_1,\gamma_2$ are strictly increasing
continuous functions such that $\gamma_1(q)-q$ and $\gamma_2(p)-p$ are bounded functions.
Then it follows that the marginals $\mathsf{M}_1^\gamma=\mathsf{G}^\tau}%{{\mathbf m}_1\circ\gamma_1^{-1}$ and
$\mathsf{M}_2^\gamma=\mathsf{G}^\tau}%{{\mathbf m}_2\circ\gamma_2^{-1}$ have finite error bars with respect to $\mathsf{Q},\mathsf{P}$. If
$\gamma$ is a nonlinear function then $\mathsf{M}$ will not be covariant.
\end{example}
It is straightforward to obtain a universal uncertainty relation for the bias-free errors for any
approximate joint observable $\mathsf{M}$ of $\mathsf{Q},\mathsf{P}$. The core of the proof is to show that finite
bias-free errors for the marginals entails the existence of a covariant observable $\mathsf{G}^\tau}%{{\mathbf m}$ (obtained by
the operation of finite mean used in \cite{Werner04}) such that its marginals
are not greater than those of $\mathsf{M}$:
\begin{equation*}
\delqo{\mathsf{M}_1}\ge\delqo{\mathsf{G}^\tau}%{{\mathbf m}_1},\quad \delpo{\mathsf{M}_2}\ge\delpo{\mathsf{G}^\tau}%{{\mathbf m}_2}.
\end{equation*}
The proof of this is similar to that of Lemma 4 of \cite{BuPe07} and will be omitted.
Using inequality (\ref{unbiased-ur}), the bias-free errors is then seen to obey the trade-off relation for $\varepsilon_1,\varepsilon_2<\frac 12$:
\begin{equation*}
{\mathcal W}^0_{\varepsilon_1}(\mathsf{M}_1,\mathsf{Q})\, {\mathcal W}^0_{\varepsilon_2}(\mathsf{M}_2,\mathsf{P})\ge
{2\pi\hbar}\, K(\varepsilon_1,\varepsilon_2).
\end{equation*}
\subsection{Trade-off relations for resolution widths}
In the work \cite{BuPe07} we claimed the validity of an uncertainty relation for resolution widths;
a proof was not given explicitly as it was considered to follow closely the steps of the proof of Theorem \ref{thm}. On revisiting this relation, we found that there is no obvious way of adapting that proof.
In fact, it may well be that it is only for sufficiently close joint approximations of $Q$ and $P$
that there have to be constraints on the resolution width similar to the error uncertainty relation.
Hence we rephrase the claim as a Problem.
\vspace{6pt}
\noindent{\bf Problem.} {\em
Let observable $\mathsf{M}$ on $\brr$ be an approximate joint observable for $\mathsf{Q},\mathsf{P}$ in the sense of finite
error bar widths. State conditions on the quality of the approximation (other than the covariance of the joint observable) which entail that the resolution widths must obey the trade-off relation (for $\varepsilon_1+\varepsilon_2<1$):
\begin{equation*
\gamma_{\varepsilon_1}(\mathsf{M}_1)\,\gamma_{\varepsilon_2}(\mathsf{M}_2)\ge 2\pi\hbar\,K(\varepsilon_1,\varepsilon_2)\quad (?)
\end{equation*}
}
\vspace{6pt}
\section{Error measures III: Noise-based error}
Classical statistical error analysis is prominently based on the use of moments of probability distributions for
the quantification of measurement errors. Thus, a wide-spread approach
found in the literature of defining a measure of error is in terms of
a formal ``root mean square" deviation of an indicator variable $Z$ of the measuring
apparatus from the variable $A$ to be measured approximately. Classically, $A$ and $Z$ are given as random variables, and quantum mechanically as selfadjoint operators:
this \emph{state-dependent noise-based error} is given as
the root mean square deviation,
\begin{equation*
\eps_{\footnotesize\textsc{no}}(A,{\h M},\rho):=\langle (Z_{\rm out}-A_{\rm in})^2\rangle_{\rho\otimes\sigma}^{1/2}.
\end{equation*}
Here $Z_{\rm out}$ denotes the output (pointer) observable at the end of the interaction
phase between object system and probe in the measurement $\mathcal{M}$, and $A_{\rm in}$ is the input object observable to be approximately measured; the object plus probe system is initially in the state $\rho\otimes\sigma$. The choice of name reflects the fact that the operator $Z_{\rm out}-A_{\rm in}$ is commonly called {\em noise operator}.
A detailed critique of this attempted quantum generalisation of the rms error is given in \cite{BLW2013a}; the main deficiency is that this quantity
fails to be a faithful representation of the absence or magnitude of an approximation error. Therefore this measure has to be used with care; it is operationally significant only in some special circumstances; then it may be used to provide estimates of measurement errors \cite{BLW2013a}.
Here we are concerned with the state-independent upper bound of the quantity $\eps_{\footnotesize\textsc{no}}$,
the \emph{(global) noise-based error} of a measurement $\mathcal{M}$ relative to $A$ as \cite{Appleby1998b}
\begin{equation*}
\eps_{\footnotesize\textsc{no}}(A,\mathcal{M}):=\sup_{\rho} \eps_{\footnotesize\textsc{no}}(A,\mathcal{M},\rho)
\end{equation*}
where the supremum is taken over all states $\rho$ for which the right hand side is well-defined.
We will say that the observable $\mathsf{C}$ defined by ${\h M}$ is a \emph{finite-noise approximation} to $A$ if
$\mathsf{C}$ has finite global noise-based error relative to $A$.
The noise-based error $\eps_{\footnotesize\textsc{no}}(A,\mathcal{M},\rho)$ can be expressed in terms of the observable $\mathsf{C}$ actually measured by ${\h M}$
\cite{Ozawa04}:
\begin{equation}\label{OzaVeps1}
\eps_{\footnotesize\textsc{no}}(A,{\h M},\rho)^2= \tr{\rho(\mathsf{C}[x^2] -\mathsf{C}[x]^2)} + \tr{\rho (\mathsf{C}[x]-A)^2}.
\end{equation}
In order to apply this error measure in the case of joint approximate measurements, we note that if a measurement scheme ${\h M}$
defines an observable $\mathsf{M}$ on $\brr$, then its marginal observables $\mathsf{M}_1,\mathsf{M}_2$ can be taken as approximators for, say,
position $Q$ and momentum $P$, respectively. In this case the noise-based errors are defined via (\ref{OzaVeps1}) with $\mathsf{C}=\mathsf{M}_1$ for $\eps_{\footnotesize\textsc{no}}(Q,{\h M},\rho)\equiv\eps_{\footnotesize\textsc{no}}(Q,\mathsf{M}_1,\rho)$ and with $\mathsf{C}=\mathsf{M}_2$ for $\eps_{\footnotesize\textsc{no}}(P,{\h M},\rho)\equiv\eps_{\footnotesize\textsc{no}}(P,\mathsf{M}_2,\rho)$.
We denote the global errors for the marginals of a general phase space observable by $\eps_{\footnotesize\textsc{no}}(Q,\mathsf{M}_1)$ and $\eps_{\footnotesize\textsc{no}}(P,\mathsf{M}_2)$, respectively. Then the following general result has been shown \cite{Appleby1998b}.
\begin{theorem}\label{thm:fin-st-err}
Let ${\h M}$ be a measurement realizing an observable $\mathsf{M}$ on $\brr$. Then the
global noise-based errors obey the following trade-off relation.
\begin{equation*
\eps_{\footnotesize\textsc{no}}(Q,\mathsf{M}_1)\,\eps_{\footnotesize\textsc{no}}(P,\mathsf{M}_2)\ge \frac \hbar 2.
\end{equation*}
The lower bound is realised for a covariant phase space observable $\mathsf{G}^\tau}%{{\mathbf m}$ with $\tau}%{{\mathbf m}$ being
the minimum uncertainty state operator with zero means of position and momentum.
\end{theorem}
\section{Connections}\label{connections}
We show that the concept of approximation based on
finite error bars generalises the notions of finite noise approximation
and metric approximations.
\begin{proposition}\label{prop:finite-dist-approx}
Any observable $\mathsf{E}_1$ on $\R$ that satisfies the condition
$\Delta_\alpha(\mathsf{E}_1,\mathsf{E})<\infty$ (for some $\alpha\in[1,\infty)$) for a sharp observable $\mathsf{E}$ on $\R$ is an
approximation to $\mathsf{E}$ in the sense of finite error bars. In that case
the following inequality holds:
\begin{equation}\label{eqn:fin-dist-errbar}
\dee{\mathsf{E}_1}\le \frac{2}{\varepsilon^{\frac1\alpha}}\,\Delta_\alpha(\mathsf{E}_1,\mathsf{E})\,.
\end{equation}
\end{proposition}
\begin{proof} The proof is a straightforward adaptation of the proof for the case $\alpha=1$ given in \cite[Prop. 5]{BuPe07}.\\
Using the definition of $\mathcal{D}_\alpha(\rho^{\mathsf{E}_1},\rho^\mathsf{E})$ and equation (\ref{DKanto}), we are given that
\begin{equation}\label{+}
\big|\tr{\rho \mathsf{E}_1[\Phi]}-\tr{\rho \mathsf{E}(\Psi)}\big|\le \Delta_\alpha(\mathsf{E}_1,\mathsf{E})^\alpha=:c^\alpha,
\end{equation}
which holds for all $\rho\in S$ and all functions $\Psi,\Phi$ satisfying the constraint
\begin{equation}\label{++}
\biggl|\Phi(y)-\Psi(x)\biggr| \le |x-y|^\alpha,\quad x,y\in \R.
\end{equation}
Let $\varepsilon\in(0,1)$ and $\delta>0$ be given. Put
$w=\delta+2n$, with $n\in\N,\ n^\alpha\ge c^\alpha/\varepsilon$.
Consider an interval $J_{q;\delta}$ and a state $\rho$ with $\rho^\mathsf{E}(J_{q;\delta})=1$. Define
the functions $\Psi_n=\Phi_n\equiv h_n$ via
\[
h_n(x):=\left\{\begin{array}{ll} n^\alpha&{\rm if}\ \ |x-q|\le\delta/2;\\
\biggl[n+\delta/2-|x-q|\biggr]^\alpha&{\rm if}\ \ \delta/2<|x-q|\le \delta/2+n;\\
0&{\rm if}\ \ \delta/2+n<|x-q|.\\
\end{array}
\right.
\]
It is not hard to verify that $\Psi_n,\Phi_n$ satisfy (\ref{++}). Condition $(\ref{+})$ for $\Psi_n=\Phi_n=h_n$ entails for
$g_n=h_n/n^\alpha$ that $\big|\tr{\rho {\mathsf{E}_1}[g_n]}-\tr{\rho \mathsf{E}[g_n]}\big|\le c^\alpha/n^\alpha$. We
then have
$\chi_{J_{q;\delta}}\le g_n\le \chi_{J_{q;w}}$.
Now $\rho^\mathsf{E}(J_{q;\delta})=1$ implies $\tr{\rho \mathsf{E}(g_n)}=1$, and so, using the assumption
$n^\alpha\ge c^\alpha/\varepsilon$, we obtain
\[
\tr{\rho \mathsf{E}_1(J_{q;w})}\ge\tr{\rho \mathsf{E}_1(g_n)}\ge\tr{\rho \mathsf{E}(g_n)}-c^\alpha/n^\alpha\ge 1-\varepsilon.
\]
To prove the inequality (\ref{eqn:fin-dist-errbar}), we note that on putting
$w=\delta+2c/(\varepsilon^{1/\alpha})$, one still obtains $\tr{\rho \mathsf{E}_1(J_{q;w})}\ge
1-\varepsilon$. This yields $\dele{\mathsf{E}_1}\le\delta+2\Delta_\alpha(\mathsf{E}_1,\mathsf{E})/\varepsilon^{1/\alpha}$, and on
letting $\delta$ approach 0, then (\ref{eqn:fin-dist-errbar}) follows.
\end{proof}
\begin{proposition}\label{prop:fin-st-err-err-bar}
Any observable $\mathsf{E}_1$ on $\R$ that satisfies the condition of finite global noise-based
error relative to a sharp observable with selfadjoint operator $A$ and associated spectral measure $\mathsf{E}$ (such that $A=\mathsf{E}[x]$),
$\eps_{\footnotesize\textsc{no}}(A,\mathsf{E}_1)<\infty$, is an
approximation to $\mathsf{E}$ in the sense of finite error bars.
In that case, the following inequality holds:
\begin{equation*
\dee{\mathsf{E}_1}\le 2\eps_{\footnotesize\textsc{no}}(A,\mathsf{E}_1)\,\left(1+\sqrt{\frac{2}{\varepsilon}}\right).
\end{equation*}
\end{proposition}
\begin{proof}
We use the facts that $A^2=\mathsf{E}[x]^2=\mathsf{E}[x^2]$ and
$\Delta(\mathsf{E},\rho)=\Delta(A,\rho)$.
We begin by rewriting the definition of $\eps_{\footnotesize\textsc{no}}$ for general states $\rho$, denoted
$\eps_{\footnotesize\textsc{no}}(\mathsf{E}_1,\mathsf{E},\rho)$, and expressing the condition
of bounded errors: for all $\rho$ and $c:=\eps_{\footnotesize\textsc{no}}(\mathsf{E}_1,\mathsf{E})<\infty$,
\begin{eqnarray*}
\eps_{\footnotesize\textsc{no}}(A,\mathsf{E}_1,\rho)^2&=&\tra{\rho(\mathsf{E}_1[x]-A)^2}+\tra{\rho(\mathsf{E}_1[x^2]-\mathsf{E}_1[x]^2)}\\
&=&\tra{\rho(\mathsf{E}_1[x]-A)^2}+\Delta(\mathsf{E}_1,\rho)^2-\Delta(\mathsf{E}_1[x],\rho)^2\ \le\ c^2.
\end{eqnarray*}
(This follows readily from the corresponding condition stipulated for all vector states.)
The first term can be estimated as follows: using the inequality
\begin{eqnarray*}
|{\rm cov}_\rho(\mathsf{E}_1[x],A)|&=\tfrac 12 \left|\tra{\rho \mathsf{E}_1[x]\,A}+\tra{\rho A \mathsf{E}_1[x]}
-2\tra{\rho \mathsf{E}_1[x]}\tra{\rho A}\right|\\
&\le \Delta(\mathsf{E}_1[x],\rho)\Delta(A,\rho),
\end{eqnarray*}
we see that
\begin{eqnarray*}
&\tr[\rho(\mathsf{E}_1[x]-A)^2]=\Delta(\mathsf{E}_1[x]-A,\rho)^2+\left(\tra{\rho(\mathsf{E}_1[x]-A)}\right)^2\\
&\ =\Delta(\mathsf{E}_1[x],\rho)^2+\Delta(A,\rho)^2-2{\rm cov}_\rho(\mathsf{E}_1[x],A)
+\left(\tra{\rho(\mathsf{E}_1[x]-A)}\right)^2\\
&\ \ge\left(\Delta(\mathsf{E}_1[x],\rho)-\Delta(A,\rho)\right)^2+\left(\tra{\rho(\mathsf{E}_1[x]-A)}\right)^2.
\end{eqnarray*}
The boundedness of $\eps_{\footnotesize\textsc{no}}(\mathsf{E}_1,A,\rho)$ then gives:
\begin{eqnarray*}
\big(\Delta(\mathsf{E}[x],\rho)&-\Delta(A,\rho)\big)^2+\left(\tra{\rho \mathsf{E}_1[x]}-\tra{\rho A}\right)^2\\
&\ +\left(\Delta(\mathsf{E}_1,\rho)^2-\Delta(\mathsf{E}_1[x],\rho)^2\right)\le \eps_{\footnotesize\textsc{no}}(\mathsf{E}_1,A,\rho)\le c^2.
\end{eqnarray*}
Each of the three bracketed terms is nonnegative and hence bounded above by $c^2$.
This implies:
\[
\Delta(\mathsf{E}_1[x],\rho)^2-c^2\le\Delta(\mathsf{E}_1,\rho)^2\le\Delta(\mathsf{E}_1[x],\rho)^2+c^2
\]
\[
\Delta(A,\rho)-c\le\Delta(\mathsf{E}_1[x],\rho)\le\Delta(A,\rho)+c
\]
\begin{equation}\label{sharp3}
\tra{\rho A}-c\le\tra{\rho \mathsf{E}_1[x]}\le\tra{\rho A}+c
\end{equation}
the first two inequalities taken together yield:
\begin{equation}\label{sharp4}
\Delta(\mathsf{E}_1,\rho)^2\le(\Delta(A,\rho)+c)^2+c^2
\end{equation}
Now observe that the variance on the l.h.s. is the variance of the distribution $\prob:=\rho^{\mathsf{E}_1}$.
We use the following variant of Chebyshev's inequality, valid for any $w>0$:
\begin{eqnarray*}
\Delta(\prob)^2&=\int (x-\prob[x])^2\prob(dx)\\
&\ge\left\{\begin{array}{l}
\int_{\R\setminusJ_{q;w}}(x-\prob[x])^2\prob(dx)
\ge \left(\tfrac{w}2-|\prob[x]-q|\right)^2\big(1-\prob(J_{q;w})\big)\\
\hfill {\rm if\ } \prob[x]\inJ_{q;w}\,;\\
\int_{J_{q;w}}(x-\prob[x])^2\prob(dx)
\ge \left(\tfrac{w}2-|\prob[x]-q|\right)^2\prob(J_{q;w}) \\
\hfill {\rm if\ } \prob[x]\not\inJ_{q;w}\, .
\end{array}\right.
\end{eqnarray*}
We will only be using cases of large $w$ where $\prob[x]\inJ_{q;w}$ so that we obtain:
\begin{equation}\label{sharp5}
\left(\tfrac{w}2-|\prob[x]-q|\right)^2\big(1-\prob(J_{q;w})\big) \le\Delta(\prob)^2
\end{equation}
Combining (\ref{sharp4}) and (\ref{sharp5}) yields:
\begin{equation}\label{sharp6}
\big(\tfrac w2-|\rho^{\mathsf{E}_1}[x]-q|\big)^2\,\big(1-\rho^{\mathsf{E}_1}(J_{q;w})\big)\le
(\Delta(A,\rho)+c)^2+c^2
\end{equation}
We will only use this in the case of states $\rho$ for which $\rho^\mathsf{E}(J_{q;\delta})=1$. In this case we
have $\Delta(A,\rho)\le\delta$ and $|\tra{\rho A}-q|\le\delta$, and using (\ref{sharp3}) we also obtain:
\[
|\rho^{\mathsf{E}_1}[x]-q| \le |\rho^{\mathsf{E}_1}[x]-\tra{\rho A}| + |\tra{\rho A}-q|\le c+\delta.
\]
We will also use only large (finite) $w$ so that we can assume
\[
\tfrac w2-|\rho^{\mathsf{E}_1}[x]-q| \ge \tfrac w2-(\delta+c)>0.
\]
Note that this entails, in particular, that $\rho^{\mathsf{E}_1}[x]\inJ_{q;w}$, so that the use of (\ref{sharp6})
is justified. Under these conditions (\ref{sharp6}) entails
\[
\big(\tfrac w2-(\delta+c)\big)^2\big(1-\rho^{\mathsf{E}_1}(J_{q;w})\big) \le (\delta+c)^2+c^2
\]
Now, for any $\varepsilon$ one can choose $w$ large enough such that
\[
\big({\tfrac w2}-(\delta+c)\big)^2=\frac{(\delta+c)^2+c^2}{\varepsilon}
\]
Then $(\sharp 7)$ implies that $1-\rho^{\mathsf{E}}(J_{q;w})\le\varepsilon$. Moreover, since
$\dele{E_1}\le w$, we also have
\[
\dele{\mathsf{E}_1}\le\frac 2{\sqrt{\varepsilon}}\sqrt{(\delta+c)^2+c^2}+2(\delta+c),
\]
which in the limit $\delta\to 0$ yields
\[
\dee{\mathsf{E}_1}\le\left(1+\frac{\sqrt{2}}{\sqrt{\varepsilon}}\right)\,2\eps_{\footnotesize\textsc{no}}(A,\mathsf{E}_1).
\]
\end{proof}
An interesting open question is whether
finite global noise-based error also implies finite Wasserstein distances.
\section{Conclusion}
We have reviewed several measures of error and intrinsic unsharpness for
measurements of position and momentum (or other observables supported on the
real line) and given a detailed investigation of their properties
and the relations between them.
We then have studied criteria for approximate (joint) measurements of
position and momentum, based on three different kinds of error measures:
Wasserstein $\alpha$-distances, error bar width (with or without bias), and global noise-based error.
We have established two inequalities relating Wasserstein $\alpha$-distance and global noise-based error
to error bar width, respectively, and have concluded that the criterion of finite error
bars is satisfied whenever
the Wasserstein $\alpha$-distance or the global noise error is finite. Thus the criterion for approximate
joint measurability of position and momentum in terms of finite error bars is the most general among the
three. It is satisfied by \emph{all} covariant phase space observables whereas
for some of these observables the $\alpha$-distances or global noise errors may be infinite.
For each of the three types of error measures we have reviewed a universal joint-measurement uncertainty
relation. Put in geometric terms, these relations state that
the marginals $\mathsf{M}_1,\mathsf{M}_2$ of an observable $\mathsf{M}$ on phase space cannot both be
arbitrarily close to $\mathsf{Q},\mathsf{P}$, respectively.
We also considered the resolution width of an observable on $\R$, introduced in
\cite{CaHeTo06}, and posed the question under which assumptions on the quality of
approximations for approximate joint measurements of position and momentum the
resolution widths of the marginals obey a Heisenberg-type uncertainty relation.
\section*{Acknowledgements}
It is a pleasure to thank Pekka Lahti for helpful comments on various manuscript versions of this work.
\section*{References}
\bibliographystyle{unsrt}
|
\section{Introduction}
The Graph Colouring Problem for Office Blocks was raised by BAE Systems at the
53rd European Study Group with Industry in 2005~\cite{AGG+05}. Consider an
office complex with space rented by several independent organisations. It is
likely that each organisation uses its own wireless network (WLAN) and ask for
a safe utilisation of it. A practical way to deal with this issue is to use a
so-called ``stealthy wallpaper'' in the walls and ceilings shared between
different organisations, which would attenuate the relevant frequencies. Yet,
the degree of screening produced will not be sufficient if two distinct
organisations have adjacent offices, that is, two offices in face-to-face
contact on opposite sides of just one wall or floor-ceiling. In this case, the
WLANs of the two organisations need to be using two different channels (the
reader is referred to the report by Allwright \emph{et al.}~\cite{AGG+05} for
the precise technical motivations).
This problem can be modeled as a graph coloring problem by building a
\emph{conflict graph} corresponding to the office complex: to each
organisation corresponds a vertex, and two vertices are adjacent if the
corresponding territories share a wall, floor, or ceiling area. The goal is to
assign a color (frequency) to each vertex (organisation) such that adjacent
vertices are assigned distinct colors. In addition, not every graph may occur
as the conflict graph of an existing office complex. However, the structure of
such conflict graphs is not clear and various fundamental questions related to
the problem at hands were asked. Arguably, one of the most natural questions
concerns the existence of bounds on the chromatic number of such conflict
graphs. More specifically, which additional constraints one should add to the
model to ensure ``good'' upper bounds on the chromatic number of conflict
graphs? These additional constraints should be meaningful regarding the
practical problem, reflecting as much as possible real-world situations.
Indeed, as noted by Tietze~\cite{Tie05}, complete graphs of arbitrary size are
conflict graphs, that is, for every integer $n$, there can be $n$
organisations whose territories all are in face-to-face contact with each
other. The reader is referred to the paper by Reed and Allwright~\cite{ReAl08}
for a description of Tietze's construction. Besicovitch~\cite{Bes47} and
Tietze~\cite{Tie65} proved that this is still the case if the territories are
asked to be convex polyhedra.
An interesting condition is when the territories are required to be
\emph{rectangular parallelepipeds} (sometimes called \emph{cuboids}), that is,
a $3$-dimensional solid figure bounded by six rectangles aligned with a fixed
set of Cartesian axes. For convenience, we shall call \emph{box} a rectangular
parallelepiped. When all territories are boxes, the \emph{clique number} of
any conflict graph, that is, the maximum size of a complete subgraph, is at
most $4$. However, Reed and Allwright~\cite{ReAl08} and also later Magnant
and Martin~\cite{MaMa11} designed arrangements of boxes that yield conflict
graphs requiring an arbitrarily high number of colours.
\noindent On the other hand, if the building is assumed to have floors (in the
usual way) and each box is \emph{$1$-floor}, i.e. restricted to be within one
floor, then the chromatic number is bounded by $8$: on each floor, the
obtained conflict graph is planar and hence can be coloured using $4$
colours~\cite{ApHa77,AHK77}. It is natural to ask whether this bound is
tight. As noted during working sessions in Oxford (see the acknowledgments),
it can be shown that up to $6$ colours can be needed, by using an arrangement
of boxes spanning three floors. Such a construction is shown in
Figure~\ref{fig:6}.
\begin{figure}[!ht]
\begin{center}
\begin{tikzpicture}[line width = 1mm]
\tikzset{BarreStyle/.style = {opacity=.4,line width=4 mm,line cap=round,color=#1}}
\draw (0,0) -- (0,4) -- (5,4) -- (5,0) -- cycle ;
\foreach \i in {1,2,3,4} { \draw (\i,1) -- (\i,3); }
\draw (2,2)--(3,2);
\draw (0,1)--(5,1);
\draw (0,3)--(5,3);
\draw [line width=.3mm,densely dotted] (-.5,-.5)--(-.5,4.5)--(5.5,4.5)--(5.5,-.5)--cycle;
\draw [line width=.3mm,densely dotted] (2.5,4.4)--(2.5,-.45);
\draw [line width=.5mm, loosely dashed] (.5,.5)--(.5,3.5)--(4.5,3.5)--(4.5,.5)--cycle;
\end{tikzpicture}
\end{center}
\caption{An arrangement of $1$-floor boxes spanning three floors and requiring
six colours. The solid, dotted, and dashed lines indicate the middle, top, and
bottom floors, respectively.}\label{fig:6}
\end{figure}
The purpose of this note is to show that the upper bound is actually tight.
More precisely, we shall build an arrangement of $1$-floor boxes that spans
two floors and yields a conflict graph requiring $8$ colours. From now on, we
shall identify a box arrangement with its conflict graph for convenience. In
particular, we assign colors directly to the boxes and define the
\emph{chromatic number of an arrangement} as that of the associated conflict
graph.
\begin{theorem}
\label{thm:8}
There exists an arrangement of $1$-floor boxes spanning two floors with
chromatic number $8$.
\end{theorem}
The boxes considered in Theorem~\ref{thm:8} have one of their geometrical
measures bounded: their height is at most one floor. We also discuss bounds
on the chromatic number of box arrangements with respect to some other
geometrical measures: the side lengths, the surface area and the volume. More
precisely, assuming that boxes have integer coordinates, we obtain the
following.
\begin{theorem}
\label{thm:geom}
We consider a box arrangement $A$ with integer coordinates.
\begin{enumerate}
\item If there exists one fixed dimension such that every box in
$A$ has length at most $\ell$ in this dimension, then
$A$ has chromatic number at most $4(\ell+1)$.
\item If for each box, there is one dimension such that the
length of this box in this dimension if at most $\ell$, then $A$
has chromatic number at most $12(\ell+1)$.
\item If the total surface area of each box in $A$ is
at most $s$, then $A$ has chromatic number at most
$9\sqrt[3]{4s}+13$.
\item If the volume of each box in $A$ is at most $v$,
then $A$ has chromatic number at most $24\sqrt[4]{6v}+13$.
\end{enumerate}
\end{theorem}
\noindent In the next section, we give the proof of Theorem~\ref{thm:8} and in
the last section we indicate how to obtain the bounds given in
Theorem~\ref{thm:geom}.
\section{Proof of Theorem~\ref{thm:8}}
We shall construct an arrangement of $1$-floor boxes that is not
$7$-colorable. Before that, we need the following definition. Consider an
arrangement $A$ and let $A_1$, $A_2$ and $A_3$ be (non-necessarily disjoint)
subsets of the boxes in $A$. Given a proper coloring $c$ of $A$, let $C_i$ be
the set of colors used for the boxes in $A_i$, for $i\in\{1,2,3\}$. The
\emph{signature $\sigma_A(c)$ of $A$ with respect to $c$} is defined to be
$a_xb_yc$, where $a\coloneqq\abs{C_1}$, $x\coloneqq\abs{C_1 \cup C_2}$,
$b\coloneqq\abs{C_2}$, $y\coloneqq\abs{C_2 \cup C_3}$, and
$c\coloneqq\abs{C_3}$. The collection of all signatures can be endowed with a
partial order: if $s=a_xb_yc$ and $s'=a'_{x'}b'_{y'}c'$ are two signatures,
then $s\le s'$ if $a\le a'$, $x\le x'$, $b\le b'$, $y\le y'$ and $c\le c'$.
To build the desired arrangement, we use the arrangement $X$ of $1$-floor
boxes described in Figure~\ref{fig-X} as a building brick. The arrangement $X$
has three specific \emph{regions}, $X_1$, $X_2$ and $X_3$. We also abusively
write $X_1$, $X_2$ and $X_3$ to mean the subsets of boxes of $X$ respectively
intersecting the regions $X_1$, $X_2$ and $X_3$ (note that some boxes may
belong to several subsets). We start by giving some properties of the
signatures of $X$ with respect to proper colorings and according to the three
subsets $X_1$, $X_2$ and $X_3$.
\begin{figure}[!ht]
\begin{center}
\begin{tikzpicture}[line width=1mm,rotate=90]
\tikzset{BarreStyle/.style = {opacity=.3,line width=6 mm,color=#1}}
\draw (0,0) -- (3,0) -- (3,5) -- (0,5) -- cycle ;
\draw (0,1) -- (3,1) ;
\draw (0,2) -- (3,2) ;
\draw (2,3) -- (3,3) ;
\draw (0,4) -- (3,4) ;
\draw (0,5) -- (3,5) ;
\draw (1,2) -- (1,4) ;
\draw (2,2) -- (2,4) ;
\draw [BarreStyle=red] (.5,4.8) to (.5,.2) ;
\draw [BarreStyle=blue] (1.5,4.8) to (1.5,.2) ;
\draw [BarreStyle=green] (2.5,4.8) to (2.5,.2) ;
\draw node at (.5,-.4) {$X_1$};
\draw node at (1.5,-.4) {$X_2$};
\draw node at (2.5,-.4) {$X_3$};
\end{tikzpicture}
\end{center}
\caption{The gadget $X$ with the regions $X_1$, $X_2$ and $X_3$.}
\label{fig-X}
\end{figure}
\begin{claim}\label{claim-X}
For every proper coloring $c$ of $X$,
\begin{enumerate}
\item $\sigma_X(c) \geq 3_32_44$,\label{X33244}
\item $\sigma_X(c) \geq 3_43_32$, or\label{X34332}
\item $\sigma_X(c) \geq 2_33_43$.\label{X23343}
\end{enumerate}
\end{claim}
The proof of this assertion does not need any insight, we thus omit it.
However, the reader interested in checking its accuracy should first note that
one can restrict to the cases where the three vertical (in Figure~\ref{fig-X})
boxes are respectively colored either $1$, $2$ and $1$; or $1$, $1$ and $2$;
or $1$, $2$ and $3$.
\begin{figure}[!ht]
\begin{center}
\begin{tikzpicture}[line width = 1mm,scale=.78]
\tikzset{BarreStyle/.style = {opacity=.3,line width=6 mm,color=#1}}
\draw (0,0)--(15,0)--(15,5)--(0,5)--cycle;
\foreach \i in {1,3,4,5,6,8,9,10,11,13,14} { \draw (\i,0)--(\i,5) ; }
\draw (1,2)--(3,2);
\draw (1,3)--(3,3);
\draw (2,3)--(2,5);
\draw (6,1)--(8,1);
\draw (6,3)--(8,3);
\draw (7,3)--(7,5);
\draw (11,4)--(13,4);
\draw (11,2)--(13,2);
\draw (12,0)--(12,2);
\draw [BarreStyle=red] (0.2,1.5) to (14.8,1.5) ;
\draw [BarreStyle=blue] (0.2,2.5) to (14.8,2.5) ;
\draw [BarreStyle=green] (0.2,3.5) to (14.8,3.5) ;
\draw node at (-.5,1.5) {$Y_1$};
\draw node at (-.5,2.5) {$Y_2$};
\draw node at (-.5,3.5) {$Y_3$};
\draw [decorate,decoration={brace,amplitude=5pt},line width=.3mm] (0.1,5.2) -- (4.9,5.2) node[midway,above=.4em] {$X^1$};
\draw [decorate,decoration={brace,amplitude=5pt},line width=.3mm] (5.1,5.2) -- (9.9,5.2) node[midway,above=.4em] {$X^2$};
\draw [decorate,decoration={brace,amplitude=5pt},line width=.3mm] (10.1,5.2) -- (14.9,5.2) node[midway,above=.4em] {$X^3$};
\end{tikzpicture}
\end{center}
\caption{The gadget $Y$ with the regions $Y_1$, $Y_2$ and $Y_3$.}
\label{fig-Y}
\end{figure}
Now, the arrangement $Y$ is obtained from three copies $X^1$, $X^2$ and $X^3$
of the arrangement $X$. We define three regions $Y_1$, $Y_2$ and $Y_3$ on $Y$
as depicted in Figure~\ref{fig-Y}. As previously, we also write $Y_1$, $Y_2$
and $Y_3$ for the subsets of boxes intersecting the region $Y_1$, $Y_2$ and
$Y_3$, respectively. We set $X_i^j\coloneqq Y_i\cap X^j$ for
$(i,j)\in\{1,2,3\}^2$.
\begin{claim}
\label{claim-Y}
In any proper coloring of $Y$, at least four colors are used in one
of the three regions.
\end{claim}
\begin{proof}
Suppose on the contrary that there is a proper coloring $c$ of $Y$ with at
most three colors in each of $Y_1$, $Y_2$ and $Y_3$. For $i\in\{1,2,3\}$, the
restriction of $c$ to $X^i$ is a proper coloring of $X^i$, which we identify
to $c$. The condition on $c$ implies that none of $\sigma_{X^1}(c)$,
$\sigma_{X^2}(c)$ and $\sigma_{X^3}(c)$ fulfills inequality~\ref{X33244} of
Assertion~\ref{claim-X}. In particular, note that exactly $3$ different
colours appear on $X_2^1$, and they also appear on $X_2^2$ and on $X_2^3$.
Since $X_1^2=X_2^2$, these three colours appear on $Y_1$. Similarly, since
$X_3^3=X_2^3$, these three colours appear on $Y_3$.
Assume now that $\sigma_{X^1}(c)$ satisfies~\ref{X34332}. Then exactly three
colours appear on $X_1^1$, one of which does not appear on $X_2^1$ as
$\abs{c(X_1^1\cup X_2^1)}\ge4$. Thus in total at least four colours appear on
$X_1^1\cup X_1^2\subset Y_1$, which contradicts our assumption on $c$. It
remains to deal with the case where $\sigma_{X^1}(c)$ fulfills~\ref{X23343} of
Assertion~\ref{claim-X}. Thanks to the symmetry of~\ref{X34332}
and~\ref{X23343} with respect to the regions $X_1$ and $X_3$, the same
reasoning as above applied to $X_3^3$ instead of $X_1^2$ yields that four
colours appear on $Y_3$, a contradiction.
\end{proof}
To finish the construction, we need the following definition. Consider
two copies $Y^1$ and $Y^2$ of $Y$. The regions $Y_i^1$ and $Y_j^2$
\emph{fully overlap} if every box in $Y_i^1$ is in face-to-face
contact with every box in $Y_j^2$. Observe that for every pair
$(i,j)\in\{1,2,3\}$, there exists a $2$-floor arrangement of $Y^1$ and
$Y^2$ such that $Y_i^1$ and $Y_j^2$ fully overlap: it is obtained by
rotating $Y^2$ ninety degrees, adequately scaling it (i.e. stretching it
horizontally) and placing it on top of $Y^1$.
We are now in a position to build the desired arrangement $Z$ spanning
two floors. To this end, we use several copies of $Y$. The first
floor of $Z$ is composed of seven parallel copies $Y^1,\ldots,Y^7$ of
$Y$ (drawn horizontally in Figure~\ref{fig-Z}). The second floor of
$Z$ is composed of fifteen parallel copies of $Y$ (drawn vertically in
Figure~\ref{fig-Z}): for each $j\in\{1,2,3\}$ and each
$i\in\{2,\ldots,6\}$, a copy $Y(i,j)$ of $Y$ is placed such that the
first region of $Y(i,j)$ fully overlap the regions
$Y_j^1,\ldots,Y_{j}^{i-1}$, the second region of $Y(i,j)$ fully
overlaps the region $Y_j^i$, and the third region of $Y(i,j)$ fully
overlaps the regions $Y_j^{i+1},\ldots,Y_{j}^{7}$.
\begin{figure}
\begin{tikzpicture}[line width = .4mm,scale=.62,every node/.style={scale=.62}]
\tikzset{BarreStyle/.style = {opacity=.3,line width= 6 mm,color=#1}}
\foreach \i in {1,...,7} {
\draw (0,2*\i)--(18.15,2*\i)--(18.15,-1+2*\i)--(0,-1+2*\i)--cycle ;
\draw node at (-.8,-.5+2*\i) {\large $Y^\i$} ;
\draw [BarreStyle=red] (.2,1.5+2*\i-2) to (6,1.5+2*\i-2) ;
\draw [BarreStyle=blue] (6.2,1.5+2*\i-2) to (12,1.5+2*\i-2) ;
\draw [BarreStyle=green] (12.2,1.5+2*\i-2) to (18,1.5+2*\i-2) ;
}
\tikzset{BarreStyle/.style = {opacity=.3,line width=4 mm,color=#1}}
\foreach \i in {0,1,2} {
\foreach \j in {0,1,2,3,4} {
\draw (.5+1.1*\j+6*\i,.5)--(.5+1.1*\j+6*\i,14.5)--(1.25+6*\i+1.1*\j,14.5)--(1.25+6*\i+1.1*\j,.5)--cycle ;
\draw [BarreStyle=red] (.87+1.1*\j+6*\i,2.3+2*\j) to (.87+1.1*\j+6*\i,0.7) ;
\draw [BarreStyle=blue] (.87+1.1*\j+6*\i,2.7+2*\j) to (.87+1.1*\j+6*\i,4.3+2*\j) ;
\draw [BarreStyle=green] (.87+1.1*\j+6*\i,14.3) to (.87+1.1*\j+6*\i,4.7+2*\j) ;
}
}
\foreach \i in {1,2,3} {
\foreach \j in {2,...,6} {
\draw node at (\j*1.1-2.2+.3+.55+\i*6-6,0) {\footnotesize $Y(\j,\i)$} ;
}
}
\end{tikzpicture}
\caption{Schematic view of the arrangement $Z$.}
\label{fig-Z}
\end{figure}
Consider a proper coloring of $Z$. Assertion~\ref{claim-Y} ensures that each
copy of $Y$ in $Z$ has a region for which at least four different colours are
used. In particular, there exists $j\in\{1,2,3\}$ such that three regions
among $Y_j^1,\ldots,Y_j^7$ are colored using four colours. Let these regions
be $Y_j^{i_1}$, $Y_j^{i_2}$ and $Y_{j}^{i_3}$ with $1\le i_1<i_2<i_3\le7$.
Now, consider the arrangement $Y(i_2,j)$. By Assertion~\ref{claim-Y}, there
exists $k\in\{1,2,3\}$ such that the $k$-th region of $Y(i_2,j)$ is also
colored using at least four different colors. Consequently, as this region
and the region $Y_j^{i_k}$ fully overlap, they are colored using at least
eight different colors. This concludes the proof.
\section{Bounds with respect to geometrical measures}
In this part, we provide bounds on the chromatic number of boxes arrangements
provided that the boxes satisfy some geometrical constraints. Namely, we prove
Theorem~\ref{thm:geom}, which is recalled here for the reader's ease.
\begin{theo2}
We consider a box arrangements $A$ with integer
coordinates.
\begin{enumerate}
\item If there exists one fixed dimension such that every box in $A$ has
length at most $\ell$ in this dimension, then $A$ has chromatic number
at most $4(\ell+1)$.\label{toutes}
\item If for each box, there is one dimension such the length of this box
in this dimension if at most $\ell$, then $A$ has chromatic number at
most $12(\ell+1)$.\label{une}
\item If the total surface area of each box in $A$ is at most $s$, then $A$
has chromatic number at most $9\sqrt[3]{4s}+13$.\label{surface}
\item If the volume of each box in $A$ is at most $v$, then $A$ has
chromatic number at most $24\sqrt[4]{6v}+13$.\label{volume}
\end{enumerate}
\end{theo2}
\begin{proof}\mbox{}
\noindent
\ref{toutes} The conflict graph corresponding to an arrangement where the
boxes have height at most $\ell$ can be vertex partitioned into $\ell+1$
planar graphs $P_0,\ldots,P_\ell$. Indeed if the distance between the levels
of two boxes is at least $\ell+1$, then these two boxes are not adjacent. So
the planar graphs are obtained by assigning, for each $x$, all the boxes that
have their floor at level $x$ to be in the graph $P_k$ where $k\coloneqq x
\mod (\ell+1)$. Consequently, the whole conflict graph has chromatic number at
most $4(\ell+1)$.
\smallskip
\noindent\ref{une} The boxes can be partitioned into three sets according to
the dimension in which the length is bounded. In other words, $A$ is
partitioned into $U_1$, $U_2$ and $U_3$ such that for each $i\in\{1,2,3\}$,
all boxes in $U_i$ have length at most $\ell$ in dimension $i$.
Consequently,~\ref{toutes} ensures that each of $U_1$, $U_2$ and $U_3$ has
chromatic number at most $4(\ell+1)$ and, therefore, $A$ has chromatic number
at most $3\cdot 4(\ell+1)=12(\ell+1)$.
\smallskip
\noindent\ref{surface} For each box, the minimum length taken over all three
dimensions is at most $\sqrt{s}$, and thus~\ref{une} implies that the
chromatic number of $A$ is $\bigo{\sqrt{s}}$. However, one can be more
careful. Let us fix a positive integer $\ell$, to be made precise later. The
set of boxes is partitioned as follows. Let $U$ be the set of boxes with
lengths in every dimension at least $\ell$ and let $R$ be the set all
remaining boxes, that is, $R\coloneqq A\setminus U$. By~\ref{une}, the
arrangement $R$ has chromatic number at most $12\ell$. Now consider a box $B$
in $U$ with dimensions $x$, $y$ and $z$, each being at least $\ell$. We shall
give an upper bound on the number of boxes of $U$ that can be adjacent to $B$.
The surface of a face of a box in $U$ is at least $\ell^2$. So in $U$ there
are at most $s/\ell^2$ that have a face totally adjacent to a face of $B$.
Some boxes of $U$ could also be adjacent to $B$ without having a face totally
adjacent to a face of $B$. In this case, such a box is adjacent to an edge of
$B$. For an edge of length $w$, there are at most $w/\ell+1$ such boxes. So
the number of boxes of $U$ adjacent to $B$ but having no face totally adjacent
to a face of $B$ is at most $4(x+y+z)/\ell+12$. Since $\ell\le\min\{x,y,z\}$,
we deduce that $2\ell(x+y+z)\le 2xy+2yz+2xz\le s$. Hence the total number of
boxes in $U$ that are adjacent to $B$ is at most
$s/\ell^2+2s/\ell^2+12=3s/\ell^2+12$. Consequently, by degeneracy, $U$ has
chromatic number at most $3s/\ell^2+13$. Therefore, $A$ has chromatic number
at most $12\ell+3s/\ell^2+13$. Setting $\ell\coloneqq\sqrt[3]{s/2}$ yields
the upper bound $9\sqrt[3]{4.s}+13$.
\smallskip
\noindent\ref{volume} Once again, for a fixed parameter $\ell$ to be made
precise later, the set of boxes is partitioned into two parts: the part $U$,
composed of all the boxes with lengths in every dimension at least $\ell$ and
the part $R$, composed of all the remaining boxes. By~\ref{une}, we know that
$R$ has chromatic number at most $12\ell$. Let $B$ be a box in $U$ with
dimensions $x$, $y$ and $z$. Since $\ell\le \min\{x,y,z\}$, the volume $v_B$
of $B$ satisfies that $6v\ge6v_B=6xyz\ge2(\ell xy+\ell xz+\ell yz)=\ell s_B$,
where $s_B$ is the total surface area of $B$. So every box in $U$ has total
surface area at most $6v/\ell$ and thus~\ref{surface} implies that $U$ has
chromatic number at most $9\sqrt[3]{4.6v/\ell}+13$. Therefore, $A$ has
chromatic number at most $9\sqrt[3]{24v/\ell}+12\ell+13$. Setting $\ell$ to be
$\sqrt[4]{3v/8}$ yields the upper bound $24\sqrt[4]{6v}+13$.
\end{proof}
In the previous theorem, we are mainly concerned with the order of magnitude
of the functions of the different parameters. However, even in this context,
we do not have any non trivial lower bound on the corresponding chromatic
numbers.
\section*{Acknowledgments.} The third author thanks Louigi Addario-Berry,
Frédéric Havet, Ross Kang, Colin McDiarmid and Tobias Müller for stimulating
discussions on the topic of this paper during a stay in Oxford, in 2005.
|
\section{Introduction}
\label{intro}
Since quantum mechanics principles are introduced into cryptography, quantum cryptography attracts more and more attention. Due to the characteristic of quantum unconditional security, many quantum cryptography protocols, such as quantum key distribution (QKD) [1-3], quantum direct communication (QDC) [4-7], quantum secret sharing (QSS) [8-10], quantum teleportation (QT) [11,12], have been proposed to solve various secure problems.
Recently, quantum private comparison (QPC) has become an important branch in quantum cryptography. Based on the properties of quantum mechanics, the participants can determine whether their secret inputs are equal or not without disclosing their own secrets to each other. In 2009, Yang et al. [13] put forward a pioneering QPC scheme based on Bell states and a hash function. Since then, a large number of QPC protocols utilizing the entangled states, such as EPR pairs, GHZ state, etc., have been proposed [14-23].
However, these QPC protocols [13-23] are feasible in the ideal scenario but not secure in practical scenario where faults (including noise and error) are existent in the quantum channel and measurement. In order to solve the problem, in 2013, Li et al. [24] present a novel QPC scheme based on GHZ states and error-correcting code (ECC) against noise. But, through analyzing Li et al.'s QPC scheme, we find it is unsecure under a special attack, called the twice-Hadamard-CNOT attack. To be specific, if any malicious party performed the twice-Hadamard-CNOT attack, he (she) can get another's secret input, which goes against the QPC's principles [25]. In order to fix the loophole, a simple solution by adopting a permutation operator before TP distributes the particles to the participants, is proposed.
The rest of this paper is constructed as follows. At first, Li et al.'s QPC protocol is briefly reviewed in Sect. 2. And in Sect. 3, we analyze the security of Li et al.'s QPC protocol by introducing the twice-Hadmard-CNOT attack, and give an improvement to fix the problem. Finally, a short collusion is drawn in Sect. 4.
\section{Review of Li et al.'s QPC protocol}
In the Ref. [24], in order to guarantee the QPC protocols secure in the practical scenario, Li et al. present a new two-party QPC scheme by using error-correcting code. The whole protocol consists of eight steps as below.
\begin{enumerate}[(1)]
\item
Alice, Bob and Calvin prepare a [$m,n$] error-correcting code which uses $m$ bits codeword to encode $n$ bits word and can correct $l$ error bits in codeword with the error-correcting function $D(x^{m})$ according to the fault rate of the noise scenario. We suppose the error-correcting code's generator matrix is $G$, and check matrix is $Q$. Then they encode $X=(x_{0},x_{1},...,x_{n-1})$ and $Y=(y_{0},y_{1},...,y_{n-1})$ to $X'=(x'_{0},x'_{1},...,x'_{m-1})$ and $Y'=(y'_{0},y'_{1},...,y'_{m-1})$ with the generator matrix $G$, respectively. There are
\begin{equation}
X'=X\cdot G,
\end{equation}
\begin{equation}
Y'=Y\cdot G.
\end{equation}
\item
Calvin prepares $m$ triplet GHZ states all in
\begin{equation}
\begin{split}
|\psi\rangle_{CAB}&=\frac{1}{\sqrt{2}}(|000\rangle+|111\rangle)_{CAB} \\
&=\frac{1}{2}(|+++\rangle+|+--\rangle+|-+-\rangle+|--+\rangle)_{CAB},
\end{split}
\end{equation}
where $|0\rangle$ and $|1\rangle$ are measured in $Z$ basis, $|+\rangle$ and $|-\rangle$ are measured in $X$ basis, and $|\pm\rangle=\frac{1}{\sqrt{2}}(|0\rangle\pm|1\rangle)$. Calvin divides these $m$ GHZ states into three sequences $S_{A}$, $S_{B}$ and $S_{C}$, which includes the first, the second, and the third particles of all GHZ states, respectively.
\item
Calvin prepares some decoy photons prepared in states ${|0\rangle,|1\rangle,|+\rangle,|-\rangle}$ in random. He inserts these decoy photons in $S_{A}$ and $S_{B}$ at random positions to form sequences $S^{*}_{A}$ and $S^{*}_{B}$ respectively. Calvin retains the quantum sequence $S_{C}$ and sends the sequence $S^{*}_{A}$ to Alice, $S^{*}_{B}$ to Bob.
\item
When Alice and Bob receive $S^{*}_{A}$ and $S^{*}_{B}$, Calvin announces the positions and measurement base of these decoy photons. Alice and Bob measure them in the same base and announce their outcome. If the error rate exceeds a rational threshold, Calvin aborts the protocol and restarts from Step (1). Otherwise, there is no eavesdropper, and the protocol continues to the next step.
\item
Alice and Bob recover $S_{A}$ and $S_{B}$ respectively by discarding the decoy photons. Then Alice, Bob and Calvin measure $S_{A}$, $S_{B}$ and $S_{C}$ in $X$ basis, respectively. If the measurement result is $|+\rangle$ ($|-\rangle$), then they encode it as the classical bit 0 (1). Thus, each of Alice, Bob and Calvin will obtain $m$ bits from $S_{A}$, $S_{B}$ and $S_{C}$, respectively. We denote each of them as $k^{A}_{i}$, $k^{B}_{i}$ and $k^{C}_{i}$ (i=0,1,...,m-1).
\item
Alice and Bob calculate $x^{''}_{i}=k^A_{i}\oplus x^{'}_{i}$ and $y^{''}_{i}=k^{B}_{i}\oplus y^{'}_{i}$. They announce $X^{''}=(x^{''}_{0},x^{''}_{1},...,x^{''}_{m-1})$ and $Y^{''}=(y^{''}_{0},y^{''}_{1},..,y^{''}_{m-1})$ to Calvin.
\item
Calvin calculates $c^{'}_{i}=k^C_{i}\oplus x^{''}_{i}\oplus y^{''}_{i}$, and gets $m$ bits sequence $C^{'}=(c^{'}_{0},c^{'}_{1},...,c^{'}_{m-1})$.
\item
Then Calvin uses the check matrix $Q$ to check whether the number of error bits exceeds the threshold $l$. If it does, Calvin aborts the protocol and restarts from Step (1). Otherwise, he gets $n$ bits sequence $C^{*}$ by decoding $C^{'}$ with error-correcting function $D(C^{'})$. If there is at least one bit 1 in $C^{*}$, Calvin announces $X\neq Y$. Otherwise, he announces $X=Y$.
\end{enumerate}
In Ref. [24], the authors claimed the scheme was secure even in the practical scenario. However, we will show how a malicious party gets the other's secret input by launching the twice-Hadmard-CNOT attack in the next section.
\section{Twice-Hadmard-CNOT attack on Li et al.'s QPC protocol and the improvement}
\subsection{Twice-Hadmard-CNOT attack on Li et al.'s QPC protocol}
As analyzed in Ref.[24], due to the decoy photons adopted in the Li et al.'s QPC protocol, some well-known attacks, such as intercept-resend attack, measurement-resend attack, and entanglement-resend attack, can be detected via the checking mechanism. Unfortunately, we find Li et al.'s QPC protocol cannot resist a special attack, i.e., the twice-Hadmard-CNOT attack. To be specific, if any party (Alice or Bob) performs the twice-Hadmard-CNOT attack, he/she can get the other's secret input. The detailed procedure of the twice-Hadmard-CNOT attack is depicted as follows.
Without loss of generality, we suppose Bob is malicious, who wants to get Alice's secret input. At first, Bob prepares $m$ auxiliary particles all in the state $|0\rangle_{e}$, then, the GHZ state and an auxiliary particle compose a composite system:
\begin{equation}
\begin{split}
|\eta\rangle&=|\psi\rangle_{CAB}|0\rangle_{e} \\
&=\frac{1}{\sqrt{2}}(|0000\rangle+|1110\rangle)_{CABe} \\
&=\frac{1}{2}(|+++0\rangle+|+--0\rangle+|-+-0\rangle+|--+0\rangle)_{CABe},
\end{split}
\end{equation}
where the subscript $C$, $A$, $B$ represent the particles in the hand of Calvin, Alice, Bob, respectively, and the subscript $e$ represents the auxiliary particle.
In Step (3), when Calvin sends the sequence $S^{*}_{A}$ to Alice, Bob may intercept $S^{*}_{A}$ and execute a Hardmard ($H$) operation on every particle in $S^{*}_{A}$ to form sequence $S^{**}_{A}$. Then, he performs a controlled-NOT (CNOT) operation $C_{Ae}$ on every particle in $S^{**}_{A}$ and the corresponding auxiliary particle $e$. Here, the particle in $S^{**}_{A}$ is the control qubit, while particle $e$ is the target qubit. After that, Bob performs another $H$ operation on every particle in $S^{**}_{A}$ in order to restore sequence $S^{**}_{A}$ to $S^{*}_{A}$, and sends $S^{*}_{A}$ to Alice. What should be noted is that the transmitting sequence $S^{*}_{A}$ remains unchanged.
Since Calvin announces the decoy photons' positions in Step (4), Bob can discard the auxiliary particles $e$ which correspond to the decoy photons in $S^{*}_{A}$. And in Step (5), after Bob recovers $S_{B}$ by discarding the decoy photons, he executes an $H$ operation on every particle in $S_{B}$ to form sequence $S^{**}_{B}$. Then Bob performs a CNOT operation $C_{Be}$ on every particle (control particle) in $S^{**}_{B}$ and the corresponding auxiliary particle $e$ (target particle). After that, he performs a $H$ operation on every particle in $S^{**}_{B}$, which aims to restore $S^{**}_{B}$ to $S_{B}$, and every auxiliary particle, respectively. Now, the state of the composite system is changed into
\begin{equation}
|\eta^{'}\rangle=\frac{1}{2}(|++\rangle_{Ce}(|++\rangle+|--\rangle)_{AB}+|--\rangle_{Ce}(|+-\rangle+|-+\rangle)_{AB}).
\end{equation}
From Eq.(5), a rule can be concluded: if $e$ is $|+\rangle$, particle $A$ and particle $B$ are in the same state; and if $e$ is $|-\rangle$, they are in the different states.
After Alice announces $X^{''}=(x^{''}_{0},x^{''}_{1},...,x^{''}_{m-1})$ in Step (6), Bob measures them in the $X$ basis, and get the measurement result $k^{e}_{i}(i=0,1,...,m-1)$. According to $k^{e}_{i}$ and the rule from Eq.(5), Bob can obtain Alice's states $k^{A}_{i}(i=0,1,...,m-1)$. Since $X^{'}$ has been announced by Alice, Bob can calculate $x^{''}_{i}\oplus k^{A}_{i}=(k^{A}_{i}\oplus x^{'}_{i})\oplus k^{A}_{i}=x^{'}_{i}$, that is to say, he can get Alice's secret input $x^{'}_{i}$.
For sake of clearness, the above procedure of the twice-Hadmard-CNOT attack can be intuitively demonstrated by the following figure (See Fig. 1).
\begin{figure}
\includegraphics{Fig1.eps}
\caption{The circuit diagram of the process of twice-Hadmard-CNOT attack. Here, $|\psi\rangle_{CAB}$ is a GHZ state in the state $(|000\rangle+|111\rangle)/\sqrt{2}$ shared by Calvin, Alice and Bob. $H$ is Hadmard operation, and \protect\includegraphics{Fig2.eps} represents the controlled-NOT gate in which the top line denotes the control qubit, the bottom line the target qubit.}
\end{figure}
\subsection{The improvement}
In order to fix the loophole of Li et al.'s QPC protocol, we meliorate it by applying a permutation operator before TP sends the particle sequences to all the participants. To be specific, Step (3), Step (4) and Step (5) in the original protocol need to be revised as follows.
\begin{enumerate}
\item [$(3)^{*}$]
Calvin prepares two group $r$-length decoy photons sequences at random in $\{|+\rangle,|-\rangle,|0\rangle,|1\rangle\}$, namely $R_{A}$, $R_{B}$, and concatenates them with $S_{A}$ ($S_{B}$) to form an extended sequence $S^{'}_{A}=R_{A}||S_{A}$ ($S^{'}_{B}=R_{B}||S_{B}$), respectively. Then Calvin applies a permutation operator $\Pi_{(r+m)}$ on $S^{'}_{A}$ ($S^{'}_{B}$) to create a new sequence $\Pi_{(r+m)}S^{'}_{A}=S^{*}_{A}$ ($\Pi_{(r+m)}S^{'}_{B}=S^{*}_{B}$), and sends the new sequence $S^{*}_{A}$ to Alice, $S^{*}_{B}$ to Bob.
\item[$(4)^{*}$]
When Alice and Bob receive $S^{*}_{A}$ and $S^{*}_{B}$, Calvin announces the coordinates of the decoy qubits $\Pi_{r}$ ($\Pi_{r}\subset\Pi_{r+m}$) send by him, and the corresponding measurement bases. Note that Calvin does not disclose the actual order of the message qubits. Then Alice and Bob measure these decoy qubits in the same bases and announce their outcomes. If the error rate exceeds a rational threshold, Calvin aborts the protocol and restarts from Step (1); otherwise, there is no eavesdropper, and the protocol continues to the next step.
\item[$(5)^{*}$]
Alice and Bob discard their decoy photons, and denote the left qubits in their hands as $S^{''}_{A}$ and $S^{''}_{B}$ (i.e.,$S^{''}_{A}=\Pi_{m}S_{A},S^{''}_{B}=\Pi_{m}S_{B}$), respectively. Then Alice, Bob and Calvin measure $S^{''}_{A}$, $S^{''}_{B}$ and $S_{C}$ in the $X$ basis, respectively, and obtain $m$ bits, $k^{A}_{i}$, $k^{B}_{i}$ and $k^{C}_{i}$ (i=0,1,...,m-1 ), respectively. After they have completed measurement operation, Calvin immediately announces the actual order of the message qubits $\Pi_{m}$ ($\Pi_{m}\subset\Pi_{r+m}$). Using this information, Alice (Bob) can rearrange $k^{A}_{i}$ ($k^{B}_{i}$) in correspondence with the original order of $S_{A}$ ($S_{B}$).
\end{enumerate}
Let us examine the security of our improvement under the twice-Hadmard-CNOT attack. Similarly, we suppose the malicious Bob aims to get Alice's secret input. In Step $(3)^{*}$, Since Calvin disrupts the sequences $S^{'}_{A}$ and $S^{'}_{B}$ by using $\Pi_{(r+m)}$ to get $S^{*}_{A}$, $S^{*}_{B}$, it means the order of transmited sequences $S^{*}_{A}$, $S^{*}_{B}$ is fully disturbed. And after discarding the decoy qubits $R_{A}$ and $R_{B}$ in Step $(5)^{*}$, Alice and Bob can only get $S^{''}_{A}$ and $S^{''}_{B}$, and cannot recover the original sequences $S_{A}$ and $S_{B}$ because that $\Pi_{m}$ is only known by Calvin. So, even if Bob launch the twice-Hadmard-CNOT attack, he cannot get the final state $|\eta^{'}\rangle=\frac{1}{2}(|++\rangle_{Ce}(|++\rangle+|--\rangle)_{AB}+|--\rangle_{Ce}(+-\rangle+|-+\rangle)_{AB})$. That is to say, Bob cannot get the correlations between Alice's and Bob's qubits by measuring the auxiliary particles $e$.
For simplicity, we take a simple two-bit private comparisonas example without considering error-correcting code and the decoy photos. In the case, suppose Alice's input is 10, Bob's input is 11, and Calvin prepares two GHZ states all in the state as Eq.(3),
\begin{equation}
\begin{split}
|\psi\rangle_{C_{1}A_{1}B_{1}}\otimes|\psi\rangle_{C_{2}A_{2}B_{2}}
&=1/2(|+++\rangle+|+--\rangle+|-+-\rangle+|--+\rangle)_{C_{1}A_{1}B_{1}}\\
&\otimes1/2(|+++\rangle+|+--\rangle+|-+-\rangle+|--+\rangle)_{C_{2}A_{2}B_{2}}
\end{split}
\end{equation}
In Step $(3)^{*}$, Calvin executes a permutation operator $\Pi_{m=2}$ on sequence $S^{'}_{A}$ and $S^{'}_{B}$, then the system may be changed into
\begin{equation}
\begin{split}
\Pi_{m=2}(|\psi\rangle_{C_{1}A_{1}B_{1}}\otimes|\psi\rangle_{C_{2}A_{2}B_{2}})&=|\psi\rangle_{C_{1}A_{2}B_{2}}\otimes|\psi\rangle_{C_{2}A_{1}B_{1}}\\
&=1/4\{|+\rangle_{C_{1}}(|++\rangle+|--\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|+\rangle_{C_{2}}(|++\rangle+|--\rangle)_{A_{1}B_{1}}\\
&\quad\;+|+\rangle_{C_{1}}(|+-\rangle+|-+\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|-\rangle_{C_{2}}(|++\rangle+|--\rangle)_{A_{1}B_{1}}\\
&\quad\;+|-\rangle_{C_{1}}(|++\rangle+|--\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|+\rangle_{C_{2}}(|+-\rangle+|-+\rangle)_{A_{1}B_{1}}\\
&\quad\;+|-\rangle_{C_{1}}(|+-\rangle+|-+\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|-\rangle_{C_{2}}(+-\rangle+|-+\rangle)_{A_{1}B_{1}}
\}
\end{split}
\end{equation}
After Bob performs the twice-Hadmard-CNOT attack, the composite system which consists of GHZ state and auxiliary particle becomes
\begin{equation}
\begin{split}
|\eta^{'}\rangle_{1}\otimes|\eta^{'}\rangle_{2}
&=1/4\{|+\rangle_{C_{1}}|+\rangle_{e_{1}}(|++\rangle+|--\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|+\rangle_{C_{2}}|+\rangle_{e_{2}}(|++\rangle+|--\rangle)_{A_{1}B_{1}}\\ &\quad\;+|+\rangle_{C_{1}}|-\rangle_{e_{1}}(|+-\rangle|-+\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|-\rangle_{C_{2}}|+\rangle_{e_{2}}(|++\rangle+|--\rangle)_{A_{1}B_{1}}\\
&\quad\;+|-\rangle_{C_{1}}|+\rangle_{e_{1}}(|++\rangle+|--\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|+\rangle_{C_{2}}|-\rangle_{e_{2}}(|+-\rangle+|-+\rangle)_{A_{1}B_{1}}\\
&\quad\;+|-\rangle_{C_{1}}|-\rangle_{e_{1}}(|+-\rangle+|-+\rangle)_{A_{2}B_{2}}\\
&\quad\;\otimes|-\rangle_{C_{2}}|-\rangle_{e_{2}}(|+-\rangle+|-+\rangle)_{A_{1}B_{1}}
\}.
\end{split}
\end{equation}
From above equation, we cannot get the correlations between the states of $\{A_{1},B_{1}\}$ or $\{A_{2},B_{2}\}$ according to the the final state of the auxiliary particle $e_{1}$ or $e_{2}$. That means, Bob cannot steal Alice's input through measuring the auxiliary particles. So, we can say the improvement can resist the twice-Hadmard-CNOT attack.
\section{Conclusione}
In all the QPC protocols, in order to ensure the protocol's security, we must guarantee any participant only knows his (her) own secret input without obtaining another' secret input. In this paper, we firstly review and analyze Li et al.'s two-party QPC protocol, and find it cannot resist the twice-Hadmard-CNOT attack, i.e., if one participant Bob launches this attack, he (she) can get the other's secret input without being detected. For avoiding this loophole, we adopt the permutation operator to rearranges the quantum sequences sent to Alice and Bob from Charlie. The delicate analysis shows the security of our improvement can be guaranteed well and truly.\\\\
\textbf{Acknowledgments}\quad\;\small\textmd{This work is supported by the National Nature Science Foundation of China (Grant Nos. 61103235, 61373131 and 61373016), the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), and State Key Laboratory of Software Engineering, Wuhan University (SKLSE2012-09-41).}
|
\section{Introduction}
\label{Introduction}
Since the discovery of the late-time accelerated expansion of the Universe,
there has been much effort into understanding the perplexing nature of the
cosmic acceleration and of gravity itself. In a first direction, one can
introduce the dark-energy concept, in the context of scalar fields, as these
are popular building blocks used to construct models of present-day
cosmological acceleration. They are appealing because such fields are
ubiquitous in theories of high energy physics beyond the standard model and,
in particular, are present in theories which include extra spatial
dimensions, such as those derived from string theories. Thus, the
modification of the Universe content \cite{Copeland:2006wr} is materialized,
in general, by the addition of
extra dynamical scalar fields, for instance canonical (quintessence)
\cite{Ratra:1987rm,Wetterich:1987fm,Liddle:1998xm,
Guo:2006ab,Dutta:2009yb,Harko:2013gha},
phantom \cite{Caldwell:1999ew,Caldwell:2003vq,Nojiri:2003vn,Onemli:2004mb,
Saridakis:2008fy}, their combination
\cite{Cai:2009zp,Guo:2004fq,Setare:2008pz},
K-essence \cite{ArmendarizPicon:2000ah,Chimento:2003ta}, Galileon
\cite{Nicolis:2008in}, etc.
On the other hand, one can modify the gravitational sector
\cite{Capozziello:2011et}, for instance constructing
$f(R)$ gravity
\cite{Capozziello:2002rd,Chiba:2003ir,Allemandi:2005qs,Nojiri:2006gh,
Nojiri:2007as,
Amendola:2006we,Starobinsky:2007hu,LanahanTremblay:2007sg,Boehmer:2007kx,
Harko:2011nh}, $f(T)$
gravity
\cite{ Bengochea:2008gz,Linder:2010py,Chen:2010va,Iorio:2012cm},
Weyl-Cartan-Weitzenb\"{o}ck gravity \cite{W1,W2}, Gauss-Bonnet gravity
\cite{Nojiri:2005vv,Koivisto:2006xf}, Ho\v{r}ava-Lifshitz gravity
\cite{Horava:2009uw,Kiritsis:2009sh,Saridakis:2009bv,Bogdanos:2009uj},
nonlinear massive gravity
\cite{deRham:2010ik,Hinterbichler:2011tt,Cai:2012ag,deRham:2014zqa}, etc,
which apart from the cosmological motivation has the additional advantage of
an
expected improved ultra-violet and quantum behavior of the theory
\cite{Stelle:1976gc}.
Furthermore, one could also construct combinations of these two directions,
such as in scalar-tensor theories
\cite{Uzan:1999ch,Amendola:1999qq,Fujii:2003pa}, in generalized Galileons
\cite{DeFelice:2010nf,Deffayet:2011gz,DeFelice:2011bh,Leon:2012mt}, etc. The
common feature of all the above theories is that they incorporate additional
degrees of freedom relatively to General Relativity and the standard
model.
Another interesting approach is that one could handle the gravitational and
matter sectors ``democratically'', that is modifying the matter part in the
Lagrangian too, along with its coupling to gravity
\cite{Bertolami:2007gv,Bertolami:2008ab,Bertolami:2008zh,Bertolami:2009ic,
Harko:2008qz,Harko:2010mv,Harko:2012hm,Wang:2012rw,Harko:2011kv,
Haghani:2013oma, Odintsov:2013iba, Harko:2014sja, Harko:2014aja}.
Generally, these theories lead to non-geodesic motion, which takes place in
the
presence of an extra force orthogonal to the four-velocity. The Newtonian
limit of the equation of motion was also considered, and a procedure for
obtaining the energy-momentum tensor of the matter was presented. On the
other hand, the gravitational field equations are equivalent to the Einstein
equations of the $f(R)$ model in empty spacetime, but differ from them, as
well as from standard General Relativity, in the presence of matter.
Therefore, the predictions of these models could lead to some major
differences,
as compared to the predictions of standard General Relativity, or its
extensions ignoring the role of matter, in several problems of current
interest, such as cosmology, gravitational collapse or the generation of
gravitational waves. The study of these phenomena may also provide some
specific signatures and effects, which could distinguish and discriminate
between the various theories of modified gravity.
Having these in mind, one could try to construct theories with additional
auxiliary fields which can alter the dynamics, however being themselves
non-dynamical and thus without altering the degrees of freedom
\cite{Pani:2013qfa,Banados:2013vya,Guo:2014bxa}. Interestingly enough, the
requirements of satisfying the
weak equivalence principle and allowing for a covariant Lagrangian
formulation, lead the field equations of these theories to include, in
general, higher-order derivatives of the matter fields. Although this feature
places tight observational constraints on these theories, it is clear
that they correspond to novel modified classes.
The plan of the work is the following: In Section \ref{model}, we present
the gravitational modification with non-dynamical auxiliary fields,
extracting the corresponding cosmological equations. In Section
\ref{Solutions}, we solve the cosmological field equations and we provide the
cosmological
behavior for various matter equations of state. Finally, in Section
\ref{Conclusions} we briefly discuss and conclude our paper.
\section{Cosmology with higher-derivative matter fields }
\label{model}
In this section we extract the cosmological equations of the scenario at
hand. In a first subsection, we briefly present the underlying gravitational
theory, namely gravity with auxiliary fields. Then, we apply it in a
cosmological framework, providing the Friedmann equations and defining the
basic observables.
\subsection{Gravity with auxiliary fields}
The key point of the scenario at hand is that one introduces auxiliary
fields whose equations of motion are not dynamical \cite{Pani:2013qfa}. Thus,
one can use these equations to eliminate these auxiliary fields. In this way,
the equations of motion for the usual matter fields are not modified (which
is a significant advantage when it comes to the observational test of the
theory), however the gravitational equations of motion do change, and in
particular they acquire higher order derivatives of the matter fields.
More specifically, the field equations can be written, without loss of
generality, as $G_{ab}+\Lambda g_{ab}=T_{ab}+S_{ab}({\bf g},{\bf T})$, where
the second rank tensor $S_{ab}({\bf g},{\bf T})$ depends explicitly on the
metric ${\bf g}$ and the matter fields (${\bf T}$ in general). The tensor
$S_{ab}$ is imposed to vanish in vacuum and be divergence-free, in virtue of
the Bianchi identity and the conservation of the matter fields, namely
$\nabla_a T^{ab}=0$.
Thus, in the present context, these theories with auxiliary fields consist in
modifying the Einstein field equation through the addition of a
divergence-free tensor, that vanishes in the vacuum, and depends on the
metric, the matter energy-momentum tensor, and its derivatives.
These conditions place strict conditions on the second-rank tensor under
consideration, hence, independently of the specific initial action, the
resulting
Einstein equations become \cite{Pani:2013qfa}
\begin{eqnarray}
G_{ab} &=& T_{ab} - \Lambda g_{ab} -
\beta_1 \Lambda\, g_{ab}\,T + \frac{1}{4}\left(1-2\beta_1
\Lambda\right)(\beta_1-\beta_4)\, g_{ab}\, T^2 \notag \\
&&
+ \left[\beta_4\left(1-2\beta_1\, \Lambda\right) - \beta_1\right] \, T\,
T_{ab} + \frac{1}{2}\,\beta_4 \, g_{ab}\, T_{cd}\, T^{cd} - 2\beta_4\,
T^c\,_a\,T_{cb} \notag \\
&& + \beta_1\, \nabla_a\nabla_b\,T - \beta_1\,
g_{ab}\, \Box\,T - \beta_4 \, \Box\,T_{ab} + 2\beta_4\,
\nabla^c\nabla_{(a}\,T_{b)c}
+\ldots\,, \label{fieldeq}
\end{eqnarray}
where $g_{ab}$ is the spacetime metric, $\Lambda$ the cosmological constant,
$T_{ab}$ the matter energy-momentum tensor, $T$ its trace, and we have set
Newton's constant to $1/8\pi $ (we refer the reader to \cite{Pani:2013qfa}
for more details). The
coefficients $\beta_1$ and $\beta_4$ are the model parameters, and it is
clear that when $\beta_1=\beta_4=0$ we recover the standard Einstein
equations. In the above equation one can clearly see that the
introduction of auxiliary fields indeed leads to gravitational equations
which include higher derivatives of the energy momentum tensor and its
trace.
\subsection{Cosmological equations}
\label{Cosmology}
In order to apply the above modified gravitational theory in cosmology we
consider the usual homogeneous and isotropic geometry, given by the flat
Friedmann-Robertson-Walker (FRW) metric
\begin{equation}
ds^2= - dt^2 + a^2(t)\,\delta_{ij} dx^i dx^j,
\end{equation}
where $a(t)$ is the scale factor. Additionally, concerning the matter
energy-momentum tensor, we use the standard form of a perfect fluid, namely
$T_{ab}=(\rho+p) U_{a} U_{b}+p \, g_{ab}$, where $\rho$ and $p$ are
respectively the matter energy density and pressure.
Inserting these in the field equations (\ref{fieldeq}), we obtain the
modified Friedmann equations as
\begin{eqnarray}\label{eqH}
3H^2 &=& (\rho +\Lambda)+3H \left[
\beta_1 ( \dot{\rho} - 3\dot{p}) - \beta_4 ( \dot{\rho} +
2\dot{p}) \right] - \beta_4 \ddot{\rho} \nonumber \\
&&
+ \frac{1}{4}\Big[ 3 \beta_1 (1+
2\beta_4 \Lambda)
(\rho^2 -3 p^2)+3\beta_4 (\rho^2 + p^2) - 12 \beta_1 (\beta_1 + \beta_4)
\Lambda \rho p
\nonumber \\
&&
- 6 (\beta_1 - \beta_4) \rho p
-4\beta_1 \Lambda (\rho - 3p) + 2\beta_1^2
\Lambda (\rho^2 + 9p^2)
\Big],
\end{eqnarray}
and
\begin{eqnarray}\label{eqdH}
\dot{H}\left[1+\beta
_4\left(\rho +p\right)\right]&=& - \frac{1}{2}(\rho + p)
- 3H^2 \beta_4 (\rho + p)
- \frac{H}{2} \left[ \beta_1 (\dot{\rho} - 3 \dot{p}) - \beta_4
(\dot{\rho} + 3 \dot{p}) \right]
\nonumber \\
&&
-\frac{1}{2}\Big[ (3\beta_1 + \beta_4)
\ddot{p} -(\beta_1 + \beta_4) \ddot{\rho} + \beta_4 (\rho^2 + p^2)
\nonumber \\
&& + \beta_1
(1 + 2 \beta_4 \Lambda) (\rho^2 -3
p^2) -2 (\beta_1 -\beta_4 +2
\beta_1 \beta_4 \Lambda) \rho p \Big],
\end{eqnarray}
respectively. The above equations can be re-written in the form of the usual
Friedmann
equations as
\begin{equation}
\label{Fr11}
3H^{2}=\rho +\rho _{DE},
\end{equation}%
\begin{equation}
\label{Fr22}
-2\dot{H}= \rho +\rho _{DE}+p+p_{DE} ,
\end{equation}%
where we have introduced the effective dark-energy sector with energy
density and pressure, $\rho_{DE}$ and $p_{DE}$, respectively, defined as
\begin{eqnarray}
\rho _{DE}&=&\Lambda + \frac{1}{4}\Big[ 3 \beta_1 (1+ 2\beta_4 \Lambda)
(\rho^2 -3 p^2)
+3\beta_4 (\rho^2 + p^2) - 12 \beta_1 (\beta_1 + \beta_4) \Lambda \rho p - 6
(\beta_1 - \beta_4) \rho p
\nonumber \\
&&-4\beta_1 \Lambda (\rho - 3p) + 2\beta_1^2 \Lambda (\rho^2 + 9p^2)
\Big]
+3H \left[ \beta_1 ( \dot{\rho} - 3\dot{p}) - \beta_4 ( \dot{\rho} +
2\dot{p}) \right] - \beta_4 \ddot{\rho},
\label{rhoDE}
\end{eqnarray}
and
\begin{eqnarray}
p_{DE} &=&\frac{1}{4\left[ \beta
_{4}(p+\rho )+1\right] }
\Bigg\{-\beta
_{4}\rho ^{3}\left[ 2\beta _{1}\left( \beta _{1}+3\beta _{4}\right) \Lambda
+3\left( \beta _{1}+\beta _{4}\right) \right] -\left( \beta _{1}-3\beta
_{4}\right) \rho ^{2}\left( 2\beta _{1}\Lambda -1\right)
\nonumber\\
&& +4\rho \left( \beta
_{1}\Lambda -\beta _{4}\Lambda +6\beta _{4}H^{2}\right) +4 p
\left[ 6\beta
_{4}H^{2}
- \left( 3\beta _{1}+\beta _{4}\right) \Lambda \right]
\nonumber\\
&& +
\beta_{4} p\rho^2 \left[ \beta _{1}\left( 10\beta
_{1}\Lambda +6\beta _{4}\Lambda +3\right) -9\beta _{4}\right] +
2 p\rho \left[ 6\beta
_{1}\left( \beta _{1}-\beta _{4}\right) \Lambda -\left( \beta _{1}+3\beta
_{4}\right) \right]
\nonumber\\
&& +3p^{2}\left[ \beta _{4}\rho
\left( -2\left( \beta _{1}-5\beta _{4}\right) \beta _{1}\Lambda +5\beta
_{1}-3\beta _{4}\right) -\left( \beta _{1}+\beta _{4}\right) \left( 6\beta
_{1}\Lambda +1\right) \right]
\nonumber\\
&& -4\Lambda
+3\beta _{4}p^{3}\left[ 6\left( \beta _{4}-\beta _{1}\right) \beta
_{1}\Lambda +3\beta _{1}-\beta _{4}\right]
\nonumber\\
&& + 4H\left[ \left( \beta _{1}-\beta
_{4}\right) \dot{\rho} (t)-3\left( \beta _{1}+\beta _{4}\right)
\dot{p}(t)\right] +4\left( 3\beta _{1}+\beta
_{4}\right)
\ddot{p}%
-4\left( \beta _{1}+\beta _{4}\right) \ddot{\rho}\Bigg\}.
\end{eqnarray}
Thus, we can calculate the equation-of-state parameter $w_{DE}$ of the
effective dark-energy sector as
$w_{DE}\equiv p_{DE}/\rho _{DE}$, or using the Friedmann equations
(\ref{Fr11}) and (\ref{Fr22}) as
\begin{equation}
w_{DE}=-\frac{2\dot{H} +p+3H^2}{3H^2-\rho }.
\label{wDE}
\end{equation}
Moreover, one can see that given the matter energy conservation
\begin{equation} \label{cons}
\dot{\rho}+3H\left( \rho +p\right) =0,
\end{equation}
the effective dark energy is also conserved, that is
\begin{equation} \label{cons2}
\dot{\rho}_{DE}+3H\left( \rho_{DE} +p_{DE}\right) =0.
\end{equation}
Finally, we can define the deceleration parameter $q$, namely
\begin{equation}
\label{qq}
q=-1-\frac{\dot{H}}{H^2}=\frac{1}{2}+\frac{3}{2}\frac{p+p_{DE}}{\rho +\rho
_{DE}}.
\end{equation}
The sign of $q$ indicates the decelerating/accelerating nature
of the cosmological expansion; cosmological models with
$q < 0$ are accelerating, while those having $q > 0$ experience
a decelerating evolution. Note that in terms of $q$ the dark-energy
equation-of-state parameter
reads
\begin{equation}
w_{DE}=\frac{(2q-1)H^2-p}{3H^2-\rho},
\label{wDE2}
\end{equation}
or, alternatively,
\begin{equation}\label{wdei}
w_{DE}=\frac{1}{3}(2q-1)\left(\frac{\rho }{\rho _{DE}}+1\right)-\frac{p}{\rho
_{DE}}.
\end{equation}
In order to be able to handle the cosmological equations of the scenario at
hand, namely the Friedmann equations (\ref{Fr11}) and (\ref{Fr22}) and the
conservation equations (\ref{cons}) (or (\ref{cons2})), we proceed as
follows. Firstly, from equation (\ref{cons}) we easily obtain
\begin{equation}
H=-\frac{\dot{\rho}}{3\left( \rho +p\right) }, \label{H1}
\end{equation}%
and
\begin{equation}
\dot{H}=\frac{\dot{\rho}\left( \dot{p}+\dot{\rho}\right) -(p+\rho
)\ddot{\rho%
}}{3(p+\rho )^{2}}. \label{dH1}
\end{equation}%
Then, substitution into equations
(\ref{eqH}) and (\ref{eqdH}) provides the equations
\begin{eqnarray}
\label{rho21}
\beta _4\ddot{\rho} &=&
\Lambda +\rho +\frac{\dot{\rho} \left[(3 \beta _1+2
\beta _4) \dot{p}+(\beta _4-\beta _1) \dot{\rho}\right]}{p+\rho}
-\frac{\dot{\rho}^2}{3 (p+\rho )^2}
\nonumber\\
&& +\frac{1}{4} \Bigg\{-2 \rho \left\{2
\beta _1 \Lambda +3 p [2 \beta _1
\Lambda (\beta _1+\beta _4)+\beta _1-\beta _4]\right\} +\rho ^2 \left[2
\beta
_1
\Lambda (\beta _1+3 \beta _4)+3 (\beta _1+\beta
_4)\right]
\nonumber\\
&& \ \ \ \ \ \ \ \ \ \ \ + 3p \left[4 \beta _1 \Lambda +p (6 \beta _1
\Lambda (\beta _1-\beta
_4)-3 \beta _1+\beta _4)\right]\Bigg\} ,
\end{eqnarray}
and
\begin{eqnarray}
\label{rho22}
\ddot{\rho} &=& -\left(3\beta _{1}+5\beta
_{4}+\frac{2}{p+\rho }\right)^{-1}\Bigg\{
3p\left[ 2\rho (2\beta
_{1}\beta _{4}\Lambda +\beta _{1}-\beta _{4})-1\right] +3p^{2}\left[ \beta
_{1}(6\beta _{4}\Lambda +3)-\beta _{4}\right]
\nonumber\\
&&
- \frac{3(3\beta _{1}+\beta _{4})(p+\rho
)^{2}\ddot{p}+\dot{p}\dot{\rho}\left[
(3\beta _{1}+5\beta _{4})(p+\rho )+2\right] -\dot{\rho}^{2}\left[ (\beta
_{1}-5\beta _{4})(p+\rho )-2\right] }{(p+\rho )^{2}}
\nonumber\\
&&\ \ \ \ \ \ \ \ \ \ \ -3\rho ^{2}(2\beta _{1}\beta
_{4}\Lambda +\beta _{1}+\beta _{4})-3\rho \Bigg\} ,
\end{eqnarray}
Finally, eliminating $\ddot{\rho}$ between equations (\ref{rho21}) and
(\ref{rho22}),
through equalizing the right hand sides, we obtain a differential equation
for $\rho$. Hence, if we additionally provide the matter equation of state,
assumed to be of barotropic form $p=p(\rho)$, this differential equation can
be solved to give $\rho(t)$. Lastly, $H(t)$ is then obtained from equation
(\ref{H1})
and then $\rho_{DE}(t)$ from equation (\ref{rhoDE}), $w_{DE}$ from equation
(\ref{wDE}) and
$q$ from equation (\ref{qq}).
\section{Cosmological solutions}
\label{Solutions}
In this section we explore the cosmological scenario with higher-matter
derivatives, for three particular cases of matter equations of state, namely
dust ($p=0$), radiation ($p=\rho /3$) and stiff fluid ($p=\rho$).
\subsection{Dust cosmological models}
As a first example we consider that the matter is in the form of dust,
namely with
$p=0$. In this case equations (\ref{rho21}) and (\ref{rho22}) are
respectively simplified to
\begin{equation}
\label{rho31}
\ddot{\rho }= \frac{3 \rho ^2 \left\{4 \Lambda +\rho \left\{4-4 \beta _1
\Lambda +\rho [2 \beta _1 \Lambda (\beta _1+3 \beta _4)+3
(\beta _1+\beta _4)]\right\}
\right\}-4 \dot{\rho }^2 \left[3 (\beta
_1-\beta _4) \rho +1\right]}{12 \beta _4 \rho ^2},
\end{equation}
and
\begin{eqnarray}\label{rho32}
\ddot{\rho }= \frac{3 \rho ^3 \left[\rho \left(2 \beta _1 \beta _4 \Lambda
+\beta _1+\beta _4\right)+1\right]+\dot{\rho}^2 \left[2-(\beta _1-5
\beta _4) \rho \right]}{\rho \left[(3 \beta _1+5 \beta _4) \rho
+2\right]}.
\end{eqnarray}
Hence, as we mentioned in the previous section, equality of the
right-hand-sides provides the differential
equation for $\rho(t)$. If for convenience
we introduce the dimensionless variables
\begin{equation}
\tau=\frac{t}{2\beta _{4}^{1/2}}, \qquad \theta =\beta _{4} \rho , \qquad
k=\frac{\beta
_{1}}{\beta _{4}}, \qquad h= \beta _4^{1/2} H, \qquad \lambda = \beta
_{4}\Lambda ,
\end{equation}
the differential equation takes the following form
\begin{eqnarray}
\label{dens1}
\frac{d\theta }{d\tau }&=&-\theta
\left\{\theta
\left[3k(\theta +3\theta
k+3)+5\right]+2\right\}^{-1/2}
\Bigg\{3\theta \left(\theta +3\theta k+4\right)[3\theta
(k+1)+2]
\nonumber\\
&& +6\lambda \Big\{ \theta \left\{
k\left[ \theta ^{2}(3k^2+14k+3)-4\theta (k+1)+2\right] +10\right\}
+4\Big\}
\Bigg\}^{1/2},
\end{eqnarray}%
where for physical reasons (monotonically decreasing energy density) we
have adopted the minus sign for the square root.
Additionally, calculating $H(t)$ from equation (\ref{H1}) and inserting
into equation (\ref{qq}) we extract the analytical expression for the
deceleration parameter as
\begin{eqnarray}
&&q = \Bigg\{ \left[6k\theta (\theta +3k\theta
+3)+4\right]\Big\{ \theta
\left[ \theta^{2}k^2(6k\lambda+28\lambda +9)+\theta^2 (6k\lambda
+12k+3)
\right.
\nonumber\\
&&
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
\ \ \left. +2\theta (9k+7-4\lambda k^2-\lambda k )+
4\lambda(k+5) +8\right]
+8\lambda \Big\} \Bigg\} ^{-1} \times
\nonumber \\
&&\Bigg\{
3\theta^5(3k^2+k) \left[k^2(6k\lambda+28\lambda
+9)+6k(\lambda +2)+3\right]-\theta^2
\left[8k(89k+77)\lambda-112\right]
\nonumber \\
&&\ \ \ -\theta^3 \left[ 8k^2\lambda(21k+80) +72k^2
+2k(108\lambda -87)
-42\right]
\nonumber\\
&&\ \ \ +3\theta^4
\left[8k(15\lambda k^3 + 50\lambda k^2+11\lambda k)+8k(3k+1)\right]
+8\theta \lambda(5-44k) +16\theta
-32\lambda \Bigg\}.
\label{qgen}
\end{eqnarray}%
Note that for the specific case where the cosmological
constant is absent (namely
$\lambda =0$), that is when the effective dark energy sector constitutes
solely from the higher matter derivatives, we obtain
\begin{eqnarray}
\label{qm1}
q=\frac{1}{2}+\frac{3 (9 k \theta+4)}{6 k \theta (\theta+3 k \theta+3)+4}-
\frac{6}{\theta+3 k \theta+4}-\frac{3}{3 (k+1) \theta +2}.
\end{eqnarray}
\begin{figure*}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmo1a.eps} %
\includegraphics[width=7.65cm]{cosmo1b.eps}
\caption{{\it{Time evolution of the scaled matter energy density
$\protect\theta $ of the dust-matter
Universe, as a function of
$\protect\tau$, for $\protect\lambda =0$ (left panel) and for
$%
\protect\lambda =0.004$ (right panel), for different values of $k$: $k=1$
(solid curve), $k=3$ (dotted curve), $k=5$ (short dashed curve), $k=9$
(dashed curve) and $k=16$ (long dashed curve). The initial value of the
energy density is $\protect\theta (0)=0.20$. }}}
\label{fig1}
\end{figure*}
\begin{figure*}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmo2a.eps} %
\includegraphics[width=7.65cm]{cosmo2b.eps}
\caption{{\it{Time evolution of the scale factor $a $ of the dust-matter
Universe, as a function of
$\protect\tau$,
for $\protect\lambda =0$ (left panel) and for $\protect\lambda =0.004$
(right panel), for different values of $k$: $k=1$ (solid curve), $k=3$
(dotted curve), $k=5$ (short dashed curve), $k=9$ (dashed curve) and $k=16$
(long dashed curve). The initial value of the energy density is $\protect%
\theta (0)=0.20$. }}}
\label{fig2}
\end{figure*}
\begin{figure*}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmo3a.eps} %
\includegraphics[width=7.65cm]{cosmo3b.eps}
\caption{{\it{Time evolution of the Hubble function $h $ of the dust-matter
Universe, as a function of
$\protect\tau$, for $\protect\lambda =0$ (left panel) and for
$\protect\lambda =0.004$
(right panel), for different values of $k$: $k=1$ (solid curve), $k=3$
(dotted curve), $k=5$ (short dashed curve), $k=9$ (dashed curve) and $k=16$
(long dashed curve). The initial value of the energy density is $\protect%
\theta (0)=0.20$. }}}
\label{fig3}
\end{figure*}
\begin{figure*}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmo4a.eps} %
\includegraphics[width=7.65cm]{cosmo4b.eps}
\caption{{\it{Time evolution of the deceleration parameter $q$ of the
dust-matter
Universe, as a function of
$\protect\tau$, for $\protect\lambda =0$ (left panel) and for $\protect%
\lambda =0.004$ (right panel), for different values of $k$: $k=1$ (solid
curve), $k=3$ (dotted curve), $k=5$ (short dashed curve), $k=9$ (dashed
curve) and $k=16$ (long dashed curve). The initial value of the energy
density is $\protect\theta (0)=0.20$. }}}
\label{fig4}
\end{figure*}
\begin{figure*}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmo5a.eps} %
\includegraphics[width=7.65cm]{cosmo5b.eps}
\caption{{\it{Time evolution of the dark-energy equation-of-state parameter
$w_{DE} $
of the dust-matter
Universe, as a function of
$\protect\tau$, for $\protect\lambda =0$ (left panel) and for $\protect%
\lambda =0.004$ (right panel), for different values of $k$: $k=1$ (solid
curve), $k=3$ (dotted curve), $k=5$ (short dashed curve), $k=9$ (dashed
curve) and $k=16$ (long dashed curve). The initial value of the energy
density is $\protect\theta (0)=0.20$. }}}
\label{fig4a}
\end{figure*}
In order to present the above behavior more transparently, we
numerically solve equation (\ref{dens1}), and then we calculate the various
physical quantities as we described above. In Figs.~\ref{fig1}-\ref{fig4a}
we present the time evolution of the matter energy density, of the scale
factor, of the Hubble function, of the deceleration parameter and of the
dark-energy equation-of-state parameter, for the cases $\lambda =0$ and
$\lambda =0.004$, and for different
values of $k$.
First of all, as expected, the cosmological dynamics depends on the value of
$\lambda$. In particular, since in the scenario at hand the effective dark
energy sector is attributed to the higher derivatives of the matter fields,
when an explicit cosmological constant is absent ($\lambda=0$) we expect the
dark-energy equation-of-state parameter to become asymptotically zero at
late times, and the Universe to be non-accelerating. This behavior can be
clearly seen in the left graphs of Figs. \ref{fig4} and \ref{fig4a}. On the
other hand, when $\lambda\neq0$, the explicit cosmological constant dominates
inside the effective dark energy sector, and in this case the
quintessence-like Universe at
late times transits to the accelerating phase, tending asymptotically to the
de Sitter evolution, with $w_{DE}\rightarrow-1$. This behavior can be
observed in the right graphs of Figs. \ref{fig4} and \ref{fig4a}.
The above asymptotic behaviors can be also analytically extracted from the
cosmological equations. In particular, setting $\lambda=0$ in
(\ref{qm1}) leads immediately to $\lim_{\theta
\rightarrow 0} q =1/2$. On the other hand,
for $\lambda > 0$, expression (\ref{qgen}) leads to
$\lim_{\theta
\to 0} q=-1$.
We close this subsection by mentioning that more complex behaviors of
$w_{DE}$ can also be achieved. In particular, according to (\ref{wdei}), if
during the cosmological expansion for some $\rho =\rho _{cr}$
the condition
\begin{equation}
\frac{1}{3}(2q-1)\left(\frac{\rho _{cr}}{\rho
_{DE}}+1\right)
=w_{DE}^{(cr)}=-1,
\end{equation}
is satisfied, then at the corresponding time the phantom divide crossing
will be realized.
\subsection{The radiation dominated phase}
Let us now investigate the cosmological behavior of a radiation
dominated
Universe, with the matter pressure satisfying the equation of state $
p=\rho /3$. The gravitational field equations (\ref{eqH}), (\ref{eqdH})
and (\ref{cons}) give
\begin{equation} \label{rad1}
\ddot{\rho}= \frac{60 \beta _4 \rho \dot{\rho }^2+64 \beta _4 \rho ^4+48
\Lambda \rho ^2-9 \dot{\rho }^2+48 \rho ^3}{48 \beta _4 \rho ^2},
\end{equation}
\begin{equation} \label{rad2}
\ddot{\rho }= \frac{30 \beta _4 \rho \dot{\rho }^2+32 \beta _4 \rho ^4+9
\dot{\rho }^2+24 \rho ^3}{3 \rho (8 \beta _4 \rho +3)},
\end{equation}
and
\begin{equation}
\dot{\rho }+4H\rho=0,
\end{equation}
respectively. It is interesting to note that for the radiation-dominated
Universe of this subsection the dynamics is determined solely by the
parameter $\beta_4$, with the parameter $\beta _1$ eliminated from the
equations. The consistency
condition requiring the equality of the right-hand-sides of equations
(\ref{rad1}) and (\ref{rad2}) provides the basic evolution equation
describing the
dynamical behavior, namely
\begin{equation}
\dot{\rho }=-4\rho \sqrt{ \frac{ 3 \Lambda +\left(3+8 \beta _4
\Lambda\right) \rho +4 \beta _4 \rho ^2 }{3 (4 \beta _4 \rho +3)}}.
\end{equation}
In the case $\Lambda =0$ and $\beta_4=0$, we obtain
\begin{equation}
\dot{\rho }=-\frac{4 \rho ^{3/2}}{\sqrt{3}},
\end{equation}
which yields the general solution
\begin{equation}
\rho =\frac{3\rho _0}{\left(2\sqrt{\rho _0}t+\sqrt{3}\right)^{2}},
\end{equation}
where we have used the initial condition $\rho (0)=\rho _0$. Therefore, in
this case we re-obtain the cosmological behavior of the radiation filled
Universes in standard cosmology, with
\begin{eqnarray}
H(t)&=& \frac{\sqrt{\rho _0}}{2\sqrt{\rho
_0}t+\sqrt{3}} ,
\\
a(t)&=&a_0\sqrt{2t\sqrt{\frac{\rho_0}{3}}+1} ,
\end{eqnarray}
and $q=1$, respectively.
\subsection{Stiff fluid cosmology}
One of the most common equations of state of high energy density cosmological
matter, which has been used extensively to study the properties of the
early Universe, is the linear barotropic equation of state, given by $p =
(\gamma - 1)\rho $, with $\gamma = \mathrm{constant} \in [1,2]$. A very
important subcase is its so-called causal limit, corresponding to $\gamma =
2$, which gives the Zeldovich, or stiff-fluid equation of state $p = \rho $
\cite{60}.
The Zeldovich equation of state applies to densities significantly
higher than nuclear densities, $\rho > 10\rho _{\rm nuc}$, with $\rho _{\rm
nuc} =
10^{14}$ g/cm$^3$. From a field theoretical point of view the Zeldovich
equation of state can be obtained by constructing a relativistic
Lagrangian that allows bare nucleons to interact attractively via scalar
meson exchange and repulsively via the exchange of a more massive vector
meson \cite{60}. On the other hand, in the non-relativistic limit both the
quantum and
classical field theories yield Yukawa-type potentials. At the highest
densities
the vector-meson exchange dominates, and by using a mean field approximation
one can show that in the extreme limit of infinite densities the pressure
tends to the energy density, namely $p\rightarrow \rho $ \cite{60}. In this
limit the sound
speed $c_s^2 = dp/d\rho \rightarrow 1$, and hence this equation of state
satisfies the causality condition, with the speed of sound less than the
speed of light.
For a stiff fluid the field equations (\ref{eqH}), (\ref{eqdH}) and (%
\ref{cons})
provide respectively
\begin{equation}
\ddot{\rho }= \frac{12 \beta _1 \rho \dot{\rho }^2-36 \beta _1 \rho ^4+18
\beta _4 \rho \dot{\rho }^2+36 \beta _4 \rho ^4-\dot{\rho }^2+12 \rho ^3}{12
\beta _4 \rho ^2},
\label{eq11}
\end{equation}
\begin{equation}
\ddot{\rho }= \frac{-\beta _1 \rho \dot{\rho }^2+12 \beta _1 \rho ^4-5 \beta
_4 \rho \dot{\rho }^2-12 \beta _4 \rho ^4-\dot{\rho }^2-6 \rho ^3}{\rho (6
\beta _1 \rho -2 \beta _4 \rho -1)},
\label{eq22}
\end{equation}
and
\begin{equation}
\dot{\rho }+6H\rho=0.
\label{eq33}
\end{equation}
We mention that we have set to zero the explicit cosmological constant
(namely $\Lambda =0$), since it is negligible comparing to the high
densities that are needed to justify the Zeldovich stiff equation of state.
The basic evolution equation that describes the energy density
evolution, obtained by equating the right hand sides of (\ref{eq11}) and
(\ref{eq22}), is given by
\begin{equation}
\label{finZ}
\dot{\rho }=-2 \sqrt{3} \sqrt{\frac{\rho ^3 \left[6 (\beta _1-\beta _4) (3
\beta _1+\beta _4) \rho ^2-(9 \beta _1+\beta _4) \rho +1\right]}{2 \rho
\left[-9
\beta _1-2 \beta _4+12 (\beta _1+\beta _4) (3 \beta _1+\beta _4) \rho
\right]+1}}.
\end{equation}
By introducing the set of dimensionless quantities $\left(\tau, \theta ,
k,h\right)$, defined as
\begin{equation}
\tau=\frac{2\sqrt{3}}{\beta _4^{1/2}}\;t, \qquad \theta = \beta _4 \rho,
\qquad k=
\frac{\beta _1}{\beta _4}, \qquad h=\frac{\beta _4^{1/2}}{2\sqrt{3}}H,
\end{equation}
equation (\ref{finZ}) takes the final form
\begin{equation}
\dot{\theta}=-\theta ^{3/2}\sqrt{\frac{\theta \left[6 (k-1) (3 k+1) \theta
-9 k-1\right]+1}{ 2 \theta \left[12 (k+1) (3 k+1) \theta -9 k-2\right]+1}}.
\end{equation}
Thus, using (\ref{eq33}) the Hubble function becomes
\begin{equation}
h=\frac{1}{6} \sqrt{\frac{\theta \left[\theta (6 (k-1) (3 k+1) \theta -9
k-1)+1\right]}{2 \theta \left[12 (k+1) (3 k+1) \theta -9 k-2\right]+1}},
\end{equation}
which leads to a scale factor of the form
\begin{equation}
a=a_0\theta ^{-1/6}.
\end{equation}
Additionally, using (\ref{qq}) the deceleration parameter becomes
\begin{equation}
q=2+\frac{3 \left[(9 k+1) \theta -2\right]}{\theta \left[6 (k-1) (3 k+1)
\theta -9 k-1\right]+1}
+\frac{6-6 (9 k+2) \theta}{2 \theta \left[12 (k+1) (3 k+1) \theta -9 k-2%
\right]+1}.
\end{equation}
We numerically evolve the scenario at hand for various parameter values, and
in Figs.~\ref{fig5}-\ref{fig7} we respectively depict the time evolution
of the matter energy density, of the scale
factor, of the scaled Hubble function, of the deceleration parameter and of
the dark-energy equation-of-state parameter.
\begin{figure*}[tbp]
\centering
\includegraphics[width=8cm]{cosmoZa.eps} %
\caption{{\it{Time evolution of the matter energy density $\protect\theta $
of the stiff-fluid Universe as a function of $\protect\tau $, for different
values of $k$: $k=-1
$ (solid curve), $k=-3$ (dotted curve), $k=-5$ (short dashed curve), $k=-9$
(dashed curve) and $k=-16$ (long dashed curve). The initial value of the
energy density is $\protect\theta (0)=1.20$. }}}
\label{fig5}
\end{figure*}
\begin{figure*}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmoZb.eps}
\includegraphics[width=7.65cm]{cosmoZc.eps} %
\caption{{\it{Time evolution of the scale factor $a$ (left panel) and of the
Hubble function $h$ (right panel) of the stiff-fluid
Universe as a function of $\protect\tau$, for different values of $k$:
$k=-1$
(solid curve), $k=-3$ (dotted curve), $k=-5$ (short dashed curve), $k=-9$
(dashed curve) and $k=-16$ (long dashed curve). The initial value of the
energy density is $\protect\theta (0)=1.20$. }}}
\label{fig6}
\end{figure*}
\begin{figure}[tbp]
\centering
\includegraphics[width=7.65cm]{cosmoZd.eps}
\includegraphics[width=7.65cm]{cosmoZe.eps} %
\caption{{\it{
Time evolution of the deceleration parameter $q$ (left panel) and of the
dark-energy equation of state parameter $w_{DE} $ (right panel) of the
stiff-fluid Universe as a function of $\protect\tau$, for different values of
$k$: $k=-1$ (solid curve), $k=-3$ (dotted curve), $k=-5$ (short dashed
curve), $k=-9$ (dashed curve) and $k=-16$ (long dashed curve). The initial
value of the energy density is $\protect\theta (0)=1.20$. }}}
\label{fig7}
\end{figure}
As we observe, for the considered numerical values of the model parameters
the cosmological evolution of the Universe is strongly decelerating for all
times, with the numerical values of the deceleration parameter in the range
$1.2<q<2.2$. This is an expected result, since we have neglected the presence
of the cosmological constant $\Lambda $ in the field equations. In physical
terms it is also expected, since from the cosmological point of view the
stiff-fluid regime lasts for a very short period during the evolution of
the early Universe, in which the Zeldovich density conditions are satisfied.
During this period the Universe is expanding at a slow rate, with the Hubble
function monotonically decreasing in time. Furthermore, the dark-energy
equation-of-state parameter has small negative values, being very
close to zero for all considered dimensionless times.
At this point, let us remind that in
standard General Relativity the evolution of the stiff-fluid Universe
is given by $\rho _{GR}=\rho _0/a^6$, $\dot{H}_{
GR}=-\rho_{GR}$, and $q_{GR}=2$. Hence, we deduce that the presence of the
higher derivative matter terms in the gravitational field equations
leads to a significant departure from the standard cosmological dynamics.
This feature could lead to strong nucleosynthesis constraints on the present
scenario, and used to distinguish it from other modified gravity classes.
\section{Conclusions}
\label{Conclusions}
In the present work we have investigated the cosmological implications of a
new class of modified gravity, which arises from the introduction of
non-dynamical auxiliary fields. This feature leads the equations of motion to
contain higher order derivatives of the matter fields. Although this could
place tight observational constraints on this theory, as long as these
constraints are satisfied the above class corresponds to a novel modified
gravitational theory and cosmology that is worthy to explore. In particular,
extracting the
gravitational field equations and imposing a flat, homogeneous and
isotropic geometry, we obtained the Friedmann equations, in which
the effective dark-energy sector contains higher derivatives of the matter
energy density and pressure.
One important feature of the present scenario is the conservation of the
matter energy-momentum tensor, which imposes strong constraints on the
cosmological evolution. In particular, due to this conservation the
density - scale-factor relation is the same as in standard General
Relativity,
which in the case of a barotropic cosmological matter fluid with $p=(\gamma
-1)\rho $ takes the form $\rho \propto a^{-3\gamma }$. However, since the
evolution of the scale factor is now different, the
matter-density time evolution differs from the standard General Relativistic
one.
In the case where the matter sector is dust-like ($p=0$), the Universe
approaches asymptotically the de Sitter stage
with $H\approx H_0={\rm constant}$ and with the matter content decreasing
according to $\rho \approx \rho _0\exp\left(-3H_0t\right)$,
where both the deceleration parameter $q$ and the dark-energy
equation-of-state parameter $w_{DE}$ tend to $-1$, due to the domination of
the explicit cosmological constant. This result is independent of the model
parameters as expected, since in all cases the decrease of the matter
energy density due to the expansion leads all the terms in the
matter-dependent effective dark-energy sector to disappear, apart from the
cosmological constant one. Hence, in the presence of
the cosmological constant the de Sitter stage is an attractor solution of the
field equations. On the other hand, in the absence of an explicit
cosmological constant, the Universe results in a non-accelerating,
matter-dominated Universe.
In the case of radiation matter $p=\rho /3$, and in the absence of an
explicit cosmological constant we obtained a non-accelerating Universe,
similar to the radiation-dominated phase of standard General Relativity.
Additionally, in the case of a stiff matter $p = \rho $, which is expected
to be realized in the very early Universe, we found that the Universe is
expanding at a slow rate, with the effective dark-energy
equation-of-state parameter having small negative values.
The above variety of cosmological evolutions reveals that the theory with
higher matter derivatives is indeed new, and deserves further investigation.
More specifically, it would be interesting and necessary to perform a
detailed cosmological perturbation analysis, in order to examine the
stability properties of the theory, as well as to confront it with
perturbation-related observables such as the large-scale structure and
the growth index. Moreover, one could perform a full phase space analysis,
which will reveal the global, asymptotic behavior of the scenario.
Furthermore, another important avenue of analysis
would be to use data from Type Ia Supernovae (SNIa), Baryon
Acoustic Oscillations (BAO), and Cosmic Microwave Background
(CMB) observations, in order to impose constraints on the theory. The above
investigations may also provide specific signatures and effects, which could
distinguish between this theory and other alternatives of modified
gravity. We aim to explore in detail these issues in upcoming
publications.
\begin{acknowledgments}
FSNL acknowledges financial support of the Funda\c{c}\~{a}o para a
Ci\^{e}ncia e Tecnologia through an Investigador FCT Research contract, with
reference IF/00859/2012, funded by FCT/MCTES (Portugal), and grants
CERN/FP/123618/2011 and EXPL/FIS-AST/1608/2013. The research of ENS is
implemented within the framework of the Operational Program ``Education and
Lifelong Learning'' (Actions Beneficiary: General Secretariat for Research
and Technology), and is co-financedby the European Social Fund (ESF) and the
Greek State.
\end{acknowledgments}
|
\section{Introduction}
This paper is the continuation of the study in (\cite{CW3}) and (\cite{CW4}). In~\cite{CW3}, we developed
a weak compactness theory for non-collapsed Ricci flows with bounded scalar curvature and bounded half-dimensional curvature integral. This weak compactness theory is applied in~\cite{CW4} to prove the Hamilton-Tian conjecture of
complex dimension 2 and its geometric consequences.
However, the assumption of half dimensional curvature integral is restrictive.
It is not available for high dimensional anti-canonical K\"ahler Ricci flow, i.e., K\"ahler Ricci flow on a Fano manifold $(M,J)$, in the class $2\pi c_1(M,J)$.
In this paper, by taking advantage of the extra structures from K\"ahler geometry, we drop this curvature integral condition.
The present paper is inspired by two different sources.
One source is the structure theory of K\"ahler Einstein manifolds which was developed over last 20 years by many people,
notably, Anderson, Cheeger, Colding, Tian and more recently, Naber, Donaldson and Sun. The recent progress of the structure theory of
K\"ahler Einstein manifolds supplies many additional tools for our approach.
The other source is the seminal work of Perelman on the Ricci flow(c.f.~\cite{Pe1},~\cite{SeT}).
Actually, it was pointed out by Perelman already that his idea in \cite{Pe1} can be applied to study K\"ahler Ricci flow.
He said that
``\textit{present work has also some applications to the
Hamilton-Tian conjecture concerning K\"ahler-Ricci flow on K\"ahler manifold with
positive first Chern class: these will be discussed in a separate paper}".
\noindent We cannot help to wonder how far he will push the subject of Ricci flow if he continued to publicize
his works on arxiv.
Although ``\textit{this separate paper}" never appears, his fundamental estimates of K\"ahler Ricci flow on Fano manifolds is the base of our present
research. Besides Perelman's estimates,
we also note that the following technical results in the Ricci flow are important to the formation of this paper
over a long period of time: the Sobolev constant estimate by Q.S. Zhang(\cite{Zhq1}) and R. Ye (\cite{Ye}),
and the volume ratio upper bound estimate by Q.S. Zhang(\cite{Zhq3}) and Chen-Wang(\cite{CW5}). Some other important estimates can be found in the summary of~\cite{CW2}.
Our key observation is that there is a ``canonical neighborhood" theorem for anti-canonical K\"ahler Ricci flows.
The idea of ``canonical neighborhood" originates from Theorem 12.1 of Perelman's paper~\cite{Pe1}.
For every 3-dimensional Ricci flow, Perelman showed that the space-time neighborhood of a high curvature point can be approximated by
a $\kappa$-solution, which is a model Ricci flow solution.
To be precise, a $\kappa$-solution is a 3-dimensional, $\kappa$-noncollapsed, ancient Ricci flow solution with bounded, nonnegative curvature operator.
By definition, it is not clear at all that the moduli of $\kappa$-solutions has compactness under (pointed-) smooth topology (modulo diffeomorphisms).
Perelman genuinely proved the compactness by delicate use of Hamilton-Ivey estimate and the geometry of nonnegatively curved 3-manifolds.
In light of the compactness of the moduli of $\kappa$-solutions, by a maximum principle type argument,
Perelman developed the ``canonical neighborhood" theorem,
which is of essential importance to his celebrated
solution of the Poincar\'{e} conjecture(c.f.~\cite{KL},~\cite{MT},~\cite{CZ}).
The idea of ``canonical neighborhood" is universal and can be applied in many different geometric settings.
In particular, there is a ``canonical neighborhood" theorem
for the anti-canonical K\"ahler Ricci flows, where estimates of many quantities, including scalar curvature, Ricci potential and Sobolev constant,
are available. Clearly, a ``canonical neighborhood" should be a neighborhood in space-time, behaving like a model space-time,
which is more or less the blowup limit of the given flow.
Therefore, it is natural to expect that the model space-time is the scalar flat Ricci flow solutions,
which must be Ricci flat, due to the
equation $ \frac{\partial}{\partial t} R= \Delta R + 2|Ric|^2$, satisfied by the scalar curvature $R$.
For this reason, the model space and model space-time can be identified, since the evolution on time direction is trivial.
It is also natural to expect that the model space has some K\"ahler structure. In other words, the model space should
be K\"ahler Ricci flat space, or Calabi-Yau space.
Now the first essential difficulty appears. A good model space should have a compact moduli.
For example, in the case of 3-dimensional Ricci flow, the moduli space of $\kappa$-solutions, which are the model space-times, has compactness in the smooth topology.
However, the moduli space of all the non-collapsed smooth Calabi-Yau space-times
is clearly not compact under the smooth topology.
A blowdown sequence of Eguchi-Hanson metrics is an easy example.
For the sake of compactness,
we need to replace the smooth topology by a weaker topology,
the pointed-Cheeger-Gromov topology.
At the same time,
we also need to enlarge the class of model spaces from complete Calabi-Yau manifolds to the Calabi-Yau spaces with mild singularities(c.f.~Definition~\ref{dfn:SC27_1}),
which we denote by $\widetilde{\mathscr{KS}}(n,\kappa)$.
Similar to the compactness theorem of Perelman's $\kappa$-solutions,
we have the compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{theoremin}[\textbf{Compactness of model moduli}]
$\widetilde{\mathscr{KS}}(n,\kappa)$ is compact under the pointed Cheeger-Gromov topology.
Moreover, each space $X \in \widetilde{\mathscr{KS}}(n,\kappa)$ is a Calabi-Yau conifold.
\label{thmin:HE21_1}
\end{theoremin}
The notion of conifold is well known to string theorist as
some special Calabi-Yau 3-folds with singularities(c.f.~\cite{Green}).
In this paper, by abusing notation,
we use it to denote a space whose singular part admits cone type tangent spaces.
The precise definition is given in Definition~\ref{dfn:HD20_1}.
Note that Calabi-Yau conifold is a generalization of Calabi-Yau orbifold.
The strategy to prove the compactness of
$\widetilde{\mathscr{KS}}(n,\kappa)$ follows the same route of the weak compactness theory
of K\"ahler Einstein manifolds, developed by Cheeger, Gromoll, Anderson, Colding, Tian, Naber, etc.
However, the analysis foundation on the singular spaces need to be carefully checked, which is done in section 2.
Theorem~\ref{thmin:HE21_1} is motivated by section 11 of Perelman's seminal paper~\cite{Pe1},
where Perelman proved the compactness of moduli space of $\kappa$-solutions and showed that
$\kappa$-solutions have many properties which are not obvious from definition.
By trivial extension, each
$X \in \widetilde{\mathscr{KS}}(n,\kappa)$ can be understood as a space-time $X \times (-\infty, \infty)$
satisfying Ricci flow equation.
Intuitively, the rescaled space-time structure in a given anti-canonical K\"ahler Ricci flow
should behave similar to that of $X \times (-\infty, \infty)$ for some $X \in \widetilde{\mathscr{KS}}(n,\kappa)$,
when the rescaling factor is large enough.
In order to make sense that two space-times are close to each other,
we need the Cheeger-Gromov topology for space-times, a slight generalization of the Cheeger-Gromov
topology for metric spaces.
When restricted on each time slice, this topology is the same as the usual Cheeger-Gromov topology.
Between every two different time slices, there is a natural homeomorphism map connecting them.
Therefore, the above intuition can be realized if we can show a blowup sequence of Ricci flow
space-times from a given K\"ahler Ricci flow converges to a limit space-time $X \times (-\infty, \infty)$,
in the pointed Cheeger-Gromov topology for space-times.
However, it is not easy to obtain the homeomorphism maps between different time slices in the limit.
Although it is quite obvious to guess that the homeomorphism maps among different time slices
are the limit of identity maps,
there exists serious technical difficulty to show the existence and regularity of the limit maps.
The difficulty boils down to a fundamental improvement of Perelman's pseudolocality theorem(Theorem 10.1 of~\cite{Pe1}).
Recall that Perelman's pseudolocality theorem says that Ricci flow cannot ``quickly" turn an almost Euclidean region into a very curved one.
It is a short-time, one-sided estimate in nature. We need to improve it to a long-time, two-sided estimate.
Not surprisingly, the rigidity of K\"ahler geometry plays an essential role for such an improvement.
The two-sided, long-time pseudolocality is an estimate in the time direction.
Modulo this time direction estimate and the weak compactness in the space direction, we can take limit for
a sequence of Ricci flows blown up from a given flow.
Then the canonical neighborhood theorem can be set up if we can show that the limit space-time locates in
$\widetilde{\mathscr{KS}}(n,\kappa)$,
following the same route as that in the proof of Theorem 12.1 of~\cite{Pe1}. \\
From the above discussion, it is clear that the strategy to prove the canonical neighborhood theorem is simple.
However, the technical difficulty hidden behind this simple strategy is not that simple.
We observe that the anti-canonical
K\"ahler Ricci flow has many additional structures, all of them should be used to carry out the proof of
the canonical neighborhood theorem.
In particular, over every anti-canonical K\"ahler Ricci flow, there is a natural anti-canonical polarization, which should play an important role, as done in~\cite{CW4}.
Although we are aiming at the anti-canonical case,
in this paper, however, we shall consider flows with more general polarizations.
We call $\mathcal{LM}=\left\{ (M^n, g(t), J, L, h(t)), t \ \in (-T,T) \subset \ensuremath{\mathbb{R}} \right\}$ a polarized K\"ahler Ricci flow if
\begin{itemize}
\item $\mathcal{M}=\left\{ (M^n, g(t), J), t \ \in (-T,T) \right\}$ is a K\"ahler Ricci flow solution.
\item $L$ is a Hermitian line bundle over $M$, $h(t)$ is a family of smooth metrics on $L$ whose curvature is $\omega(t)$, the metric form compatible with
$g(t)$ and the complex structure $J$.
\end{itemize}
Clearly, the first Chern class of $L$ is $[\omega(t)]$, which does not depend on time. So a polarized K\"ahler Ricci flow stays in a fixed integer K\"ahler class.
The evolution equation of $g(t)$ can be written as
\begin{align}
\frac{\partial}{\partial t} g_{i\bar{j}}=-R_{i\bar{j}}+ \lambda g_{i\bar{j}}, \label{eqn:K07_1}
\end{align}
where $\lambda=\frac{c_1(M)}{c_1(L)}$.
Since the flow stays in the fixed class, we can let $\omega_t=\omega_0 + \sqrt{-1} \partial \bar{\partial} \varphi$. Then $\dot{\varphi}$
is the Ricci potential, i.e.,
\begin{align*}
\sqrt{-1} \partial \bar{\partial} \dot{\varphi} = -Ric + \lambda g.
\end{align*}
Note the choice of $\varphi$ is unique up to adding a constant. So we can always modify the choice of $\varphi$ such that $\displaystyle \sup_M \dot{\varphi}=0$.
For simplicity, we denote $\mathscr{K}(n,A)$ as the collection of all the polarized K\"ahler Ricci flows
$\mathcal{LM}$ satisfying the following estimate
\begin{align}
\begin{cases}
&T \geq 2, \\
&C_S(M)+\frac{1}{\Vol(M)}+|\dot{\varphi}|_{C^0(M)} + |R-n\lambda|_{C^0(M)} \leq A,
\quad \textrm{for every time} \quad t \in (-T,T).
\end{cases}
\label{eqn:SK20_1}
\end{align}
Here $C_S$ means the Sobolev constant, $A$ is a uniform constant. In this paper, we study the structure of
polarized K\"ahler Ricci flows locating in the space $\mathscr{K}(n,A)$.
The motivation behind (\ref{eqn:SK20_1}) arises from the fundamental estimate of diameter, scalar curvature,
$C^1$-norm of Ricci potential, and Sobolev constant
along the anti-canonical K\"ahler Ricci flows(c.f.~\cite{SeT},~\cite{Zhq1},~\cite{Ye}).
Every polarized K\"ahler Ricci flow solution in $\mathscr{K}(n,A)$ has at least three structures: the metric space structure, the flow structure, the line bundle structure.
Same structures can be discussed on the model space-time in $\widetilde{\mathscr{KS}}(n,\kappa)$.
All the structures of a flow in $\mathscr{K}(n,A)$ can be modeled by the corresponding structures
in $\widetilde{\mathscr{KS}}(n,\kappa)$, which is the same meaning as the ``canonical neighborhood theorem".
We shall compare these structures term by term. \\
Under the (pointed-)Cheeger-Gromov topology at time $0$, let us compare the metric structure of a flow in
$\mathscr{K}(n,A)$ with a Calabi-Yau conifold in $\widetilde{\mathscr{KS}}(n,\kappa)$.
We shall show that $\mathscr{K}(n,A)$ and $\widetilde{\mathscr{KS}}(n,\kappa)$ behaves almost the same in this
perspective. Intuitively, one can think that the weak compactness theory of Ricci-flat manifolds and Einstein manifolds are almost the same.
For simplicity of notation, we use $\stackrel{G.H.}{\longrightarrow}$ to denote the convergence in Gromov-Hausdorff topology. We use $\stackrel{\hat{C}^k}{\longrightarrow}$ to denote
the $C^k$-Cheeger-Gromov topology, i.e., the convergence is in the Gromov-Hausdorff topology,
and can be improved to be in $C^k$-topology (modulo diffeomorphisms) away from singularities.
We call a point being regular if it has a neighborhood with smooth manifold structure and call a point being singular if it is not regular(c.f. Proposition~\ref{prn:HA08_1} and Remark~\ref{rmk:HA07_1}).
\begin{theoremin}[\textbf{Metric space estimates}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$. By taking subsequence if necessary, we have
\begin{align*}
(M_i, x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M}, \bar{x}, \bar{g}).
\end{align*}
The limit space $\bar{M}$ has a classical regular-singular decomposition $\mathcal{R} \cup \mathcal{S}$ with the following properties.
\begin{itemize}
\item $\left(\mathcal{R}, \bar{g}\right)$ is a smooth, open Riemannian manifold. Moreover,
$\mathcal{R}$ admits a limit K\"ahler structure $\bar{J}$ such that $\left(\mathcal{R}, \bar{g}, \bar{J} \right)$ is an open K\"ahler manifold.
\item $\mathcal{S}$ is a closed set and $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$, where $\mathcal{M}$ means Minkowski dimension(c.f.~Definition~\ref{dfn:HE08_1}).
\item Every tangent space of $\bar{M}$ is an irreducible metric cone.
\item Let $\mathrm{v}$ be the volume density, i.e.,
\begin{align}
\mathrm{v}(y)=\limsup_{r \to 0} \omega_{2n}^{-1}r^{-2n}|B(y,r)|
\end{align}
for every point $y \in \bar{M}$.
Then a point is regular if and only if $\mathrm{v}(y)=1$, a point is singular if and only if
$\mathrm{v}(y) \leq 1-2\delta_0$, where $\delta_0$ is a dimensional
constant determined by Anderson's gap theorem.
\end{itemize}
\label{thmin:SC24_1}
\end{theoremin}
It is important to note the difference between $\widetilde{\mathscr{KS}}(n,\kappa)$ and $\mathscr{K}(n,A)$.
We use $\widetilde{\mathscr{KS}}(n,\kappa)$ to denote the space of possible bubbles, or blowup limits.
Therefore, every metric space in it is a non-compact one. However, each time slice of flows in $\mathscr{K}(n,A)$ is a compact manifold.
The limit space $\bar{M}$ of Theorem~\ref{thmin:SC24_1} maybe compact and does not belong to $\widetilde{\mathscr{KS}}(n,\kappa)$.
In the study of the line bundle structure of $\mathscr{K}(n,A)$, the Bergman function plays an important role.
Actually, for every positive integer $k$ large enough such that $L^k$ is globally generated, we define the Bergman function $\mathbf{b}^{(k)}$ as follows
\begin{align}
\mathbf{b}^{(k)}(x,t)=\log \sum_{i=0}^{N_k} \norm{S_i^{(k)}}{h(t)}^2(x,t),
\label{eqn:HA03_3}
\end{align}
where $N_k=\dim_{\ensuremath{\mathbb{C}}} H^0(M, L^k)-1$, $\left\{S_i^{(k)} \right\}_{i=0}^{N_k}$ are orthonormal basis of $H^0(M, L^k)$ under the natural metrics $\omega(t)$ and $h(t)$.
Theorem~\ref{thmin:SC24_1} means that the metric structure of the center time slice of a K\"ahler Ricci flow in $\mathscr{K}(n,A)$ can be modeled by
non-collapsed Calabi-Yau manifolds with mild singularities. In particular, each tangent space of a point in the limit space is a metric cone.
The trivial line bundle structure on metric cone then implies an estimate of line bundle structure of the original manifold, due to
delicate use of H\"omander's $\bar{\partial}$-estimate, as done by Donaldson and Sun(c.f.~\cite{DS}).
\begin{theoremin}[\textbf{Line bundle estimates}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, then
\begin{align*}
\inf_{x \in M} \mathbf{b}^{(k_0)}(x,0) \geq -c_0
\end{align*}
for some positive number $c_0=c_0(n,A)$, and positive integer $k_0=k_0(n,A)$.
\label{thmin:HC08_1}
\end{theoremin}
In other words, Theorem~\ref{thmin:HC08_1} states that there is a uniform partial-$C^0$-estimate at time $t=0$. This estimate then implies variety structure of
limit space, as discussed in~\cite{Tian12} and~\cite{DS}.
Theorem~\ref{thmin:HC08_1} can be understood that the line bundle structure of $\mathscr{K}(n,A)$ is modeled after
that of $\widetilde{\mathscr{KS}}(n,\kappa)$.
Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1} deal only with one time slice.
In order to make sense of limit K\"ahler Ricci flow, we have to compare the limit spaces of different time slices.
For example, we choose $x_i \in M_i$, then we have
\begin{align*}
(M_i, x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M}, \bar{x}, \bar{g}), \quad (M_i, x_i, g_i(-1)) \longright{\hat{C}^{\infty}} (\bar{M}', \bar{x}', \bar{g}').
\end{align*}
How are $\bar{M}$ and $\bar{M}'$ related? If $\bar{x}$ is a regular point of $\bar{M}$, can we say $\bar{x}'$ is a regular point of $\bar{M}'$?
Note that Perelman's pseudolocality theorem cannot answer this question, due to its short-time, one-sided property. In order to relate different time slices,
we need to improve Perelman's pseudolocality theorem to the following long-time, two-sided estimate,
which is the technical core of the current paper.
\begin{theoremin}[\textbf{Time direction estimates}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Suppose $x_0 \in M$, $\Omega=B_{g(0)}(x_0,r)$, $\Omega'=B_{g(0)}(x_0, \frac{r}{2})$ for some $r \in (0,1)$.
At time $t=0$, suppose the isoperimetric constant estimate $\mathbf{I}(\Omega) \geq (1-\delta_0) \mathbf{I}(\ensuremath{\mathbb{C}}^n)$ holds for $\delta_0=\delta_0(n)$,
the same constant in Theorem~\ref{thmin:SC24_1}. Then we have
\begin{align*}
|\nabla^k Rm|(x,t) \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0}, \quad x \in \Omega',
\quad t \in [-1,1],
\end{align*}
where $C_k$ is a constant depending on $n,A,r$ and $k$.
\label{thmin:HC06_1}
\end{theoremin}
Theorem~\ref{thmin:HC06_1} holds trivially on each space in $\widetilde{\mathscr{KS}}(n,\kappa)$, when regarded as
a static Ricci flow solution. Therefore, it can be understood as the time direction structure, or the flow structure of
$\mathcal{LM} \in \mathscr{K}(n,A)$ is similar to that of $\widetilde{\mathscr{KS}}(n,\kappa)$.
Theorem~\ref{thmin:HC06_1} removes the major stumbling block for defining a limit K\"ahler Ricci flow,
since it guarantees that the regular-singular decomposition of the limit space is independent of time.
Therefore, there is a natural induced K\"ahler Ricci flow structure on the regular part of the limit space. We denote its completion
by a limit K\"ahler Ricci flow solution, in a weak sense. Clearly, the limit K\"ahler Ricci flow naturally inherits a limit line
bundle structure, or a limit polarization, on the regular part.
Moreover, the limit underlying space does have a variety structure due to Theorem~\ref{thmin:HC08_1}.
With these structures in hand, we are ready to discuss the convergence theorem of polarized K\"ahler Ricci flows,
which is the main structure theorem of this paper(c.f. section~\ref{sec:sflow} for meaning of the notations).
\begin{theoremin}[\textbf{Weak compactness of polarized flows}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$ satisfying $\diam_{g_i(0)}(M_i)<C$ uniformly or $\sup_{\mathcal{M}_i} |R| \to 0$.
By passing to subsequence if necessary, we have
\begin{align*}
\left(\mathcal{LM}_i, x_i \right) \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left(\overline{\mathcal{LM}}, \bar{x}\right),
\end{align*}
where $\overline{\mathcal{LM}}$ is a polarized K\"ahler Ricci flow solution on an analytic normal variety $\bar{M}$,
whose singular set $\mathcal{S}$ has Minkowski codimension at least $4$, with respect to each $\bar{g}(t)$.
Moreover, if $\bar{M}$ is compact, then it is a projective normal variety with at most log-terminal singularities.
\label{thmin:HC06_2}
\end{theoremin}
As it is developed for, our structure theory has applications in the study of anti-canonical K\"ahler Ricci flows.
Due to the fundamental estimate of Perelman and the monotonicity of his $\mu$-functional along each anti-canonical
K\"ahler Ricci flow, we can apply Theorem~\ref{thmin:HC06_2} directly and obtain the following theorem.
\begin{theoremin}[\textbf{Hamilton-Tian conjecture}]
Suppose $\left\{ (M^n, g(t)), 0 \leq t <\infty \right\}$ is an anti-canonical K\"ahler Ricci flow solution
on a Fano manifold $(M,J)$. For every $s>1$, define
\begin{align*}
&g_s(t) \triangleq g(t+s), \\
&\mathcal{M}_{s} \triangleq \{(M^n, g_s(t)), -s \leq t \leq s\}.
\end{align*}
Then for every sequence $s_i \to \infty$, by taking subsequence if necessary, we have
\begin{align}
\left(\mathcal{M}_{s_i}, g_{s_i}\right) \longright{\hat{C}^{\infty}} \left( \bar{\mathcal{M}}, \bar{g} \right), \label{eqn:SC13_111}
\end{align}
where the limit space-time $\bar{\mathcal{M}}$ is a K\"ahler Ricci soliton flow solution on a $Q$-Fano normal variety $(\bar{M},\bar{J})$.
Moreover, with respect to each $\bar{g}(t)$, there is a uniform $C$ independent of time such that
the $r$-neighborhood of the singular set $\mathcal{S}$ has measure not greater than $Cr^4$.
\label{thmin:SC24_3}
\end{theoremin}
Theorem~\ref{thmin:SC24_3} confirms the famous Hamilton-Tian conjecture, with more information
than that was conjectured(c.f.~Conjecture 9.1. of~\cite{Tian97} for the precise statement).
The two dimensional case was confirmed by the authors in~\cite{CW3}.
We note that in a recent paper \cite{TZZ2},
another approach to attack Hamilton-Tian conjecture in complex dimension $3$,
based on $L^4$-bound of Ricci curvature, was presented by Z.L. Zhang and G. Tian.
Their work in turn depends on the comparison geometry with integral Ricci bounded developed
by G.F. Wei and P. Petersen(\cite{PW1}).
For other important progress in K\"ahler Ricci flow, we refer interested readers to the following papers(far away from being complete): \cite{Sesum}, \cite{Zhq1}, \cite{Ye}, \cite{TZ2}, \cite{Zhq2}, \cite{SongWeinkove}, \cite{TZ2}, \cite{SongTian}, \cite{PhongSturm}, \cite{Tosa}, \cite{SzeKrf}, as well as references listed therein.\\
As corollaries of Theorem~\ref{thmin:SC24_3}, we can affirmatively answer some problems raised in~\cite{CW4}.
\begin{corollaryin}
Every anti-canonical K\"ahler Ricci flow is tamed, i.e., partial-$C^0$-estimate holds along the flow.
\label{clyin:SC24_4}
\end{corollaryin}
\begin{corollaryin}
Suppose $\{(M^n, g(t)), 0 \leq t < \infty \}$ is an anti-canonical K\"ahler Ricci flow on a Fano manifold $M$.
Then the flow converges to a K\"ahler Einstein metric
if one of the following conditions hold for every large positive integer $\nu$.
\begin{itemize}
\item $\alpha_{\nu,1}>\frac{n}{n+1}$.
\item $\alpha_{\nu,2}>\frac{n}{n+1}$ and $\alpha_{\nu,1} > \frac{1}{2- \frac{n-1}{(n+1) \alpha_{\nu,2}}}$.
\end{itemize}
\label{clyin:SC24_5}
\end{corollaryin}
Corollary~\ref{clyin:SC24_5} give rise to a method for searching Fano K\"ahler Einstein metrics in high dimension,
which generalize the 2-dimensional case due to Tian(c.f.~\cite{Tian90}). The quantities $\alpha_{\nu,k}$ are some algebro-geometric invarariant.
The interested readers are referred to~\cite{Tian90} for the precise definition.
Our structure theory can be applied to study a family of K\"ahler Ricci flows with some uniform initial conditions.
In this perspective, we have the following theorem.
\begin{theoremin}[\textbf{Partial-$C^0$-conjecture of Tian}]
For every positive constants $R_0,V_0$, there exists a positive integer $k_0$
and a positive constant $c_0$ with the following properties.
Suppose $(M,\omega, J)$ is a K\"ahler manifold satisfying
$Ric \geq R_0$ and $\Vol(M) \geq V_0$, $[\omega]=2\pi c_1(M,J)$.
Then we have
\begin{align*}
\inf_{x \in M} \mathbf{b}^{(k_0)}(x) > -c_0.
\end{align*}
\label{thmin:HA01_1}
\end{theoremin}
Theorem~\ref{thmin:HA01_1} confirms the partial-$C^0$-conjecture of Tian(c.f.~\cite{Tian90Kyo},\cite{Tian12}).
The low dimension case ($n \leq 3$) was proved by Jiang(\cite{Jiang}), depending on the partial-$C^0$-estimate along the flow,
developed by Chen-Wang(\cite{CW3},\cite{CW4}) in complex dimension 2
and Tian-Zhang(\cite{TZZ2}) in complex dimension 3.
In fact, a more general version of Theorem~\ref{thmin:HA01_1} is proved(c.f. Theorem~\ref{thm:HA01_3}).
As a corollary of Theorem~\ref{thmin:HA01_1}, we have
\begin{corollaryin}(c.f.~\cite{Sze})
The partial-$C^0$-estimate holds along the classical continuity path.
\label{clyin:HA01_2}
\end{corollaryin}
Following Corollary~\ref{clyin:SC24_4}, we obtain the following result, which was originally proved by
G. Sz\'{e}kelyhidi(c.f.~\cite{Sze}) along the classical continuity path.
\begin{corollaryin}
Suppose $(M,J)$ is a Fano manifold with $Aut(M,J)$ discrete. If it is stable in the sense of S.Paul(c.f.~\cite{Paul12}), then it admits a K\"ahler Einstein metric.
\label{clyin:HA03_1}
\end{corollaryin}
An important application of our structure theory is devoted to the study of the relationships among different stabilities.
By the work of Chen, Donaldson and Sun(c.f.~\cite{CDS0},~\cite{CDS1},~\cite{CDS2} and~\cite{CDS3}), a long standing stability conjecture, going back to Yau(c.f. Problem 65 of~\cite{Yau93}) and critically contributed by Tian(c.f~\cite{Tian97}) and Donaldson(c.f.~\cite{Do02}), was confirmed.
We now know a Fano manifold is K-stable if and only if it admits K\"ahler Einstein metrics. A posteriori, we see that
the K-stability is equivalent to Paul's stability if the underlying manifold has discrete automorphism group.
It is an interesting problem to prove this equivalence a priori,
which will be discussed in a separate paper(c.f.~\cite{CSW}). \\
Let us quickly go over the relationships among the theorems.
Theorem~\ref{thmin:HE21_1} is the structure theorem of the model space $\widetilde{\mathscr{KS}}(n,\kappa)$.
Theorem~\ref{thmin:SC24_1}, Theorem~\ref{thmin:HC08_1} and Theorem~\ref{thmin:HC06_1}
combined together give the canonical neighborhood structure of the polarized K\"ahler Ricci flow
in $\mathscr{K}(n,A)$, in a strong sense.
The main structure theorem in this paper is Theorem~\ref{thmin:HC06_2}, the weak compactness theorem of polarized K\"ahler Ricci flows.
It is clear that Theorem~\ref{thmin:SC24_3} and Theorem~\ref{thmin:HA01_1} are
direct applications of Theorem~\ref{thmin:HC06_2}. The proof of Theorem~\ref{thmin:HC06_2} is based on the combination of
Theorem~\ref{thmin:SC24_1}, Theorem~\ref{thmin:HC08_1} and Theorem~\ref{thmin:HC06_1}. These three theorems deal with different structures
of $\mathscr{K}(n,A)$, including the Ricci flow structure, metric space structure, line bundle structure and variety structure.
The importance of these structures decreases in order, for the purpose of developing compactness.
However, all these structures are intertwined together.
Paradoxically, the proof of the compactness of these structures does not follow the same order, due to the lack of precise estimate of Bergman functions.
Instead of proving them in order, we define a concept called ``polarized canonical radius", which guarantees the convergence of all these structures
under this radius. The only thing we need to do then is to show that this radius cannot be too small. Otherwise, we can apply a maximum principle argument
to obtain a contradiction, which essentially arise from the monotonicity of Perelman's reduced volume and localized $W$-functional.\\
This paper is organized as follows. In section 2, we discuss the model space $\widetilde{\mathscr{KS}}(n,\kappa)$, which consists of non-collapsed Calabi-Yau spaces with mild singularities. By checking analysis foundation and repeating the weak compactness theory of K\"ahler Einstein manifolds, we prove the compactness of
$\widetilde{\mathscr{KS}}(n,\kappa)$ and show that every space in it is a conifold.
In other words, we prove Theorem~\ref{thmin:HE21_1} at the end of section 2.
We also develop some a priori estimates, which will be essentially used in the following sections.
In section 3, we define the ``canonical radius" and discuss the convergence of metric structures when canonical radius is uniformly bounded from below.
In section 4, we first set up a forward, long-time pseudolocality theorem based on the existence of
partial-$C^0$-estimate. Motivated by this pseudolocality theorem, we then refine the ``canonical radius" to ``polarized canonical radius" and discuss the convergence of
flow structure and line bundle structure under the assumption that polarized canonical radius is uniformly bounded from below. Finally, at the end of section 4, we use a maximum principle argument to
show that there is an a priori bound of the polarized canonical radius.
In section 5, we prove Theorem~\ref{thmin:SC24_1}-\ref{thmin:HC06_2}, together with some other
more detailed properties of the space $\mathscr{K}(n,A)$.
Up to this section, everything is developed for general polarized K\"ahler Ricci flow.
At last, in section 6, we focus on the anti-canonical K\"ahler Ricci flows.
Applying the general structure theory, we prove Theorem~\ref{thmin:SC24_3} and Theorem~\ref{thmin:HA01_1}. \\
\noindent {\bf Acknowledgment}
Both authors are very grateful to professor Simon Donaldson for his constant support.
The second author would like to thank Song Sun and Shaosai Huang for helpful discussions and suggestions.
Thanks also go to Weiyong He, Haozhao Li,Yuanqi Wang, Guoqiang Wu, Chengjian Yao, Hao Yin, Kai Zheng
for their valuable comments.
\section{Model space---Calabi-Yau Space with mild singularities}
The Model space of a polarized K\"ahler Ricci flow in $\mathscr{K}(n,A)$ consists of the space-time blowup limits from
flows in $\mathscr{K}(n,A)$. In this section, we shall discuss the properties of the model space, from the perspective of
metric space structure and the intrinsic Ricci flow structure.
\subsection{Singular Calabi-Yau space $\widetilde{\mathscr{KS}}(n,\kappa)$}
Let $\mathscr{KS}(n)$ be the collection of all the complete $n$-dimensional Calabi-Yau (K\"ahler Ricci flat) manifolds.
By Bishop-Gromov comparison, it is clear that the asymptotic volume ratio is well defined for every manifolds in the moduli space $\mathscr{KS}(n)$.
The gap theorem of Anderson (c.f. Gap Lemma 3.1 of~\cite{An90}) implies that the asymptotic volume ratio is strictly less than $1-2\delta_0$ whenever the underlying manifold is not the flat $\ensuremath{\mathbb{C}}^n$, where
$\delta_0$ is a dimensional constant. We fix this constant and call it as Anderson constant in this paper.
Let $\mathscr{KS}(n,\kappa)$ be a subspace of $\mathscr{KS}(n)$, with every element has asymptotic volume ratio at least $\kappa$.
Clearly, $\mathscr{KS}(n,\kappa)$ is not compact under the pointed-Gromov-Hausdorff topology.
It can be compactified as a space $\overline{\mathscr{KS}}(n,\kappa)$.
However, this may not be the largest space that one can develop weak-compactness theory. So we extend the space $\overline{\mathscr{KS}}(n,\kappa)$ further to a possibly bigger compact space $\widetilde{\mathscr{KS}}(n,\kappa)$, which is defined as follows.
\begin{definition}
Let $\widetilde{\mathscr{KS}}(n,\kappa)$ be the collection of length spaces $(X,g)$ with the following properties.
\begin{enumerate}
\item $X$ has a disjoint regular-singular decomposition $X=\mathcal{R} \cup \mathcal{S}$, where $\mathcal{R}$ is the regular part, $\mathcal{S}$ is the singular part.
A point is called regular if it has a neighborhood which is isometric to a totally geodesic convex domain of some smooth Riemannian manifold. A point is called singular
if it is not regular.
\item The regular part $\mathcal{R}$ is a nonempty, open Ricci-flat manifold of real dimension $m=2n$.
Moreover, there exists a complex structure $J$ on $\mathcal{R}$ such that $(\mathcal{R}, g, J)$ is a K\"ahler manifold.
\item $\mathcal{R}$ is weakly convex, i.e., for every point $x \in \mathcal{R}$, there exists a measure ($2n$-dimensional Hausdorff measure) zero set
$\mathcal{C}_x \supset \mathcal{S}$ such that every point in $X \backslash \mathcal{C}_x$ can be connected to $x$
by a unique shortest geodesic in $\mathcal{R}$. For convenience, we call $\mathcal{C}_x$ as the cut locus of $x$.
\item $\dim_{\mathcal{M}} \mathcal{S} < 2n-3$, where $\mathcal{M}$ means Minkowski dimension.
\item Let $\mathrm{v}$ be the volume density function,i.e.,
\begin{align}
\mathrm{v}(x) \triangleq \lim_{r \to 0} \frac{|B(x,r)|}{\omega_{2n} r^{2n}} \label{eqn:SC16_1}
\end{align}
for every $x \in X$. Then $\mathrm{v}\equiv 1$ on $\mathcal{R}$ and $\mathrm{v} \leq 1-2\delta_0$ on $\mathcal{S}$.
In other words, the function $\mathrm{v}$ is a criterion function for singularity.
Here $\delta_0$ is the Anderson constant.
\item The asymptotic volume ratio $\mathrm{avr}(X) \geq \kappa$. In other words, we have
\begin{align*}
\lim_{r \to \infty} \frac{|B(x,r)|}{\omega_{2n}r^{2n}} \geq \kappa
\end{align*}
for every $x \in X$.
\end{enumerate}
Let $\widetilde{\mathscr{KS}}(n)$ be the collection of metric spaces $(X,g)$ with all the above properties except the last one. Since Euclidean space is a special element,
we define
\begin{align*}
\widetilde{\mathscr{KS}}^{*}(n) \triangleq \widetilde{\mathscr{KS}}(n) \backslash \{(\ensuremath{\mathbb{C}}^n, g_{\ensuremath{\mathbb{E}}})\},
\quad \widetilde{\mathscr{KS}}^{*}(n,\kappa) \triangleq \widetilde{\mathscr{KS}}(n,\kappa) \backslash \{(\ensuremath{\mathbb{C}}^n, g_{\ensuremath{\mathbb{E}}})\}.
\end{align*}
\label{dfn:SC27_1}
\end{definition}
Note that the $\kappa$ in $\widetilde{\mathscr{KS}}(n)$ means the asymptotic area ratio is at least $\kappa$. If we drop $\kappa$, the space $\widetilde{\mathscr{KS}}(n)$ may contain
compact spaces. The default measure is always the $2n$-dimensional Hausdorff measure, unless we mention otherwise.
We use $\dim_{\mathcal{H}}$ to denote Hausdorff dimension, $\dim_{\mathcal{M}}$ to denote Minkowski dimension, or the
box-counting dimension. Since Minkowski dimension is not as often used as Hausdorff dimension, let us recall the definition of it quickly(c.f.~\cite{Falco}).
\begin{definition}
Suppose $E$ is a bounded subset of $X$.
$E_r$ is the $r$-neighborhood of $E$ in $X$. Then
the upper Minkowski dimension of $E$ is defined as the limit:
$\displaystyle \dim_{\mathcal{H}} X-\liminf_{r \to 0^+} \frac{\log |E_r|}{\log r}$.
We say $\dim_{\mathcal{M}} E \leq \dim_{\mathcal{H}} X-k$
if the upper Minkowski dimension of $E$ is not greater than $2n-k$.
Namely, we have
\begin{align*}
\liminf_{r \to 0^+} \frac{\log |E_r|}{\log r} \geq k.
\end{align*}
If $E$ is not a bounded set, we say $\dim_{\mathcal{M}} E \leq \dim_{\mathcal{H}} X-k$ if
$\dim_{\mathcal{M}} E \cap B \leq \dim_{\mathcal{H}} X-k$ for each unit geodesic ball $B \subset X$
satisfying $B\cap E \neq \emptyset$.
\label{dfn:HE08_1}
\end{definition}
In general, it is known that Hausdorff dimension is not greater than Minkowski dimension. Hence, we always have
$\dim_{\mathcal{H}}\mathcal{S} \leq \dim_{\mathcal{M}} \mathcal{S}$.
In our discussion, $X$ clearly has Hausdorff dimension $2n$. Therefore, $\dim_{\mathcal{M}} \mathcal{S}<2n-3$
implies that for each nonempty intersection $B(x_0,1) \cap \mathcal{S}$, its $r$-neighborhood has measure
$o(r^{3})$ for sufficiently small $r$. By virtue of the high codimension of $\mathcal{S}$ and the Ricci-flatness of $\mathcal{R}$,
in many aspects, each metric space $X \in \widetilde{\mathscr{KS}}(n,\kappa)$ can be treated as an intrinsic Ricci-flat space. We shall see that the geometry of $X$ is almost the same as that of Calabi-Yau manifold.
\begin{proposition}[\textbf{Bishop-Gromov volume comparison}]
Suppose $x_0 \in X$, $0<r_a<r_b<\infty$, and $\delta>0$. Then we have
\begin{align}
&\frac{|B(x_0,r_a)|}{r_a^{2n}} \geq \frac{|B(x_0,r_b)|}{r_b^{2n}}, \label{eqn:GD08_2}\\
&\frac{|B(x_0,r_a+\delta)|-|B(x_0,r_a)|}{(r_a+\delta)^{2n}-r_a^{2n}} \geq \frac{|B(x_0,r_b+\delta)|-|B(x_0,r_b)|}{(r_b+\delta)^{2n}-r_b^{2n}}. \label{eqn:GD08_3}
\end{align}
\label{prn:HD19_1}
\end{proposition}
\begin{proof}
We first prove (\ref{eqn:GD08_2}) for the case $x_0 \in \mathcal{R}$.
Away from the cut locus $\mathcal{C}_{x_0}$, which is measure-zero, every point can be connected to $x_0$ by a unique smooth geodesic.
Therefore, every point $y \in X \backslash \mathcal{C}_{x_0}$ can be identified with a point $(\gamma'(0), L) \in \ensuremath{\mathbb{R}}^{2n}$, where $\gamma$ is
the shortest geodesic connecting $x_0$ and $y$, with $\gamma(0)=x_0$, $L$ is the length of $\gamma$. In this way, we constructed a
polar coordinate system around $x_0$. Since $|B(x_0,r)|=|B(x_0,r) \backslash \mathcal{C}_{x_0}|$, by calculating the volume element evolution
along each $\gamma$ in polar coordinate, we obtain the volume comparison same as the Riemannian case.
This is more or less standard. For example, one can check the details from~\cite{ZhuSH}, or the survey~\cite{Wei}.
Now we show (\ref{eqn:GD08_2}) for $x_0 \in \mathcal{S}$. Let $x_i \in \mathcal{R}$ and $x_i \to x_0$. Fix $r>0$. Note that
\begin{align}
\lim_{i \to \infty} |B(x_i,r)| = |B(x,r)|. \label{eqn:GD01_0}
\end{align}
Actually, for each $\epsilon>0$ and large $i$, we have $B(x_i,r-\epsilon) \subset B(x,r) \subset B(x_i, r+\epsilon)$ and hence
\begin{align}
&|B(x_i,r-\epsilon)|-|B(x,r)| \leq |B(x_i,r)| -|B(x,r)|\leq |B(x_i,r+\epsilon)|-|B(x,r)|, \notag \\
& | |B(x_i,r)| -|B(x,r)| | \leq |B(x_i,r+\epsilon)-B(x_i,r-\epsilon)|. \label{eqn:GD01_1}
\end{align}
Note that $x_i$ is a regular point for each $i>1$, by standard Bishop-Gromov comparison, we have
\begin{align*}
|B(x_i,r+\epsilon)-B(x_i,r-\epsilon)| \leq 2n\omega_{2n} \{(r+\epsilon)^{2n}-(r-\epsilon)^{2n}\} \leq C(n,r) \epsilon.
\end{align*}
Therefore, taking limit of (\ref{eqn:GD01_1}) as $i \to \infty$ and then let $\epsilon \to 0$, we obtain (\ref{eqn:GD01_0}).
Consequently, we have
\begin{align}
\lim_{i \to \infty} \omega_{2n}^{-1}r_a^{-2n}|B(x_i,r_a)|= \omega_{2n}^{-1}r_a^{-2n}|B(x_0,r_a)|, \quad
\lim_{i \to \infty} \omega_{2n}^{-1}r_b^{-2n}|B(x_i,r_b)|=\omega_{2n}^{-1}r_b^{-2n}|B(x_0,r_b)|. \label{eqn:GD01_2}
\end{align}
Again, $x_i$ is a regular point for each $i>1$, so (\ref{eqn:GD08_2}) was proved for $x_i$ and can be written as
\begin{align*}
\omega_{2n}^{-1}r_a^{-2n}|B(x_i,r_a)| \geq \omega_{2n}^{-1}r_b^{-2n}|B(x_i,r_b)|.
\end{align*}
Plugging the above inequality into (\ref{eqn:GD01_2}), we obtain (\ref{eqn:GD08_2}) for the singular point $x_0$.
The proof of (\ref{eqn:GD08_3}) is similar. We first prove (\ref{eqn:GD08_3}) for regular point $x_0$ and then use approximation to prove it for
singular $x_0$. For regular $x_0$, in polar coordinates, (\ref{eqn:GD08_3}) can be proved the same as the smooth Riemannian manifold
case(c.f. Theorem 3.1 of~\cite{ZhuSH}). In the approximation step, it is important to have volume continuity of annulus. However, this can be proved similar
to (\ref{eqn:GD01_0}), by using triangle inequalities.
\end{proof}
\begin{corollary}[\textbf{Volume doubling}]
$X$ is a volume doubling metric space. More precisely, for every $x_0 \in X$ and $r>0$, we have
\begin{align*}
\frac{|B(x_0,2r)|}{|B(x_0,r)|} \leq \kappa^{-1}.
\end{align*}
\label{cly:HD20_1}
\end{corollary}
\begin{corollary}[\textbf{``Area ratio" monotonicity}]
For each $x_0 \in X$, there is a function $A(r)$, the ``area ratio", defined almost everywhere on $(0,\infty)$ such that
\begin{align}
&|B(x_0,r)|=\int_0^r A(s) s^{2n-1} ds, \quad \forall \; r>0. \label{eqn:GD08_4}\\
& \frac{|B(x_0,r_b)|}{r_b^{2n}}-\frac{|B(x_0,r_a)|}{r_a^{2n}}=\int_{r_a}^{r_b} \frac{2n}{r} \left( \frac{A(r)}{2n} - \frac{|B(x_0,r)|}{r^{2n}}\right) dr, \quad \forall \; 0<r_a<r_b. \label{eqn:GD08_5}
\end{align}
Furthermore, $A$ is non-increasing on its domain. In other words, we have $ A(r_a) \geq A(r_b)$ whenever $A(r_a)$, $A(r_b)$ are well defined and $0<r_a<r_b$.
\label{cly:GD08_1}
\end{corollary}
\begin{proof}
From the approximation process in the proof of Proposition~\ref{prn:HD19_1}, we see that even for $x_0 \in \mathcal{S}$, the inequalities
\begin{align*}
0 \leq \frac{d}{dr} |B(x_0, r)| \leq 2n \omega_{2n} r^{2n-1}, \quad
-\frac{2n}{r} \leq \frac{d}{dr} \left\{\frac{|B(x_0, r)|}{\omega_{2n} r^{2n}} \right\} \leq 0,
\end{align*}
hold in the barrier sense. In particular, $ |B(x_0, r)|$ and $\omega_{2n}r^{-2n}|B(x_0,r)|$ are monotone, uniformly Lipschitz functions of $r$ on each compact sub-interval of $(0,\infty)$.
Therefore, they have bounded derivatives almost everywhere. By abuse of notation, we denote the derivatives of $|B(x_0,r)|$ by $|\partial B(x_0, r)|$.
Let $A(r)$ be $r^{1-2n}|\partial B(x_0, r)|$. Clearly, $A(r)$ is defined almost everywhere on $(0,\infty)$.
Intuitively, $A(r)$ is the area ratio of geodesic sphere. By absolute continuity of $|B(x_0,r)|$ and $r^{-2n}|B(x_0,r)|$, (\ref{eqn:GD08_4}) and (\ref{eqn:GD08_5}) are nothing but the Newton-Leibniz formula.
We now show the monotonicity of $A$. Actually, suppose $A(r_a)$ and $A(r_b)$ are well defined. Then we have
\begin{align*}
&A(r_a)=\lim_{\epsilon \to 0^{+}} \frac{|B(x_0,r_a+\epsilon)|-|B(x_0,r_a)|}{r_a^{2n-1}\epsilon}=\lim_{\epsilon \to 0^{+}} \frac{2n \left\{|B(x_0,r_a+\epsilon)|-|B(x_0,r_a)| \right\}}{(r_a+\epsilon)^{2n}-r_a^{2n}},\\
&A(r_b)=\lim_{\epsilon \to 0^{+}} \frac{|B(x_0,r_b+\epsilon)|-|B(x_0,r_b)|}{r_b^{2n-1}\epsilon}=\lim_{\epsilon \to 0^{+}} \frac{2n \left\{|B(x_0,r_b+\epsilon)|-|B(x_0,r_b)| \right\}}{(r_b+\epsilon)^{2n}-r_b^{2n}}.
\end{align*}
Following from (\ref{eqn:GD08_3}) and the above identities, we obtain $A(r_a) \geq A(r_b)$ by taking limits.
\end{proof}
\begin{proposition}[\textbf{Segment inequality}]
For every nonnegative function $f \in L_{loc}^1(X)$, define
\begin{align*}
\mathcal{F}_f(x_1,x_2) \triangleq \inf_{\gamma} \int_0^{l} f(\gamma(s))ds,
\end{align*}
where the infimum is taken over all minimal geodesics $\gamma$, from $x_1$ to $x_2$ and $s$ denotes the arc length.
Suppose $p \in X$, $r>0$, $A_1$, $A_2$ are two subsets of $B(p,r)$.
Then we have
\begin{align}
\int_{A_1 \times A_2} \mathcal{F}_{f}(x_1,x_2) \leq 4^n r (|A_1|+|A_2|) \int_{B(p, 3r)} f.
\label{eqn:HD17_3}
\end{align}
\label{prn:HD17_1}
\end{proposition}
\begin{proof}
Fix a smooth point $x_1$, then away from cut locus, every point can be connected to $x_1$ by a unique geodesic.
Since $X \times X$ is equipped with the product measure, it is clear that away from a measure-zero set, every point
$(x_1, x_2) \in X \times X$ has the property that $x_1$ and $x_2$ are smooth and can be joined by
a unique smooth shortest geodesic.
Then the proof of (\ref{eqn:HD17_3}) is reduced to the same status as the Riemannian manifold case.
The interested readers can find the details in the work of Cheeger and Colding in~\cite{CCWarp}.
\end{proof}
Due to the work of Cheeger and Colding (c.f. Remark 2.82 of~\cite{CCWarp}),
the segment inequality implies the $(1,2)$-Poincar\'{e} inequality in general.
In our particular case, the Poincar\'{e} constant can be understood more precisely.
\begin{proposition}[\textbf{Bound of Poincar\'{e} constant}]
Suppose $f \in L_{loc}^1(X)$, $h$ is an upper gradient of $f$ in the sense of Cheeger(c.f.~Definition~\ref{dfn:HD18_1}).
Then for every geodesic ball $B(p,r) \subset X$ and real number $q \geq 1$, we have
\begin{align}
\fint_{B(p,r)} |f-\underline{f}| \leq 2 \cdot 6^{2n} \cdot r \left(\fint_{B(p,3r)} h^q \right)^{\frac{1}{q}},
\label{eqn:HD18_1}
\end{align}
where $\fint$ means the average, $\underline{f}$ is the average of $f$ on $B(p,r)$.
In particular, there is a uniform $(1,2)$-Poincar\`e constant on $X$.
\label{prn:HC29_2}
\end{proposition}
\begin{proof}
This is standard. For example, one can check~\cite{CCWarp} and references therein for the details.
\end{proof}
\begin{proposition}[\textbf{Bound of Sobolev constant}]
There is a uniform isoperimetric constant on $X$.
Consequently, a uniform $L^2$-Sobolev inequality hold on $X$.
\label{prn:HC29_3}
\end{proposition}
\begin{proof}
Due to the uniform non-collapsing condition and the weak convexity and Ricci-flatness of $\mathcal{R}$, the argument of Croke (c.f.~\cite{Croke}) applies.
So there is a uniform isoperimetric constant on $X$.
The $L^2$-Sobolev constant follows from the isoperimetric constant(c.f.~\cite{SchYau}).
\end{proof}
Note that for each $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, we lose smooth structure around $\mathcal{S}$.
In orbifold case, one can recover the smooth structure at a local ``covering" space. For our $X$, it is not known whether one has such a property. However, the good news is that the smooth structure does not play an essential role in many aspects. In the next subsection, we shall see that the analysis on $X$ is almost the same as that on manifold.
\subsection{Sobolev space, Dirichlet form and heat semigroup}
On a metric measure space, one can define
Sobolev space $H_{1,2}(X)$ following Cheeger(\cite{Cheeger99}), or $N^{1,2}(X)$
following Shanmugalingam(\cite{ShanNew}).
However, these two definitions coincide whenever volume doubling property and uniform $(1,2)$-Poincar\'{e} inequality holds,
in light of Theorem 4.10 of~\cite{ShanNew}, or the discussion on page 440 of~\cite{Cheeger99}.
In particular, for the space $(X,g,d\mu)$ which we are interested in, we have $N^{1,2}(X)=H_{1,2}(X)$ as Banach spaces.
For simplicity, we shall only use the notation $N^{1,2}(X)$ and follow the route of Cheeger.
\begin{definition}
Suppose $\Omega \subset X$. Let $f: \Omega \to [0,\infty]$ be an extended function.
An extended real function $h: \Omega \to [0,\infty]$ is called an upper gradient of $f$ on $\Omega$
if for every two points
$z_1,z_2 \in \Omega$ and all continuous rectifiable curves $c: [0,l] \to \Omega$, parameterized by arc length $s$, with $z_1,z_2$ end points, we have
\begin{align*}
|f(z_1)-f(z_2)| \leq \int_0^l h(c(s))ds.
\end{align*}
\label{dfn:HD18_1}
\end{definition}
\begin{definition}
The Sobolev space $N^{1,2}(X)$ is the subspace of $L^2(X)$ consisting of functions $f$ for which the norm
\begin{align}
\norm{f}{N^{1,2}}^2=\norm{f}{L^2}^2 + \inf_{f_i} \liminf_{i \to \infty}\norm{h_i}{L^2}^2 < \infty,
\label{eqn:HD13_1}
\end{align}
where the limit infimum is taken over all upper gradients $h_i$ of the functions $f_i$, which satisfies $\norm{f_i -f}{L^2(X)} \to 0$.
\end{definition}
Note that the above $N^{1,2}$-norm is equivalent to Cheeger's definition(c.f. equation (2.1) of ~\cite{Cheeger99}).
With this norm, we know $N^{1,2}(X)$
is complete(c.f. Theorem 2.7 of~\cite{Cheeger99}).
Clearly, it follows directly from the definition that zero function $f \in L^2(X)$ is the zero function in $N^{1,2}(X)$.
It is not surprising that $N^{1,2}(X)$ is the classical Sobolev space whenever $X$ is a smooth manifold.
This can be easily proved following the same argument of Theorem 4.5 of~\cite{ShanNew}, where the same conclusion was proved whenever
$X$ is a domain of Euclidean space. In particular, as Banach spaces, we have
\begin{align}
N^{1,2}(\mathcal{R}) \cong W^{1,2}(\mathcal{R}),
\end{align}
where $W^{1,2}(\mathcal{R})$ is the classical Sobolev space on the smooth manifold $\mathcal{R}$.
\begin{proposition}[\textbf{Smooth approximation}]
Suppose $\Omega$ is an open set of $X$, $f \in N^{1,2}(\Omega)$.
Then there is a sequence of $f_i \in C^{\infty}(\Omega \backslash \mathcal{S}) \cap N^{1,2}(\Omega)$,
$\supp f_i \subset \Omega \backslash \mathcal{S}$ such that
\begin{align}
\lim_{i \to \infty} \norm{f_i-f}{N^{1,2}(\Omega)}=0.
\label{eqn:HD15_1}
\end{align}
Moreover, if $f$ is also nonnegative, we can choose the approximation $f_i$ nonnegative.
If $\Omega$ is bounded, then $\supp f_i$ is a compact subset of $\overline{\Omega} \backslash \mathcal{S}$.
\label{prn:HD04_1}
\end{proposition}
\begin{proof}
It suffices to show the proof for the case when both $\diam(\Omega)$ and $\norm{f}{L^{\infty}}$ are bounded.
For otherwise, we can apply the bounded result for the truncated function $\min\{k, \max\{-k, f\}\}$ on $\Omega \cap B(x_0,k)$ for each $k$ and then use a
standard diagonal sequence argument, to reduce to the bounded case.
Since $\mathcal{S}$ has measure zero, $\Omega \backslash \mathcal{S}$ is a smooth manifold, we have
\begin{align*}
\norm{f_i-f}{N^{1,2}(\Omega)}=\norm{f_i-f}{N^{1,2}(\Omega \backslash \mathcal{S})}
=\norm{f_i-f}{W^{1,2}(\Omega \backslash \mathcal{S})}.
\end{align*}
Therefore, (\ref{eqn:HD15_1}) is equivalent to
\begin{align}
\lim_{i \to \infty} \norm{f_i-f}{W^{1,2}(\Omega \backslash \mathcal{S})}=0.
\label{eqn:HD15_2}
\end{align}
However, this sequence of $f_i$ can be constructed following a standard method,
as indicated by the proof of Theorem 2 of section 5.3.2 of Evans' book~\cite{Evans}.
For the convenience of the readers, we include a detailed construction of $f_i$ here.
For each positive integer $i$, define
\begin{align*}
\Omega_i \triangleq \{y \in \Omega| d(y,\mathcal{S})>2^{-i}\}, \quad
V_i \triangleq \Omega_{i+3} \backslash \overline{\Omega}_{i+1}, \quad
W_i \triangleq \Omega_{i+4} \backslash \overline{\Omega}_{i}.
\end{align*}
Also, choose open sets $V_0$ and $W_0$ such that
\begin{align*}
\overline{\Omega}_4 \cap \Omega \supset V_0 \supset \overline{\Omega}_2 \cap \Omega, \quad W_0 \supset \overline{\Omega}_6 \cap \Omega \supset V_0.
\end{align*}
Then we have
\begin{align*}
\Omega \backslash \mathcal{S}= \bigcup_{i=0}^{\infty} V_i=\bigcup_{i=0}^{\infty} W_i, \quad \overline{V}_i \cap \Omega_i \subset W_i, \quad \forall \; i \geq 0.
\end{align*}
Clearly, by composing with $d(\cdot, \mathcal{S})$, we can choose Lipschitz cutoff functions $\zeta_i$ such that $\zeta_i=1$ on $V_i$
and $\supp \zeta_i \subset W_i$, $|\nabla \zeta_i|<2^{i+5}$.
Set $ \eta_i \triangleq \frac{\zeta_i}{\sum_j \zeta_j}$.
Clearly, $\eta_i$ is a kind of partition of unity subordinate to the covering $\bigcup_{i} W_i$. In other words, we have
\begin{align*}
\begin{cases}
0\leq \eta_i \leq 1, & \eta_i \in C_c^{1}(W_i), \quad \forall \; i \geq 1, \\
\sum_{i} \eta_i =1, & \textrm{on} \; \Omega \backslash \mathcal{S}.
\end{cases}
\end{align*}
Note that $\eta_0$ is special. It is only in $C^{1}(W_0)$ in general. However, it vanishes around $\partial W_0 \cap \Omega$.
For each $i \geq 0$, note that
$V_i \cap V_j =\emptyset$ if $|i-j| \geq 2$, $W_i \cap W_j= \emptyset$ if $|i-j| \geq 4$. Therefore, we have
\begin{align*}
0\leq \eta_i<1, \quad |\nabla \eta_i|<2^{i+10}.
\end{align*}
For each $i \geq 1$, we see that $\eta_i f \in W_0^{1,2}(W_i)$. Note that $W_i \subset \mathcal{R}$.
Applying convolution with smooth mollifiers(c.f. Theorem 1 of section 5.3.1 of~\cite{Evans}), we can choose a smooth function
$h_i \in C_c^{\infty}(W_i)$ such that
\begin{align*}
\norm{h_i-\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}^2=\norm{h_i-\eta_i f}{W^{1,2}(W_i)}^2 < 9^{-i-1}\epsilon^2.
\end{align*}
For $i=0$, we can choose $h_0 \in C^{\infty}(W_0)$ which vanishes around $\partial W_0 \cap \Omega$ such that the above inequality hold.
For each large $k$, we define $ H_k \triangleq \sum_{i=0}^k h_i$. Then $H_k \in C^{\infty}(\cup_{i=0}^k W_i)
\subset C^{\infty}(\Omega \backslash \mathcal{S})$.
Moreover, we have estimate
\begin{align}
\norm{H_k-f}{W^{1,2}(\Omega \backslash \mathcal{S})}
&=\norm{\sum_{i=0}^k h_i -\sum_{i=1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}
=\norm{\sum_{i=0}^k (h_i-\eta_i f) -\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})} \notag\\
&\leq \sum_{i=0}^k \norm{h_i-\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}
+ \norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}.
\label{eqn:HD03_2}
\end{align}
However, the first term on the right hand side of the above inequality can be bounded as follows.
\begin{align}
\sum_{i=0}^k \norm{h_i-\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}
<\sum_{i=0}^{k} 3^{-i-1} \epsilon<\frac{1}{2}\epsilon.
\label{eqn:HD03_3}
\end{align}
On the other hand, note that $\sum_{i=k+1}^{\infty} \eta_i =1$ on $\bigcup_{i=k+5}^{\infty}W_i$, and it is supported on
$\bigcup_{i=k+1}^{\infty}W_i$. Thus, we have
\begin{align}
\norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}^2
&\leq \norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\bigcup_{i=k+5}^{\infty}W_i)}^2
+\norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2 \notag\\
&=\norm{f}{W^{1,2}(\bigcup_{i=k+5}^{\infty}W_i)}^2
+\norm{\sum_{i=k+1}^{k+8}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2.
\label{eqn:HD03_1}
\end{align}
For simplicity of notation, define $\chi_k \triangleq \sum_{i=k+1}^{k+8} \eta_i$. Clearly, $0\leq \chi_k \leq 1$.
We have
\begin{align*}
\norm{\sum_{i=k+1}^{k+8}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
&=\norm{\chi_k f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
=\int_{\bigcup_{i=k+1}^{k+4}W_i} \chi_k^2 f^2 + \left| \left \langle \chi_k \nabla f + f \nabla \chi_k \right \rangle \right|^2\\
&\leq \int_{\bigcup_{i=k+1}^{k+4}W_i} f^2 + 2\chi_k^2 |\nabla f|^2 + 2f^2|\nabla \chi_k|^2\\
&\leq \left( 2 \int_{\bigcup_{i=k+1}^{k+4}W_i} f^2+|\nabla f|^2 \right)
+2\norm{f}{L^{\infty}(\Omega)}^2 \int_{\bigcup_{i=k+1}^{k+4}W_i} |\nabla \chi_k|^2.
\end{align*}
It is easy to see that $|\nabla \chi_k|<2^{k+20}$ by estimate of $\eta_k$.
By virtue of Minkowski codimension assumption, we obtain
\begin{align*}
\left|\bigcup_{i=k+1}^{k+4}W_i \right|< \left|\bigcup_{i=k+1}^{\infty}W_i \right|< C2^{-3k}
<C \left(2^{-k+5} \right)^3,
\end{align*}
which in turn implies that
\begin{align*}
\norm{\sum_{i=k+1}^{k+8}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
\leq 2\norm{f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
+ C\norm{f}{L^{\infty}(\Omega)}^2 2^{-k}.
\end{align*}
Plug the above inequality into (\ref{eqn:HD03_1}), we obtain
\begin{align*}
\norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega)}^2
\leq 2 \norm{f}{W^{1,2}(\bigcup_{i=k+1}^{\infty}W_i)}^2 + C\norm{f}{L^{\infty}(\Omega)}^2 2^{-k}.
\end{align*}
Together with (\ref{eqn:HD03_2}) and (\ref{eqn:HD03_3}), the above inequality implies that
\begin{align*}
\norm{H_k-f}{W^{1,2}(\Omega)} <\frac{1}{2}\epsilon
+2 \norm{f}{W^{1,2}(\bigcup_{i=k+1}^{\infty}W_i)}^2 + C\norm{f}{L^{\infty}(\Omega)}^2 2^{-k}.
\end{align*}
Recall that $f \in W^{1,2}(\Omega \backslash \mathcal{S})$, $|\bigcup_{i=k+1}^{\infty}W_i| \to 0$ as $k \to \infty$. So we can choose $k$ large
enough such that
\begin{align*}
\norm{H_k-f}{W^{1,2}(\Omega \backslash \mathcal{S})} <\epsilon.
\end{align*}
Let $\epsilon=\frac{1}{i}$, we denote the corresponding $H_k$ in the above inequality by $f_i$.
Clearly, $f_i$ is supported on $\Omega \backslash \mathcal{S}$ and is smooth.
Moreover, (\ref{eqn:HD15_2}), consequently (\ref{eqn:HD15_1}), follows from the above inequality.
It follows from the construction that $f_i \geq 0$ whenever $ f \geq 0$. Also, from the construction,
if $\Omega$ is bounded, $\supp f_i$ is a compact subset of $\overline{\Omega} \backslash \mathcal{S}$.
\end{proof}
\begin{corollary}[\textbf{Smooth functions with compact supports}]
$C_c^{\infty}(\mathcal{R}) \cap N^{1,2}(X)$ is dense in $N^{1,2}(X)$.
\label{cly:HE10_1}
\end{corollary}
\begin{proof}
Fix $f \in N^{1,2}(X)$, without loss of generality, we may assume that $f \in C^{\infty}(\mathcal{R})$ and $f$ vanishes around $\mathcal{S}$,
by Proposition~\ref{prn:HD04_1}.
Fix $x_0 \in \mathcal{R}$ and let $r(x)=d(x,x_0)$.
For each large $k$, let $\phi_k=\phi(r(x)-k)$, where $\phi$ is a smooth cutoff function on real axis such that $\phi \equiv 1$ on $(-\infty, 0)$ and
$\phi \equiv 0$ on $(1,\infty)$. Moreover, $|\phi'| \leq 2$.
Note that $\supp f \cap \overline{B(x_0, k+1)}$ is a compact subset of $B(x_0,k+2) \backslash \mathcal{S}$. By convolution with mollifier if necessary,
we can assume $\phi_k$ is smooth and on $\supp f \cap \supp \phi_k$, $\supp \phi_k \subset B(x_0,k+2)$, $\phi_k \equiv 1$ on $B(x_0,k-1)$.
Moreover, $|\nabla \phi_k|<4$ and $0 \leq \phi_k<2$.
Therefore, $\phi_k f \in C_c^{\infty}(\mathcal{R})$. It is easy to calculate
\begin{align*}
\norm{f-\phi_k f}{N^{1,2}(X)}^2&=\int_X (1-\phi_k)^2f^2 d\mu + \int_X \left| \nabla \{(1-\phi_k) f\} \right|^2 d\mu\\
&\leq \int_{X \backslash B(x_0,k-1)} (1-\phi_k)^2 f^2 du + 2\int_{X \backslash B(x_0, k-1)} \left\{(1-\phi_k)^2 |\nabla f|^2 + f^2|\nabla \phi_k|^2 \right\} d\mu\\
&\leq \int_{X \backslash B(x_0,k-1)} f^2 du + 2\int_{X \backslash B(x_0, k-1)} \left\{|\nabla f|^2 +16 f^2 \right\} d\mu\\
&\leq 33 \int_{X \backslash B(x_0, k-1)} \left\{|\nabla f|^2 + f^2 \right\} d\mu.
\end{align*}
Clearly, the right hand side of the above inequality goes to $0$ as $k \to \infty$, since $f \in N^{1,2}(X)$. Therefore, every $f \in N^{1,2}(X)$ can
be approximated by smooth functions with compact supports.
\end{proof}
In light of Proposition~\ref{prn:HD04_1}, we can define $N_0^{1,2}(\Omega)$ as the completion of all the functions in
$C_c^{\infty}(\Omega \backslash \mathcal{S}) \cap N^{1,2}(X)$, under the $N^{1,2}(\Omega)$-norm.
Note that a function $f$ in $N_0^{1,2}(\Omega)$ may not have compact support, with respect to $\Omega$.
However, $f|_{\partial \Omega}=0$, in the sense of traces.
\begin{proposition}[\textbf{Global continuous approximation}]
For each $f \in C_c(X)$, i.e., a continuous function with compact support, there exists a sequence of $f_i \in C_c(X) \cap N^{1,2}(X)$ such that
$\displaystyle \lim_{i \to \infty} \norm{f_i-f}{C(X)} \to 0$.
\label{prn:HD24_1}
\end{proposition}
\begin{proof}
For each $\epsilon>0$, $x \in X$, define $\phi_{\epsilon,x}$ to be the character equation of the geodesic ball
$B(x,\epsilon)$.
In other words, $\phi_{\epsilon, x} \equiv 1$ on $B(x,\epsilon)$ and $0$ on $X \backslash B(x,\epsilon)$. Define
$\psi_{\epsilon,x}$ to be $\frac{\phi_{\epsilon,x}}{|B(x,\epsilon)|}$. Clearly, we have
\begin{align}
\int_X \psi_{\epsilon,x}(y)d\mu_y=1.
\label{eqn:HD26_1}
\end{align}
Similar to Euclidean case, we define approximation functions as convolution of $f$ and $\psi_{\epsilon,\cdot}$ as follows:
\begin{align*}
f_{\epsilon}(x) \triangleq (\psi_{\epsilon}* f)(x)= \int_X f(y) \psi_{\epsilon, x}(y) d\mu_y.
\end{align*}
Fix $\epsilon>0$. Suppose $x_1,x_2$ are two points in $X$ with distance $\rho \in (0,\epsilon)$. Then we calculate
\begin{align*}
|f_{\epsilon}(x_1)-f_{\epsilon}(x_2)| &\leq \int_X |f|(y) \left| \psi_{\epsilon,x_1}(y)-\psi_{\epsilon,x_2}(y)\right| d\mu_y\\
&\leq \norm{f}{C(X)} \int_X \left| \frac{\phi_{\epsilon,x_1}}{|B(x_1,\epsilon)|} - \frac{\phi_{\epsilon,x_2}}{|B(x_2,\epsilon)|}\right| d\mu_y\\
&=\frac{\norm{f}{C(X)} }{|B(x_1,\epsilon)||B(x_2,\epsilon)|} \int_X \left\vert\phi_{\epsilon,x_1} |B(x_2,\epsilon)|
-\phi_{\epsilon,x_2} |B(x_1,\epsilon)| \right\vert d\mu_y\\
&\leq C(n,\kappa) \norm{f}{C(X)} \epsilon^{-4n}
\int_X \left\vert\phi_{\epsilon,x_1} |B(x_2,\epsilon)|
-\phi_{\epsilon,x_2} |B(x_1,\epsilon)| \right\vert d\mu_y.
\end{align*}
Notice that
\begin{align*}
&\quad \int_X \left\vert\phi_{\epsilon,x_1} |B(x_2,\epsilon)|
-\phi_{\epsilon,x_2} |B(x_1,\epsilon)| \right\vert d\mu_y\\
&=\int_X \left \vert \phi_{\epsilon,x_1}\left\{ |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\}
+|B(x_1,\epsilon)| \cdot (\phi_{\epsilon,x_1}-\phi_{\epsilon,x_2}) \right\vert d\mu_y\\
&\leq \int_X\phi_{\epsilon,x_1} \left \vert |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\vert d\mu_y
+|B(x_1,\epsilon)| \int_X |\phi_{\epsilon,x_1} -\phi_{\epsilon,x_2}| d\mu_y \\
&=|B(x_1,\epsilon)| \left\{ \left \vert |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\vert
+ \int_X |\phi_{\epsilon,x_1} -\phi_{\epsilon,x_2}| \right\}\\
&=|B(x_1,\epsilon)| \left\{ \left \vert |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\vert
+|B(x_1,\epsilon) \backslash B(x_2,\epsilon)|+|B(x_2,\epsilon) \backslash B(x_1,\epsilon)| \right\}\\
&\leq 2 |B(x_1,\epsilon)| \left\{
|B(x_1,\epsilon) \backslash B(x_2,\epsilon)|+|B(x_2,\epsilon) \backslash B(x_1,\epsilon)| \right\}.
\end{align*}
By Bishop-Gromov volume comparison and non-collapsing condition, we have
\begin{align*}
&|B(x_2,\epsilon) \backslash B(x_1,\epsilon)|
\leq \left| B(x_1,\epsilon+\rho) \backslash B(x_1,\epsilon-\rho)\right| \leq C(n,\kappa) \epsilon^{2n-1} \rho, \\
&|B(x_1,\epsilon) \backslash B(x_2,\epsilon)|
\leq \left| B(x_2,\epsilon+\rho) \backslash B(x_2,\epsilon-\rho)\right| \leq C(n,\kappa) \epsilon^{2n-1} \rho.
\end{align*}
Thus, for each $\rho \in (0,\epsilon)$, we have estimate
\begin{align*}
|f_{\epsilon}(x_1)-f_{\epsilon}(x_2)|
\leq \frac{C(n,\kappa) \norm{f}{C(X)}}{\epsilon} \rho,
\end{align*}
which means that the Lipschitz constant of $f_{\epsilon}$ is uniformly bounded, for each fixed $\epsilon$.
In particular, $f_{\epsilon}$ locates in $C_c(X) \cap N^{1,2}(X)$.
It follows from (\ref{eqn:HD26_1}) that
\begin{align*}
\left| f_{\epsilon}(x) -f(x) \right|&=\left| \int_X \{f(y)-f(x)\} \psi_{\epsilon, x}(y) d\mu_y \right|
\leq \int_{B(x,\epsilon)} |f(y)-f(x)| \psi_{\epsilon,x}(y) d\mu_y
\leq \sup_{y \in B(x,\epsilon)} |f(y)-f(x)|.
\end{align*}
Note that $f$ is uniformly continuous since $f$ is continuous and $\supp(f)$ is contained in a compact subset of $X$.
Hence the right hand side of the above inequality converges to zero uniformly as $\epsilon \to 0$.
Therefore, $\psi_{2^{-i}} *f$ is a sequence of functions in $C_c(X) \cap N^{1,2}(X)$ and converges to $f$ in $C(X)$-norm.
\end{proof}
For each open set $\Omega \subset X$,
there is a restriction map $\pi: N^{1,2}(\Omega) \to N^{1,2}(\Omega \backslash \mathcal{S})$ in the obvious way.
Note that $\Omega \backslash \mathcal{S}=\Omega \cap \mathcal{R}$ is a smooth manifold,
hence $N^{1,2}(\Omega \backslash \mathcal{S})=W^{1,2}(\Omega \backslash \mathcal{S})$.
In general, the map $\pi$ is not surjective. However, in our special setting, $\mathcal{S}$ has high codimension,
we have much more information.
\begin{proposition}(\textbf{Identity is isometry})
Suppose $\Omega$ is an open set of $X$, then
the restriction map $\pi: N^{1,2}(\Omega) \to N^{1,2}(\Omega \backslash \mathcal{S})=W^{1,2}(\Omega \backslash \mathcal{S})$ is an isomorphic isometry.
\label{prn:HD13_1}
\end{proposition}
\begin{proof}
If we proved $\pi$ is an isomorphism, it is clear that $\pi$ is an isometry since $\mathcal{S}$ has measure zero.
Thus, we only need to focus on the proof of isomorphism. For simplicity, we assume $\Omega=X$.
Then $\Omega \backslash \mathcal{S}=X \backslash \mathcal{S}=\mathcal{R}$.
\textit{Injectivity}: Suppose $\pi(f)=0$. Then $\norm{f}{L^2(X)}=0$ since $\mathcal{S}$ has measure zero.
Due to the fact $f \in N^{1,2}(X)$, $\norm{f}{L^2(X)}=0$ implies that $\norm{f}{N^{1,2}(X)}=0$.
Therefore, $f$ is the zero element in $N^{1,2}(X)$.
\textit{Surjectivity}: For every $0 \neq \tilde{f} \in W^{1,2}(\mathcal{R})$, from the proof of Proposition~\ref{prn:HD04_1},
there is a sequence of smooth functions $\tilde{f}_i$ supported on $\mathcal{R}$ such that
$\displaystyle \norm{\tilde{f}_i -\tilde{f}}{W^{1,2}(\mathcal{R})} \to 0$.
In particular, $\tilde{f}_i$ is a Cauchy sequence in $W^{1,2}(\mathcal{R})$.
Since $\displaystyle \norm{\tilde{f}_i -\tilde{f}}{N^{1,2}(X)}=\norm{\tilde{f}_i -\tilde{f}}{W^{1,2}(\mathcal{R})}$,
it is clear that $\tilde{f}_i$ is a Cauchy sequence in $N^{1,2}(X)$.
Therefore, there is a function $f \in N^{1,2}(X)$, as the limit of $\tilde{f}_i$, by completeness of $N^{1,2}(X)$.
After we obtain $f$, it is clear that $\norm{f_i-f}{N^{1,2}(X)} \to 0$, which forces that
\begin{align*}
\norm{f_i-\pi(f)}{W^{1,2}(\mathcal{R})} \to 0.
\end{align*}
Therefore, $\pi(f)=\tilde{f}$.
\end{proof}
In light of Proposition~\ref{prn:HD13_1}, we can regard $N^{1,2}(X)$ as the same Banach space as $W^{1,2}(\mathcal{R})$.
However, $W^{1,2}(\mathcal{R})$ is a Hilbert space. This induces a natural inner product structure on $W^{1,2}(\mathcal{R})$ as follows:
\begin{align*}
\langle\langle f_1, f_2\rangle\rangle
= \int_{\mathcal{R}} \left\{ \pi(f_1) \pi(f_2) +\langle \nabla \pi(f_1), \nabla \pi(f_2)\rangle \right\}d\mu, \quad \forall \; f_1, f_2 \in N^{1,2}(X).
\end{align*}
For simplicity of notation, we shall not differentiate $f$ and $\pi(f)$. Under this convention, we have
\begin{align*}
\langle\langle f_1, f_2\rangle\rangle
= \int_{\mathcal{R}} \left\{ f_1f_2 +\langle \nabla f_1, \nabla f_2 \rangle \right\}d\mu, \quad \forall \; f_1, f_2 \in N^{1,2}(X).
\end{align*}
Therefore, $N^{1,2}(X)$ is isomorphic to $W^{1,2}(\mathcal{R})$ as a Hilbert space. For every $f_1,f_2 \in N^{1,2}(X)$, we define
a nonnegative, symmetric, bilinear form $\mathscr{E}$ as follows
\begin{align}
\mathscr{E}(f_1,f_2) \triangleq \int_{\mathcal{R}} \langle \nabla f_1, \nabla f_2 \rangle d\mu.
\label{eqn:HD26_2}
\end{align}
We want to show that $\mathscr{E}$ is a Dirichlet form.
Actually, it is clear that $\norm{f}{N^{1,2}(X)}^2=\norm{f}{L^2(X)}^2+\mathscr{E}(f,f)$.
Since $N^{1,2}(X)$ is complete, we know that $\mathscr{E}$ is closed by definition.
On the other hand, since $W^{1,2}(\mathcal{R})$ is dense
in $L^2(\mathcal{R})=L^2(X)$, $\mathcal{S}$ has measure zero,
it follows directly that $N^{1,2}(X)$ is dense in $L^2(X)$.
Furthermore, it is clear that
\begin{align}
\mathscr{E}(\min\{1,\max\{0,f\}\}, \min\{1,\max\{0,f\}\}) \leq \mathscr{E}(f,f), \quad \forall \; f \in N^{1,2}(X).
\label{eqn:HE26_1}
\end{align}
Therefore, $\mathscr{E}$ is a closed, nonnegative, symmetric, bilinear form on $N^{1,2}(X)$,
which is a dense subspace of $L^2(X)$, with unit contraction property (\ref{eqn:HE26_1}).
It follows from a standard definition (c.f.~\cite{FuOsTa} for definition of Dirichlet form) that $\mathscr{E}$ is a Dirichlet form.
Not surprisingly, this Dirichlet form $\mathscr{E}$ is much better than general Dirichlet form since
the underlying space $X$ has rich geometry. In fact, suppose $u \in N_0^{1,2}(\Omega)$ for some open set $\Omega \subset X$,
it is clear that $u \equiv 0$ on $\Omega$ if and only if $ \mathscr{E}(u,u)=0$.
This means that $\mathscr{E}$ is irreducible by direct definition.
Also, for every constant $c$, we have $\mathscr{E}(u,v)=0$, whenever $v \equiv c$ in a neighborhood of the support set of $u$.
This means that $\mathscr{E}$ is strongly local.
Furthermore, it follows from Corollary~\ref{cly:HE10_1} that $N^{1,2}(X) \cap C_c(X)$ is dense in $N^{1,2}(X)$ with $N^{1,2}$-norm.
On the other hand, Proposition~\ref{prn:HD24_1} implies that $N^{1,2}(X) \cap C_c(X)$ is dense in $C_c(X)$ with uniform supreme norm.
Consequently, $N^{1,2}(X) \cap C_c(X)$ is a core of $\mathscr{E}$ and $\mathscr{E}$ is a regular Dirichlet form, following from the definition verbatim.
Putting all the above information together, we obtain the following property.
\begin{proposition}[\textbf{Existence of excellent Dirichlet form}]
On the Hilbert space $L^2(X)$, there exists a Dirichlet form $\mathscr{E}$ defined on a dense subspace
$N^{1,2}(X) \subset L^2(X)$, by formula (\ref{eqn:HD26_2}). Furthermore, the Dirichlet form $\mathscr{E}$
is irreducible, strongly local and regular.
\label{prn:HD26_1}
\end{proposition}
With respect to the Dirichlet form $\mathscr{E}$, one can obtain much geometric and analytic information. A good reference is
the nice paper~\cite{KoZhou}, by P. Koskela and Y. Zhou. We now focus on some elementary properties.
Note that there is a unique generator (c.f. Chapter 1 of~\cite{FuOsTa}) of $\mathscr{E}$, which we denote by $\mathcal{L}$.
In other words, $\mathcal{L}$ is a self-adjoint and non-positive definite operator in $L^2(X)$ with domain $Dom(\mathcal{L})$
which is dense in $N^{1,2}(X)$ such that
\begin{align}
\mathscr{E}(f,h)=-\int_X h \cdot \mathcal{L}f d\mu, \quad \forall \; f \in Dom(\mathcal{L}), \; h\in N^{1,2}(X).
\label{eqn:HD27_1}
\end{align}
Note that $C_c^{\infty}(\mathcal{R})$ is a dense subset of $Dom(\mathcal{L})$. Suppose $f \in C_c^{\infty}(\mathcal{R})$, $h \in N^{1,2}(X)=N^{1,2}(\mathcal{R})$,
it is clear that
\begin{align*}
\mathscr{E}(f,h)=\int_{\mathcal{R}} \langle \nabla f, \nabla h\rangle d\mu= -\int_X h \cdot \Delta f d\mu.
\end{align*}
Therefore, $\mathcal{L}$ is nothing but the extension of the classical Laplacian operator, with domain as the largest dense subset of $N^{1,2}(X)$ such that the integration by parts, i.e., equation (\ref{eqn:HD27_1}), holds.
For this reason, we shall just denote $\mathcal{L}$ by $\Delta$ in the future.
Based on the generator operator $\Delta$, there is an associated heat semigroup $\left(P_t\right)_{t \geq 0}= \left( e^{t\Delta}\right)_{t \geq 0}$, which acts on
$L^2(X)$ with the following properties(c.f. Chapter 1 of~\cite{FuOsTa}).
\begin{itemize}
\item Semi-group: $P_0=Id$; $P_t \circ P_s=P_{t+s}$, for every $t, s \geq 0$.
\item Generator: $\displaystyle \lim_{t \to 0^+} \norm{\frac{1}{t}(P_t f-f) -\Delta f}{L^2(X)}=0$, for every $f \in L^2(X) \cap Dom(\Delta)$.
\item $L^2$-contractive: $\norm{P_t f}{L^2(X)}^2 \leq \norm{f}{L^2(X)}^2$, for every $f \in L^2(X)$, $t>0$.
\item Strong continuous: $\displaystyle \lim_{t \to 0^{+}}\norm{P_t f-f}{L^2(X)}=0$, for every $f \in L^2(X)$.
\item Markovian: $\norm{P_t f}{L^{\infty}(X)} \leq \norm{f}{L^{\infty}(X)}$, for every $f \in L^2(X)\cap L^{\infty}(X)$, $t>0$.
\item Heat solution: $\Delta P_t f= \frac{\partial}{\partial t} P_t f$, for every $f \in L^2(X)$ and $t>0$.
\end{itemize}
The above properties are well known in semigroup theory on Banach spaces(c.f. Section 7.4 of \cite{Evans}). Actually, for every $f \in L^2(X)$,
one can also show that $P_t f$ is the unique square-integrable solution with initial value $f$(c.f. Proposition 1.2 of~\cite{Stu95} and references therein).
We call $(P_t)_{t \geq 0}$ as the heat semigroup as usual.
Associated with this heat semigroup, there exists a nonnegative kernel function, or fundamental solution, $p(t,x,y)$,
such that
\begin{align*}
P_t(f)(y)=\int_X f(x) p(t,x,y)d\mu_x, \quad \forall \; f \in L^2(X), \; t>0.
\end{align*}
Moreover, $p$ satisfies the symmetry $p(t,x,y)=p(t,y,x)$. Interested readers are referred to Proposition 2.3 and the discussion in Section 2.4(C) of~\cite{Stu95}
for more detailed information.
As usual, we call $p(t,x,y)$ as the heat kernel.
\begin{definition}
Suppose $u \in N_{loc}^{1,2}(\Omega)$. Define
\begin{align}
\int_{\Omega} \varphi \Delta u \triangleq -\mathscr{E}(u,\varphi)
\label{eqn:HD29_2}
\end{align}
for every $\varphi \in N_c^{1,2}(\Omega)$, i.e., $\varphi \in N^{1,2}(\Omega)$ and has compact support set in $\Omega$.
Similarly, (\ref{eqn:HD29_2}) can be applied if $u \in N^{1,2}(\Omega)$ and $\varphi \in N_0^{1,2}(\Omega)$.
\label{dfn:HD29_1}
\end{definition}
Suppose $u \in N_{loc}^{1,2}(\Omega) \cap C^{2}(\Omega \backslash \mathcal{S})$, then $\Delta u|_{\Omega \backslash \mathcal{S}}$
is a continuous function.
By taking value $\infty$ on $\mathcal{S}$, we can regard $\Delta u$ as an extended function on $\Omega$. Suppose $\Delta u \in L_{loc}^2(\Omega)$, then
for every smooth test function $\varphi_i \in N_c^{1,2}(\Omega)$, we have
\begin{align*}
\int_{\Omega \backslash \mathcal{S}} \varphi_i \Delta u=-\int_{\Omega \backslash \mathcal{S}} \langle \nabla u, \nabla \varphi_i \rangle.
\end{align*}
Let $\varphi$ be the limit of $\varphi_i$ in $N^{1,2}(\Omega)$. Taking limit of the above equation shows that
\begin{align*}
\int_{\Omega} \varphi \Delta u=-\int_{\Omega} \langle \nabla u, \nabla \varphi \rangle.
\end{align*}
Therefore, whenever $u \in N_{loc}^{1,2}(\Omega) \cap C^{2}(\Omega \backslash \mathcal{S})$ and the classical $\Delta u$ is in $L_{loc}^2(\Omega)$,
we see that the LHS and RHS of (\ref{eqn:HD29_2}) holds in the classical sense. Similar argument applies
if $u \in N^{1,2}(\Omega) \cap C^2(\Omega \backslash \mathcal{S})$, $\Delta u \in L^2(\Omega)$, $\varphi \in N_0^{1,2}(\Omega)$.
Therefore, Definition~\ref{dfn:HD29_1} is justified.
Now we assume $u \in N_c^{1,2}(\Omega)$.
Then in the weak sense, for every $\varphi \in N_c^{1,2}(\Omega)$, we can define
$\int_{\Omega} \varphi \Delta u$. It is not hard to see that $\int_{\Omega} \varphi \Delta u$ makes sense even if $\varphi$ is in $N^{1,2}(\Omega)$ only.
In fact, let $\chi$ be a cutoff function with value $1$ on $\Omega'$ and vanishes
around $\partial \Omega$, where $\Omega'$ contains the support of $u$.
By Definition~\ref{dfn:HD29_1}, we have
\begin{align*}
\int_{\Omega} (\chi \varphi) \Delta u= -\mathscr{E}(u, \chi \varphi)=-\int_{\Omega} \langle \nabla u, \nabla (\chi \varphi)\rangle
=-\int_{\Omega'} \langle \nabla u, \nabla \varphi \rangle=-\int_{\Omega} \langle \nabla u, \nabla \varphi \rangle
=-\mathscr{E}(u,\varphi).
\end{align*}
The above calculation does not depend on the particular choice of $\chi$. Consequently, we can define
$\int_{\Omega} \varphi \Delta u$ as $-\mathscr{E}(u,\varphi)$.
Summarizing the above discussion, we have the following property.
\begin{proposition}[\textbf{Integration by parts}]
Suppose $\Omega$ is a domain in $X$, $f_1 \in N_c^{1,2}(\Omega)$, $f_2 \in N^{1,2}(\Omega)$.
Then we have
\begin{align}
\int_{\Omega} f_2 \Delta f_1 d\mu = -\int_{\Omega} \langle \nabla f_1, \nabla f_2 \rangle d\mu=\int_{\Omega} f_1 \Delta f_2 d\mu.
\label{eqn:HD13_2}
\end{align}
Furthermore, if $f_2 \in N^{1,2}(\Omega) \cap C^2(\Omega \backslash \mathcal{S})$ and $\Delta f_2 \in L^2(\Omega \backslash \mathcal{S})$
as a classical function, then we can understand the integral
$\int_{\Omega} f_1 \Delta f_2 d\mu=\int_{\Omega \backslash \mathcal{S}} f_1 \Delta f_2 d\mu$ in the classical sense.
Similarly, if $f_1 \in N_c^{1,2}(\Omega) \cap C^2(\Omega \backslash \mathcal{S})$,
$\Delta f_1 \in L^2(\Omega \backslash \mathcal{S})$ as a classical function, then
$\int_{\Omega} f_2 \Delta f_1 d\mu=\int_{\Omega \backslash \mathcal{S}} f_2 \Delta f_1 d\mu$ can be understood in
the classical sense. If both $f_1$ and $f_2$ locate in $N_0^{1,2}(\Omega)$, then (\ref{eqn:HD13_2}) also holds.
\label{prn:HD13_2}
\end{proposition}
\begin{definition}
Suppose $u \in N_{loc}^{1,2}(\Omega)$, $f \in L_{loc}^2(\Omega)$, we say $\Delta u \geq f$ in the weak sense whenever
\begin{align}
\int_{\Omega}(-\Delta u +f)\varphi= \mathscr{E}(u,\varphi) + \int_{\Omega} f\varphi \leq 0
\label{eqn:HD29_1}
\end{align}
for every nonnegative test function $\varphi \in N_c^{1,2}(\Omega)$. We call $u$ subharmonic if $\Delta u \geq 0$ in the weak sense.
We call $u$ superharmonic if $-u$ is subharmonic. We call $u$ harmonic if $u$ is both subharmonic and superharmonic.
\label{dfn:HD29_2}
\end{definition}
Due to Proposition~\ref{prn:HD04_1}, for a $u \in N_{loc}^{1,2}(\Omega)$, in order to check (\ref{eqn:HD29_1}) for all $\varphi \in N_c^{1,2}(\Omega)$,
it suffices to check all smooth nonnegative test functions with supports in $\Omega \backslash \mathcal{S}$.
It is important to notice that the restriction of $\Delta$ on $\mathcal{R}$ is the classical Laplacian on Riemannian manifold.
In fact, if a function $u$ is harmonic in the above sense, then $u |_{\mathcal{R}}$ is harmonic function in the distribution sense.
By standard improving regularity theory of elliptic equations, we know our $u$ is smooth and $\Delta u=0$ in the classical sense.
Similarly, one can follow the standard route to define heat solution (sub solution, super solution) for the heat operator
$\square= \left( \frac{\partial}{\partial t} -\Delta \right)$ in the weak sense. We leave these details to interested readers.
It is quite clear that a weak heat solution is a smooth function when restricted on $\mathcal{R} \times (0,T]$, by
standard improving regularity theory of heat equations(c.f. Chapter 7 of~\cite{Evans}).
\subsection{Harmonic functions and heat flow solutions on model space}
Suppose $K$ is a compact subset of $\Omega \backslash \mathcal{S}$, it is clear that $K$ is also a compact subset of $\Omega$.
However, the reverse is not true. If $K$ is a compact subset of $\Omega$, then $K\backslash \mathcal{S}$ may not be a compact
subset of $\Omega \backslash \mathcal{S}$.
For this reason, we see that $N_{loc}^{1,2}(\Omega) \subset N_{loc}^{1,2}(\Omega \backslash \mathcal{S})$ and not equal if $\mathcal{S} \neq \emptyset$, even if
$\mathcal{S}$ has very high codimension.
However, if we restrict our attention only on bounded subharmonic functions, then the above difference will vanish.
\begin{proposition}[\textbf{Extension of bounded subharmonic functions}]
Suppose $\Omega$ is a bounded open domain in $X$, $u$ is a bounded subharmonic function on $\Omega \backslash \mathcal{S}$.
Then $u \in N_{loc}^{1,2}(\Omega)$ and it is subharmonic on $\Omega$.
\label{prn:HD16_2}
\end{proposition}
\begin{proof}
It suffices to prove that $u \in N_{loc}^{1,2}(\Omega)$.
Note that by definition, we only have $u \in N_{loc}^{1,2}(\Omega \backslash \mathcal{S})$, which is a superset of $N_{loc}^{1,2}(\Omega)$.
In fact, for each small $r>0$, one can construct a Lipschitz cutoff function
\begin{align}
\chi(x)=\phi \left(\frac{d(x,\mathcal{S})}{r} \right),
\label{eqn:HD16_1}
\end{align}
where $\phi$ is a cutoff function on $[0,\infty)$ which is equivalent to $1$ on $[0,1]$, $0$ on $[2, \infty)$, and $|\phi'| \leq 2$.
By the assumption of Minkowski codimension of $\mathcal{S}$, we have
\begin{align}
| \Omega \cap \supp \chi| \leq Cr^3, \quad \int_{\Omega} |\nabla \chi|^2 \leq C r^{-2}\int_{\Omega \cap \{\nabla \chi \neq 0\}} \chi^2 \leq Cr.
\label{eqn:HD16_2}
\end{align}
Fix a relatively compact subset $\Omega' \subset \Omega$, we can find a cutoff function $\eta$ which is identically $1$ on $\Omega'$ and vanishes around
$\partial \Omega$. Moreover, $|\nabla \eta| \leq C$, which depends on $\Omega'$ and $\Omega$.
By adding a constant if necessary, we can assume $u \geq 0$.
Note that $u$ is subharmonic on $\Omega \backslash \mathcal{S}$, $u\eta^2 (1-\chi)^2$ can be chosen as a test function. It follows from definition that
\begin{align*}
0&\leq \int_{\Omega \backslash \mathcal{S}} (\Delta u) u\eta^2(1-\chi)^2
=-\int_{\Omega \backslash \mathcal{S}} \langle \nabla u, \nabla (u\eta^2(1-\chi)^2) \rangle \\
& =-\int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 (1-\chi)^2
+ \int_{\Omega \backslash \mathcal{S}} u \langle \nabla u, -2(1-\chi)^2\eta \nabla \eta+ 2\eta^2(1-\chi) \nabla \chi \rangle.
\end{align*}
Note that $u,\eta,\nabla \eta$ are bounded. Then H\"{o}lder inequality applies.
\begin{align*}
\frac{1}{2} \int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 (1-\chi)^2
&\leq C+ C\int_{\Omega \backslash \mathcal{S}}\eta^2 (1-\chi) |\nabla u||\nabla \chi|\\
&\leq C+ \frac{1}{4}\int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 (1-\chi)^2 + C\int_{\Omega \backslash \mathcal{S}} \eta^2 |\nabla \chi|^2.
\end{align*}
Recall the definition of $\chi$ in (\ref{eqn:HD16_1}) and estimate (\ref{eqn:HD16_2}).
Let $r \to 0$, the above inequality yields that
\begin{align*}
\int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 \leq C,
\end{align*}
which forces that $\displaystyle \int_{\Omega' \backslash \mathcal{S}} |\nabla u|^2 \leq C$.
Hence $u \in W^{1,2}(\Omega' \backslash \mathcal{S})$ since $u$ is bounded. This is the same to say $u \in N^{1,2}(\Omega')$.
By the arbitrary choice of $\Omega'$, we have proved that $u \in N_{loc}^{1,2}(\Omega)$.
\end{proof}
We now move on to the discussion of heat kernels.
\begin{proposition}[\textbf{Heat Kernel estimates}]
The exists a unique heat kernel $p(t,x,y)$ of $X$, with respect to the Dirichlet form $\mathscr{E}=\langle \nabla \cdot, \nabla \cdot \rangle$.
Moreover, $p(t,x,y)$ satisfies the following properties.
\begin{itemize}
\item Stochastically completeness. In other words, we have
\begin{align*}
\int_X p(t,x,y) d\mu_x=1
\end{align*}
for every $x \in X$.
\item The Gaussian estimate holds. In other words, there exists a constant $C$ depending only on $n,\kappa$ such that
\begin{align}
\frac{1}{C}t^{-n} e^{-\frac{d^2(x,y)}{3t}} \leq p(t,x,y) \leq Ct^{-n} e^{-\frac{d^2(x,y)}{5t}}
\label{eqn:HC31_7}
\end{align}
for every $x,y \in X$ and $t>0$.
\item For each positive integer $j$, there is a constant $C=C(n,\kappa,j)$ such that
\begin{align}
\left| \left( \frac{\partial}{\partial t} \right)^j p(t,x,y)\right| \leq Ct^{-n-j} e^{-\frac{d^2(x,y)}{5t}}
\label{eqn:HD08_1}
\end{align}
for every $x,y \in X$ and $t>0$.
\label{prn:HC29_7}
\end{itemize}
\label{prn:HD07_1}
\end{proposition}
\begin{proof}
Since $X$ satisfies the doubling property and has a uniform $(1,2)$-Poincar\'{e} constant, by Corollary~\ref{cly:HD20_1}
and Proposition~\ref{prn:HC29_2},
the existence of the heat Kernel follows from the work of Sturm (c.f.Proposition 2.3 of~\cite{Stu95}).
The uniqueness of the heat kernel follows from the uniqueness of the heat semigroup.
The stochastic completeness is guaranteed by the doubling property, see Theorem 4 and the following remarks of~\cite{Stu94}.
The Gaussian estimate follows from Corollary 4.2 and Corollary 4.10 in Sturm's paper~\cite{Stu96}, where $C$
depends on the volume doubling condition and the $(1,2)$-Poincar\'{e} constant.
The heat kernel derivative estimate, inequality (\ref{eqn:HD08_1}), follows from Corollary 2.7. of ~\cite{Stu95}, whose proof follows the same line
as Theorem 6.3 of~\cite{Saloff}.
\end{proof}
In general, the estimates of heat kernel only hold almost everywhere with respect to the measure $d\mu$.
However, in the current situation, when restricted on $\mathcal{R} \times (0,\infty)$, $p$ is clearly a smooth function.
Therefore, (\ref{eqn:HC31_7}) and (\ref{eqn:HD08_1}) actually hold true everywhere away from $\mathcal{S}$.
Note that $\Delta p=\frac{\partial}{\partial t} p$ clearly locates in $L^2(X) \cap C^{\infty}(\mathcal{R})$.
Hence integration by parts (Proposition~\ref{prn:HD13_2}) applies. Then by standard radial cutoff function construction and direct calculation,
Proposition~\ref{prn:HD07_1}
yields the following estimates immediately.
\begin{corollary}[\textbf{Off-diagonal integral estimates of heat kernel}]
For every $r>0, t>0$, and $x_0 \in X$, we have
\begin{align}
&\int_{X \backslash B(x_0,2r)} p^2 d\mu_x<Ct^{-n}e^{-\frac{r^2}{5t}}, \label{eqn:HD08_7}\\
&\int_{X \backslash B(x_0,2r)} |\nabla p(t,x,x_0)|^2 d\mu_x
< C \left( \frac{1}{t} + \frac{1}{r^2}\right) t^{-n} e^{-\frac{r^2}{5t}}, \label{eqn:HD08_4}
\end{align}
for some $C=C(n,\kappa)$. Consequently, we have
\begin{align}
\int_0^{t} \int_{X \backslash B(x_0,2r)} \left( p^2+|\nabla p|^2 \right) d\mu_x ds
< C \int_0^t \left(1+\frac{1}{s} + \frac{1}{r^2}\right) s^{-n} e^{-\frac{r^2}{5s}} ds.
\label{eqn:HD08_3}
\end{align}
\end{corollary}
By virtue of Proposition~\ref{prn:HD04_1}, smooth functions are dense in $N^{1,2}(X)$. Then it is easy to see that $(X,g,d\mu)$, together with the heat process,
has non-negative Ricci curvature in the sense of Bakry-Emery, i.e., $X \in CD(0,\infty)$ by the notation of Bakry-Emery(c.f.~\cite{BaEm},~\cite{Bak}).
The following Proposition is nothing but part of Proposition 2.1 of~\cite{Bak}.
The rigorous proof is tedious and is postponed in the appendix.
\begin{proposition}[\textbf{Weighted Sobolev inequality}]
For every function $f \in N^{1,2}(X)$, every $t>0$ and every $y \in X$, we have
\begin{align}
\int_X f^2(x) p(t,x,y) d\mu_x - \left(\int_X f(x) p(t,x,y)d\mu_x \right)^2 \leq 2t \int_X |\nabla f|^2(x) p(t,x,y) d\mu_x.
\label{eqn:HD05_3}
\end{align}
In other words, for every $t>0$, with respect to the probability measure $p(t,x,y)d\mu_x$, $L^2$-Sobolev inequality holds
with the uniform Sobolev constant $\frac{1}{2t}$.
\label{prn:HD07_2}
\end{proposition}
On a Riemannian manifold with proper geometry bound, the heat kernel can be regarded as a solution starting from a $\delta$-function.
This property also holds for every $X \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{proposition}[\textbf{$\delta$-function property of heat kernel}]
Suppose $w$ is a function on $[0,t] \times X$, differentiable along the time direction,
$w(s,\cdot) \in N_c^{1,2}(X)$ for each $s \in [0,t]$.
Moreover, we assume $\displaystyle \limsup_{s \to 0^{+}} \norm{w(s,\cdot)}{L^1(X)}<\infty$, $w$ is continuous at $(0,x_0)$.
Then we have
\begin{align}
&-w(0,x_0) +\int_{X} w(t,x)p(t,x,x_0)d\mu_x=\int_0^{t} \int_X \left\{ \left(\frac{\partial}{\partial s}+\Delta \right) w(s,x) \right\} p(s,x,x_0) d\mu_x ds.
\label{eqn:HD09_1}
\end{align}
Consequently, equation (\ref{eqn:HD09_1}) holds for functions $w(s,x)+a(s)$ where $a$ is a differentiable function of time.
\label{prn:HD09_1}
\end{proposition}
\begin{proof}
Clearly, (\ref{eqn:HD09_1}) holds if $w(s,\cdot) \equiv a(s)$. Therefore, it suffices to show (\ref{eqn:HD09_1}) when $w(s,\cdot) \in N_c^{1,2}(X)$ for each $s$.
For simplicity of notation, we denote $p(t,x,x_0)$ by $p$ and assume $d\mu_x$ as the default measure. It follows from integration by parts that
\begin{align*}
\frac{d}{dt} \int_{X} wp
=\int_X \left\{ \left(\frac{\partial}{\partial t}+\Delta \right) w \right\} p + \int_X w \left\{\left(\frac{\partial}{\partial t}-\Delta \right) p\right\}
=\int_X \left\{ \left(\frac{\partial}{\partial t}+\Delta \right) w \right\} p.
\end{align*}
For each $\epsilon>0$, we can find $\delta$ small enough such that $|w(s,x)-w(0,x_0)|<\epsilon$ whenever $0<s<\delta^2$ and $d(x,x_0)<\delta$.
Then the heat kernel estimate implies that
\begin{align*}
&\quad \left|-w(0,x_0)+\lim_{t \to 0^+} \int_X w(t,x)p(t,x,x_0)d\mu_x \right|
= \left|\lim_{t \to 0^+} \int_X \left\{w(t,x)-w(0,x_0)\right\}p(t,x,x_0)d\mu_x \right|\\
&\leq \lim_{t \to 0^{+}} \left\{ |w(0,x_0)|\int_{X \backslash B(x_0,\delta)} p(t,x,x_0)d\mu_x
+ \int_{X \backslash B(x_0,\delta)} |w(t,x)| p(t,x,x_0)d\mu_x +\epsilon \right\}
\leq \epsilon.
\end{align*}
By arbitrary choice of $\epsilon$, we have $\displaystyle \lim_{t \to 0^+} \int_X w(t,x)p(t,x,x_0)d\mu_x =w(0,x_0)$. Plugging this relationship into the integration of previous equation,
we obtain (\ref{eqn:HD09_1}).
\end{proof}
Based on the excellent properties of heat kernels, from Proposition~\ref{prn:HC29_7} to Proposition~\ref{prn:HD09_1}, we are ready to generalize the celebrated Cheng-Yau estimate (c.f.~\cite{ChYau}) to our setting.
We basically follow the paper~\cite{KoRaSh}.
However, due to the essential importance of this estimate and the excellent geometry of our underlying space, we write down a simplified proof here.
\begin{proposition} [\textbf{Cheng-Yau type gradient estimate}]
Suppose $\Omega=B(x_0,4r)$ for some $r>0$ and $x_0 \in \mathcal{R}$.
Suppose $u \in L^{\infty}(\Omega) \cap N^{1,2}(\Omega)$ and $u$ satisfies the equation
\begin{align}
\Delta u=h
\label{eqn:HE03_3}
\end{align}
for some $h \in C^{\frac{1}{2}}(\Omega)$.
Then we have
\begin{align}
|\nabla u|(x_0) \leq \frac{C}{r} \left( \norm{u}{L^{\infty}(\Omega)} + r^{\frac{5}{2}}[h]_{C^{\frac{1}{2}}(\Omega)} + r^2|h(x_0)| \right)
\label{eqn:HD01_1}
\end{align}
for a constant $C=C(n,\kappa)$, where $\displaystyle [h]_{C^{\frac{1}{2}}(\Omega)}=\sup_{x,y \in \Omega} \frac{|h(x)-h(y)|}{d^{\frac{1}{2}}(x,y)}$.
\label{prn:HC29_6}
\end{proposition}
\begin{proof}
Without loss of generality, we assume $r=1$ and $\norm{u}{L^{\infty}(\Omega)}=1$.
Let $\chi$ be a Lipschitz cutoff function such that $\chi \equiv 1$ on $B(x_0,1)$ and vanishes outside $B(x_0, 2)$ such that $|\nabla \chi| \leq 2$.
Define
\begin{align*}
a(t) &\triangleq P_t(u\chi)(x_0), &J(t) &\triangleq \frac{1}{t} \int_0^{t} \int_X |\nabla w(s,x)|^2 p(s,x,x_0) d\mu_x ds, \quad \forall \; t>0,\\
w(t,x) &\triangleq u(x)\chi(x)-a(t), & J(0) &\triangleq \lim_{t \to 0^{+}} J(t)=|\nabla w(0, x_0)|^2=|\nabla u(x_0)|^2.
\end{align*}
From the definition of $w(t,x)$, it is clear that $\displaystyle \int_X w(t,x) p(t,x,x_0)d\mu_x=0$. Applying (\ref{eqn:HD05_3}), we have
\begin{align}
&\int_X w^2(t,x) p(t,x,x_0)d\mu_x \leq 2t\int_X |\nabla w|^2p(t,x,x_0) d\mu_x, \label{eqn:HD09_3}\\
&|a(t)|= \left|\int_X u\chi p(t,x,x_0)d\mu_x\right| \leq \int_X |u\chi| p(t,x,x_0)d\mu_x \leq \int_X p(t,x,x_0)d\mu_x \leq 1, \label{eqn:HF20_1}\\
&|w(t,x)|= |u(x)\chi(x) -a(t)| \leq |u(x)\chi(x)|+|a(t)| \leq 2, \quad \forall \; x \in X. \label{eqn:HD06_5}
\end{align}
It follows from the definition of $J$ that
\begin{align}
|\nabla u(x_0)|^2=J(0)=-\int_0^{1} J'(t) dt +J(1).
\label{eqn:HD06_6}
\end{align}
However, in light of (\ref{eqn:HD09_3}), we have
\begin{align*}
J'(t)&=-\frac{1}{t} J(t)+\frac{1}{t} \int_X |\nabla w|^2p(t,x,x_0) d\mu_x
\geq-\frac{1}{t} J(t)+\frac{1}{2t^2} \int_X w^2(t,x) p(t,x,x_0)d\mu_x \notag\\
&=\frac{1}{t^2} \left(-tJ(t) +\frac{1}{2}\int_X w^2(t,x) p(t,x,x_0)d\mu_x \right)
=\frac{F(t)}{t^2},
\end{align*}
where we used the definition $F(t) \triangleq -tJ(t) +\frac{1}{2}\int_X w^2(t,x) p(t,x,x_0)d\mu_x$.
Therefore, by the previous inequalities, equation (\ref{eqn:HD06_6}) can be rewritten as
\begin{align}
|\nabla u(x_0)|^2 \leq -\int_0^1 \frac{F(s)}{s^2} ds -F(1) +\frac{1}{2} \int_X w^2p
\leq 2-\int_0^1 \frac{F(s)}{s^2} ds -F(1).
\label{eqn:HD09_7}
\end{align}
Therefore, $|\nabla u(x_0)|$ follows from the estimate of $F(t)$.
We now focus on the estimate of $F(t)$. Applying (\ref{eqn:HD09_1}) to $w^2$, we obtain
\begin{align}
\int_0^t \int_X \left\{ \left( \frac{\partial}{\partial s} + \Delta \right) w^2(s,x)\right\}p(s,x,x_0)d\mu_x ds
=\int_X w^2(t,x)p(t,x,x_0) d\mu_x,
\label{eqn:HD06_7}
\end{align}
since $w(0,x_0)=0$. Note that the application of (\ref{eqn:HD09_1}) can be justified.
Actually, from its definition, $w^2(x,t)=u\chi (u\chi -2a(t)) +a^2(t)$. The first part of the right hand side of this equation is a function in $N_c^{1,2}(X)$ for each $t$. It is Lipschitz continuous around $(0,x_0)$,
since $x_0$ is a smooth point and the standard improving regularity theory of elliptic functions applies here.
The second part is a differentiable function of time.
Therefore, (\ref{eqn:HD09_1}) applies for $w^2$.
On the other hand, the fact that $u$ is a weak solution of (\ref{eqn:HE03_3}) implies that
\begin{align*}
|\nabla w|^2=\frac{1}{2}\Delta w^2-w\Delta w
=\frac{1}{2}\Delta w^2-w(u\Delta \chi +\chi h+2 \langle \nabla u, \nabla \chi \rangle)
\end{align*}
in the weak sense.
Plugging the above equation and (\ref{eqn:HD06_7}) into the definition of $J(t)$, we have
\begin{align*}
tJ(t)
&=\int_0^t\int_X \left\{ \frac{1}{2}\left( \frac{\partial}{\partial s}+\Delta\right) w^2
-w(u\Delta \chi+\chi h +2 \langle \nabla u, \nabla \chi \rangle) \right\} p(s,x,x_0)d\mu_x ds \notag\\
&=\int_0^t\int_X \left\{ \frac{1}{2}w^2
-w(u\Delta \chi +\chi h+2 \langle \nabla u, \nabla \chi \rangle) \right\} p(s,x,x_0)d\mu_x ds.
\end{align*}
For the simplicity of notation, we will denote $p(s,x,x_0)$ by $p$ only. Also, we will drop integration elements when they are clear.
From the definition of $F(t)$, the above equation can be written as
\begin{align*}
&\quad F(t)-\int_0^t\int_X w\chi h p=\int_0^{t}\int_X w(u\Delta \chi+2 \langle \nabla u, \nabla \chi \rangle) p.
\end{align*}
Recall that $w=u\chi -a$ and $\nabla w=u\nabla \chi + \chi \nabla u$. Integrating by parts gives us
\begin{align*}
F(t)-\iint w\chi h p =\iint \langle -up\nabla w-uw \nabla p +pw \nabla u, \nabla \chi \rangle
=\iint \langle -ap \nabla u -uw \nabla p, \nabla \chi \rangle - pu^2|\nabla \chi|^2.
\end{align*}
By (\ref{eqn:HF20_1}) and (\ref{eqn:HD06_5}), we have $|a| \leq 1$ and $|uw| \leq 2$.
By the choice of $\chi$ and the H\"{o}lder inequality, it is clear that
\begin{align}
&\quad \left|F(t) +\int_0^t\int_X ( u^2|\nabla \chi|^2 -w\chi h) p \right|
\leq 2 \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} (p|\nabla u| + |\nabla p|) \notag\\
&\leq C\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla u| ^2+1\right) \right)^{\frac{1}{2}}
\cdot \left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla p| ^2+p^2\right) \right)^{\frac{1}{2}} \notag\\
&\leq C\left(1+ \norm{h}{L^{\infty}(\Omega)} \right) \sqrt{t}\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla p| ^2+p^2\right) \right)^{\frac{1}{2}} . \label{eqn:GD05_1}
\end{align}
Note that in the last step of the above inequality, we used the following Caccioppoli-type inequality:
\begin{align*}
\int_{B(x_0,2)} |\nabla u|^2 \leq C \int_{B(x_0,4)} (u^2+h^2) \leq C(n,\kappa) \left(1+\norm{h}{L^2(\Omega)} \right)^2
\leq C(n,\kappa) \left(1+\norm{h}{L^{\infty}(\Omega)} \right)^2,
\end{align*}
which can be proved by multiplying equation (\ref{eqn:HE03_3}) on both sides by $\tilde{\chi}^2 u$ and doing integration by parts, for some cutoff function $\tilde{\chi}$.
By inequality (\ref{eqn:HD08_3}), the last term in (\ref{eqn:GD05_1}) can be controlled by $C(\kappa, \beta) t^{\beta}$ for any positive number $\beta$.
For the simplicity of later calculation, we choose $\beta=\frac{3}{4}$. Note that $\norm{h}{L^{\infty}(\Omega)}<|h(x_0)|+[h]_{C^{\frac{1}{2}}(\Omega)}$.
Let $L=1+ |h(x_0)|+[h]_{C^{\frac{1}{2}}(\Omega)}$, then we have
\begin{align}
|F(t)| &\leq \left| \int_0^t\int_X w\chi h p \right| + \left| \int_0^t\int_{B(x_0,2) \backslash B(x_0,1)} u^2|\nabla \chi|^2 p \right|+ CLt^{\frac{5}{4}} \notag\\
&\leq \left| \int_0^t\int_X w\chi h p \right| +4 \int_0^t\int_{B(x_0,2) \backslash B(x_0,1)} p +CLt^{\frac{5}{4}}.
\label{eqn:HD09_9}
\end{align}
The second term on the right hand side of the above inequality can be absorbed by the last term, due to the exponential decay of $p$ and Euclidean volume growth condition(c.f. Proposition~\ref{prn:HD07_1} and Proposition~\ref{prn:HD19_1}).
On the other hand, since $pw$ has zero integeral, we have
\begin{align*}
\left| \int_0^t\int_X w\chi h p \right|&= \left| \int_0^t\int_X (\chi h-h(x_0)) pw \right|
\leq \left| \int_0^t\int_{B(x_0,2)} (\chi h-h(x_0)) pw \right| + \left| \int_0^t\int_{X \backslash B(x_0,2)} h(x_0)pw \right|.
\end{align*}
However, as $\chi h-h(x_0)$ vanishes at $x_0$, we have $ |\chi h-h(x_0)| \leq [h]_{C^{\frac{1}{2}}(\Omega)} \cdot d^{\frac{1}{2}}(x,x_0)$.
Consequently, we have
\begin{align*}
&\left| \int_0^t\int_{B(x_0,2)} (\chi h-h(x_0)) pw d\mu_x ds\right| < C [h]_{C^{\frac{1}{2}}(\Omega)} \int_0^t s^{\frac{1}{4}} ds< C [h]_{C^{\frac{1}{2}}(\Omega)} t^{\frac{5}{4}} < CLt^{\frac{5}{4}},\\
&\left| \int_0^t\int_{X \backslash B(x_0,2)} h(x_0)pw d\mu_x ds\right|<2|h(x_0)| \int_0^t\int_{X \backslash B(x_0,2)} p d\mu_x ds < C|h(x_0)| t^{\frac{5}{4}}<CLt^{\frac{5}{4}}.
\end{align*}
Plugging the above inequalities into (\ref{eqn:HD09_9}), we obtain
\begin{align}
|F(t)|<CLt^{\frac{5}{4}}, \quad \int_0^{1}\frac{|F(s)|}{s^2} ds< CL \int_0^1 s^{-\frac{3}{4}} ds<CL.
\label{eqn:HD09_5}
\end{align}
Recall that $L=1+ |h(x_0)|+[h]_{C^{\frac{1}{2}}(\Omega)}$. Plugging the above inequalities into (\ref{eqn:HD09_7}), we obtain the desired estimate of $|\nabla u(x_0)|$.
\end{proof}
Combining Proposition~\ref{prn:HC29_6} and the $C^{\alpha}$-estimate (c.f. Theorem 4.1 of~\cite{Saloffnote}) for bounded heat solutions, one can derive the Li-Yau type gradient estimate.
Alternatively, for bounded heat solution $u$, one can follow the proof of Lemma~\ref{lma:HH10_1} to obtain the uniform bound of $|\nabla u|$ by De-Giorgi iteration process.
Another interesting application of Cheng-Yau type inequality is the following Liouville theorem.
\begin{corollary}[\textbf{Liouville theorem}]
Suppose $u$ is a bounded harmonic function on $\mathcal{R}$, then $u \equiv C$.
\label{cly:HE08_1}
\end{corollary}
\begin{proof}
By virtue of Proposition~\ref{prn:HD16_2}, the extension property of subharmonic functions, we know that $u \in N_{loc}^{1,2}(X)$.
In light of Proposition~\ref{prn:HD04_1}, the dense property of smooth functions in $N_0^{1,2}(\Omega)$, it is clear that $\Delta u=0$ on $X$ in the weak sense.
Fix $x_0 \in \mathcal{R}$ and a large $r>0$, by Cheng-Yau estimate in Proposition~\ref{prn:HC29_6},
we have
\begin{align*}
|\nabla u|(x_0) < \frac{C}{r}
\end{align*}
for some uniform constant $C$. Let $r \to \infty$, we see that $|\nabla u|(x_0)=0$. It follows that $\nabla u \equiv 0$ on $\mathcal{R}$
by the arbitrary choice of $x_0 \in \mathcal{R}$. Consequently, $u \equiv C$ on $\mathcal{R}$.
\end{proof}
\begin{proposition}[\textbf{Estimates for Dirichlet problem solution}]
Suppose $\Omega$ is a bounded open set of $X$, $f $ is a continuous function in $N^{1,2}(\Omega)$. Then we have the following properties.
\begin{itemize}
\item There is a unique solution $u \in N^{1,2}(\Omega)$ solving the Dirichlet problem
\begin{align}
\Delta u=0, \quad \textrm{in} \; \Omega; \quad
(u-f)|_{\partial \Omega}=0
\label{eqn:HE03_2}
\end{align}
in the weak sense of traces. In other words, $\Delta u=0$ in the weak sense and $u-f \in N_0^{1,2}(\Omega)$.
\item Weak maximum principle holds for $u$, i.e.,
\begin{align}
\sup_{x \in \Omega} u(x)=\sup_{x \in \partial \Omega} u(x), \quad \inf_{x \in \Omega} u(x)=\inf_{x \in \partial \Omega} u(x).
\end{align}
\item Strong maximum principle holds for $u$, i.e., if there is an interior point $x_0 \in \Omega$ such that
$\displaystyle u(x_0)= \sup_{x \in \Omega} u(x)$ or $\displaystyle u(x_0)=\inf_{x \in \Omega} u(x)$, then $u$ is a constant.
\end{itemize}
\label{prn:HD11_1}
\end{proposition}
\begin{proof}
The existence and uniqueness of the Dirichlet problem follows from Theorem 7.12 and Theorem 7.14 of Cheeger's work~\cite{Cheeger99},
where a much more general case was considered. The weak maximum principle follows from the uniqueness.
Also, the weak maximum principle was proved by Shanmugalligam in~\cite{ShanHar}.
The strong maximum principle follows from elliptic Harnack estimates, which is a consequence of the volume doubling and $(1,2)$-Poincar\'{e} inequality.
This is due to the work of K.T. Sturm in~\cite{Stu96}. The manifold case was obtained by A. Grigor'yan in~\cite{Grigor}, and L. Saloff-Coste in~\cite{Saloffnote}.
More information can be found in the beautiful survey~\cite{Saloffsur} by L. Saloff-Coste.
We write down an elementary proof here for the convenience of the readers, based on the excellent underlying geometry.
Here we follow Corollary 6.4 of~\cite{KiSh}.
Without loss of generality, we assume $u \geq 0$ on $\partial \Omega$ and $u$ is not a constant.
It suffices to show that $u>0$ in $\Omega$. Clearly, by classical harmonic function theory on Riemannian manifold and continuity of
$u$(c.f.~Proposition~\ref{prn:HD16_1}), it is clear that
$u>0$ on $\Omega \cap \mathcal{R}$. Therefore, we only need to show that $u>0$ on $\Omega \cap \mathcal{S}$.
We argue by contradiction. If this statement were wrong, we can find a point $y_0 \in \Omega \cap \mathcal{S}$ such that
$u(y_0)=0$. Choose $r$ small enough such that $B(y_0,2r) \subset \Omega$.
For each small $\epsilon$, choose $\tau$ small enough such that
\begin{align*}
\left| \Omega_{\tau} \cap B(y_0,2r) \right| < \epsilon |B(y_0,2r)|,
\end{align*}
where $\Omega_{\tau}=\{x \in \Omega| u(x) \leq \tau\}$. Note that $\tau$ can be chosen since $\Omega_0 \cap B(y_0,2r)$ is a subset of
$\mathcal{S}$ which has zero measure, and $u$ is continuous.
Now consider the function $\tau-u$, which is obviously harmonic. Let $(\tau-u)^{+}$ be $\max\{\tau-u, 0\}$. Then
$(\tau-u)^{+}$ is a bounded, continuous, subharmonic function in $N^{1,2}(B(y_0,2r))$. So Moser(or Nash-Moser-De-Giorgi) iteration applies to obtain
\begin{align*}
\sup_{B(y_0,r)} |(\tau-u)^{+}|^2 \leq C \int_{B(y_0,2r)} |(\tau-u)^{+}|^2=C \int_{\Omega_{\tau} \cap B(y_0,2r)} |(\tau-u)^{+}|^2 \leq C\epsilon r^{2n} \tau^2
\end{align*}
for some $C=C(n,\kappa)$. Choose $\epsilon$ small enough such that $C\epsilon^2 r^{2n}<\frac{1}{4}$. Then we have
\begin{align*}
\sup_{B(y_0,r)} |(\tau-u)^{+}| <\frac{\tau}{2},
\end{align*}
which implies that $u>\frac{\tau}{2}$ on $B(y_0,r)$. In particular, $u(y_0) \geq \frac{\tau}{2}>0$, which contradicts the assumption $u(y_0)=0$.
\end{proof}
Clearly, the essential stuff in the proof of the strong maximum principle of Proposition~\ref{prn:HD11_1} is a delicate use of elliptic Moser iteration.
In equation (\ref{eqn:HE03_2}), if we replace the operator $\Delta$ by $\square$, the heat operator, then one can easily obtain a strong maximum principle for heat equation solutions, based on a parabolic Moser iteration. The details are left to the interested readers.
On the other hand, if we replace the right hand side of equation (\ref{eqn:HE03_2}) by a function $h$, we can also obtain uniqueness and existence of solutions.
\begin{proposition}[\textbf{Existence and uniqueness of Poisson equation solution}]
Suppose $\Omega$ is a bounded open set of $X$, $f \in N^{1,2}(\Omega)$, $h \in L^2(\Omega)$.
Then there exists a unique $u \in N^{1,2}(\Omega)$ such that
\begin{align}
\Delta u=h, \quad \textrm{in} \; \Omega; \quad
(u-f)|_{\partial \Omega}=0.
\label{eqn:HE03_1}
\end{align}
\label{prn:HE03_1}
\end{proposition}
\begin{proof}
First, let us consider the Poisson equation
\begin{align*}
\Delta v=h, \; \textrm{in} \; \Omega; \quad v \in N_0^{1,2}(\Omega).
\end{align*}
By standard functional analysis, the existence of the above equation is guaranteed by Riesz representation theorem.
The uniqueness follows from the irreducibility of $\mathscr{E}$.
Second, it is obvious that there is a bijective map between the solution $u$ of (\ref{eqn:HE03_1})
and harmonic solution $w$ of (\ref{eqn:HE03_2}) by $u=w+v$. Therefore, the existence and uniqueness of
(\ref{eqn:HE03_1}) follows from Proposition~\ref{prn:HD11_1}.
\end{proof}
Combining the strong maximum principle for harmonic functions in Proposition~\ref{prn:HD11_1} with the heat kernel estimates, we obtain the
strong maximum principle for subharmonic functions.
\begin{proposition} [\textbf{Strong maximum principle for subharmonic functions}]
Suppose $\Omega$ is a bounded domain of $X$, $u$ is a continuous subharmonic function in $N^{1,2}(\Omega)$. Then we have
\begin{align}
\sup_{x \in \Omega} u(x) = \sup_{x \in \partial \Omega} u(x).
\label{eqn:HC29_2}
\end{align}
In other words, the weak maximum principle holds for subharmonic functions. Moreover, if $\displaystyle \sup_{x \in \Omega}u(x)$ is achieved at
some point $x_0 \in \Omega$, then $u$ is a constant. Namely, the strong maximum principle holds for subharmonic functions.
\label{prn:HC29_5}
\end{proposition}
\begin{proof}
The weak maximum principle is well known in literature.
For example, see Lemma 4 of~\cite{Stu94} and the reference therein.
The strong maximum principle can be proved as that in Proposition~\ref{prn:HD11_1}. Actually, $-w$ is a nonnegative superharmonic function on $\Omega$.
If $w$ is not a constant, then we can regard $-w$ as $u$ in the second part of the proof of Proposition~\ref{prn:HD11_1}.
Then everything goes through since only Moser iteration for subharmonic function is used there.
\end{proof}
\begin{proposition}[\textbf{Removing singularity of harmonic functions}]
Suppose $\Omega$ is an open domain in $X$, $u$ is a bounded harmonic function on $\Omega \backslash \mathcal{S}$.
Then $u$ can be regarded as a harmonic function on $\Omega$. Moreover, on each compact subset of $\Omega$, $u$ is uniformly Lipschitz continuous.
In particular, $u$ can be extended continuously over the singular set $\Omega \cap \mathcal{S}$.
\label{prn:HD16_1}
\end{proposition}
\begin{proof}
In light of Proposition~\ref{prn:HD16_2}, we see that $u \in N_{loc}^{1,2}(\Omega)$. Since $\Delta u=0$ on $\Omega \backslash \mathcal{S}$, we see that
\begin{align*}
\int_{\Omega} \langle \nabla u, \nabla \varphi \rangle=0
\end{align*}
for every smooth test function supported on $\Omega \backslash \mathcal{S}$. However, such functions are dense in $N_0^{1,2}(\Omega)$, so the above
equation actually holds form every $\varphi \in N_c^{1,2}(\Omega)$. Therefore, $u$ is harmonic in $\Omega$ by definition.
The Lipschitz continuity follows from Proposition~\ref{prn:HC29_6} and the density and weak convexity of $\mathcal{R}$.
\end{proof}
Note that the weak convexity of $\mathcal{R}$ is important for that $u$ can be extended over singularities. For otherwise, the limit of $u(x_i)$ for $x_i \to x_0$ may depends
on the choice of sequence $\{x_i\}$, where $x_0 \in \mathcal{S}, x_i \in \mathcal{R}$. If $\mathcal{R}$ is not convex, there is an easy counter example
of Proposition~\ref{prn:HD16_1}. Let $X$ be the union of two cones $C(S^3/\Gamma)$, for some finite group $\Gamma \in ISO(S^3)$, by identifying two vertices.
In this case, $\mathcal{S}$ is the isolated vertex $O$. Let $u$ be $1$ on one branch and $0$ on the other, then it is clear that $u$ is a harmonic function
on $X \backslash \mathcal{S}$. However, $u$ can not take a value at $O$ so that $u$ is continuous. Of course, convexity of $\mathcal{R}$ is only a sufficient
condition to guarantee the continuity extension. It can be replaced by other weaker conditions.
Moreover, based on Proposition~\ref{prn:HC29_6}, one can obtain uniform gradient estimate of $|\nabla p(t,\cdot, x_0)|$, which depends only
on $t,n$ and $\kappa$. Hence the heat kernel $p(t,\cdot, x_0)$ is a continuous function on $X \times (0,\infty)$.
Therefore, by approximation, the estimate in Proposition~\ref{prn:HD07_1} holds on every point on
$X$, even if this point is singular.
\subsection{Approximation functions of distance}
\begin{proposition}[\textbf{Almost super-harmonicity of distance function}] Suppose $x_0 \in X$, $r(x)=d(x,x_0)$. Then we have
\begin{align}
\Delta r^2 \leq 4n
\label{eqn:HC29_1}
\end{align}
in the weak sense. In other words, for every nonnegative $\chi \in N_c^{1,2}(X)$, we have
\begin{align}
-\int_X \left\langle \nabla r^2, \nabla \chi \right\rangle \leq \int_X 4n \chi.
\label{eqn:HD16_3}
\end{align}
\label{prn:HC29_1}
\end{proposition}
\begin{proof}
Let us first assume $x_0 \in \mathcal{R}$.
Clearly, away from the generalized cut locus, we have
$\Delta r^2 \leq 4n$ in the classical sense. Therefore, $\Delta r^2 \leq 4n$ on $\mathcal{R}$ in the distribution sense,
same as the smooth Riemannian manifold case. Since smooth cutoff functions supported on $\mathcal{R}$
are dense in $N_0^{1,2}(X)$(c.f. Corollary~\ref{cly:HE10_1} and Proposition~\ref{prn:HD04_1}, where $codim(\mathcal{S})>2$ is essentially used),
we see that for every $\chi \in N_c^{1,2}(X)$, inequality (\ref{eqn:HD16_3}) holds true.
Now suppose $x_0 \in \mathcal{S}$, we can choose regular points $x_i \to x_0$. Let $r_i=d(x_i, \cdot)$, then for each nonnegative function $\chi \in N_c^{1,2}(X)$, we have
\begin{align*}
-\int_X \left\langle \nabla r_i^2, \nabla \chi \right\rangle \leq \int_X 4n \chi.
\end{align*}
Let $\Omega$ be a bounded open set containing the support of $\chi$.
Then $r_i^2$ weakly converges to a unique limit in $N^{1,2}(\Omega)$,
$r_i^2$ strongly converges to $r^2$ in $L^2(\Omega)$. This means that
$r^2$ is the weak limit of $r_i^2$ in $N^{1,2}(\Omega)$. It follows that
\begin{align*}
-\int_X \left\langle \nabla r^2, \nabla \chi \right\rangle
=-\lim_{i \to \infty} \int_X \left\langle \nabla r_i^2, \nabla \chi \right\rangle
\leq \int_X 4n \chi.
\end{align*}
\end{proof}
In view of Proposition~\ref{prn:HC29_1}, we can obtain many rigidity theorems.
\begin{lemma}[\textbf{Cheeger-Gromoll type splitting}]
Suppose $X$ contains a straight line $\gamma$. Then there is a length space $N$ such that
$X$ is isometric to $N \times \ensuremath{\mathbb{R}}$ as metric space product.
\label{lma:HE04_1}
\end{lemma}
\begin{proof}
The proof is almost the same as the classical one. However, we shall take this as an opportunity to check the analysis tools developed in previous subsections.
Actually, fix $x_0 \in \gamma$, we can divide $\gamma$ into two rays $\gamma^{+}$ and $\gamma^{-}$.
Accordingly, there are Buseman functions $b^{+}$ and $b^{-}$. Proposition~\ref{prn:HC29_1} implies that $\Delta r \leq \frac{2n-1}{r}$ in the weak sense,
which in turn forces both $b^{+}$ and $b^{-}$ to be subharmonic functions. By triangle inequality, we know $b^{+}+b^{-} \geq 0$ globally and
achieve $0$ on $x_0$. It follows from strong maximum principle, by Proposition~\ref{prn:HC29_5}, that $b^{+}+b^{-} \equiv 0$.
Consequently, $b^{+}$ is harmonic. Then Weitzenb\"{o}ck formula implies that in the weak sense, we have
\begin{align*}
0=\frac{1}{2}\Delta |\nabla b^{+}|^2= |Hess_{b^{+}}|^2.
\end{align*}
Since $b^{+}$ is harmonic, it is harmonic on $\mathcal{R}=X \backslash \mathcal{S}$. By standard improving regularity theory of harmonic functions on
smooth manifold, we see that $b^{+}$ is a smooth function on $\mathcal{R}$ satisfying
\begin{align*}
|\nabla b^{+}|^2 \equiv 1, \quad |Hess_{b^{+}}|^2 \equiv 0.
\end{align*}
Up to this step, everything is the same as the classical case.
However, since the regular part $\mathcal{R}$ is not complete, the following argument is slightly different.
On $\mathcal{R}$, since $\mathcal{L}_{\nabla b^{+}} g=2Hess_{b^{+}}=0$, we see that $\nabla b^{+}$ is a Killing field.
The flow generated by $\nabla b^{+}$ preserves metrics, and in particular the volume element.
By the high codimension of $\mathcal{S}$, weak convexity of $\mathcal{R}$ and the essential gap of volume density between regular and singular points, one can obtain that the flow generated by $\nabla b^{+}$ preserves regularity.
The full details will be explained as follows.
Let $\varphi_t$ be the time $t$ flow map generated by $\nabla b^{+}$ when it is well defined. In other words, we have $\frac{d}{dt} \varphi_t(x)=\left. \nabla b^{+} \right|_{\varphi_t(x)}$ whenever $\varphi_t(x) \in \mathcal{R}$.
\begin{claim}[\textbf{Existence of flows away from small sets}]
For each fixed $x_0 \in X$ and $A>0$, there is a set $E_A$ such that $\varphi_t(x)$ exists and locates in $\mathcal{R}$ for all
$x \in B(x_0, A) \backslash E_A$ and $t \in [-A, A]$. Moreover, we have
\begin{align}
\dim_{\mathcal{M}} E_A \leq \dim_{\mathcal{M}} \mathcal{S} +1<2n-2. \label{eqn:MC14_4}
\end{align}
Consequently, there is a measure-zero set $E$ such that $\varphi_t(x)$ exists and locates in $\mathcal{R}$ for all $x \in X \backslash E$ and $t \in (-\infty, \infty)$.
\label{clm:MA20_1}
\end{claim}
Fix $x_0 \in X$ and choose $\xi$ to be a very small positive number, $A$ to be a large positive number.
Let $q_0$ be the Minkowski codimension of $\mathcal{S}$, i.e., $q_0=2n-\dim_{\mathcal{M}} \mathcal{S}$,
$\epsilon$ be a very small number to be used in the volume estimate related to Minkowski codimension.
Note that $d(\cdot, \mathcal{S})$ is a Lipshitz function with Lipshitz constant $1$.
By perturbing $d(\cdot, \mathcal{S})$, we can find a smooth hyper surface $\Sigma_{\xi}$ (c.f. Corollary~\ref{cor:MA20_1} for more details)
such that
\begin{align*}
&\left| B(x_0, 3 A) \cap \Sigma_{\xi} \right|_{\mathcal{H}^{2n-1}} \leq C \xi^{q_0-\epsilon-1}, \\
& \frac{1}{H} \xi <d(x, \mathcal{S})< H \xi, \quad \forall \; x \in \Sigma_{\xi} \cap B(x_0, 2A).
\end{align*}
Note that the $\xi^{q_0-\epsilon}$ in the first inequality comes from the fact that $\dim_{\mathcal{M}}\mathcal{S}=2n-q_0$ and the application of co-area formula.
The constant $C$ in the first inequality depends both on $\epsilon$ and the set $B(x_0, 3A)$.
The constant $H$ in the second inequality depends on $\kappa, n$ and comes from the perturbation technique. $H$ will be fixed in the following discussion.
$C$ may vary from line to line, as usual.
Define a set
\begin{align}
E_{A,\xi} \triangleq \left\{ x\in \overline{B(x_0, A)} \left| d\left( \varphi_t(x), \mathcal{S} \right) \leq \frac{1}{H} \xi, \quad \textrm{for some} \; t \in [-A, A] \right. \right\}.
\label{eqn:MA20_0}
\end{align}
Now we decompose $E_{A,\xi}$ into two parts $I$ and $II$ as follows.
\begin{align*}
I= \{x \in \overline{B(x_0, A)} | d(x, \mathcal{S}) \leq 10 \xi\} \cap E_{A,\xi}, \quad
II=\{x \in \overline{B(x_0, A)} | d(x, \mathcal{S}) >10 \xi \} \cap E_{A,\xi}.
\end{align*}
By the Minkowski dimension assumption of $\mathcal{S}$, we know that
\begin{align}
|I| \leq \left| \left\{ \left. x \in \overline{B(x_0, A)} \right| d(x, \mathcal{S}) \leq 10 \xi \right\} \right| \leq C \xi^{q_0-\epsilon},
\label{eqn:MA20_1}
\end{align}
where $|\cdot|$ means the $2n$-dimensional Hausdorff measure.
Taking every point $y \in II$, one can flow it to a point on $\Sigma_{\xi}$ at some time $t \in [-A, A]$ by the definition of $E_{A,\xi}$.
Since $|\nabla b^{+}|=1$, it is clear that $d(y, \varphi_t(y)) \leq |t|$. Then triangle inequality implies that
\begin{align*}
d(x_0, \varphi_t(y)) < d(x_0, y) + d(y, \varphi_t(y)) \leq A + |t| \leq 2A < 3A
\end{align*}
Therefore, the set $II$ can be locally regarded as a bundle over $\Sigma_{\xi} \cap B(x_0, 3A)$. Note that along the flow line of the Killing field $\nabla b^+$, $\varphi_t$ preserves local isometry.
We equip the set $\left\{\Sigma_{\xi} \cap B(x_0, 3A) \right\} \times [-A, A]$ with the obvious product measure. Consider the map
\begin{align*}
\varphi: \Omega \subset \left\{\Sigma_{\xi} \cap B(x_0, 3A) \right\} \times [-A, A] &\mapsto X, \\
(x,t) &\mapsto \varphi_t(x).
\end{align*}
Here $\Omega$ is the maximal subset of $\left\{\Sigma_{\xi} \cap B(x_0, 3A) \right\} \times [-A, A]$ such that $\varphi(x,t)=\varphi_t(x)$ is well defined.
It is clear that $\varphi$ decrease volume whenever the flow line is not perpendicular to $\Sigma_{\xi}$. It follows that
\begin{align*}
|II| \leq |\Omega| \leq \left| \left\{\Sigma_{\xi} \cap B(x_0, 3A) \right\} \right|_{\mathcal{H}^{2n-1}} \cdot 3A \leq CA \xi^{q_0-\epsilon-1}.
\end{align*}
Combining the above inequality with (\ref{eqn:MA20_1}), we have
\begin{align}
|E_{A,\xi}| \leq |I|+|II| \leq C \xi^{q_0-\epsilon} + C A \xi^{q_0-\epsilon-1} \leq C\xi^{q_0-\epsilon-1}, \label{eqn:MA20_3}
\end{align}
where the last $C$ depends on $n,\kappa$ and $B(x_0, 3A)$.
We observe that
\begin{align}
\left\{ x \left| d(x, E_{A, \xi}) < H^{-1}\xi \right. \right\} \subset E_{2A, 2\xi}. \label{eqn:MC14_0}
\end{align}
In fact, if $x$ locates in the $H^{-1}\xi$-neighborhood of $E_{A, \xi}$, then we can find a point $y \in E_{A, \xi}$ such that
$d(x, E_{A, \xi})=d(x,y)=\delta<H^{-1}\xi$. So we can find a shortest geodesic connecting $x$ to $y$ satisfying
$\gamma(0)=x$ and $\gamma(\delta)=y$, with $|\gamma|=\delta$.
Note that we can assume $\gamma$ is a smooth geodesic. For otherwise, we have
$d(x, \mathcal{S})\leq \delta <H^{-1}\xi$, which automatically implies $x \in E_{2A, 2\xi}$ by the definition in (\ref{eqn:MA20_0}).
For the same reason, triangle inequality guarantees us to assume
\begin{align}
d(\gamma, \mathcal{S})>H^{-1}\xi. \label{eqn:MC14_1}
\end{align}
As $y \in E_{A, \xi}$, following its definition in (\ref{eqn:MA20_0}), we can find a $t_0 \in [-A, A]$ such that
\begin{align}
d(\varphi_{t_0}(y), \mathcal{S}) \leq H^{-1}\xi. \label{eqn:MC14_2}
\end{align}
Let $s_0$ be the smallest positive value such that $d(\varphi_{s_0}(\gamma), \mathcal{S}) \leq H^{-1}\xi$.
The combination of (\ref{eqn:MC14_1}) and (\ref{eqn:MC14_2}) then yields that $0<|s_0| \leq |t_0|$.
Note that $\varphi_{s}(\gamma)$ is well defined on $(-|s_0|, |s_0|)$. So the length of $\varphi_s(\gamma)$ is the same as the length of $\gamma$
for each $s \in (-|s_0|, |s_0|)$.
Now we are ready to estimate the distance
between $\varphi_{s_0}(x)$ and the singular set $\mathcal{S}$:
\begin{align*}
d(\varphi_{s_0}(x), \mathcal{S}) \leq |\varphi_{s_0}(\gamma)| + d(\varphi_{s_0}(\gamma), \mathcal{S})
=|\gamma| + d(\varphi_{s_0}(\gamma), \mathcal{S})
<H^{-1} \xi + H^{-1} \xi=\frac{2\xi}{H}.
\end{align*}
Note that $E_{A, \xi} \subset \overline{B(x_0, A)}$. Triangle inequality implies that $x \in \overline{B(x_0, 2A)}$.
Recall also that $s_0 \in [-A, A] \subset [-2A, 2A]$. In light of definition equation (\ref{eqn:MA20_0}), the above inequality
implies that $x \in E_{2A, 2\xi}$. So we finish the proof of (\ref{eqn:MC14_0}).
Let $\xi=\xi_i \to 0$ and define
\begin{align}
E_{A} \triangleq \cap_{i=1}^{\infty} E_{A,\xi_i}. \label{eqn:MC14_3}
\end{align}
We obtain a set $E_{A} \subset \overline{B(x_0, A)}$ such that $\varphi_t(x) \in \mathcal{R}$ for each point
$x \in \overline{B(x_0, A)} \backslash E_{A}$ and $t \in [-A,A]$.
Furthermore, from defintion equation (\ref{eqn:MC14_3}), it is clear that $E_A \subset E_{A, \xi}$, which together with (\ref{eqn:MC14_0})
implies that the $H^{-1}\xi$ neighborhood of $E_A$ is contained in $E_{2A, 2\xi}$.
Consequently, it follows from (\ref{eqn:MA20_3}) (replacing $A$ by $2A$, $\xi$ by $2\xi$) that
\begin{align*}
\left| \left\{ x \left | d(x, E_A) < H^{-1} \xi \right.\right\} \right|_{\mathcal{H}^{2n}} \leq C \xi^{q_0-\epsilon-1}=C \left( H^{-1} \xi \right)^{q_0-\epsilon-1},
\end{align*}
where the last $C$ depends on $n,\kappa, H, \epsilon$ and the set $B(x_0, 3A)$, but independt of the choice of $\xi$.
Since $\epsilon$ can be any small number, the above inequality implies (\ref{eqn:MC14_4}) by the definition of
Minkowski dimension(c.f. Definition~\ref{dfn:HE08_1}).
Then we set $A=A_i \to \infty$ and define $E \triangleq \cup_{i=1}^{\infty} E_{A_i}$. As a union of countably many measure-zero sets, $E$ is clearly a measure-zero set.
Clearly, for each $x \in X \backslash E$, and each $t \in (-\infty, \infty)$, we can always find a large $A_i$ such that $x \in B(x_0, A_i)$ and $t \in (-A_i, A_i)$.
Then it follows that $\varphi_t(x)$ exists and locates in $B(x_0, 2A_i) \cap \mathcal{R} \subset \mathcal{R}$.
So we finish the proof of Claim~\ref{clm:MA20_1}.
\begin{claim}[\textbf{Flow lines preserve regularity}]
Suppose $x_0$ is a regular point, then the whole flow line of $\nabla b^+$ initiated from $x_0$ is defined for all time and stays in $\mathcal{R}$.
\label{clm:GD11_1}
\end{claim}
Suppose Claim~\ref{clm:GD11_1} was wrong. Without loss of generality, we can assume that the flow image of $x_0$ hits singularity the first time at $T_0>0$.
Let $\varphi_{t}$ be the time $t$ flow map generated by $\nabla b^{+}$. Fix $\epsilon$ very small, then $y_0=\varphi_{T_0-\epsilon}(x_0)$ is a regular point.
By Claim~\ref{clm:MA20_1}, we see that $\varphi_{T_0-\epsilon}$ is well defined away from a measure-zero set $E$.
Furthermore, it preserves volume element and length element. For each $r \in (0,1)$, we define $\chi_r$ as
\begin{align*}
\chi_r(x)=
\begin{cases}
r-d(x,x_0), & \textrm{if} \; d(x,x_0)<r \\
0, &\textrm{if} \; d(x,x_0) \geq r.
\end{cases}
\end{align*}
The function $f_{r,\epsilon}=\chi_r \circ \varphi_{T_0-\epsilon}^{-1}$ is defined on $X \backslash E$. Actually, by choosing $A >>T_0+1$, we know that
$f_{r,\epsilon}$ is defined on $B(y_0, A) \backslash E_{A}$ and vanishes outside $B(y_0, 0.5A)$. This is a simple application of triangle inequality.
If $x \in B(y_0, A) \backslash E_{A}$ and $d(x,y_0)>0.5A>5(T_0+1)$, then we have
\begin{align*}
d(\varphi_{T_0-\epsilon}^{-1}(x), x_0) &\geq -d(\varphi_{T_0-\epsilon}^{-1}(x), x) + d(x, x_0)
\geq -d(\varphi_{T_0-\epsilon}^{-1}(x), x) -d(y_0, x_0)+d(x, y_0)\\
&\geq -2(T_0-\epsilon) + 5 (T_0+1)> 3T_0+5>r.
\end{align*}
Then $\chi_r(\varphi_{T_0-\epsilon}^{-1}(x))=0$ by definition of $\chi_r$.
By the local isometry property of $\varphi_{T_0-\epsilon}$, we obtain that $|\nabla (\chi_r \circ \varphi_{T_0-\epsilon}^{-1})| \leq 1$ on $B(y_0, A) \backslash E_{A}$.
Recall that $E_A$ has codimension at least $2$ by (\ref{eqn:MC14_4}). Then it is clear that $f_{r,\epsilon} \in N_0^{1,2}(X)$.
Note that $f_{r,\epsilon}$
has a version $\tilde{f}_{r,\epsilon}$ which is globally Lipshitz with Lipshitz constant $1$. In other words, one can find a measure-zero set $F$ such that $\tilde{f}_{r,\epsilon}=f_{r,\epsilon}$
on $X \backslash F$. The function $\tilde{f}_{r,\epsilon}$ can be obtained as follows. Let $f_{r,\epsilon}(t)$ be the heat flow solution initiated from $f_{r,\epsilon} \in N^{1,2}(X)$.
Then $|\nabla f_{r,\epsilon}(t)|$ is a heat subsolution.
Note that here we used the condition $\dim_{\mathcal{M}} \mathcal{S}<2n-3$ and the weak convexity of $\mathcal{R}$, to guarantee that $|\nabla f_{r,\epsilon}(t)| \in N_{loc}^{1,2}(X)$.
Further details can be found in Appendix~\ref{app:A}.
By maximum principle(on the space possibly has singularities), we see that $|\nabla f_{r,\epsilon}(t)| \leq 1$ for each
$t>0$. Let $t_i \to 0$, then the limit of $f_{r,\epsilon}(t_i)$ can be chosen as $\tilde{f}_{r,\epsilon}$.
Under the help of $\tilde{f}_{r,\epsilon}$, we shall see that $\varphi_{T_0-\epsilon}$ is an isometry from $X \backslash E$ to $X \backslash E$, by further ajusting $E$ with an extra measure-zero set if necessary.
Actually, if $x \in \partial B(x_0,r) \backslash E$, then $y=\varphi_{T_0-\epsilon}(x)$ is a regular point. Note that $\tilde{f}_{r,\epsilon}(y_0)=r$
and $\tilde{f}_{r,\epsilon}(y)=0$. Since $\tilde{f}_{r,\epsilon}$ has uniform global Lipshitz constant $1$, we have $d(y,y_0) \geq r$.
Therefore, $\varphi_{T_0-\epsilon}$ is a distance-expanding map from $X \backslash E$ to $X \backslash E$. By reversing the position of $x_0,y_0$ and using $-\nabla b^{+}$ to generate flow, it is clear that
$\varphi_{T_0-\epsilon}$ is distance-shrinking. Combining these two directions, we obtain
\begin{align*}
\varphi_{T_0-\epsilon}^{-1} \left( B(y_0, r) \backslash E \right) = B(x_0, r) \backslash E.
\end{align*}
In particular, we see that
\begin{align*}
|B(x_0,r)|=|B(y_0,r)|=|B(\varphi_{T_0-\epsilon}(x_0), r)|.
\end{align*}
Using triangle inequality and letting $\epsilon \to 0$, we have $|B(x_0,r)|=|B(\varphi_{T_0}(x_0), r)|$ for each $r \in (0, 1)$.
However, $x_0$ is regular. For some $r \in (0, 1)$, we have $\omega_{2n}^{-1}r^{-2n}|B(x_0,r)|>1-\frac{\delta_0}{100}$.
Same volume ratio estimate hold for the ball $B(\varphi_{T_0}(x_0),r)$.
Therefore, $\varphi_{T_0}(x_0)$ is forced to be a regular point by Anderson's gap theorem(c.f. Corollary~\ref{cly:SL23_1}).
This contradicts the assumption of $T_0$. Therefore, the proof of Claim~\ref{clm:GD11_1} is complete. \\
Let $N$ be the level set $b^{+}=0$, $N'=N \cap \mathcal{R}$. Then it is clear that
$\mathcal{R}=N' \times \ensuremath{\mathbb{R}}$. Taking metric completion on both sides, we obtain $X= N \times \ensuremath{\mathbb{R}}$ as metric space product.
\end{proof}
Lemma~\ref{lma:HE04_1} should be a special case of Gigli~\cite{Gigli13}. Its local version is the following lemma.
\begin{lemma}[\textbf{Metric cone rigidity}]
Suppose $x_0 \in X$, $\Omega=B(x_0,1)$. Then the following conditions are equivalent.
\begin{enumerate}
\item Volume ratio same on scales $0.5$ and $1$, i.e., we have
\begin{align}
2^{2n}|B(x_0,0.5)|=|B(x_0,1)|. \label{eqn:HE10_1}
\end{align}
\item $\Omega$ is a volume cone, i.e., for every $0<r_1<r_2<1$, we have
\begin{align}
r_1^{-2n}|B(x_0,r_1)|=r_2^{-2n}|B(x_0,r_2)|. \label{eqn:HE04_1}
\end{align}
\item $\frac{r^2}{2}$ is the unique weak solution of the Poisson equation
\begin{align}
\Delta u=2n, \; \textrm{in} \; \Omega; \quad \left.\left(u-\frac{r^2}{2} \right)\right|_{\partial \Omega}=0.
\label{eqn:HE03_4}
\end{align}
\item $\frac{r^2}{2}$ induces local metric cone structure on $\Omega$. In other words, on $\Omega \backslash \mathcal{S}$, we have
\begin{align*}
Hess_{\frac{r^2}{2}} -g \equiv 0.
\end{align*}
\end{enumerate}
\label{lma:HE04_2}
\end{lemma}
\begin{proof}
\textit{1 $\Rightarrow$ 2:}
Let $A(r)$ be the ``area" ratio function in Corollary~\ref{cly:GD08_1}, i.e., $ A(r)= \frac{|\partial B(x_0,r)|}{r^{2n-1}}$.
By Corollary~\ref{cly:GD08_1}, we know $A(r)$ is defined almost everywhere and is non-increasing on its domain.
Note that
\begin{align*}
\frac{d}{dr} \left( \frac{|B(x_0,r)|}{r^{2n}}\right)=\frac{|\partial B(x_0,r)|}{r^{2n}} -\frac{2n}{r} \frac{|B(x_0,r)|}{r^{2n}}
=\frac{2n}{r} \left\{\frac{A(r)}{2n} -\frac{|B(x_0,r)|}{r^{2n}} \right\} \leq 0.
\end{align*}
Combining (\ref{eqn:HE10_1}) and (\ref{eqn:GD08_5}), we have $ \frac{A(r)}{2n}-\frac{|B(x_0,r)|}{r^{2n}} \equiv 0$ for a.e. $r \in (0.5, 1)$.
In particular, we have
\begin{align*}
|B(x_0,1)|=\frac{A(1)}{2n}=\int_0^{1} A(1) r^{2n-1}dr,
\end{align*}
where $A(1)$ is understood as $\displaystyle \lim_{r \to 1^{-}} A(r)$.
On the other hand, it follows from (\ref{eqn:GD08_4}) that
\begin{align*}
|B(x_0,1)|=\int_0^{1} A(r) r^{2n-1} dr.
\end{align*}
So we have
\begin{align*}
\int_0^{1} (A(r)-A(1)) r^{2n-1} dr=0.
\end{align*}
Note that $A(r)$ is a non-increasing function. So the above equality means that
\begin{align*}
A(1) \equiv A(r) \equiv \lim_{r \to 0^{+}} A(r).
\end{align*}
It follows that $|B(x_0,r)| = \frac{A(1)}{2n}r^{2n}$ for every $0<r<1$. In particular, $B(x_0,1)$ is a volume cone.
\textit{2 $\Rightarrow$ 3:} Suppose $u$ is the unique solution of the Poisson equation (\ref{eqn:HE03_4}), we need to show that
$u\equiv \frac{r^2}{2}$. By uniqueness of weak solutions, it suffices to show that $\frac{r^2}{2}-u$ is harmonic on $\Omega$, i.e., for every
$\varphi \in N_c^{1,2}(\Omega)$, we have $\int_{\Omega} \varphi \Delta \left( \frac{r^2}{2}-u\right)=0$.
By rescaling, we can also assume $0 \leq \varphi \leq 1$.
Fix such a $\varphi$, we can choose $\epsilon$ small such that the support of $\varphi$ is contained in $B(x_0, 1-\epsilon)$. Define
\begin{align*}
\eta(x)=
\begin{cases}
1, & \textrm{if} \quad d(x,x_0)<1-\epsilon, \\
\frac{1-d(x,x_0)}{\epsilon}, & \textrm{if} \quad 1-\epsilon \leq d(x,x_1) \leq 1.
\end{cases}
\end{align*}
Note that $\eta \in N_0^{1,2}(\Omega)$, $\frac{r^2}{2}-u$ is superharmonic on $\Omega$. It follows from integration by parts that
\begin{align*}
\int_X \eta \Delta\left( \frac{r^2}{2} -u\right) =-2n\int_{\Omega} \eta +\frac{1}{\epsilon} \int_{B(x_0,1) \backslash B(x_0,1-\epsilon)} r \geq -O(\epsilon),
\end{align*}
where we used volume cone condition in the last step.
Thus, we have
\begin{align*}
0 \geq \int_{\Omega} \varphi \Delta \left( \frac{r^2}{2}-u\right)
=\int_{\Omega} \eta \Delta \left( \frac{r^2}{2}-u\right)+\int_{\Omega} (\varphi-\eta) \Delta \left( \frac{r^2}{2}-u\right)
\geq \int_{\Omega} \eta \Delta \left( \frac{r^2}{2}-u\right) \geq -O(\epsilon).
\end{align*}
Let $\epsilon \to 0$, we obtain $\int_{\Omega} \varphi \Delta \left( \frac{r^2}{2}-u\right)=0$. Consequently, $\frac{r^2}{2}-u$ is harmonic by the arbitrary choice
of $\varphi$.
\textit{3 $\Rightarrow$ 4:}
Since $\frac{r^2}{2}$ solves the Poisson equation with right hand side a constant, by standard bootstrapping argument for elliptic equation,
we see that $\frac{r^2}{2}$ is a smooth function on $\Omega \backslash \mathcal{S}$.
Clearly, we have $\left|\nabla \frac{r^2}{2}\right|^2=\frac{r^2}{2}$. Taking Laplacian on both sides, Weitzenb\"{o}ck formula yields that
$\displaystyle \left|Hess_{\frac{r^2}{2}} \right|^2=2n$, which in turn implies that
\begin{align*}
\left|Hess_{\frac{r^2}{2}} -g\right|^2= \left|Hess_{\frac{r^2}{2}} \right|^2-2\Delta \frac{r^2}{2}+2n
= \left|Hess_{\frac{r^2}{2}} \right|^2-2n=0.
\end{align*}
Therefore, on $\Omega \backslash \mathcal{S}$, we have $Hess_{\frac{r^2}{2}} -g\equiv 0$ in the classical sense.
Consequently, $\nabla \frac{r^2}{2}$ is a conformal Killing field.
Similar to the proof of Claim~\ref{clm:GD11_1}, one can show that the flow generated by $\nabla \frac{r^2}{2}$ preserves regularity.
Hence it is clear that $\Omega \backslash \mathcal{S}$ has a local metric cone structure, whose completion implies that $\overline{\Omega}$ is a unit ball in a metric cone.
\textit{4 $\Rightarrow$ 1:}
For each $0<r<1$, note that $|B(x_0,r)|=|B(x_0,r) \backslash \mathcal{S}|$. Note the flow generated by $\nabla \frac{r^2}{2}$ preserves regularity.
More precisely, we have
\begin{align*}
\mathcal{L}_{\nabla \frac{r^2}{2}} g=g, \quad \mathcal{L}_{\nabla \frac{r^2}{2}} d\mu= 2n d\mu.
\end{align*}
Then (\ref{eqn:HE04_1}) follows from the integration of the above equation along flow lines.
\end{proof}
\begin{lemma}[\textbf{K\"ahler cone splitting}]
Suppose $X \in \widetilde{\mathscr{KS}}^*(n,\kappa)$ is a metric cone with vertex $x_0$. Then we can find a metric cone $C(Z)$ with vertex $z^*$ such that
\begin{align*}
X= \ensuremath{\mathbb{C}}^{n-k} \times C(Z), \quad x_0=(0, z^*), \quad 2 \leq k \leq n.
\end{align*}
Moreover, there is no straight line in $C(Z)$ passing through $z^*$.
\label{lma:HD30_1}
\end{lemma}
\begin{proof}
It suffices to show that if $X$ splits off a real straight line $\ensuremath{\mathbb{R}}$, then it splits off a complex line $\ensuremath{\mathbb{C}}$.
In fact, if there is a straight line passing through $x_0$, we can find a function $h$ which is the Buseman function determined by the line.
Therefore, $\nabla h$ is a parallel vector field with $|\nabla h| \equiv 1$. The K\"ahler condition implies that
$J \nabla h$ is another parallel vector field satisfying $|J \nabla h| \equiv 1$ on $\mathcal{R}=X \backslash \mathcal{S}$.
On the regular set, define function
\begin{align*}
u=\left\langle J\nabla h, \nabla \frac{r^2}{2} \right\rangle,
\end{align*}
where $r$ is the distance to the vertex $x_0$.
Metric cone condition implies that $Hess_{\frac{r^2}{2}}=g$. Since $J\nabla h$ is parallel, we see that
\begin{align*}
\nabla u= Hess_{\frac{r^2}{2}} (J\nabla h, \cdot)=J \nabla h.
\end{align*}
Recall that $Hess_h \equiv 0$. Taking gradient of the above equation implies that $Hess_u \equiv 0$. This forces that $\nabla u=J \nabla h$
is also a splitting direction. Note that although $u$ is only defined on $\mathcal{R}$, which is not complete.
However, we can bypass this difficulty as done in the proof of Lemma~\ref{lma:HE04_1}, since $\nabla u$ is a Killing field preserving regularity.
Therefore, we obtain a splitting factor $\ensuremath{\mathbb{C}}$. Since $J\nabla u=-\nabla h$, the space spanned by $\nabla u$ and
$J \nabla u$ is closed under the $J$-action. This induces the $J$-action closedness of the split linear space, which then must be $\ensuremath{\mathbb{C}}^{n-k}$
for some integer $k$. Because $X$ is not $\ensuremath{\mathbb{C}}^n$, we know the singular set is not empty, whose dimension restriction forces that $k \geq 2$.
\end{proof}
For each rigidity property in Lemma~\ref{lma:HE04_1}-Lemma~\ref{lma:HD30_1}, there should exist an ``almost" version.
For example, Lemma~\ref{lma:HE04_2} basically says that a volume cone implies a metric cone.
Hence the ``almost" version is that for a unit geodesic ball $B(x_0,1)$ whose volume ratio function $r^{-2n}|B(x_0,r)|$ is very close to a constant
function on $[0,1]$, then after proper rescaling, each ball $B(x_0, r)$ is very close to $B(x_0,1)$ in the Gromov-Hausdorff topology.
The basic idea is expressed clearly in~\cite{CCWarp}.
We only interpret what they did. Actually, if volume ratio is almost a constant, then it is expected that $|Hess_{\frac{r^2}{2}}-g|$ has a small
$L^2$-norm. However, since the regularity of distance function $r$ is bad, one should replace $\frac{r^2}{2}$ by an approximation function, which is
very close to $\frac{r^2}{2}$ in $N^{1,2}$-norm on one hand, and has excellent regularity on the other hand.
Such approximation function is nothing but the solution of the Poisson equation (\ref{eqn:HE03_4}).
For the purpose of developing ``almost" rigidity properties, one need some technical preparation, which will be listed as Lemmas.
Note that the space $\widetilde{\mathscr{KS}}(n,\kappa)$ has scaling invariance.
Therefore, we can always let the scale we are interested in to be $1$, to simplify the notations.
In view of Proposition~\ref{prn:HC29_1}, we can define many auxiliary radial functions, as in the classical case for Riemannian manifold(c.f.~\cite{CCWarp}).
For each $0<r<R<\infty$, define
\begin{align}
&\underline{U}(r) \triangleq \frac{r^2}{4n},
\quad
\underline{G}(r) \triangleq \frac{r^{2-2n}}{2n(2n-2)\omega_{2n}}, \\
&\underline{U}_R \triangleq \frac{r^2-R^2}{4n}, \quad \underline{G}_R \triangleq \frac{r^{2-2n}-R^{2-2n}}{2n(2n-2)\omega_{2n}}, \\
&\underline{L}_R \triangleq \frac{r^{2-2n}R^{2n}-R^{2}}{2n(2n-2)} + \frac{r^2-R^2}{4n}.
\label{eqn:HE08_1}
\end{align}
Then by Proposition~\ref{prn:HC29_1} and direct calculation, we have the following lemma.
\begin{lemma}[\textbf{Existence of good radial comparison functions}]
Suppose $x_0 \in X$. Let $r(x)=d(x,x_0)$ and
define $\underline{U}_1(x)=\underline{U}_1(r(x))$, $\underline{G}_1=\underline{G}_1(r(x))$ and $\underline{L}_1(x)=\underline{L}(r(x))$
as done in (\ref{eqn:HE08_1}). Then we have
\begin{align*}
&\Delta \underline{U}_1 \leq 1, \textrm{on} \; X; \quad \underline{U}_1|_{\partial B(x_0,1)}=0. \\
&\Delta \underline{G}_1 \geq 0, \textrm{on} \; B(x_0,1) \backslash \{x_0\}; \quad \underline{G}_1 |_{\partial B(x_0,1)}=0. \\
&\Delta \underline{L}_1 \geq 0, \textrm{on} \; B(x_0,1) \backslash \{x_0\}; \quad \underline{L}_1 |_{\partial B(x_0,1)}=0.
\end{align*}
\label{lma:HE08_1}
\end{lemma}
Lemma~\ref{lma:HE08_1} is used to improve the maximum principle. Same as that done by Abresch-Gromoll (c.f. Proposition 2.3 of~\cite{AbGr}), we
obtain the following estimate of excess function.
\begin{lemma}[\textbf{Abresch-Gromoll type estimate}]
Suppose $x_0 \in X$,
$\gamma$ is a line segment centered at $x_0$ with length $2$, end points $p_{+}$ and $p_{-}$.
Let $e(x)$ be the excess function $d(x,p_{+})+d(x,p_{-})-2$.
Then we have
\begin{align}
\sup_{x \in B(x_0,\epsilon)} e(x) \leq C \epsilon^{\frac{2n}{2n-1}} \label{eqn:HE08_2}
\end{align}
for each $\epsilon \in (0, 1)$ and some universal constant $C=C(n)$.
\label{lma:HE05_1}
\end{lemma}
Lemma~\ref{lma:HE08_1} can also be applied to construct good cutoff functions.
\begin{lemma}[\textbf{Cutoff functions on annulus}]
Suppose $x_0 \in X$, $0<\rho<1<\infty$.
Then there exists a function $\phi: X \to [0,1]$ such that
\begin{align*}
&\phi \in C^{\infty}(B(x_0,1) \backslash \mathcal{S}), \quad \supp \phi \Subset B(x_0, 1), \quad \phi \equiv 1 \; \textrm{on} \; B(x_0,\rho), \\
&|\nabla \phi| \leq c(n, \rho), \qquad \quad |\Delta \phi| \leq c(n, \rho), \quad \textrm{on} \; B(x_0,1) \backslash \mathcal{S}.
\end{align*}
Furthermore, for each pair $\rho_1, \rho_2$ satisfying $0<\rho_1<\frac{1}{2}<2<\rho_2<\infty$, there exists a function $\phi: X \times [0,1]$ such that
\begin{align*}
&\phi \in C^{\infty}((B(x_0,\rho_2) \backslash \overline{B(x_0, \rho_1 )}) \cap \mathcal{R}),
\quad \supp \phi \Subset B(x_0, \rho_2) \backslash \overline{B(x_0,\rho_1)}, \\
&\phi \equiv 1 \; \textrm{on} \; B\left(x_0,\frac{\rho_2}{2} \right) \backslash \overline{B(x_0, 2\rho_1)}, \\
&|\nabla \phi| \leq c(n, \rho_1,\rho_2), \qquad \quad |\Delta \phi| \leq c(n, \rho_1,\rho_2),
\quad \textrm{on} \; (B(x_0,\rho_2) \backslash \overline{B(x_0, \rho_1 )}) \cap \mathcal{R}.
\end{align*}
\label{lma:HE04_3}
\end{lemma}
The proof of Lemma~\ref{lma:HE04_3} is based on the maximum principle, solvability of Poisson equation and the fact that $\Delta \underline{L}_1 \geq 1$
and $\Delta \underline{U}_{R'} \geq 1$ for each $R'>0$.
With these properties, one can compare $\underline{L}_1$ with the Poisson equation solution $f$ which has same boundary value as $\underline{L}_1$.
Then construct cutoff function based on the value of $f$. Since the proof follows that of~\cite{CCWarp} verbatim, we omit the details here.
\begin{lemma}[\textbf{Harmonic approximation of local Buseman function}]
There exists a constant $c=c(n)$ with the following properties.
Suppose $x_0 \in X$,
$\gamma$ is a line segment centered at $x_0$ with length $2$, end points $p_{+}$ and $p_{-}$, $\epsilon$ is an arbitrary small positive number,
say $0<\epsilon<0.1$.
In the ball $B(x_0,4\epsilon)$, define local Buseman functions
\begin{align*}
b_{+}(x)=d(x, p_{+})-d(x_0, p_{+}), \quad b_{-}(x)=d(x,p_{-})-d(x_0, p_{-}).
\end{align*}
Let $u_{\pm}$ be the harmonic functions in $B(x_0,4\epsilon)$ such that $\left.\left( u_{\pm}-b_{\pm} \right) \right|_{\partial B(x_0,4\epsilon)}=0$.
Let $u$ be one of $u_{\pm}$ and $b$ be the corresponding $b_{\pm}$ respectively. Then we have
\begin{itemize}
\item $|u-b| \leq c\epsilon^{1+\alpha}$.
\item $\fint_{B(x,\epsilon)} |\nabla (u-b)|^2 \leq c\epsilon^{\alpha}$.
\item $\fint_{B(x,\epsilon)} |Hess_{u}|^2 < c\epsilon^{-2+\alpha}$.
\end{itemize}
Here $\alpha=\alpha(n)$ is a universal constant, which can be chosen as $\frac{1}{2n-1}$.
\label{lma:HE05_2}
\end{lemma}
\begin{proof}
For simplicity, we assume $u=u_{+}$ and $b=b_{+}$.
The pointwise estimate of $|u-b|$ follows from maximum principle and the the excess estimate (\ref{eqn:HE08_2}), same as traditional case.
We proceed to show the integral estimate of $|\nabla (u-b)|$.
Note that $b(x)=r(x)-d(x_0,p_{+})$, where $r(x)=d(x, p_{+})$. It follows from rescaling that
\begin{align*}
\fint_{B(x_0,4\epsilon)} |\Delta b| = \fint_{B(x_0,4\epsilon)} |\Delta r| < C \epsilon^{-1}.
\end{align*}
Actually, by the fact $\Delta r \leq \frac{2n-1}{r}$, the estimate of $\int |\Delta r|$ is reduced to the estimate of
$\int \Delta r$. However, $\int \Delta r$ can be bounded by integration by parts,
modulo some technical discussion around the generalized cut locus and singular set $\mathcal{S}$. Due to the high
codimension of $\mathcal{S}$, the integral of $\Delta r$ around of $\mathcal{S}$ can be ignored. Then we return
to the smooth manifold case, which is discussed clearly in~\cite{Cheeger01}.
Clearly, $u-b \in N_0^{1,2}(B(x_0,4\epsilon))$. Hence integration by parts, Proposition~\ref{prn:HD13_2},
applies and we have
\begin{align*}
\fint_{B(x_0,4\epsilon)} |\nabla (u-b)|^2 &= \fint_{B(x_0,4\epsilon)} (u-b) \Delta (b-u)= \fint_{B(x_0,4\epsilon)} (u-b) \Delta b
\leq C\epsilon^{1+\alpha} \fint_{B(x_0,4\epsilon)} |\Delta b| < C\epsilon^{\alpha}.
\end{align*}
Note that $u$ is harmonic in $B(x_0,4\epsilon)$. Weitzenb\"{o}ck formula implies that
\begin{align*}
\frac{1}{2} \Delta \left(|\nabla u|^2-1\right)=\frac{1}{2} \Delta |\nabla u|^2=|Hess_{u}|^2 \geq 0
\end{align*}
in the classical sense on $B(x_0, 4\epsilon) \backslash \mathcal{S}$. By extension property of subharmonic function, Proposition~\ref{prn:HD16_2}, we see that
$|\nabla u|^2 \in N_{loc}^2(B(x_0, 4\epsilon))$. Let $\phi$ be a cutoff function vanishes on $\partial B(x_0, 4\epsilon)$ and equivalent to $1$ on $B(x_0,\epsilon)$,
with $\epsilon |\nabla \phi|$
and $\epsilon^2 |\Delta \phi|$ bounded as in Lemma~\ref{lma:HE04_3}. Clearly, $\phi \in N_c^{1,2}(B(x_0, 4\epsilon))$.
Therefore, it follows from integration by parts, Proposition~\ref{prn:HD13_2}, that
\begin{align*}
2\fint_{B(x_0, 4\epsilon)} \phi |Hess_u|^2=\fint_{B(x_0, 4\epsilon)} \phi \Delta \left(|\nabla u|^2-1\right)=\fint_{B(x_0, 4\epsilon)} \left(|\nabla u|^2-1 \right) \Delta \phi.
\end{align*}
Consequently, we obtain
\begin{align*}
\fint_{B(x_0,\epsilon)} |Hess_{u}|^2 \leq \fint_{B(x_0, 4\epsilon)} \phi |Hess_u|^2 \leq C \epsilon^{-2} \fint_{B(x_0, 4\epsilon)} \left||\nabla u|^2-1\right| \leq C\epsilon^{\alpha-2}.
\end{align*}
\end{proof}
Note that Lemma~\ref{lma:HE05_2} implies almost splitting property already. Therefore, it is generalization of Lemma~\ref{lma:HE04_1}, the splitting property.
Not surprisingly, one can use Lemma~\ref{lma:HE05_2} to prove Lemma~\ref{lma:HE04_1}, at least formally.
Actually, if there is a line with length $2L$ centered at $x_0$, then in the unit ball $B(x_0,1)$, it follows from Lemma~\ref{lma:HE05_2} that
\begin{align*}
|u-b|<cL^{-\alpha}, \quad \fint_{B(x_0,1)} |\nabla(u-b)|<cL^{-\alpha}, \quad \fint_{B(x_0,1)} |Hess_u|^2<cL^{-\alpha}.
\end{align*}
Let $L \to \infty$, we see that $Hess_{u} \equiv 0$ on $\mathcal{R}$.
From the proof of Lemma~\ref{lma:HE05_2}, it is clear that the key to obtain smallness of $|Hess_u|^2$ is the integration by parts, which is checked in our case.
For smooth Riemannian manifold, the approximation in Lemma~\ref{lma:HE05_2} was improved by Colding and Naber in~\cite{ColdNa}.
The essential difference is that they chose parabolic approximation functions, instead of harmonic approximations.
Suppose $\gamma$ is a line segment with length $2$, centered around $x_0$, with end points $p_{+}$ and $p_{-}$.
Then one can construct cutoff functions $\psi$ such that it vanishes outside
$B(x_0, 8)$ and inside $B(p_{+}, 0.1)$ and $B(p_{-}, 0.1)$, and equals $1$ on $B(x_0,4) \backslash (B(p_{+}, 0.2) \cup B(p_{-}, 0.2))$. Moreover, we have pointwise
bound of $|\Delta \psi|$ and $|\nabla \psi|$. Then for $b_{\pm}$, we can run heat flow starting from $\psi b_{\pm}$ to obtain solution $h_{t,\pm}$.
Then the function $h_{t,\pm}$ is a better approximation function of $b_{\pm}$ on the scale around $\sqrt{t}$.
The extra technical tools needed for Colding-Naber's argument beyond the harmonic approximation consists of an a priori bound of heat kernel, and the construction of
cutoff function with the properties as mentioned above. However, in light of Proposition~\ref{prn:HD07_1} and Lemma~\ref{lma:HE04_3}, both tools are available in our setting.
Therefore, we can develop our version of the parabolic approximation estimate, Theorem 2.19 of~\cite{ColdNa}, in the current case.
\begin{lemma}[\textbf{Parabolic approximation of local Buseman function}]
There exist two constants $c=c(n), \bar{\epsilon}=\bar{\epsilon}(n)$ with the following properties.
Suppose $x_0 \in X$, $\gamma$ is a line segment whose center point locates in $B(x_0,0.2)$, with end points $p_{+}$ and $p_{-}$, with length $2$.
Let $h_{t}$ be the heat approximation of $b$ which is one of $b_{\pm}$.
Suppose the excess value $d(x_0,p_{+})+d(x_0,p_{-})-2<\epsilon^2$ for some $\epsilon \in (0,\bar{\epsilon})$.
Then there exists $\lambda \in [0.5, 2]$ such that
\begin{itemize}
\item $|h_{\lambda \epsilon^2} - b| \leq c \epsilon^2$.
\item $\fint_{B(x,\epsilon)} ||\nabla h_{\lambda \epsilon^2}|^2-1| \leq c\epsilon$.
\item $\int_{0.1}^{1.9} \fint_{B(x,\epsilon)} ||\nabla h_{\lambda \epsilon^2}|^2-1| \leq c\epsilon^2$.
\end{itemize}
Most importantly, we have
\begin{align*}
\int_{0.1}^{1.9} \fint_{B(\gamma(s), \epsilon)} |Hess_{h_{\lambda \epsilon^2}}|^2 \leq c.
\end{align*}
\label{lma:HE04_4}
\end{lemma}
Note that we did not formulate the parabolic approximation in the most precise way. For example, $\gamma$ need not to be a geodesic, an $\epsilon$-geodesic
suffices. Interested readers are referred to \cite{ColdNa} for the most general version.
According to the discussion form Lemma~\ref{lma:HE04_3} to Lemma~\ref{lma:HE04_4}, it is quite clear that the integral estimate of approximation functions
can be obtained in the same way as the Riemannian manifold case, provided the following properties.
\begin{itemize}
\item Almost super-harmonicity of distance functions, Proposition~\ref{prn:HC29_1}.
\item Bishop-Gromov volume comparison, Proposition~\ref{prn:HD19_1}.
\item Strong maximum principle for subharmonic functions, Proposition~\ref{prn:HC29_5}.
\item Integration by parts, Proposition~\ref{prn:HD13_2}.
\item Existence of excellent cutoff function, Lemma~\ref{lma:HE04_3}.
\end{itemize}
Since all of these properties are checked in our situation, we can follow the route of Cheeger-Colding to obtain the following properties, almost line by line.
\begin{lemma}[\textbf{Approximation slices}]
Suppose $x_0 \in X$, $\gamma_1, \gamma_2, \cdots, \gamma_k$ are $k$ line segments with length $2L>>2$ such that the center point of
$\gamma_k$ locates in $B(x_0,1)$ for each $k$. Furthermore, these lines are almost perpendicular to each other, i.e., the Gromov-Hausdorff distance between
$\gamma_1 \cup \gamma_2 \cdots \cup \gamma_k$ and $\tilde{\gamma}_1 \cup \tilde{\gamma}_2 \cup \cdots \tilde{\gamma}_k$ is bounded by
$L\psi(L^{-1})$, where $\tilde{\gamma}_i$ is the line segment on the $i$-th coordinate axis of $\ensuremath{\mathbb{R}}^k$, centered at the origin and with length $2L$, $\psi$ is a nonnegative monotonically increasing function satisfying $\psi(0)=0$.
Suppose the end points of $\gamma_k$ are $p_{i,+}$ and $p_{i,-}$.
Let $b_{i,\pm}$ be the corresponding local Buseman functions with respect to $\gamma_i$. Let $u_i$ be the harmonic function
on $B(x_i,4)$ with the same value as $b_{i,+}$ on $\partial B(x_0,4)$. Then we have
\begin{align*}
\int_{B(x_0,1)} \left\{ \sum_{1 \leq i \leq k} |\nabla u_i -1|^2 + \sum_{1 \leq i<j\leq k}|\langle \nabla u_i, \nabla u_j\rangle| + \sum_{1 \leq i \leq k} |Hess_{u_i}|^2\right\}
\leq \bar{\psi}(L^{-1}),
\end{align*}
where $\bar{\psi}$ is also a nonnegative monotonically increasing function satisfying $\bar{\psi}(0)=0$, depending on $\psi$.
\label{lma:HE07_1}
\end{lemma}
Let $\vec{u}=(u_1, u_2, \cdots, u_k)$, we can regard $\vec{u}$ as an almost submersion from $B(x_0,1)$ to its image on
$\ensuremath{\mathbb{R}}^k$. Consequently, slice argument can be set up as that in~\cite{CCT}. The slice argument together with the Chern-Simons theory can improve the behavior of the singular set $\mathcal{S}$. A more fundamental application of the slice argument is to set up the following volume convergence property, as done in~\cite{ColdVol}.
\begin{proposition}[\textbf{Volume continuity}]
For every $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$ and $\epsilon>0$, there is a constant $\xi=\xi(X,\epsilon)$ such that
\begin{align*}
\left| \log \frac{|B(y_0,1)|}{|B(x_0,1)|} \right|<\epsilon
\end{align*}
for any $(Y,y_0,h) \in \widetilde{\mathscr{KS}}(n,\kappa)$ satisfying $d_{PGH}((X,x_0,g), (Y,y_0,h))<\xi$.
\label{prn:HD22_1}
\end{proposition}
Recall that $d_{PGH}$ means the pointed-Gromov-Hausdorff distance. In Proposition~\ref{prn:HD22_1}, the inequality $d_{PGH}((X,x_0,g), (Y,y_0,h))<\xi$ means that the Gromov-Hausdorff distance between $B(x_0, \xi^{-1}) \subset X$ and $B(y_0, \xi^{-1}) \subset Y$ is less than $\xi$.
Applying the same argument as in~\cite{CCWarp}, we obtain the rigidity of almost volume cones.
\begin{proposition}[\textbf{Almost volume cone implies almost metric cone}]
For each $\epsilon>0$, there exists $\xi=\xi(n,\epsilon)$ with the following properties.
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$ satisfies $\displaystyle \frac{|B(x_0,2)|}{ |B(x_0,1)|} \geq (1-\epsilon)2^{2n}$,
then there exists a metric cone over a length space $Z$, with vertex $z^*$ such that
\begin{align*}
\diam(Z) < \pi + \xi, \quad
d_{GH} \left( B(x_0,1), B(z^*, 1)\right) <\xi.
\end{align*}
Furthermore, $\displaystyle \lim_{\epsilon \to 0} \xi(n,\epsilon)=0$.
\label{prn:HD22_2}
\end{proposition}
Similar to Lemma 9.14 of~\cite{CCT}, we obtain the almost K\"ahler cone splitting, based on Proposition~\ref{prn:HD22_2}.
\begin{proposition}[\textbf{Almost K\"ahler cone splitting}]
For each $\epsilon>0$, there exists $\xi=\xi(n,\epsilon)$ with the following properties.
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x_0 \in X$, $b$ is a Lipschitz function on $B(x_0,2)$ satisfying
\begin{align*}
\sup_{B(x_0,2) \backslash \mathcal{S}} |\nabla b| \leq 2, \quad \fint_{B(x_0,2) \backslash \mathcal{S}} |Hess_b|^2 \leq \epsilon^2.
\end{align*}
Suppose also $\displaystyle \frac{|B(x_0,2)|}{|B(x_0,1)|} \geq (1-\epsilon) 2^{2n}$, i.e., $B(x_0,1)$ is an almost volume cone.
Then there exists a Lipschitz function $\tilde{b}$ on $B(x_0,1)$ such that
\begin{align*}
\sup_{B(x_0,1) \backslash \mathcal{S}} \left|\nabla \tilde{b}\right| \leq 3,
\quad \fint_{B(x_0,1) \backslash \mathcal{S}} \left|\nabla \tilde{b}-J \nabla b\right|^2 \leq \xi.
\end{align*}
\label{prn:HE08_1}
\end{proposition}
\subsection{Volume radius}
Anderson's gap theorem implies that one can improve regularity of the very interior part of a geodesic ball whenever the volume ratio of the geodesic ball is very close
to the Euclidean one. This suggests us to define the volume radius as follows.
\begin{definition}
Let $\delta_0$ be the Anderson constant.
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x_0 \in X$.Then we define
\begin{align*}
&\Omega_{x_0} \triangleq \left\{ r| r>0, r^{-2n}|B(x_0,r)|\geq (1-\delta_0) \omega_{2n} \right\}.\\
&\mathbf{vr}(x_0) \triangleq
\begin{cases}
\sup \Omega_{x_0}, &\textrm{if} \; \Omega_{x_0} \neq \emptyset, \\
0, &\textrm{if} \; \Omega_{x_0}=\emptyset.
\end{cases}
\end{align*}
We call $\mathbf{vr}(x_0)$ the volume radius of the point $x_0$.
\label{dfn:SC17_1}
\end{definition}
According to this definition, a point is regular if and only if its volume radius is positive.
On the other hand, if the space is not $\ensuremath{\mathbb{C}}^n$, then every point has a finite volume radius by a generalized Anderson's gap theorem.
\begin{proposition}[\textbf{Euclidean space by $\mathbf{vr}$}]
Suppose $X \in \widetilde{\mathscr{KS}}(n)$ and $\mathbf{vr}(x_0)=\infty$ for some $x_0 \in X$, then $X$ is isometric to the Euclidean space $\ensuremath{\mathbb{C}}^{n}$.
\label{prn:SC17_1}
\end{proposition}
\begin{proof}
Fix an arbitrary point $x \in X$, then volume comparison implies that
\begin{align*}
\mathrm{v}(x) \geq \lim_{r \to \infty} \omega_{2n}^{-1} r^{-2n} |B(x,r)|=\mathrm{avr}(X) \geq 1-\delta_0.
\end{align*}
Therefore, $x$ is a regular point. Since $x$ is arbitrarily chosen, we see that $X \in \mathscr{KS}(n)$. Then the statement follows from Anderson's gap theorem.
\end{proof}
A local version of Proposition~\ref{prn:SC17_1} is the following local Harnack inequality of $\mathbf{vr}$.
\begin{proposition}[\textbf{Local Harnack inequality of volume radius}]
There is a constant $\tilde{K}=\tilde{K}(n)$ with the following properties.
Suppose $x \in X \in \widetilde{\mathscr{KS}}^{*}(n)$, $r=\mathbf{vr}(x)>0$, then we have
\begin{align}
\tilde{K}^{-1}r \leq \mathbf{vr} \leq \tilde{K}r
\label{eqn:SC26_8}
\end{align}
in the ball $B(x, \tilde{K}^{-1}r)$. Moreover, for every $\rho \in (0, \tilde{K}^{-1}r)$, $y \in B(x, \tilde{K}^{-1}r)$, we have
\begin{align}
&\omega_{2n}^{-1}\rho^{-2n} |B(y,\rho)| \geq 1-\frac{\delta_0}{100}, \label{eqn:SC26_9}\\
& |Rm|(y) \leq \tilde{K}^{2} r^{-2}, \label{eqn:SC27_3}\\
& inj(y) \geq \tilde{K}^{-1} r. \label{eqn:SC27_4}
\end{align}
\label{prn:SC26_7}
\end{proposition}
\begin{proof}
It follows from Bishop volume comparison, Anderson's gap theorem and a compactness argument.
Actually, by adjusting $\tilde{K}$ if necessary, it suffices to show (\ref{eqn:SC27_3}).
We argue by contradiction. Suppose (\ref{eqn:SC27_3}) were wrong, by point-selecting and rescaling, we can find a sequence of $L_i \to \infty$
and Ricci-flat spaces $(X_i, x_i, g_i) \in \widetilde{\mathscr{KS}}^{*}(n)$ such that
\begin{align*}
|Rm|(x_i)=1, \quad \sup_{x \in B(x_i, L_i)} |Rm|(x) \leq 2, \quad \omega_{2n}^{-1} L_i^{-2n} |B(x_i, L_i)| \geq 1-\delta_0.
\end{align*}
Improving regularity property of Ricci-flat metrics implies higher order estimate of $Rm$ in the balls $B(x_i, L_i-1)$.
Therefore, we can take smooth convergence limit(c.f.~\cite{Ha95}):
\begin{align*}
(X_i, x_i, g_i) \longright{C^{\infty}-Cheeger-Gromov} (X_{\infty}, x_{\infty}, g_{\infty}).
\end{align*}
The limit space satisfying $|Rm|(x_{\infty})=1$ and $\textrm{avr}(X_{\infty}) \geq 1-\delta_0$, which is impossible by Anderson's gap theorem or Proposition~\ref{prn:SC17_1}.
\end{proof}
On a Ricci-flat geodesic ball, it is well known that $|Rm|$ bound implies
bound of $|\nabla^k Rm|$ for each positive integer $k$ in a smaller geodesic ball.
So (\ref{eqn:SC27_3}) immediately yields the following corollary.
\begin{corollary}[\textbf{Improving regularity property of volume radius}]
There is a small positive constant $c_a=c_a(n)$ with the following properties.
Suppose $x \in X \in \widetilde{\mathscr{KS}}^{*}(n)$, $\mathbf{vr}(x) \geq r>0$, then we have
\begin{align}
r^{2+k}|\nabla^k Rm|(y) \leq c_a^{-2}, \quad \forall \; y \in B(x,c_a r), \quad 0 \leq k \leq 5.
\label{eqn:SL23_8}
\end{align}
\label{cly:SL23_1}
\end{corollary}
In Ricci bounded geometry, harmonic radius (c.f.~\cite{An90}) plays an important role. A point $x$ is defined to have harmonic radius at least $r$ if on the smooth geodesic ball $B(x,r)$, there exists a harmonic diffeomorphism
$\Psi=(u_1, u_2, \cdots, u_{2n}): B(x,r) \to \Omega \subset \ensuremath{\mathbb{R}}^{2n}$ such that
\begin{align*}
\frac{1}{2}\delta_{ij} \leq g_{ij}=g(\nabla u_i, \nabla u_j) \leq 2 \delta_{ij}, \quad
r^{\frac{3}{2}}\norm{g_{ij}}{C^{1,\frac{1}{2}}} \leq 2.
\end{align*}
Then harmonic radius is defined as the supreme of all the possible $r$'s mentioned above. For convenience, we use
$\mathbf{hr}$ to denote harmonic radius. This definition can be easily moved to our case when the underlying space
is in $\widetilde{\mathscr{KS}}^*(n)$. We define $\mathbf{hr}$ to be $0$ on the singular part of the underlying space.
It is clear from the definitions and Proposition~\ref{prn:SC26_7} that volume radius and harmonic radius can bound each other, i.e., they are equivalent. The following Proposition is obvious.
\begin{proposition}[\textbf{Equivalence of volume and harmonic radius}]
Suppose $x \in X \in \widetilde{\mathscr{KS}}^{*}(n)$, then we have
\begin{align*}
\frac{1}{C} \mathbf{hr}(x) \leq \mathbf{vr}(x) \leq C \mathbf{hr}(x)
\end{align*}
for some uniform constant $C=C(n)$.
\label{prn:HA09_1}
\end{proposition}
Note that the regularity requirement of the underlying space to define volume radius is much weaker than that to define harmonic radius a priori. Therefore, Proposition~\ref{prn:HA09_1} already implies a regularity improvement.
We shall set up the compactness theory based on volume radius,
since volume radius may be applicable to more general metric measure spaces.
Let $X \in \widetilde{\mathscr{KS}}^*(n)$ and decompose it as $X=\mathcal{R} \cup \mathcal{S}$. Then $\mathbf{vr}$ is a positive finite function on $\mathcal{R}$ and equals $0$ on $\mathcal{S}$.
\begin{proposition}[\textbf{Rigidity of volume ratio}]
Suppose $X \in \widetilde{\mathscr{KS}}(n)$. If for two concentric geodesic balls $B(x_0,r_1) \subset B(x_0,r_2)$
centered at a regular point $x_0$, we have
\begin{align}
\omega_{2n}^{-1}r_1^{-2n}|B(x_0,r_1)|= \omega_{2n}^{-1}r_2^{-2n}|B(x_0,r_2)|,
\label{eqn:SL23_1}
\end{align}
then the ball $B(x_0,r_2)$ is isometric to a geodesic ball of radius $r_2$ in $\ensuremath{\mathbb{C}}^n$.
Furthermore, if $X \in \mathscr{KS}(n)$, then we can further conclude that $X$ is Euclidean.
\label{prn:SC17_3}
\end{proposition}
\begin{proof}
From the proof of Lemma~\ref{lma:HE04_2}, it is clear that $B(x_0,r_2)$ is a volume cone with constant volume
ratio $\omega_{2n}$.
Observe the change of volume element along each smooth geodesic emanating from $x_0$, in the polar coordinate.
By the volume density gap between regular and singular points,
the optimal volume ratio of $B(x_0,1)$ forces that it does not contain any singular point.
Then the situation is the same as the smooth Riemannian case. Clearly, a smooth Ricci-flat geodesic ball with volume
ratio $\omega_{2n}$ is isometric to a Euclidean ball of the same radius.
If $X\in \mathscr{KS}(n)$, by analyticity of metric tensor, it is clear that $X$ is flat and hence $\ensuremath{\mathbb{C}}^n$ due to its non-collapsing property at infinity.
\end{proof}
\begin{proposition}[\textbf{Continuity of volume radius}]
$\mathbf{vr}$ is a continuous function on $X$ whenever $X \in \widetilde{\mathscr{KS}}(n)$.
\label{prn:HE07_2}
\end{proposition}
\begin{proof}
Since $\mathbf{vr} \equiv \infty$ on $\ensuremath{\mathbb{C}}^n$, which is obvious continuous. So we can assume $X \in \widetilde{\mathscr{KS}}^{*}(n)$ without loss of generality.
By Proposition~\ref{prn:SC17_1}, we know $\mathbf{vr}$ is a finite function on $X$.
So we assume $\mathbf{vr}$ is a function with value in $[0,\infty)$. It is also easy to see that $\mathbf{vr}$ is continuous at singular points.
We know that a point $x_0$ is singular if and only if $\mathbf{vr}(x_0)=0$. Clearly, for every sequence $x_i \to x_0$, we must have
$\displaystyle \lim_{i \to \infty} \mathbf{vr}(x_i)=0$.
Otherwise, we have a sequence $x_i$ converging to $x_0$ and
$\displaystyle \lim_{i \to \infty} \mathbf{vr}(x_i) \geq \xi>0$. However, we note that $x_0 \in B(x_i, \tilde{K}^{-1} \xi)$ for
large $i$. Therefore, $x_0$ is forced to be regular by the improving regularity property of volume radius. Contradiction.
Therefore, discontinuity point must admit positive $\mathbf{vr}$ if it does exist. Suppose $x_0$ is a discontinuous point of $\mathbf{vr}$.
Then we can find a sequence of points $x_i \in X$ such that
\begin{align*}
&x_0=\lim_{i \to \infty} x_i, \\
&0<\mathbf{vr}(x_0)=r_0<\infty, \\
&\lim_{i \to \infty} \mathbf{vr}(x_i) \neq r_0.
\end{align*}
Clearly, $\log \mathbf{vr}(x_i)$ are uniformly bounded by Proposition~\ref{prn:SC26_7}.
So we can assume $\mathbf{vr}(x_i)$ converge to a positive number $\bar{r}$. By volume continuity, we clearly have
\begin{align*}
\omega_{2n}^{-1}r_0^{-2n}|B(x_0,r_0)|=1-\delta_0=\lim_{i \to \infty} \omega_{2n}^{-1}\mathbf{vr}(x_i)^{-2n}|B(x_i,\mathbf{vr}(x_i))|=
\omega_{2n}^{-1}\bar{r}^{-2n}|B(x_0,\bar{r})|.
\end{align*}
Since $\bar{r} \neq r_0$, we obtain from Proposition~\ref{prn:SC17_3} that $B(x_0,r_0)$ is a ball in a metric cone centered at the vertex. Note that $x_0$ is a regular
point since $\mathbf{vr}(x_0)>0$. Therefore, $B(x_0,r_0)$ is the standard ball in $\ensuremath{\mathbb{C}}^{n}$ with radius $r_0$.
Consequently, the normalized volume ratio of $B(x_0,r_0)$ is $1$, which contradicts the fact that $\mathbf{vr}(x_0)=r_0$ and the definition of volume radius.
\end{proof}
The volume radius has better property. It satisfies Harnack inequality in the interior of a shortest geodesic.
The H\"older continuity estimate of Colding-Naber (c.f.~\cite{ColdNa}) can be interpreted by volume radius as follows.
\begin{proposition}[\textbf{Global Harnack inequality of volume radius}]
For every small constant $c$, there is a constant $\epsilon=\epsilon(n,\kappa,c)$ with the following properties.
Suppose $(X,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x,y \in X$, $\gamma$ is shortest, unit speed geodesic connecting $x$ and $y$, with smooth interior parts.
Suppose $\gamma(0)=x, \gamma(L)=y$, $L \leq r$.
If $\mathbf{vr}(y)>c r$, then we have
\begin{align}
\mathbf{vr}(\gamma(t))> \epsilon r, \quad \forall \; t \in [cL, L].
\label{eqn:SC26_4}
\end{align}
In particular, if $\min\{\mathbf{vr}(x), \mathbf{vr}(y)\}>cr$, then we have
\begin{align}
\mathbf{vr}(\gamma(t))>\epsilon r, \quad \forall \; t \in [0,L].
\label{eqn:SC26_44}
\end{align}
\label{prn:SC26_2}
\end{proposition}
\begin{proof}
Clearly, (\ref{eqn:SC26_44}) follows from (\ref{eqn:SC26_4}). Therefore, it suffices to prove (\ref{eqn:SC26_4}) only.
Up to a normalization, we can assume $r=L=1$.
So $\gamma$ is the shortest geodesic connecting $x,y$ such that $\gamma(0)=x, \gamma(1)=y$.
By assumption, we have $\mathbf{vr}(y)>c$.
By local Harnack inequality of volume radius, Proposition~\ref{prn:SC26_7},
there exists $\bar{\epsilon}=\bar{\epsilon}(n, c)$ such that $\mathbf{vr}>\bar{\epsilon}$ for each $\gamma(t)$
with $t \in [1-\bar{\epsilon}, 1]$. Clearly, in the middle part of $\gamma$, i.e., for every $t \in [\bar{\epsilon}, 1-\bar{\epsilon}]$, we have
$|\Delta r|< \frac{C}{\bar{\epsilon}}$ for a universal $C=C(n)$, where $r$ is the distance to $\gamma(0)$.
Because of the segment inequality (Proposition~\ref{prn:HD17_1}) and the parabolic approximation (Lemma~\ref{lma:HE04_4}),
we can follow the proof of Proposition 3.6 and Theorem 1.1 of~\cite{ColdNa} verbatim.
Similar to the statement in the proof of Theorem 1.1 on page 1213 of~\cite{ColdNa}, we can find constants $\bar{s}=\bar{s}(n,c,\bar{\epsilon}), \bar{r}=\bar{r}(n,c,\bar{\epsilon})$
such that for every $t_1, t_2 \in [\bar{\epsilon},1-\bar{\epsilon}]$ satisfying $|t_1-t_2|<\bar{s}$ and every $r \in (0,\bar{r})$, we have
\begin{align*}
1-\frac{\delta_0}{100} \leq \frac{|B(\gamma(t_1), r)|}{|B(\gamma(t_2),r)|} \leq 1+\frac{\delta_0}{100}.
\end{align*}
Then it is easy to see that if the volume radius is uniformly bounded below at $t_1$, it must be uniformly bounded below at $t_2$.
Actually, suppose the volume radius at $\gamma(t_1)$ is greater than $r_1$ for some $r_1 \in (0,\bar{r})$,
by inequality (\ref{eqn:SC26_9}) in Proposition~\ref{prn:SC26_7}, we have
$\omega_{2n}^{-1}r^{-2n}|B(\gamma(t_1),r)| \geq 1-\frac{\delta_0}{100}$ for every $r \in [0, \frac{r_1}{\tilde{K}}]$.
Put this information into the above inequality implies that
\begin{align*}
\omega_{2n}^{-1}r^{-2n}|B(\gamma(t_2),r)| \geq \frac{1-\frac{\delta_0}{100}}{1+\frac{\delta_0}{100}}>1-\delta_0,
\quad \forall \; r \in \left[0, \frac{r_1}{\tilde{K}} \right].
\end{align*}
Therefore, the volume radius of $\gamma(t_2)$ is at least $\frac{r_1}{\tilde{K}}$. From this induction, it is clear that
\begin{align*}
\mathbf{vr}(\gamma(t)) \geq
\tilde{K}^{-\frac{1-\bar{\epsilon}-t}{\bar{s}}} \mathbf{vr}(\gamma(1-\bar{\epsilon}))
>\bar{\epsilon} \tilde{K}^{-\frac{1-\bar{\epsilon}-t}{\bar{s}}}.
\end{align*}
Let $\epsilon$ be the number on the right hand side of the above inequality when $t=\bar{\epsilon}$.
Then $\epsilon=\epsilon(\bar{\epsilon}, \tilde{K}, \bar{s})=\epsilon(n,\kappa,c)$
and we finish the proof of (\ref{eqn:SC26_4}).
\end{proof}
In general, if $X$ is only a metric space, we even do not know whether $\mathbf{vr}$ is semi-continuous. The continuity of $\mathbf{vr}$ on $X$ whenever
$X \in \widetilde{\mathscr{KS}}(n,\kappa)$ makes $\mathbf{vr}$ a convenient tool to study the geometry of $X$. By Proposition~\ref{prn:SC26_7},
one can improve regularity on a scale proportional to $\mathbf{vr}$. So it is convenient to decompose the space $X$ according to the function $\mathbf{vr}$.
\begin{definition}
Suppose $(X,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$. Define
\begin{align}
&\mathcal{F}_{r}(X) \triangleq \left\{ x \in X | \mathbf{vr}(x) \geq r \right\}, \label{eqn:SB25_1}\\
&\mathcal{D}_{r}(X) \triangleq \left( \mathcal{F}_{r}(X) \right)^{c}=\left\{ x \in X | \mathbf{vr}(x) < r \right\}. \label{eqn:SB25_2}
\end{align}
We call $\mathcal{F}_{r}(X)$ the $r$-regular part of $X$, $\mathcal{D}_{r}(X)$ the $r$-singular part of $X$.
\label{dfn:SB25_1}
\end{definition}
From Definition~\ref{dfn:SB25_1}, it is clear that
\begin{align}
& \mathcal{R}(X)= \bigcup_{r>0} \mathcal{F}_{r}(X), \label{eqn:SB25_4}\\
& \mathcal{S}(X)= \bigcap_{r>0} \mathcal{D}_{r}(X). \label{eqn:SB25_5}
\end{align}
We observe that the volume radius of each point is related to its distance to singular set by the following property.
\begin{proposition}[\textbf{$\mathbf{vr}$ bounded from above by distance to $\mathcal{S}$}]
Suppose $(X,x,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then
\begin{align}
& \left\{ x | d(x, \mathcal{S}) \geq r \right\} \supset \mathcal{F}_{\tilde{K}r}, \label{eqn:SB25_14} \\
& \left\{x | d(x,\mathcal{S}) < r \right\} \subset \mathcal{D}_{\tilde{K}r}. \label{eqn:SB25_13}
\end{align}
\end{proposition}
\begin{proof}
Choose an arbitrary point $x \in \mathcal{F}_{\tilde{K}r}$, then $ {\mathbf{vr}}(x) \geq \tilde{K} r.$ It follows from Proposition~\ref{prn:SB25_2} that
$\mathbf{vr}(y) \geq r>0$ for every point $y \in B(x, r)$. Therefore, every point in $B(x,r)$ is regular. So $d(x,\mathcal{S})>r$. This proves
(\ref{eqn:SB25_14}) by the arbitrary choice of $x \in \mathcal{F}_{\tilde{K}r}$. Taking complement of (\ref{eqn:SB25_14}), we obtain (\ref{eqn:SB25_13}).
\end{proof}
\subsection{Compactness of $\widetilde{\mathscr{KS}}(n, \kappa)$}
As a model space, $\widetilde{\mathscr{KS}}(n,\kappa)$ should have compactness.
However, we need first to obtain a weak compactness, then we improve regularity further to obtain the
genuine compactness.
It is not hard to see the weak compactness theory of Anderson-Cheeger-Colding-Tian-Naber can be generalized to apply on $\widetilde{\mathscr{KS}}(n,\kappa)$
without fundamental difficulties, almost verbatim.
Actually, the key of Anderson-Cheeger-Colding-Tian-Naber theory is that one can approximate the distance function by harmonic function, or heat flow solution,
which have much better regularity for developing integral estimates. These estimates are justified by the technical preparation in previous subsections.
\begin{proposition}[\textbf{Weak compactness}]
Suppose $(X_i,x_i,g_i) \in \widetilde{\mathscr{KS}}(n,\kappa)$, by taking subsequences if necessary,
we have
\begin{align*}
(X_i, x_i, g_i) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{X},\bar{x},\bar{g})
\end{align*}
for some length space $\bar{X}$ which satisfies all the properties of spaces in
$\widetilde{\mathscr{KS}}(n,\kappa)$ except the 3rd and 4th property, i.e., the weak convexity of $\mathcal{R}$
and the Minkowski dimension estimate $\mathcal{S}$.
However, the Hausdorff dimension of $\mathcal{S}$ is not greater than $2n-4$.
\label{prn:HA05_1}
\end{proposition}
\begin{proof}
Note that each space in $\widetilde{\mathscr{KS}}(n,\kappa)$ satisfies volume doubling property. Therefore, if
there exists a sequence $(X_i,x_i,g_i) \in \widetilde{\mathscr{KS}}(n,\kappa)$, by standard ball packing argument, it is clear that
\begin{align*}
(X_i, x_i, g_i) \stackrel{G.H.}{\longrightarrow} (\bar{X},\bar{x},\bar{g})
\end{align*}
for some length space $\bar{X}$. Then let us list the properties satisfied by $\bar{X}$.
By Proposition~\ref{prn:HD22_1}, $\bar{X}$ inherits a natural measure from the limit process, which is a measure compatible with
the limit metric structure, as that in~\cite{CC1}. Then the volume convergence follows, almost tautologically.
It follows directly from this property and the volume comparison that $\bar{X}$ satisfies
Property 6 in Definition~\ref{dfn:SC27_1}.
In the limit space $\bar{X}$, we can define regular points as the collection
of points where every tangent space is $\ensuremath{\mathbb{R}}^{2n}$, singular points as those points which are not regular.
Let $\mathcal{R}(\bar{X})$ and $\mathcal{S}(\bar{X})$ be the regular and singular part of $\bar{X}$ respectively.
We automatically obtain the regular-singular decomposition $\bar{X}=\mathcal{R}(\bar{X}) \cup \mathcal{S}(\bar{X})$.
By a version of Anderson's gap theorem(c.f. Proposition~\ref{prn:SC17_1}) and volume convergence,
a blowup argument shows that each regular point
has a small neighborhood which has a smooth manifold structure. Clearly, this manifold is Ricci-flat with an attached
limit K\"ahler structure. So we proved Property 1 and Property 2, except the non-emptiness of $\mathcal{R}$.
Note that each tangent space of $\bar{X}$ is a volume cone, due to the volume convergence
and Bishop-Gromov volume comparison, which can be established as that in~\cite{CC1}.
The it follows from Proposition~\ref{prn:HD22_2} that every volume cone is actually a metric cone. Then an induction argument can be applied, like that in~\cite{CC1}, to obtain the stratification of singularities $\mathcal{S}=\mathcal{S}_1\cup \mathcal{S}_2 \cdots \cup \mathcal{S}_{2n}$, where $\mathcal{S}_k$
is the union of singular points whose tangent space can split-off at least $(2n-k)$-straight lines.
In particular, generic points of $\bar{X}$ have tangent spaces $\ensuremath{\mathbb{R}}^{2n}$. In other words, generic points
are regular, so $\mathcal{R} \neq \emptyset$ and we finish the proof of Property 2.
The K\"ahler condition
guarantees that each tangent cone exactly splits off $\ensuremath{\mathbb{C}}^k$, by Proposition~\ref{prn:HE08_1}, as done in~\cite{CCT}.
So the stratification of singular set can be improved as
$\mathcal{S}=\mathcal{S}_2 \cup \mathcal{S}_4 \cup \cdots \mathcal{S}_{2n}$.
By Lemma~\ref{lma:HE07_1}, we can apply slice argument as that in \cite{CCT} and \cite{Cheeger03}.
Consequently, Chern-Simons theory implies that codimension 2 singularity cannot appear, due to the fact that a generic slice is a smooth surface with boundary, and the Ricci curvature's restriction on such a surface is zero.
Actually, the smoothness of generic slices follows from the high codimension of the singular set (item 4 of Definition~\ref{dfn:SC27_1}) and the gradient estimates of the harmonic approximation functions(Proposition~\ref{prn:HC29_6}).
Therefore, $\mathcal{S}=\mathcal{S}_4 \cup \cdots \mathcal{S}_{2n}$,
which means $\dim_{\mathcal{H}} \mathcal{S} \leq 2n-4$.
Let $\bar{y} \in \mathcal{S}(\bar{X})$. Suppose $y_i \in X_i$ satisfies $y_i \to \bar{y}$. Then either there is a uniform $\xi$ such that
every point in each $B(y_i,\xi)$ are regular(but without uniform curvature bound as $i$ increase), or we can choose $y_i$ such that every $y_i$ is singular. In the first case,
we can use a blowup argument and Anderson's gap theorem to show that the volume density of $\bar{y}$ is strictly less
that $1-2\delta_0$. In the second case, we can use volume comparison and convergence to show
$\mathrm{v}(\bar{y}) \leq 1-2\delta_0$. So we proved Property 5.
We have checked all the properties of $\bar{X}$ as claimed. We now need to improve the convergence topology from Gromov-Hausdorff topology.
However, this improvement follows from volume convergence and the improving regularity property of volume radius, Corollary~\ref{cly:SL23_1}.
\end{proof}
From the above argument, it is clear that no new idea is needed beyond the traditional theory,
when technical lemmas and propositions in the previous sections are available.
Actually, weak compactness can be established under even weaker conditions, which will be discussed in our forthcoming work.
Based on the weak compactness, we immediately obtain an $\epsilon$-regularity property, as that in~\cite{CCT}.
\begin{proposition}[\textbf{$\epsilon$-regularity}]
There exists an $\epsilon=\epsilon(n,\kappa)$ with the following properties.
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x_0 \in X$.
Suppose
\begin{align*}
d_{GH}\left( B(x_0,1), B((z_0^*,0), 1) \right)<\epsilon
\end{align*}
where $(z_0^*,0) \in C(Z_0) \times \ensuremath{\mathbb{R}}^{2n-3}$ for some metric cone $C(Z_0)$ with vertex $z_0^*$. Then we have
\begin{align*}
\mathbf{vr}(x_0)>\frac{1}{2}.
\end{align*}
\label{prn:HD20_1}
\end{proposition}
\begin{proof}
Otherwise, there is a sequence of $\epsilon_i \to 0$ and $x_i \in X_i$ violating the statement.
By weak compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$, we can assume $x_i \to x$ and $z_i^* \to z^*$ with the following identity holds.
\begin{align*}
d_{GH}\left( (B(x,1), B((z^*,0), 1))\right)=0.
\end{align*}
In particular, the tangent cone at $x$ is exactly the cone $C(Z) \times \ensuremath{\mathbb{R}}^{2n-3}$, which must be $\ensuremath{\mathbb{C}}^n$ by the complex rigidity.
Therefore, $B(x,1)$ is the unit ball in $\ensuremath{\mathbb{C}}^n$. Thus, the volume convergence implies that for large $i$,
$2^{2n}\left|B \left(x_i, \frac{1}{2} \right) \right|$ can be very close to $1$. In particular, $\mathbf{vr}(x_i)>\frac{1}{2}$
by the definition of volume radius. However, this contradicts our assumption.
\end{proof}
Then we are able to move the integral estimate of~\cite{CN} to $X \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{proposition}[\textbf{Density estimate of regular points}]
For every $0<p<2$, there is a constant $E=E(n,\kappa,p)$ with the following properties.
Suppose $(X,x,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then we have
\begin{align}
r^{2p-2n} \int_{B(x,r)} \mathbf{vr}(y)^{-2p} dy \leq E(n,\kappa,p). \label{eqn:SB25_15}
\end{align}
\label{prn:SB25_2}
\end{proposition}
\begin{proof}
In light of Proposition~\ref{prn:SC26_7}, it is clear that volume radius and harmonic radius are uniformly equivalent. Therefore, Proposition~\ref{prn:SB25_2} is nothing but a singular version of the Cheeger-Naber estimate(c.f. the second inequality of part 2 of Corollary 1.26 in~\cite{CN}).
As pointed out by Cheeger and Naber, their estimate holds for Gromov-Hausdorff limit for Ricci-flat manifolds.
Actually, going through their proof, it is clear that the smooth structure of the underlying space is not used.
Intuitively, if Bishop-Gromov volume comparison holds, then most geodesic balls are almost volume cones, hence almost metric cones.
However, if a cone is very close to a cone which splits off at least $(2n-3)$-lines,
then it must be Euclidean space by the $\epsilon$-regularity property. This intuition was quantified in~\cite{CN}, by the method they called quantitative calculus,
which does not depends on smooth structure by its nature.
We note that the quantitative calculus argument of~\cite{CN} works when we have the following properties.
\begin{itemize}
\item Bishop-Gromov volume comparison, by Proposition~\ref{prn:HD19_1}.
\item Weak compactness of $ \widetilde{\mathscr{KS}}(n,\kappa)$, by Proposition~\ref{prn:HA05_1}.
\item Volume convergence, by Proposition~\ref{prn:HD22_1}.
\item Almost volume cone implies almost metric cone, by Proposition~\ref{prn:HD22_2}.
\item $\epsilon$-regularity, by Proposition~\ref{prn:HD20_1}.
\end{itemize}
Since all these properties hold on $ \widetilde{\mathscr{KS}}(n,\kappa)$, the proof follows that of~\cite{CN} verbatim.
\end{proof}
An immediate consequence of Proposition~\ref{prn:SB25_2} is the following volume estimate of neighborhood of
singular set.
\begin{corollary}[\textbf{Volume estimate of singular neighborhood}]
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $0<\rho<<1$.
Then for each $0<p<2$, we have
\begin{align*}
\left| \{x| d(x,\mathcal{S})<\rho, x \in B(x_0, 1)\} \right|< C\rho^{2p},
\end{align*}
for some $C=C(n,\kappa,p)$.
\label{cly:HD19_2}
\end{corollary}
\begin{proof}
It follows from Definition~\ref{dfn:SB25_1} that
\begin{align*}
(2r)^{-2p}\left|B(x_0,1) \cap \mathcal{F}_{r} \backslash \mathcal{F}_{2r}\right|
= \int_{B(x_0,1) \cap \mathcal{F}_{r} \backslash \mathcal{F}_{2r}} (2r)^{-2p}
< \int_{B(x_0,1) \cap \mathcal{F}_{r} \backslash \mathcal{F}_{2r}} \mathbf{vr}^{-2p}<E(n,\kappa,p),
\end{align*}
which implies that
\begin{align*}
\left|B(x_0,1) \cap \mathcal{D}_{2r} \backslash \mathcal{D}_{r}\right| < 2^{2p} E r^{2p}
\Rightarrow |B(x_0,1) \cap \mathcal{D}_{2r}|<\frac{E}{1-4^{-p}} (2r)^{2p}.
\end{align*}
By virtue of (\ref{eqn:SB25_13}), we have
\begin{align*}
\left|B(x_0,1) \cap \{x |d(x,\mathcal{S})<\rho\} \right| \leq |B(x_0,1) \cap \mathcal{D}_{\tilde{K}\rho}|< \frac{E}{1-4^{-p}} \tilde{K}^{2p} \rho^{2p}<C\rho^{2p}.
\end{align*}
\end{proof}
Now we are ready to prove the compactness theorem.
\begin{theorem}[\textbf{Compactness}]
$\widetilde{\mathscr{KS}}(n,\kappa)$ is compact under the pointed Cheeger-Gromov topology.
\label{thm:HD19_1}
\end{theorem}
\begin{proof}
Suppose $(X_i, x_i,g_i) \in \widetilde{\mathscr{KS}}(n,\kappa)$, we already know, by Proposition~\ref{prn:HA05_1},
that $(X_i,x_i,g_i)$ converges to a limit space $(\bar{X},\bar{x},\bar{g})$, which satisfies almost all the properties of
spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$, except the weak convexity of $\mathcal{R}$ and the
Minkowski dimension estimate of $\mathcal{S}$.
However, fix every two points $\bar{y},\bar{z} \in \mathcal{R} \subset \bar{X}$, we can find a sequence of points
$y_i, z_i \in X_i$ such that $y_i \to \bar{y}$ and $z_i \to \bar{z}$. It is clear that
$\mathbf{vr}(y_i) \to \mathbf{vr}(\bar{y})$ and $\mathbf{vr}(z_i) \to \mathbf{vr}(\bar{z})$.
It follows from the global Harnack inequality of volume radius, Proposition~\ref{prn:SC26_2}, that each shortest geodesic
$\gamma_i$ connecting $y_i$ and $z_i$ is uniformly regular. Consequently, the limit shortest geodesic $\bar{\gamma}$
connecting $\bar{y}$ and $\bar{z}$ is a smooth shortest geodesic. Therefore, we have actually proved that
$\mathcal{R}$ is convex, rather than weakly convex.
Furthermore, if we repeatedly use the first inequality in Proposition~\ref{prn:SC26_2} and smooth convergence determined by volume radius,
one can see that a shortest geodesic $\bar{\gamma}$ with smooth interior can be obtained, even if we drop the condition $\bar{y} \in \mathcal{R}$.
In other words, if $\bar{z} \in \mathcal{R}$, $\bar{y} \in \bar{X}$, then there is a shortest geodesic $\bar{\gamma}$ connecting them, with smooth interior.
This means that $\mathcal{R}$ is strongly convex.
By convexity of $\mathcal{R}$, it is clear that the limit space $\bar{X}$ has Bishop-Gromov volume comparison.
By virtue of volume convergence and the same argument in Proposition~\ref{prn:HE07_2},
we see that $\mathbf{vr}$ is a continuous function under the pointed Cheeger-Gromov topology. In other words, for every point $\bar{z} \in \bar{X}$, and points
$z_i \in X_i$ satisfying $z_i \to \bar{z}$, we have $\displaystyle \mathbf{vr}(\bar{z})=\lim_{i \to \infty} \mathbf{vr}(z_i)$.
For each $r>0$, by density estimate, Proposition~\ref{prn:SB25_2},
we see that inequality (\ref{eqn:SB25_15}) holds for every $B(x_i,r)$ uniformly.
Taking limit, by the convergence of volume radius, we obtain (\ref{eqn:SB25_15}) holds on $(\bar{X},\bar{x},\bar{g})$, for
each $p\in (1.5, 2)$. Then it follows from Corollary~\ref{cly:HD19_2} and the definition of Minkowski dimension (c.f. Definition~\ref{dfn:HE08_1})
that $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$.
\end{proof}
\begin{theorem}[\textbf{Space regularity improvement}]
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, then $\mathcal{R}$ is strongly convex, and
$\dim_{\mathcal{M}}\mathcal{S} \leq 2n-4$.
Suppose $x_0 \in \mathcal{S}$, $Y$ is a tangent space of $X$ at $x_0$.
Then $Y$ is a metric cone in $\widetilde{\mathscr{KS}}(n,\kappa)$ with the splitting
\begin{align*}
Y=\ensuremath{\mathbb{C}}^{n-k} \times C(Z)
\end{align*}
for some $k \geq 2$, where $C(Z)$ is a metric cone without lines.
\label{thm:HE08_1}
\end{theorem}
\begin{proof}
The strong convexity of $\mathcal{R}$ and $\dim_{\mathcal{M}}\mathcal{S} \leq 2n-4$ follows from the argument in the proof of Theorem~\ref{thm:HD19_1}.
Moreover, by Theorem~\ref{thm:HD19_1}, we know each tangent space, as a pointed Gromov-Hausdorff limit,
must locate in $\widetilde{\mathscr{KS}}(n,\kappa)$. Since $Y$ is a volume cone, due to volume convergence,
the splitting of $Y$ follows from Lemma~\ref{lma:HD30_1}.
\end{proof}
Because of Theorem~\ref{thm:HD19_1} to Theorem~\ref{thm:HE08_1}, it seems reasonable to make the following definition for the simplicity of notations.
\begin{definition}
A length space $(X^n,g)$ is called a conifold of complex dimension $n$ if the following properties are satisfied.
\begin{enumerate}
\item $X$ has a disjoint regular-singular decomposition $X=\mathcal{R} \cup \mathcal{S}$, where $\mathcal{R}$ is the regular part, $\mathcal{S}$ is the singular part.
A point is called regular if it has a neighborhood which is isometric to a totally geodesic convex domain of some smooth Riemannian manifold. A point is called singular
if it is not regular.
\item The regular part $\mathcal{R}$ is a nonempty, open manifold of real dimension $2n$.
Moreover, there exists a complex structure $J$ on $\mathcal{R}$ such that $(\mathcal{R}, g, J)$ is a K\"ahler manifold.
\item $\mathcal{R}$ is strongly convex, i.e., for every two points $x \in \mathcal{R}$ and $y \in X$, one can find a shortest geodesic $\gamma$ connecting $x$, $y$ whose every interior point is in $\mathcal{R}$. In particular, $\mathcal{R}$ is geodesic convex.
\item $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$, where $\mathcal{M}$ means Minkowski dimension.
\item Every tangent space of $x \in \mathcal{S}$ is a metric cone of Hausdorff dimension $2n$. Moreover, if $Y$ is a tangent cone of $x$, then the unit ball
$B(\hat{x},1)$ centered at vertex $\hat{x}$ must satisfy
\begin{align*}
|B(\hat{x},1)|_{d\mu} \leq (1-\delta_0) \omega_{2n}
\end{align*}
for some uniform positive number $\delta_0=\delta_0(n)$.
Here $d\mu$ is the $2n$-dimensional Hausdorff measure, $\omega_{2n}$ is the volume of unit ball in $\ensuremath{\mathbb{C}}^n$.
\end{enumerate}
\label{dfn:HD20_1}
\end{definition}
Roughly speaking, a conifold is a space which is almost a manifold away from a small singular set, where every tangent space is a metric cone.
Note that we abuse notation here since the conifold has different meaning
in the literature of string theory(c.f.~\cite{Green}).
It is easy to see that every K\"ahler orbifold with singularity codimension not less than $4$ is a conifold in our sense.
With this terminology, we see that $\widetilde{\mathscr{KS}}(n,\kappa)$ is nothing but the collection of Calabi-Yau conifold
with Euclidean volume growth, i.e.,
\begin{align*}
\lim_{r \to \infty} \frac{|B(x,r)|_{d\mu}}{\omega_{2n}r^{2n}} \geq \kappa, \quad \forall \; x \in X.
\end{align*}
Then Theorem~\ref{thm:HD19_1} can be interpreted as that the moduli space of non-collapsed Calabi-Yau conifolds is compact, under the pointed Cheeger-Gromov topology. Theorem~\ref{thm:HE08_1} can be understood as that a ``weakly" Calabi-Yau conifold is really a conifold, due to an intrinsic improving regularity property
originates from the intrinsic Ricci flatness of the underlying space.
The property of the moduli space $\widetilde{\mathscr{KS}}(n,\kappa)$ is quite clear now.
\begin{proof}[Proof of Theorem~\ref{thmin:HE21_1}]
It follows directly from Theorem~\ref{thm:HD19_1}, Theorem~\ref{thm:HE08_1} and Definition~\ref{dfn:HD20_1}.
\end{proof}
Actually, along the route to prove Theorem~\ref{thm:HE08_1}, we shall be able to improve the regularity of the
spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$ even further. For example, we believe the following statement is true.
\begin{conjecture}
At every point $x_0$ of a Calabi-Yau conifold $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, the tangent space is unique.
\label{cje:HE09_1}
\end{conjecture}
The above problem is only interesting when $n>2$ and away from generic singular point.
Note that if $X$ is a limit space of a sequence of Ricci flat manifolds, then the uniqueness of tangent cone is a well known open problem, in the classical theory of Cheeger-Colding-Tian.
Clearly, similar questions can be asked for general K\"ahler Einstein conifold.
It is not hard to see that a compact K\"ahler Einstein conifold is a projective variety.
Due to its independent interest, we shall discuss this issue in another separate paper.
\subsection{Space-time structure of $\widetilde{\mathscr{KS}}(n)$}
\label{subsec:reduced}
Every space $X \in \widetilde{\mathscr{KS}}(n)$ can be regarded as a trivial Ricci flow solution.
Therefore, Perelman's celebrated work \cite{Pe1} can find its role in the study of $X$.
Let us briefly recall some fundamental functionals defined for the Ricci flow by Perelman.
Suppose $\{(X^m, g(t)), -T \leq t \leq 0\}$ is a Ricci flow solution on a smooth complete Riemannian manifold
$X$ of real dimension $m$. Suppose $x,y \in X$. Suppose $\boldsymbol{\gamma}$ is a space-time curve
parameterized by $\tau=-t$ such that
\begin{align*}
\boldsymbol{\gamma}(0)=(x,0), \quad \boldsymbol{\gamma}(\bar{\tau})=(y,-\bar{\tau}).
\end{align*}
Let $\gamma$ be the space-projection curve of $\boldsymbol{\gamma}$.
In other words, we have
\begin{align*}
\boldsymbol{\gamma}(\tau)=(\gamma(\tau), -\tau).
\end{align*}
By the way, for the simplicity of notations, we always use bold symbol of a Greek character to denote a space-time curve. The corresponding space projection will be denoted by the normal Greek character.
Following Perelman, the Lagrangian of the space-time curve $\boldsymbol{\gamma}$ is defined as
\begin{align}
\mathcal{L}(\boldsymbol{\gamma})=\int_0^{\bar{\tau}} \sqrt{\tau} \left(R+|\dot{\gamma}|^2 \right)_{g(-\tau)} d\tau.
\label{eqn:MA22_1}
\end{align}
Among all such $\boldsymbol{\gamma}$'s that connected $(x,0)$, $(y,-\bar{\tau})$ and parameterized by $\tau$,
there is at least one smooth curve $\boldsymbol{\alpha}$ which minimizes the Lagrangian.
This curve is called a shortest reduced geodesic. The reduced distance between $(x,0)$ and $(y,-\bar{\tau})$ is
defined as
\begin{align}
l((x,0),(y,-\bar{\tau}))=\frac{\mathcal{L}(\boldsymbol{\alpha})}{2\sqrt{\bar{\tau}}}.
\label{eqn:MA22_2}
\end{align}
Let $V=\dot{\alpha}$. Then $V$ satisfies the equation
\begin{align}
\nabla_V V +\frac{V}{2\tau} + 2Ric(V, \cdot) + \frac{\nabla R}{2}=0,
\label{eqn:MA22_3}
\end{align}
which is called the reduced geodesic equation. It is easy to check that $\dot{\alpha}=V=\nabla l$.
The reduced volume is defined as
\begin{align}
\mathcal{V}((x,0), \bar{\tau})=\int_{X} (4\pi \bar{\tau})^{-\frac{m}{2}} e^{-l} dv.
\label{eqn:MA22_4}
\end{align}
It is proved by Perelman that $(4\pi \tau)^{-\frac{m}{2}} e^{-l}dv$, the reduced volume element, is
monotonically non-increasing along each reduced geodesic emanating from $(x,0)$.
Suppose the Ricci flow solution mentioned above is static, i.e., $Ric \equiv 0$. Then it is easy to check that
\begin{align}
\begin{cases}
&\mathcal{L}(\boldsymbol{\alpha})=\frac{d^2(x,y)}{2\sqrt{\bar{\tau}}}, \\
&l((x,0),(y,-\bar{\tau}))=\frac{d^2(x,y)}{4\bar{\tau}}, \\
&\nabla_V V +\frac{V}{2\tau}=0, \\
&|\dot{\alpha}|^2=|V|^2=|\nabla l|^2=\tau l, \\
&\mathcal{V}((x,0), \bar{\tau})=\int_{X} (4\pi \bar{\tau})^{-\frac{m}{2}} e^{-\frac{d^2}{4\bar{\tau}}} dv.
\end{cases}
\label{eqn:SL25_6}
\end{align}
Now we assume $X \in \widetilde{\mathscr{KS}}(n)$.
By a trivial extension in an extra time direction, we obtain a static, eternal singular K\"ahler Ricci flow solution.
Since distance structure is already known, we can define reduced distance, reduced volume, etc, following the equation
(\ref{eqn:SL25_6}). Clearly, this definition coincides with the original one when $X$ is smooth.
The following theorem is important to bridge the Cheeger-Colding's structure theory to the Ricci flow theory.
\begin{theorem}[\textbf{Volume ratio and reduced volume}]
Suppose $X \in \widetilde{\mathscr{KS}}(n)$, $x \in X$. Let $X \times (-\infty, 0]$ have the obvious
static space-time structure. Then we have
\begin{align}
&\mathrm{avr}(X)=\lim_{\tau \to \infty} \mathcal{V}((x,0), \tau). \label{eqn:SL25_4}\\
&\mathrm{v}(x)=\lim_{\tau \to 0} \mathcal{V}((x,0), \tau). \label{eqn:SL25_7}
\end{align}
\label{thm:SL25_1}
\end{theorem}
\begin{proof}
The proof relies on the volume cone structure at local tangent space, or tangent space at infinity. So
the proof of (\ref{eqn:SL25_4}) and (\ref{eqn:SL25_7}) are almost the same. For simplicity, we will only
prove (\ref{eqn:SL25_4}) and leave the proof for (\ref{eqn:SL25_7}) to the readers.
Clearly, the real dimension of $X$ is $m=2n$.
For each $\epsilon$ small, we have
\begin{align*}
m \omega_m \mathrm{avr}(X)+\epsilon
> H^{-m+1}|\partial B(x,H)| >m \omega_m \mathrm{avr}(X)-\epsilon,
\end{align*}
whenever $H$ is large enough. Note that
\begin{align*}
\mathcal{V}((x,0), H^2)&=(4\pi)^{-\frac{m}{2}} H^{-m} \int_0^{\infty} |\partial B(x,r)| e^{-\frac{r^2}{4H^2}} dr, \\
1&=(4\pi)^{-\frac{m}{2}} H^{-m} \int_0^{\infty} m \omega_m r^{m-1} e^{-\frac{r^2}{4H^2}} dr.
\end{align*}
So we have
\begin{align*}
\mathcal{V}((x,0), H^2)-\mathrm{avr}(X)
=(4\pi)^{-\frac{m}{2}} H^{-m} \int_0^{\infty} \left\{|\partial B(x,r)| -m\omega_m\mathrm{avr}(X) r^{m-1}\right\}
e^{-\frac{r^2}{4H^2}} dr.
\end{align*}
We can further decompose the last integral as follows.
\begin{align*}
&\quad \left| \int_0^{\epsilon H} \left\{|\partial B(x,r)| -m\omega_m\mathrm{avr}(X)r^{m-1} \right\}
e^{-\frac{r^2}{4H^2}} dr \right|\\
&\leq m\omega_m \int_0^{\epsilon H} r^{m-1} e^{-\frac{r^2}{4H^2}} dr
=m\omega_m H^m \int_0^{\epsilon} s^{m-1}e^{-\frac{s^2}{4}} ds, \\
&\quad \left| \int_{\epsilon H}^{\infty} \left\{|\partial B(x,r)| -m\omega_m\mathrm{avr}(X)r^{m-1} \right\}
e^{-\frac{r^2}{4H^2}} dr \right|\\
&\leq \epsilon \int_{\epsilon H}^{\infty} r^{m-1}e^{-\frac{r^2}{4H^2}} dr
<\epsilon H^m \int_{0}^{\infty} e^{-\frac{s^2}{4}} ds=\epsilon H^m \pi^{\frac{1}{2}}.
\end{align*}
Therefore, we have
\begin{align*}
\left|\mathcal{V}((x,0), H^2)-\mathrm{avr}(X)\right|
<(4\pi)^{-\frac{m}{2}} \left\{m\omega_m \int_0^{\epsilon} s^{m-1}e^{-\frac{s^2}{4}} ds + \epsilon \pi^{\frac{1}{2}}\right\}.
\end{align*}
Since the above inequality holds for every $H$ large enough, we see that
\begin{align*}
\left|\lim_{\tau \to \infty} \mathcal{V}((x,0),\tau)-\mathrm{avr}(X) \right|
<(4\pi)^{-\frac{m}{2}} \left\{m\omega_m \int_0^{\epsilon} s^{m-1}e^{-\frac{s^2}{4}} ds + \epsilon \pi^{\frac{1}{2}}\right\}.
\end{align*}
Let $\epsilon \to 0$, we obtain (\ref{eqn:SL25_4}).
\end{proof}
Theorem~\ref{thm:SL25_1} says that when we study the asymptotic behavior of $X$, the volume ratio and reduced volume
play the same role. Note that volume ratio is monotone along radius direction on a manifold with nonnegative Ricci curvature, which property plays an essential role in Cheeger-Colding's theory.
Since reduced volume is monotone along Ricci flow, Theorem~\ref{thm:SL25_1} suggests that Cheeger-Colding's theory can be transplanted to the Ricci flow case.
\section{Canonical radius}
In section 2, we established the compactness of the model space $\widetilde{\mathscr{KS}}(n,\kappa)$,
following the route of Anderson-Cheeger-Colding-Tian-Naber.
It is clear that the volume ratio's monotonicity is essential to this route. However, most K\"ahler manifolds do not have this monotonicity.
For example, if we take out a time slice from a K\"ahler Ricci flow solution, there is no obvious reason at all that volume ratio monotonicity holds on it.
Therefore, in order to set up weak compactness for general K\"ahler manifolds, we have to give up the volume ratio monotonicity and search for a new route.
This will be done in this section.
\subsection{Motivation and definition}
Let us continue the discussion in Section 2.6.
As a consequence of the weak compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$, we have
density estimate of volume radius, Proposition~\ref{prn:SB25_2}.
For simplicity of notation, we fix some $p_0$ very close to $2$, say $p_0=2-\frac{1}{1000n}$.
Define
\begin{align}
\mathbf{E} \triangleq E(n,\kappa,p_0)+200 \omega_{2n} \kappa^{-1}. \label{eqn:SC17_11}
\end{align}
Here we adjust the number $E(n,\kappa,p_0)$ to a much larger number, to reserve spaces for later use.
Then Proposition~\ref{prn:SB25_2} implies
\begin{align}
r^{2p_0-2n} \int_{B(x,r)} \mathbf{vr}(y)^{-2p_0} dy < \mathbf{E}. \label{eqn:SL23_5}
\end{align}
The above inequality contains a lot of information. For example, it immediately implies that in every unit ball,
there exists a fixed sized sub-ball with uniform regularity.
\begin{proposition}[\textbf{Generic regular sub-ball}]
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then we have
\begin{align}
\mathcal{F}_{c_b r} \cap B(x_0,r) \neq \emptyset, \label{eqn:SC29_10}
\end{align}
where
\begin{align}
c_b \triangleq \left(\frac{\omega_{2n} \kappa}{4 \mathbf{E}} \right)^{\frac{1}{2p_0}}.
\label{eqn:SL23_6}
\end{align}
\label{prn:HE11_1}
\end{proposition}
\begin{proof}
Let $\mathbf{vr}$ achieve maximum value at $y_0$ in the ball closure $\overline{B(x_0,r)}$.
By inequality (\ref{eqn:SL23_5}), we have
\begin{align*}
\mathbf{vr}(y_0)^{-2p_0} \leq \fint_{B(x,r)} \mathbf{vr}(y)^{-2p_0} dy
\leq (\omega_{2n} \kappa)^{-1} r^{-2n} \int_{B(x,r)} \mathbf{vr}(y)^{-2p_0} dy
\leq (\omega_{2n} \kappa)^{-1} r^{-2p_0} \mathbf{E}.
\end{align*}
It follows that
\begin{align*}
\mathbf{vr}(y_0) \geq \left(\frac{\omega_{2n}\kappa}{\mathbf{E}} \right)^{\frac{1}{2p_0}}r
>c_br.
\end{align*}
By continuity of $\mathbf{vr}$, there must exist a point $z \in B(x_0,r)$ such that $\mathbf{vr}(z) >c_b r$.
In other words, we have $z \in \mathcal{F}_{cr} \cap B(x_0,r)$. So (\ref{eqn:SC29_10}) holds.
\end{proof}
Let $\mathbf{E}$ and $c_b$ be the constants defined in (\ref{eqn:SC17_11}) and (\ref{eqn:SL23_6}).
Then we can choose a small constant $\epsilon_b$ such that
\begin{align}
\epsilon_b \triangleq \epsilon\left(n,\kappa,\frac{c_b}{100} \right)
\label{eqn:SL23_7}
\end{align}
by the dependence in (\ref{eqn:SC26_44}) of Proposition~\ref{prn:SC26_2}.
Combining the estimates in $\widetilde{\mathscr{KS}}(n,\kappa)$, we obtain the following theorem.
\begin{theorem}[\textbf{A priori estimates in model spaces}]
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then the following estimates
hold.
\begin{enumerate}
\item Strong volume ratio estimate: $\kappa \leq \omega_{2n}^{-1}r^{-2n}|B(x_0,r)| \leq 1$.
\item Strong regularity estimate: $r^{2+k}|\nabla^k Rm|\leq c_a^{-2}$ in the ball $B(x_0, c_a r)$ for every $0 \leq k \leq 5$ whenever $\mathbf{vr}(x_0) \geq r$.
\item Strong density estimate: $\displaystyle r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}(y)^{-2p_0} dy \leq \mathbf{E}$.
\item Strong connectivity estimate: Every two points $y_1,y_2 \subset B(x_0,r) \cap \mathcal{F}_{\frac{1}{100}c_b r}(X)$
can be connected by a shortest geodesic $\gamma$ such that
$\gamma \subset \mathcal{F}_{\epsilon_b r} (X)$.
\end{enumerate}
\label{thm:SL21_1}
\end{theorem}
We shall show that a weak compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$ can be established using the estimates in Theorem~\ref{thm:SL21_1},
without knowing the volume ratio monotonicity.
For this new route of weak compactness theory, we define a scale called canonical radius with respect to $\widetilde{\mathscr{KS}}(n,\kappa)$.
Under the canonical radius, rough estimates like that in Theorem~\ref{thm:SL21_1} are satisfied.
In this section, we focus on the study of smooth complete K\"ahler manifold.
Every such a manifold is denoted by $(M^n, g, J)$, where $n$ is the complex dimension.
The Hausdorff dimension, or real dimension of $M$ is $m=2n$.
We first need to make sense of the rough volume radius, without the volume ratio monotonicity.
\begin{definition}
Denote the set $\left\{ r \left| 0<r<\rho, \omega_{2n}^{-1}r^{-2n}|B(x_0,r)|\geq 1-\delta_0 \right. \right\}$ by
$I_{x_0}^{(\rho)}$ where $x_0 \in M$, $\rho$ is a positive number.
Clearly, $I_{x_0}^{(\rho)} \neq \emptyset$ since $M$ is smooth. Define
\begin{align*}
\mathbf{vr}^{(\rho)}(x_0) \triangleq \sup I_{x_0}^{(\rho)}.
\end{align*}
For each pair $0<r \leq \rho$, define
\begin{align*}
&\mathcal{F}_{r}^{(\rho)}(M) \triangleq \left\{ x \in M | \mathbf{vr}^{(\rho)}(x) \geq r \right\}, \\
&\mathcal{D}_{r}^{(\rho)}(M) \triangleq \left\{ x \in M | \mathbf{vr}^{(\rho)}(x) < r \right\}.
\end{align*}
\label{dfn:SC24_1}
\end{definition}
\begin{definition}
A subset $\Omega$ of $M$ is called $\epsilon$-regular-connected on the scale $\rho$ if every two points $x,y \in \Omega$ can be connected by a
rectifiable curve $\gamma \subset \mathcal{F}_{\epsilon}^{(\rho)}$ and $|\gamma| < 2d(x,y)$.
For notational simplicity, if the scale is clear in the context, we shall just say
$\Omega$ is $\epsilon$-regular-connected.
\label{dfn:SC15_3}
\end{definition}
Inspired by the estimates in Theorem~\ref{thm:SL21_1}, we can define the concept of canonical radius as follows.
\begin{definition}
We say that the canonical radius (with respect to model space $\widetilde{\mathscr{KS}}(n,\kappa)$) of a point $x_0 \in M$ is
not less than $r_0$ if for every $r < r_0$, we have the following properties.
\begin{enumerate}
\item Volume ratio estimate: $\kappa \leq \omega_{2n}^{-1}r^{-2n}|B(x_0,r)| \leq \kappa^{-1}$.
\item Regularity estimate: $r^{2+k}|\nabla^k Rm|\leq 4 c_a^{-2}$ in the ball $B(x_0, \frac{1}{2}c_a r)$ for every $0 \leq k \leq 5$ whenever
$\omega_{2n}^{-1}r^{-2n}|B(x_0,r)| \geq 1-\delta_0$.
\item Density estimate: $\displaystyle r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}^{(r)}(y)^{-2p_0} dy \leq 2\mathbf{E}$.
\item Connectivity estimate: $B(x_0,r) \cap \mathcal{F}_{\frac{1}{50}c_b r}^{(r)}(M)$ is $\frac{1}{2}\epsilon_b r$-regular-connected on the scale $r$.
\end{enumerate}
Then we define canonical radius of $x_0$ to be the supreme of all the $r_0$ with the properties mentioned above.
We denote the canonical radius by $\mathbf{cr}(x_0)$.
For subset $\Omega \subset M$, we define the canonical radius of $\Omega$ as the infimum of all $\mathbf{cr}(x)$ where $x \in \Omega$.
We denote this canonical radius by $\mathbf{cr}(\Omega)$.
\label{dfn:SC02_1}
\end{definition}
\begin{remark}
In Definition~\ref{dfn:SC02_1}, the first condition(volume ratio estimate) is used to guarantee the existence of Gromov-Hausdorff limit. The second condition (regularity estimate) is for the purpose of improving regularity. The third condition (density estimate), together with the second condition(regularity estimate),
implies that the regular part is almost dense(c.f.~Theorem~\ref{thm:HE11_1}). The fourth condition(connectivity estimate) is defined to assure that the regular part is connected (c.f.~Proposition~\ref{prn:SB27_1}).
\label{rmk:SB16_1}
\end{remark}
Because of the regularity estimate of Definition~\ref{dfn:SC02_1}, it is useful to define the concept of canonical volume radius as follows.
\begin{definition}
Suppose $\rho_0=\mathbf{cr}(x_0)$. Then we define
\begin{align}
\mathbf{cvr}(x_0) \triangleq \mathbf{vr}^{(\rho_0)}(x_0). \label{eqn:SC24_4}
\end{align}
We call $\mathbf{cvr}(x_0)$ the canonical volume radius of the point $x_0$.
\label{dfn:SB25_2}
\end{definition}
\begin{remark}
For every compact smooth manifold $M$, there is an $\eta>0$ such that
every geodesic ball with radius less than $\eta$ must have normalized volume radius at least $1-\delta_0$.
Then it is easy to see that $\displaystyle r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}^{(r)}(y)^{-2p_0} dy$ is a continuous
function with respect to $x_0$ and $r$. Therefore, if $\rho_0=\mathbf{cr}(x_0)$ is a finite positive number, we have
\begin{align}
\displaystyle \rho_0^{2p_0-2n} \int_{B(x_0, \rho_0)} \mathbf{vr}^{(\rho_0)}(y)^{-2p_0} dy \leq 2\mathbf{E}.
\label{eqn:SL25_1}
\end{align}
If $r \leq \mathbf{cr}(M)$, then $\mathbf{vr}^{(r)} \leq \mathbf{cvr}$ as functions. Therefore, we have
\begin{align}
r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{cvr}(y)^{-2p_0} dy \leq
r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}^{(r)}(y)^{-2p_0} dy \leq 2\mathbf{E}.
\label{eqn:SL25_2}
\end{align}
\label{rmk:SL22_1}
\end{remark}
Let $r_0$ be $\mathbf{cvr}(x_0)$. By Definition~\ref{dfn:SB25_2}, it is clear that $r_0 \leq \mathbf{cr}(x_0)$.
If $r_0=\mathbf{cvr}(x_0)<\mathbf{cr}(x_0)$, then we have
\begin{align}
&\omega_{2n}^{-1}r_0^{-2n}|B(x_0,r_0)| = 1-\delta_0, \label{eqn:SC16_2}\\
&\omega_{2n}^{-1}r^{-2n}|B(x_0,r)| < 1-\delta_0, \quad \forall \; r \in (r_0, \mathbf{cr}(x_0)). \label{eqn:SC16_3}
\end{align}
If $r_0=\mathbf{cvr}(x_0)=\mathbf{cr}(x_0)$, then we only have
\begin{align}
\omega_{2n}^{-1} r_0^{-2n}|B(x_0,r_0)| \geq 1-\delta_0. \label{eqn:SC17_1}
\end{align}
It is possible that equality (\ref{eqn:SC16_2}) does not hold on the scale $r_0$ in this case.
\begin{remark}
The three radii functions, $\mathbf{cr}$,$\mathbf{vr}$ and $\mathbf{cvr}$ are all positive functions on the interior part of $M$.
However, we do not know whether they are continuous in general.
\label{rmk:DA25_1}
\end{remark}
We shall use canonical radius as a tool to study the weak-compactness theory of K\"ahler manifolds.
\subsection{Rough estimates when canonical radius is bounded from below}
We assume $\mathbf{cr}(M) \geq 1$ in the following discussion of this subsection.
Under this condition, we collect important estimates for the development of weak-compactness.
For simplicity of notation, we denote
\begin{align}
\mathcal{F}_{r} \triangleq \mathcal{F}_{r}^{(\mathbf{cr}(M))}, \quad
\mathcal{D}_{r} \triangleq \mathcal{D}_{r}^{(\mathbf{cr}(M))}.
\label{eqn:SL27_1}
\end{align}
Note that this definition can be regarded as the generalization of the corresponding definition for metric spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$.
It coincides the original one since $\mathbf{cr}(M)=\infty$ whenever $M \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{proposition}
For every $0<r\leq \rho_0 \leq 1$, $x_0 \in M$, we have
\begin{align}
&\left| B(x_0,\rho_0) \cap \mathcal{D}_r \right| < 4\mathbf{E} \rho_0^{2n-2p_0} r^{2p_0}, \label{eqn:SC02_2} \\
&\left| B(x_0,\rho_0) \cap \mathcal{F}_r \right| > \left(\kappa \omega_{2n} - 4\mathbf{E}r^{2p_0}\rho_0^{-2p_0} \right) \rho_0^{2n}.
\label{eqn:SC02_4}
\end{align}
In particular, there exists at least one point $z \in B(x_0,\rho_0)$ such that
\begin{align}
\mathbf{cvr}(z) > c_b \rho_0,
\label{eqn:SC02_5}
\end{align}
where $c_b=\left(\frac{\kappa \omega_{2n}}{4 \mathbf{E}} \right)^{\frac{1}{2p_0}}$.
\label{prn:SC02_2}
\end{proposition}
\begin{proof}
Recall that $\mathbf{vr}^{(\mathbf{cr}(M))} \geq \mathbf{vr}^{(\rho_0)}$. By density estimate(c.f. Definition~\ref{dfn:SC02_1}), we have
\begin{align*}
r^{-2p_0} \left|B(x_0,\rho_0) \cap \mathcal{D}_r \right|
\leq \int_{B(x_0,\rho_0) \cap \mathcal{D}_{r}}\left\{ \mathbf{vr}^{(\mathbf{cr}(M))}\right\}^{-2p_0}
\leq \int_{B(x_0,\rho_0)} \left\{ \mathbf{vr}^{(\rho_0)}\right\}^{-2p_0} \leq 2\mathbf{E} \rho_0^{2n-2p_0}.
\end{align*}
Then (\ref{eqn:SC02_2}) follows from above inequality. Recall that $\mathcal{D}_r$ is the set where $\mathbf{vr}^{(\mathbf{cr}(M))}<r$.
Together with the $\kappa$-non-collapsing condition, (\ref{eqn:SC02_2}) yields (\ref{eqn:SC02_4}).
Let $r=c_b \rho_0$, then (\ref{eqn:SC02_2}) implies
\begin{align*}
\left| B(x_0,\rho_0) \cap \mathcal{F}_{c_b \rho_0} \right| >0.
\end{align*}
In particular, $B(x_0,\rho_0) \cap \mathcal{F}_{c_b \rho_0} \neq \emptyset$. In other words, we can find a point $z \in B(x_0, \rho_0)$ satisfying
$\mathbf{vr}^{(\mathbf{cr}(M))}>c_b \rho_0$ and consequently inequality (\ref{eqn:SC02_5}).
\end{proof}
\begin{corollary}
Suppose $x_0 \in M$, $H \geq 1 \geq r$, then we have
\begin{align}
\left| B(x_0,H) \cap \mathcal{D}_r \right| \leq \left(\frac{2^{2n+2}\left|B(x_0, 2H) \right|}{\kappa \omega_{2n}} \right) r^{2p_0} \mathbf{E}.
\label{eqn:SC06_1}
\end{align}
\label{cly:SC04_2}
\end{corollary}
\begin{proof}
Try to fill the ball $B(x_0,H)$ with balls $B(y_i,\frac{1}{2})$ such that $y_i \in B(x_0,H)$ until no more such balls can squeeze in. Clearly, we have
\begin{align*}
B(x_0,H) \subset \bigcup_{i=1}^{N} B(y_i,1), \quad \bigcup_{i=1}^{N} B \left(y_i, \frac{1}{2} \right) \subset B \left(x_0,H+\frac{1}{2} \right) \subset B(x_0,2H).
\end{align*}
On one hand, the balls $B\left(y_i,\frac{1}{2}\right)$ are disjoint to each other. So we have
\begin{align}
N\kappa \omega_{2n} \left( \frac{1}{2}\right)^{2n} \leq \sum_{i=1}^{N} \left|B(y_i,1)\right|
\leq \left|B(x_0,2H)\right|, \quad
\Rightarrow \quad N \leq \frac{2^{2n} |B(x_0,2H)|}{\kappa \omega_{2n}}.
\label{eqn:SC06_2}
\end{align}
On the other hand, $B(x_0,H)$ is covered by $\displaystyle \bigcup_{i=1}^{N} B(y_i,1)$. So we have
\begin{align*}
\left| B(x_0,H) \cap \mathcal{D}_r \right| \leq \sum_{i=1}^{N} \left| B(y_i,1) \cap \mathcal{D}_r \right| \leq 4N\mathbf{E} r^{2p_0} \leq \left(\frac{2^{2n+2}\left|B(x_0, 2H) \right|}{\kappa \omega_{2n}} \right) r^{2p_0} \mathbf{E},
\end{align*}
where we used (\ref{eqn:SC06_2}) and (\ref{eqn:SC02_2}).
\end{proof}
\begin{proposition}
For every $r \leq 1$, two points $x,y \in \mathcal{F}_{r}$ can be connected by a curve $\gamma \subset \mathcal{F}_{\frac{1}{2}\epsilon_b r}$ with length $|\gamma| <3d(x,y)$.
\label{prn:SB27_1}
\end{proposition}
\begin{proof}
By rescaling if necessary, we can assume $r=1$. Then $\mathbf{cr}(M) \geq 1$.
Suppose $x,y \in \mathcal{F}_{1}$. If $d(x,y) \leq 1$, then there is a curve connecting $x,y$ and it satisfies the requirements, by the $\frac{1}{2}\epsilon_b$-regular connectivity property of the canonical radius. So we assume $H=d(x,y) >1$ without loss of generality.
Let $\beta$ be a shortest geodesic connecting $x,y$ such that $\beta(0)=x$ and $\beta(H)=y$.
Let $N$ be an integer locating in $[2H, 2H+1]$. Define
\begin{align*}
s_i=\frac{Hi}{N}, \quad x_i=\beta(s_i).
\end{align*}
Clearly, $x_0=x, x_N=y$, which are both in $\mathcal{F}_1 \subset \mathcal{F}_{\frac{c_b}{50}}$.
For each $1 \leq i \leq N-1$, $x_i$ may not locate in $\mathcal{F}_{\frac{c_b}{50}}$. However,
in the ball $B(x_i, \frac{1}{20})$, there exists a point $x_i'$ such that
\begin{align*}
\mathbf{vr}(x_i') \geq \frac{1}{2} c_b \cdot \frac{1}{20}=\frac{c_b}{40}> \frac{c_b}{50}.
\end{align*}
Clearly, we have
\begin{align*}
d(x_i', x_{i+1}') \leq d(x_i', x_i) + d(x_i, x_{i+1}) + d(x_{i+1}, x_{i+1}') \leq \frac{H}{N} + \frac{1}{10} < \frac{3}{5} \leq 1.
\end{align*}
Since $\mathbf{cr}(M) \geq 1$, one can apply $\frac{1}{2}\epsilon_b$-regular connectivity property of the canonical radius
to find a curve $\beta_i$ connecting $x_i'$ and $x_{i+1}'$ such that
$\beta_i \subset \mathcal{F}_{\frac{1}{2}\epsilon_b}$. Moreover, we have
\begin{align*}
|\beta_i| \leq 2d(x_i', x_{i+1}') \leq 2\left(\frac{H}{N} + \frac{1}{10} \right).
\end{align*}
Concatenating all $\beta_i$'s, we obtain a curve $\gamma$ connecting $x=x_0,y=x_{N}$ and $\gamma \subset \mathcal{F}_{\frac{1}{2} \epsilon_b}$.
Furthermore, we have
\begin{align*}
|\gamma|=\sum_{i=0}^{N-1} |\beta_i| \leq 2N \left( \frac{H}{N} + \frac{1}{10} \right)= 2H +\frac{1}{15}N \leq 2H + \frac{2H+1}{5} \leq \frac{12}{5}H + \frac{1}{5}
< \frac{13}{5}H <3H.
\end{align*}
\end{proof}
\begin{corollary}
For every $x \in M$, $0<r \leq 1$, we can find a curve $\gamma$ connecting $\partial B(x, \frac{r}{2})$
and $\partial B(x,r)$ such that
\begin{align*}
\gamma \subset \mathcal{F}_{\frac{1}{2}\epsilon_br}, \quad |\gamma| \leq 2r.
\end{align*}
In particular, we have
\begin{align*}
\partial B(x, r) \cap \mathcal{F}_{\frac{\epsilon_b}{2}r} \neq \emptyset.
\end{align*}
\label{cly:SL24_1}
\end{corollary}
\begin{proof}
Let $\beta$ be a shortest geodesic connecting $x$ and some point $y \in \partial B(x, \frac{9}{8}r)$.
Let $z$ be the intersection point of $\beta$ and $\partial B(x, \frac{3}{8}r)$. Let $y', z'$ be regular points
around $y,z$, i.e., we require
\begin{align*}
y' \in B\left(y,\frac{r}{8} \right) \cap \mathcal{F}_{\frac{c_b}{8}r}, \quad
z' \in B\left(z,\frac{r}{8} \right) \cap \mathcal{F}_{\frac{c_b}{8}r}.
\end{align*}
Clearly, triangle inequality implies that
\begin{align*}
d(y',z') \leq \frac{6}{8}r + \frac{1}{8}r + \frac{1}{8}r=r \leq 1.
\end{align*}
Since $\mathbf{cr}(M) \geq 1$, by connectivity estimate, there is a curve $\alpha$ connecting $y'$ and $z'$ such that
\begin{align*}
|\alpha| \leq 2r, \quad \alpha \subset \mathcal{F}_{\frac{\epsilon_b}{2}r}.
\end{align*}
Note that $z' \in B(x,\frac{r}{2})$ and $y' \in B(x,r)^{c}$. The connectedness of $M$ guarantees that
$\alpha$ must have intersection with both $\partial B(x,\frac{r}{2})$ and $\partial B(x,r)$.
So we can truncate $\alpha$ to obtain a curve $\gamma$ which connects $\partial B(x,\frac{r}{2})$ and $\partial B(x,r)$. Clearly, we have
\begin{align*}
\gamma \subset \alpha \subset \mathcal{F}_{\frac{\epsilon_b}{2}r},
\quad |\gamma| \leq |\alpha| \leq 2r.
\end{align*}
\end{proof}
\begin{proposition}
Suppose $x \in M$, $0<r \leq 1$. Then for every point $y \in \mathcal{F}_{\frac{1}{2}\epsilon_br} \cap \partial B(x,r)$.
There is a curve $\gamma$ connecting $x$ and $y$ such that
\begin{itemize}
\item $|\gamma|<10r$.
\item For each nonnegative integer $i$,
$\gamma \cap B(x,2^{-i}r) \backslash B\left(x, 2^{-i-1}r \right)$ contains a component which connects
$\partial B(x,2^{-i}r)$ and $\partial B(x, 2^{-i-1}r)$ and is contained in
$\mathcal{F}_{2^{-i-3}\epsilon_b^2r}$.
\end{itemize}
\label{prn:SL24_1}
\end{proposition}
\begin{proof}
Choose $y_i$ be a point on $\partial B(x, 2^{-i}r) \cap \mathcal{F}_{2^{-i-1}\epsilon_b r}$.
By Proposition~\ref{prn:SB27_1}, for each $i \geq 0$, there is a curve $\gamma_i$ connecting
$y_i$ and $y_{i+1}$ such that
\begin{align*}
|\gamma_i| < 9 \cdot 2^{-i-1} r, \quad \gamma_i \subset \mathcal{F}_{2^{-i-3}\epsilon_b^2 r}.
\end{align*}
Concatenate all the $\gamma_i$'s to obtain $\gamma$. Then $\gamma$ satisfies all the properties.
\end{proof}
For the purpose of improving regularity, we need to study the behavior of $\mathbf{cvr}$.
Similar to $\mathbf{vr}$ on spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$ (c.f. Proposition~\ref{prn:SC26_7}),
$\mathbf{cvr}$ satisfies a local Harnack inequality.
\begin{proposition}
There is a constant $K=K(n,\kappa)$ with the following properties.
Suppose $x \in X$, $r=\mathbf{cvr}(x)<\frac{1}{K}$, then for every point $y \in B(x, K^{-1}r)$, we have
\begin{align}
&K^{-1}r \leq \mathbf{cvr}(y) \leq Kr, \label{eqn:SC24_1}\\
&\omega_{2n}^{-1}\rho^{-2n} |B(y,\rho)| \geq 1-\frac{1}{100}\delta_0, \quad \forall \; \rho \in (0, K^{-1}r), \label{eqn:SC24_2}\\
& |Rm|(y) \leq K^{2} r^{-2}, \label{eqn:SC29_6}\\
& inj(y) \geq K^{-1} r. \label{eqn:SC29_7}
\end{align}
\label{prn:SC24_1}
\end{proposition}
\begin{corollary}
For every $r \in (0,1]$, $\mathcal{F}_{r}(M)$ is a closed set. Moreover,
$\mathbf{cvr}$ is an upper-semi-continuous function on $\mathcal{F}_{r}(M)$.
\label{cly:SC24_1}
\end{corollary}
\begin{proof}
Fix $r \in (0,1]$. Suppose $x_i \in \mathcal{F}_{r}(M)$ converges to a point $x \in M$.
Let $r_i=\mathbf{cvr}(x_i)$. We need to show $\mathbf{cvr}(x) \geq r$. Clearly, this follows directly
if $\displaystyle \lim_{i \to \infty} r_i=\infty$ by Proposition~\ref{prn:SC24_1}. Without loss of generality, we may assume
$r_i$ is uniformly bounded from above. Use Proposition~\ref{prn:SC24_1} again, we see that
$r_i$ is uniformly bounded away from zero. Let $r$ be a limit of $r_i$. Then we have
\begin{align}
|B(x,r)|=\lim_{i \to \infty} |B(x_i,r_i)| \geq \lim_{i \to \infty} (1-\delta_0) \omega_{2n} r_i^{2n}
= (1-\delta_0) \omega_{2n} r^{2n},
\end{align}
which implies $\displaystyle \mathbf{cvr}(x) \geq r= \lim_{i \to \infty} r_i$ by definition of canonical volume radius and the fact that
$\displaystyle r=\lim_{i \to \infty} r_i\leq 1 \leq \mathbf{cvr}$.
Consequently, we have $x \in \mathcal{F}_{r}(M)$. Therefore, $\mathcal{F}_{r}(M)$ is a closed set by the arbitrary choice of $\{x_i\}$.
From the above argument, we have already seen that
\begin{align}
\mathbf{cvr}(x) \geq \lim_{i \to \infty} \mathbf{cvr}(x_i),
\end{align}
which means that $\mathbf{cvr}$ is an upper-continuous function on $\mathcal{F}_{r}(M)$.
\end{proof}
Clearly, the conclusion in the above corollary is weaker than that in Proposition~\ref{prn:HE07_2}, since here we do not have a rigidity property like Proposition~\ref{prn:SC17_3}. However, even if
$\displaystyle \mathbf{cvr}(x)>\lim_{i \to \infty} \mathbf{cvr}(x_i)$, the local Harnack inequality of $\mathbf{cvr}$ guarantees that $\displaystyle \mathbf{cvr}(x) < K \lim_{i \to \infty} \mathbf{cvr}(x_i)$. So $\mathbf{cvr}$ is better than
general semi-continuous function. For example, in the decomposition $M=\mathcal{F}_r \cup \mathcal{D}_r$,
every point $y \in \partial \mathcal{F}_r$ satisfies $r \leq \mathbf{cvr}(y) \leq Kr$. In many situations, it is convenient to just regard $K=1$, i.e., $\mathbf{cvr}$ being continuous, without affecting the effectiveness of the argument.
Furthermore, up to perturbation, one can even regard $\mathbf{cvr}$ as smooth functions. Full details of the perturbation can be found in Appendix~\ref{app:B}.
\subsection{K\"ahler manifolds with canonical radius bounded from below}
Similar to the traditional theory, volume convergence is very important. However, in the current situation, the volume convergence can be proved in a much easier way.
\begin{proposition}[\textbf{Volume convergence}]
Suppose $(M_i,g_i,J_i)$ is a sequence of K\"ahler manifolds satisfying $\mathbf{cr}(M_i) \geq r_0$. Then
we have \begin{align*}
(M_i, x_i, g_i) \longright{G.H.} (\bar{M}, \bar{x}, \bar{g}).
\end{align*}
Moreover, the volume (2n-dimensional Gromov-Hausdorff measure) is continuous under this convergence, i.e., for every fixed $\rho_0>0$, we have
\begin{align*}
|B(\bar{x}, \rho_0)|= \lim_{i \to \infty} |B(x_i,\rho_0)|.
\end{align*}
\label{prn:SC12_1}
\end{proposition}
\begin{proof}
The existence of the Gromov-Hausdorff limit space follows from the volume doubling property and the standard ball-packing argument. Fix $r<<\rho_0$, then it follows from the definition of $\mathcal{F}_r$ that the convergence
on $B(x_i, \rho_0) \cap \mathcal{F}_r$ can be improved to $C^4$-topology. Then the volume converges trivially on this part. On the other hand, the volume of $B(x_i, \rho_0) \cap \mathcal{D}_r$ is bounded by $C r^{2p_0}$, which tends to
zero as $r \to 0$. So the volume convergence of geodesic balls $B(x_i,\rho_0)$ follows from the combination of the two factors mentioned above. More details are given as follows.
Let $\left( \bar{M}, \bar{x}, \bar{g} \right)$ be
the limit space. For each $r \leq r_0$, define
\begin{align}
&\mathcal{R}_{r} \triangleq \left\{ \bar{y} \in \bar{M}
\left| \textrm{There exists} \; y_i \in M_i \; \textrm{such that} \; y_i \to \bar{y} \; \textrm{and} \; \liminf_{i \to \infty} \mathbf{cvr}(y_i) \geq r \right. \right\}, \label{eqn:HA11_3}\\
&\mathcal{S}_{r} \triangleq \left( \mathcal{R}_{r} \right)^{c}, \label{eqn:HA11_4}\\
&\mathcal{R}' \triangleq \bigcup_{0<r\leq r_0} \mathcal{R}_{r}, \label{eqn:GD25_2} \\
&\mathcal{S}' \triangleq \bigcap_{0<r\leq r_0} \mathcal{S}_{r}. \label{eqn:GD25_3}
\end{align}
We now show that $\mathcal{S}'$ is a subset of $\bar{M}$ of Minkowski dimension at most $2n-2p_0$.
Without loss of generality, it suffices to show this dimension for $\mathcal{S}' \cap B(\bar{x}, \rho_0)$.
For each small $r>0$, we shall construct a covering for the set $\mathcal{S}_r \cap B(\bar{x}, \rho_0)$. Clearly, the choice
$\cup_{z \in \mathcal{S}_r \cap B(\bar{x}, \rho_0)} B(z, r)$ is a cover, but with uncoutable many balls.
By Vitali covering lemma, we can find countable many $z_k$'s such that $\cup_{z_k} B(z_k, r)$ is a disjoint union and
\begin{align}
\mathcal{S}_r \cap B(\bar{x}, \rho_0) \subset \bigcup_{z_k} B(z_k, 5r). \label{eqn:GD25_1}
\end{align}
We shall show that this covering is actually a finite covering with number of balls $N$ uniformly bounded by $Cr^{2p_0-2n}$.
Let $z_k$ be the limit point of $z_{k,i} \in M_i$. For large $i$, it follows from definition that $\mathbf{cvr}(z_{k,i}) <2r$.
By Proposition~\ref{prn:SC24_1}, we see that $B(z_{k,i}, 5r) \subset \mathcal{D}_{5Kr}$. It follows that
\begin{align*}
\bigcup_{k=1}^{} B(z_{k,i}, 0.5 r) \subset \bigcup_{k} B(z_{k,i}, 5r) \subset \left\{ B(x_i, 2\rho_0) \cap \mathcal{D}_{5Kr} \right\}.
\end{align*}
Note that $\bigcup_{k} B(z_{k,i}, 0.5 r)$ is a disjoint union.
Taking volume on the manifold $M_i$, using the volume ratio's lower bound and Proposition~\ref{prn:SC02_2}, we obtain
\begin{align*}
N \kappa \omega_{2n} (0.5 r)^{2n} \leq \sum_{k} |B(z_{k,i}, 0.5r)| \leq |B(x_i, 2\rho_0) \cap \mathcal{D}_{5Kr}| \leq C (5Kr)^{2p_0}.
\end{align*}
It follows that $N \leq Cr^{2p_0-2n}$ for some uniform constant $C$.
Therefore, the covering we choose in (\ref{eqn:GD25_1}) is a finite covering with the number of balls dominated by $Cr^{2p_0-2n}$.
Since $\mathcal{S}'$ is a subset of $\mathcal{S}_r$, we obtain a covering of $\mathcal{S}' \cap B(\bar{x},\rho_0)$ by size-$r$ balls with number at most $Cr^{2p_0-2n}$,
where $C$ is independent of $r$. Therefore, we have
\begin{align}
\dim_{\mathcal{M}} \left\{ \mathcal{S}' \cap B(\bar{x}, \rho_0) \right\} \leq 2n-2p_0.
\label{eqn:GD25_4}
\end{align}
In particular, $\mathcal{S}' \cap B(\bar{x},\rho_0)$ has $2n$-Hausdorff measure zero, or volume zero. This means we can ignore the effect of $\mathcal{S}'$ when we consider
volume convergence. On the other hand, away from $\mathcal{S}'$, the volume convergence is obvious. We therefore obtain the volume convergence property
whenever $B(x_i,\rho_0)$ converges to $B(\bar{x},\rho_0)$.
\end{proof}
Now we are able to show the weak compactness theorem.
\begin{theorem}[\textbf{Rough weak compactness}]
Same conditions as Proposition~\ref{prn:SC12_1}.
Denote $\mathcal{R} \subset \bar{M}$ as the set of regular points, i.e., the points with some small neighborhoods which have $C^4$-Riemannian manifolds structure.
Denote $\mathcal{S} \subset \bar{M}$ be the set of singular points, i.e., the points which are not regular.
Then we have the regular-singular decomposition $\bar{M}=\mathcal{R} \cup \mathcal{S}$ with the following properties.
\begin{itemize}
\item The regular part $\mathcal{R}$ is an open, path connected $C^4$-Riemannian manifold.
Furthermore, for every two points $x,y \in \mathcal{R}$, there exists a curve $\gamma$ connecting $x,y$ satisfying
\begin{align}
\gamma \subset \mathcal{R}, \quad |\gamma| \leq 3d(x,y). \label{eqn:HE11_1}
\end{align}
\item The singular part $\mathcal{S}$ satisfies the Minkowski dimension estimate
\begin{align}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-2p_0.
\label{eqn:SB13_4}
\end{align}
\end{itemize}
\label{thm:HE11_1}
\end{theorem}
\begin{proof}
Let $\left( \bar{M}, \bar{x}, \bar{g} \right)$ be
the limit space. For each $r \leq r_0$, define $\mathcal{R}_r$, $\mathcal{S}_r$ as in (\ref{eqn:HA11_3}) and (\ref{eqn:HA11_4}).
Define $\mathcal{R}', \mathcal{S}'$ as in (\ref{eqn:GD25_2}) and (\ref{eqn:GD25_3}).
Recall that the regular set $\mathcal{R} \subset \bar{M}$ is defined as the collection of points which have small neighborhoods with manifolds structure.
We shall show that $\mathcal{R}'$ is nothing but $\mathcal{R}$, i.e., $\displaystyle \mathcal{R} = \bigcup_{0<r\leq r_0} \mathcal{R}_{r}$.
Actually, by regularity estimate property of canonical radius, for every fixed $r \in (0,r_0)$, every point
$\bar{y} \in \mathcal{R}_{r}$, we see that the convergence to $B(\bar{y}, \frac{1}{3}c_a r)$ can be improved to be in the $C^{4}$-topology. Clearly, $B(\bar{y}, \frac{1}{3}c_a r)$ has a manifold structure. So $\mathcal{R}_r \subset \mathcal{R}$. Let $r \to 0$, we have $\displaystyle \bigcup_{0<r\leq r_0} \mathcal{R}_{r} \subset \mathcal{R}$.
On the other hand, suppose $\bar{y} \in \mathcal{R}$. Then there is a ball $B(\bar{y},r)$ with a manifold structure.
By shrinking $r$ if necessary, we can assume that the volume ratio of this ball is very close to the Euclidean one.
Note that the volume ($2n$-dimensional Hausdorff measure) converges when $(M_i, x_i, g_i)$ converges to $(\bar{M}, \bar{x}, \bar{g})$.
Suppose $y_i \to \bar{y}, y_i \in M_i$. Then we have $\omega_{2n}^{-1}r^{-2n}|B(y_i,r)|>1-\delta_0$ for large $i$. By definition, this means that
$\mathbf{cvr}(y_i,0) \geq r$. It follows from the regularity estimates that
$\displaystyle \bar{y} \in \mathcal{R}_{r} \subset \mathcal{R}'$.
By the arbitrary choice of $\bar{y}$, we obtain $\mathcal{R} \subset \mathcal{R}'$.
So we finish the proof of
\begin{align*}
\mathcal{R}=\mathcal{R}'=\bigcup_{0<r\leq r_0} \mathcal{R}_{r}.
\end{align*}
Combining the above equation with the definitions in (\ref{eqn:GD25_2}) and (\ref{eqn:GD25_3}), we have $\mathcal{S}=\mathcal{S}'$. Therefore, (\ref{eqn:SB13_4}) follows from $\dim_{\mathcal{M}} \mathcal{S}' \leq 2n-2p_0$,
which can be proved following (\ref{eqn:GD25_4}). Alternatively, we can prove (\ref{eqn:SB13_4}) as follows.
Fix $r<r_0$. Let $\rho_0=r_0$ and take limit of (\ref{eqn:SC02_2}), we obtain
\begin{align}
|B(\bar{y}, r_0) \cap \mathcal{S}_{r}| \leq 4\mathbf{E}r_0^{2n-2p_0} r^{2p_0}, \label{eqn:B10_1}
\end{align}
for every $\bar{y} \in \bar{M}$. Suppose $y \in \mathcal{R}_r \subset \bar{M}$.
The regularity estimate property of canonical radius yields that every point in $B(y, \frac{1}{4}c_a r)$ is regular.
So $d(y,\mathcal{S})\geq \frac{1}{4}c_a r$.
It follows that
\begin{align*}
\mathcal{R}_{r} \subset \left\{x \in \bar{M} \left| d(x, \mathcal{S}) \geq \frac{1}{4}c_a r \right.\right\}
\Leftrightarrow \mathcal{S}_{r} \supset \left\{x \in \bar{M} \left| d(x, \mathcal{S}) <\frac{1}{4}c_a r \right. \right\}.
\end{align*}
Therefore, we have
\begin{align}
\{x \in \bar{M} | d(x, \mathcal{S}) <r\} \subset \mathcal{S}_{4c_a^{-1} r}, \label{eqn:SB13_1}
\end{align}
whenever $r$ is very small. Combining (\ref{eqn:B10_1}) and (\ref{eqn:SB13_1}) yields
\begin{align*}
\left| B(\bar{y},r_0) \cap \left\{x \in \bar{M} | d(x, \mathcal{S}) <r \right\} \right| \leq 4^{2p_0+1}\mathbf{E}c_a^{-2p_0} r_0^{2n-2p_0} r^{2p_0}=C r^{2p_0}.
\end{align*}
Since the above inequality holds for every small $r$ and every $\bar{y} \in \bar{M}$, it yields (\ref{eqn:SB13_4}) directly.
It follows from definition that $\mathcal{R}$ is an open $C^4$-manifold. The path connectedness of $\mathcal{R}$ follows from (\ref{eqn:HE11_1}).
Now we proceed to show (\ref{eqn:HE11_1}). Fix $x,y \in \mathcal{R}$, let $r=\sup\{\rho | x \in \mathcal{R}_{\rho}, y \in \mathcal{R}_{\rho}\}$.
Since $r>0$, we can choose sequence $x_i, y_i \in \mathcal{F}_{\frac{r}{2}}(M_i)$ such that $x_i \to x$, $y_i \to y$.
Let $\gamma_i$ be a curve connecting $x_i, y_i$ constructed by the method described in Proposition~\ref{prn:SB27_1}.
Clearly, $\gamma_i \subset \mathcal{F}_{\frac{1}{4}\epsilon_b r}$ and $|\gamma_i|<3d(x_i,y_i)$. Note that the convergence of
$\mathcal{F}_{\frac{1}{4}\epsilon_b r}$ to its limit set is in $C^{4}$-topology. Consequently, the limit curve $\gamma$ satisfies (\ref{eqn:HE11_1}).
\end{proof}
\begin{remark}
The definition of regular points in Theorem~\ref{thm:HE11_1} is stronger than the classical one, i.e., a point is regular if and only if every tangent space at this point is
isometric to $\ensuremath{\mathbb{C}}^{n}$. Therefore, some regular points in the classical definition may be singular in our definition.
Of course, at last, our definition coincides with the classical one after we set up sufficient estimates(c.f. Remark~\ref{rmk:HA07_1}).
\label{rmk:HA01_1}
\end{remark}
\begin{remark}
Note that in Definition~\ref{dfn:SC02_1}, the definition of canonical radius, no K\"ahler condition is used. Therefore, the convergence results discussed in this subsection works naturally in the Riemannian setting.
\label{rmk:MA21_1}
\end{remark}
The properties of the limit space $\bar{M}$ in Theorem~\ref{thm:HE11_1} are not good enough. For example, we do not know if every tangent space is a metric cone, we do not know if $\mathcal{R}$
is convex. In general, one should not expect these to hold. However, if $(M_i, g_i, J_i)$ is a blowup sequence from given K\"ahler Ricci flow solutions with proper geometric bounds, we shall show that $\bar{M}$ do have the mentioned good properties.
\section{Polarized canonical radius}
In this section, we shall improve the regularity of the limit pace $\bar{M}$ in Theorem~\ref{thm:HE11_1}, under the help of K\"ahler geometry and the Ricci flow.
The Ricci flow has reduced volume and local $W$-functional monotonicity, discovered by Perelman. These monotonicities will be used to show
that each tangent space is a metric cone, and the regular part $\mathcal{R}$ is weakly convex, under some natural geometric conditions. However, the weak-compactness we developed in last section only deals with the metric structure. On $\bar{M}$, we cannot see a Ricci flow structure. In order to make use of the intrinsic monotonicity of the Ricci flow, we need a weak compactness of Ricci flows, not just the weak compactness of time slices. However, along the Ricci flow, the metric at different time slices cannot be compared obviously if no estimate of Ricci curvature is known. This is one of the fundamental difficulty to develop the weak compactness theory of the Ricci flows.
We overcome this difficulty by taking advantage of the rigidity of K\"ahler geometry.
\subsection{A rough long-time pseudolocality theorem for polarized K\"ahler Ricci flow}
Suppose $\mathcal{LM}=\left\{ (M^n, g(t), J, L, h(t)), t \ \in I \subset \ensuremath{\mathbb{R}} \right\}$ is a polarized K\"ahler Ricci flow.
Let $\mathbf{b}$ be the Bergman function with respect to $\omega(t)$ and $h(t)$, i.e.,
\begin{align*}
\mathbf{b}(x,t) =\log \sum_{k=0}^{N} \norm{S_k}{h(t)}^2, \quad \int_{M} \langle S_k, S_l\rangle \omega(t)^n = \delta_{kl},
\end{align*}
where $N=\dim (H^0(L))-1$, $\{S_k\}_{k=0}^{N}$ are holomorphic sections of $L$.
By pulling back the Fubini-Study metric through the natural holomorphic embedding, we have
\begin{align*}
\tilde{\omega}= \iota^*(\omega_{FS})= \omega+ \sqrt{-1} \partial \bar{\partial} \mathbf{b}.
\end{align*}
Let $\omega_0=\omega(0), \mathbf{b}_0=\mathbf{b}(0)$. Then
\begin{align*}
\omega(t)=\omega_0 + \sqrt{-1}\partial \bar{\partial} \varphi,
\quad
\tilde{\omega}(t)=\omega_0 + \sqrt{-1}\partial \bar{\partial} (\varphi + \mathbf{b}).
\end{align*}
Clearly, $\varphi(0)=0$.
In this section, we focus on polarized K\"ahler Ricci flow $\mathcal{LM}$ satisfying the following estimate
\begin{align}
\norm{\dot{\varphi}}{C^0(M)} + \norm{\mathbf{b}}{C^0(M)} + \norm{R}{C^0(M)} +|\lambda| +C_S(M) \leq B
\label{eqn:SK23_1}
\end{align}
for every time $t \in I$. Let $\mathbf{b}^{(k)}$ be the Bergman function of the line bundle $L^{k}$ with the naturally induced metric.
Then a standard argument implies that
\begin{align}
\norm{\dot{\varphi}}{C^0(M)} + \norm{\mathbf{b}^{(k)}}{C^0(M)} + \norm{R}{C^0(M)} +|\lambda| +C_S(M) \leq B^{(k)}
\label{eqn:SK23_2}
\end{align}
for a constant $B^{(k)}$ depending on $B$ and $k$. Define
\begin{align*}
&\tilde{\omega}^{(k)} \triangleq \frac{1}{k} \left( \iota^{(k)} \right)^* (\omega_{FS}), \\
&F^{(k)} \triangleq \Lambda_{\omega_t} \tilde{\omega}_{0}^{(k)}=n- \Delta \left(\varphi-\mathbf{b}_0^{(k)} \right).
\end{align*}
In this section, the existence time of the polarized K\"ahler Ricci flow is always infinity, i.e., $I=[0,\infty)$.
\begin{lemma}[\textbf{Integral bound of trace}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}).
Suppose $u$ is a positive, backward heat equation solution, i.e.,
\begin{align*}
\square^* u=(-\partial_{t}-\Delta +R-\lambda n) u=0,
\end{align*}
and $\int_{M} u dv \equiv 1$. Then for every $t_0>0$, we have
\begin{align}
\int_0^{t_0} \int_{M} F^{(k)}u dv dt \leq \left(n+2B^{(k)} \right)t_0 + 2B^{(k)}.
\label{eqn:SK07_2}
\end{align}
\label{lma:I29_1}
\end{lemma}
\begin{proof}
For simplicity of notation, we only give a proof of the case $k=1$ and denote $F=F^{(1)}$. Note that $\mathbf{b}=\mathbf{b}^{(1)}, B=B^{(1)}$.
The proof of general $k$ follows verbatim.
Direct calculation shows that
\begin{align*}
\int_0^{t_0} \int_{M} Fu dv &=nt_0- \int_{0}^{t_0} \int_{M} \{\Delta (\varphi-\mathbf{b}_0)\} u dv\\
&=nt_0-\int_0^{t_0}\int_{M} (\varphi-\mathbf{b}_0) (\Delta u) dv\\
&=nt_0+ \int_{0}^{t_0} \int_{M} (\varphi-\mathbf{b}_0) \left( \dot{u}-Ru +\lambda n u \right) dv\\
&=nt_0 + \int_{0}^{t_0} \left[ \frac{d}{dt}\left( \int_{M} (\varphi-\mathbf{b}_0) u dv \right) - \int_{M}\dot{\varphi} u dv\right]dt\\
&=nt_0+ \left. \int_{M} (\varphi-\mathbf{b}_0) u dv \right|_0^{t_0} - \int_0^{t_0} \int_{M} \dot{\varphi} u dv dt.
\end{align*}
Note $|\varphi| \leq Bt_0$ at time $t_0$, then (\ref{eqn:SK07_2}) follows from the above inequality and (\ref{eqn:SK23_2}).
\end{proof}
We shall proceed to improve the integral estimate (\ref{eqn:SK07_2}) of $F^{(k)}$ to pointwise estimate, under local geometry bounds.
Before we go into details, let us first fix some notations.
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow solution satisfying (\ref{eqn:SK23_1}), $x_0 \in M$.
In this subsection, we shall always assume
\begin{align}
\Omega \triangleq B_{g(0)}(x_0,r_0), \quad \Omega' \triangleq B_{g(0)}(x_0, (1-\delta)r_0),
\quad \Omega'' \triangleq B_{g(0)}(x_0, (1-2\delta)r_0).
\label{eqn:SL26_3}
\end{align}
See Figure~\ref{figure:fourballs} for intuition. Then we define
\begin{align}
w_0 \triangleq \phi \left(\frac{2(d-1+2\delta)}{\delta} \right),
\label{eqn:SL26_4}
\end{align}
where $d=d_{g(0)}(x_0, \cdot)$, $\phi$ is a cutoff function, which equals one on $(-\infty, 1]$, decreases to $0$ on $(1,2)$. Moreover, $(\phi')^2 \leq 10\phi$.
Note that such $\phi$ exists by considering the behavior of $e^{-\frac{1}{s}}$ around $s=0$.
Clearly, $w_0$ satisfies
\begin{align}
\begin{cases}
&|\nabla w_0|^2 \leq \frac{40}{\delta^2}w_0, \\
&w_0 \equiv 1, \quad \textrm{on} \quad \Omega'', \\
&w_0 \equiv 0, \quad \textrm{on} \quad (\Omega')^c.
\end{cases}
\label{eqn:SL26_5}
\end{align}
\begin{lemma}[\textbf{Pointwise bound of trace}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$, $\Omega'$ is defined by (\ref{eqn:SL26_3}).
Suppose $\frac{1}{2} \omega_0 \leq \tilde{\omega}_0^{(k)} \leq 2 \omega_0$ on $\Omega'$.
Let $w$ be a solution of heat equation $\square w=\left( \frac{\partial}{\partial t} -\Delta \right)w=0$, initiating from a cutoff function $w_0$
satisfying (\ref{eqn:SL26_5}). Then for every $t_0>0$ and $y_0 \in M$, we have
\begin{align}
F^{(k)}(y_0,t_0) w(y_0, t_0) \leq C \label{eqn:SK23_3}
\end{align}
where $C=C(B,k,\delta,t_0)$.
\label{lma:J02_1}
\end{lemma}
\begin{proof}
For simplicity of notation, we only give a proof for the case $k=1$ and denote $F=F^{(1)}$, $B=B^{(1)}$, $H=\frac{40}{\delta^2}$.
The proof of general $k$ follows verbatim.
Note that $0 \leq w_0 \leq 1$, since $w$ is the heat solution, it follows from maximum principle that $0 \leq w \leq 1$.
On the other hand, according to the choice of $w_0$, we have $|\nabla w|^2 -Hw \leq 0$ at the initial time. Direct calculation implies that
\begin{align*}
\square \left\{ e^{\lambda t}\left(|\nabla w|^2 -Hw\right) \right\}
=-e^{\lambda t}\left\{|\nabla \nabla w|^2 +|\nabla \bar{\nabla} w|^2 +Hw\right\} \leq 0.
\end{align*}
Therefore, $|\nabla w|^2 -Hw \leq 0$ is preserved along the flow by maximum principle.
In other words, we always have
\begin{align*}
w|\nabla \log w|^2 \leq H, \quad 0 \leq w \leq 1
\end{align*}
on the space-time $M \times [0,\infty)$.
In light of parabolic Schwarz lemma(c.f.~\cite{SongWeinnotes} and references therein), we obtain
\begin{align*}
\square \log F \leq B F-\lambda.
\end{align*}
Note that
\begin{align*}
\omega(t)=\omega_0+\sqrt{-1} \partial \bar{\partial} \varphi=\tilde{\omega}_0+\sqrt{-1} \partial \bar{\partial} (\varphi-\mathbf{b}_0)
=\tilde{\omega}_0+\sqrt{-1} \partial \bar{\partial}\tilde{\varphi},
\end{align*}
where we denote $\varphi-\mathbf{b}_0$ by $\tilde{\varphi}$ for simplicity of notation. It is obvious that $\dot{\tilde{\varphi}}=\dot{\varphi}$.
Direct calculation shows that
\begin{align*}
&\square \tilde{\varphi}=\dot{\varphi}-\Delta \tilde{\varphi}= F-n+\dot{\varphi}, \\
&\square (\log F - B\tilde{\varphi})\leq B(n-\dot{\varphi})-\lambda \leq B(n+\norm{\dot{\varphi}}{C^0(\mathcal{M})})+|\lambda|
\leq C.
\end{align*}
Let $u$ be the solution of $\square^*u=0$, starting from a $\delta$-function from $(y_0, t_0)$. Then we calculate
\begin{align*}
\frac{d}{dt} \int_{M} Fe^{-B\tilde{\varphi}} wu dv
&=\int_{M} \square(Fe^{-B\tilde{\varphi}}w) u dv - \int_{M} Fe^{-B\tilde{\varphi}}w \square^*u dv\\
&=\int_{M} \square(Fe^{-B\tilde{\varphi}}w) u dv\\
&=\int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}w)-|\nabla \log (Fe^{-B\tilde{\varphi}}w)|^2 \right\} udv\\
&\leq \int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}w) \right\} udv\\
&=\int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}) + \square \log w \right\} udv\\
&=\int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}) + |\nabla \log w|^2 \right\} udv\\
&\leq C \int_{M} Fe^{-B\tilde{\varphi}} w u dv + H \int_{M} Fe^{-B\tilde{\varphi}} u dv.
\end{align*}
It follows that
\begin{align*}
\frac{d}{dt} \left\{e^{-Ct} \int_{M} Fe^{-B\tilde{\varphi}} w u dv \right\} \leq H e^{-Ct} \int_{M} Fu dv.
\end{align*}
Integrate the above inequality and apply Lemma~\ref{lma:I29_1}, we have
\begin{align*}
e^{-Ct_0} F(y_0,t_0)w(y_0, t_0) e^{-B\tilde{\varphi}(y_0, t_0)} &\leq \left. \int_{M} Fe^{-B\tilde{\varphi}}wu dv \right|_{t=0} + H \int_0^{t_0} \int_{M} Fu dv dt\\
&\leq C \left. \int_{\Omega'} Fu dv \right|_{t=0}+C\\
&\leq C \left. \int_{\Omega'} u dv \right|_{t=0} + C \leq C.
\end{align*}
Therefore, (\ref{eqn:SK23_3}) follows directly from the above inequality.
\end{proof}
\begin{figure}
\begin{center}
\psfrag{t0}[c][c]{$t=0$}
\psfrag{t1}[c][c]{$t=t_0$}
\psfrag{x0}[c][c]{$x_0$}
\psfrag{A1}[c][c]{$(M,x_0,g(t_0))$}
\psfrag{A2}[c][c]{$(M,x_0,g(0))$}
\psfrag{B1}[c][c]{$\Omega=\textrm{blue}$}
\psfrag{B2}[c][c]{$\Omega'=\textrm{green}$}
\psfrag{B3}[c][c]{$\Omega''=\textrm{red}$}
\psfrag{B4}[c][c]{$B_{g(t_0)}(x_0,r)=\textrm{black}$}
\includegraphics[width=0.5 \columnwidth]{fourballs}
\caption{Different domains}
\label{figure:fourballs}
\end{center}
\end{figure}
\begin{lemma}[\textbf{Lower bound of heat solution}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$,
notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}).
Suppose $\Omega'' \subset B_{g(t)}(x_0, r)$ for some $t>0$ and $r>0$. Then in the geodesic ball $B_{g(t)}(x_0, r)$, we have
\begin{align*}
w(y,t)>c
\end{align*}
for some constant $c=c(n,B,k,\delta,r_0,r,t)$.
\label{lma:SK23_1}
\end{lemma}
\begin{proof}
By the construction of $w_0$ and maximum principle, it is clear that $0 <w \leq 1$ when $t>0$.
Let $P$ be the heat kernel function, then we can write
\begin{align}
w(x_0,t)=\int_M P(x_0,t; y,0) w_0(y) dv_y \geq \int_{\Omega''} P(x_0,t;y,0)w_0(y)dv_y.
\label{eqn:SK23_4}
\end{align}
In light of the Sobolev constant bound and scalar curvature bound, one has the on-diagonal bound
\begin{align*}
\frac{1}{C} t^{-n} \leq P(x,t;x,0) \leq C t^{-n},
\end{align*}
which combined with the gradient estimate of heat equation(c.f. Theorem 3.3 of~\cite{Zhq2}) implies that
\begin{align*}
P(x,t;y,0) \geq \frac{1}{C} t^{-n}
\end{align*}
where $C=C(B,d_{g(t)}(x,y))$. Plugging this estimate into (\ref{eqn:SK23_4}) implies that
\begin{align*}
w(x_0,t) \geq \frac{|\Omega''|}{Ct^n}.
\end{align*}
Note that $C_S$ bound forces $|\Omega''|$ is bounded from below. Since $0<w\leq 1$,
then (\ref{eqn:SK23_4}) follows from the above inequality and the gradient estimate of heat equation.
\end{proof}
The following two lemmas show that K\"ahler geometry is much more rigid than Riemannian geometry.
\begin{lemma}[\textbf{Fubini-Study approximation}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$,
notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}).
Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$. Then there exists an integer $k=k(B,r_0, \delta)$ such that
\begin{align}
\frac12 \omega_0 \leq \tilde{\omega}_0^{(k)} \leq 2\omega_0 \label{eqn:SL29_3}
\end{align}
on $\Omega'$.
\label{lma:SK25_1}
\end{lemma}
\begin{proof}
This follows essentially from the peak section method of Tian(c.f.~\cite{Tian90JDG},\cite{Lu}). We give a proof here for the convenience of the readers.
Fix arbitrary $x \in \Omega'$, $V \in T_x^{(1,0)}M$ with unit norm. In order to prove (\ref{eqn:SL29_3}), it suffices to show that
\begin{align}
\frac{1}{2} \leq \tilde{\omega}(V,JV) \leq 2. \label{eqn:SL29_1}
\end{align}
Around $x$, we can always choose a normal coordinate (K-coordinate, c.f.~\cite{Lu}) chart around $x$ such that
\begin{align*}
V=\frac{\partial}{\partial z_1}, \quad
g_{i\bar{j}}(x)=\delta_{i\bar{j}}, \quad
\frac{\partial^{p_1+p_2+\cdots +p_n}}{\partial z_1^{p_1} \partial z_2^{p_2} \cdots \partial z_n^{p_n}} g_{i\bar{j}}(x)=0
\end{align*}
for any nonnegative integers $p_1,p_2, \cdots, p_n$ with $p=p_1+p_2+p_3+\cdots p_n>0$. Moreover,
there exists a
local holomorphic frame $e_L$ of $L$ around $x$ such that the local representation $a$ of the Hermitian metric $h$ has the
properties
\begin{align*}
a(x)=1, \quad \frac{\partial^{p_1+p_2+\cdots +p_n}}{\partial z_1^{p_1}\partial z_2^{p_2} \cdots \partial z_n^{p_n}}a(x)=0
\end{align*}
for any nonnegative integers $p_1,p_2, \cdots, p_n$ with $p=p_1+p_2+p_3+\cdots p_n>0$.
Suppose $\{S_0^{k}, \cdots S_{N_k}^{k}\}$ is an orthonormal basis of $H^0(M, L^k)$,
where $N_k=\dim_{\ensuremath{\mathbb{C}}} H^0(M,L^k)-1$. Around $x$, we can write
\begin{align*}
S_0^{k}=f_0^{k} e_L, \cdots, S_{N_k}^{k}=f_{N_k}^{k} e_L.
\end{align*}
Rotating basis if necessary(c.f.~\cite{Tian90JDG}), we can assume
\begin{align*}
&f_i^{k}(x)=0, \quad \forall \; i \geq 1, \\
&\frac{\partial f_i^k}{\partial z_j}(x)=0, \quad \forall \; i \geq j+1.
\end{align*}
Recall that
\begin{align*}
\tilde{\omega}^{(k)}&=\omega_0 + \frac{1}{k} \sqrt{-1} \partial \bar{\partial} \log \sum_{j=0}^{N_k} \norm{S_j^{k}}{}^2
=\frac{1}{k} \sqrt{-1} \partial \bar{\partial} \log \sum_{j=0}^{N_k} |f_j^k|^2.
\end{align*}
So we have
\begin{align}
\tilde{\omega}^{(k)}(V,JV)
=\frac{1}{k} \frac{\partial^2 \log \sum_{j=0}^{N_k} |f_j^k|^2}{\partial z_1 \bar{\partial} z_1}
=\frac{|\frac{\partial f_1^k}{\partial z_1}|^2}{k|f_0^k|^2}.
\label{eqn:SL29_2}
\end{align}
Because of (\ref{eqn:SL29_1}) and (\ref{eqn:SL29_2}), the problem boils down to a precise estimate of
$\frac{\partial f_1^k}{\partial z_1}$ and $f_0^k$.
As pointed out by Tian in \cite{Tian90JDG}, the peak section method is local in nature.
The global information of the underlying manifold is only used in the step of H\"{o}rmander's estimate. However, in our case, we have
\begin{align*}
&\sqrt{-1} \partial \bar{\partial} \dot{\varphi}+Ric = \lambda g, \quad |\dot{\varphi}| + |\lambda| \leq B.
\end{align*}
Due to the uniformly bounded geometry (up to $C^2$-norm of $g$) inside $\Omega'$ and the uniform bound of
$\sqrt{-1} \partial \bar{\partial} \dot{\varphi}+Ric$ on the whole manifold $M$, Lemma 1.2 of~\cite{Tian90JDG}
follows directly and can be written as follows.
\textit{ For an $n$-tuple of integers $(p_1,p_2, \cdots, p_n) \in \ensuremath{\mathbb{Z}}_{+}^n$ and an integer $p'>p=p_1+p_2+\cdots+p_n$, there
exists an $k_0=k_0(n,B,r_0,\delta)$ such that for $k>k_0$, there is a unit norm holomorphic section $S \in H^0(M, L^k)$ satisfying}
\begin{align*}
\int_{M \backslash \{|z|^2 \leq \frac{(\log k)^2}{k}\}} \norm{S}{}^2 dv \leq \frac{1}{k^{2p'}}.
\end{align*}
Then the same argument as in~\cite{Tian90JDG} implies that(c.f. Lemma 3.2 of~\cite{Tian90JDG})
\begin{align}
&\left|f_0^k(x)-\sqrt{\frac{(n+k)!}{k!}} \left\{1+\frac{1}{2(k+n+1)!} (R(x)-n^2-n) \right\} \right|<\frac{C}{k^2}, \label{eqn:MC18_1}\\
&\left|\frac{\partial f_1^k}{\partial z_1}(x)-\sqrt{\frac{(n+k+1)!}{k!}}\left\{ 1+\frac{1}{2(k+n+1)}(R(x)-n^2-3n-2)\right\} \right| <\frac{C}{k^2}, \label{eqn:MC18_2}
\end{align}
for some $C=C(n,B,r_0,\delta)$. Here $R$ is the complex scalar curvature. Plugging the above estimate into (\ref{eqn:SL29_2}),
we obtain (\ref{eqn:SL29_1}), whenever $k$ is larger than a big constant, which depends only on $n,B,r_0,\delta$.
\end{proof}
\begin{lemma}[\textbf{Liouville type theorem}]
Every complete K\"ahler Ricci flat metric $\tilde{g}$ on $\ensuremath{\mathbb{C}}^n$ must be a Euclidean metric if there is a constant $C$ such that
\begin{align}
\frac{1}{C} \delta_{i\bar{j}}(z) \leq \tilde{g}_{i\bar{j}}(z) \leq C \delta_{i\bar{j}}(z), \quad \forall \; z \in \ensuremath{\mathbb{C}}^n.
\label{eqn:SL26_1}
\end{align}
\label{lma:HE11_1}
\end{lemma}
\begin{proof}
The original proof of this lemma goes back to the famous paper of E. Calabi~\cite{Ca58} and Pogorelov~\cite{Po78} on real Monge Amp\`{e}re equation. For complex
Monge Amp\`{e}re equation, this is initially due to Riebesehl-Schulz~\cite{RS84} where higher derivatives are used heavily.
We say a few words here for the convenience of the readers,
using the Schauder estimate of Evans-Krylov.
Actually, it is not difficult to see that the problem boils down to the study of a
global pluri-subharmonic function $u$ in $\ensuremath{\mathbb{C}}^n$ such that
\begin{align}
\begin{cases}
&\det \left({\partial^2 u \over {\partial z_i\partial \bar z_j}}\right) = 1, \\
&C^{-1} (\delta_{i\bar j}) < \left({\partial^2 u \over {\partial z_i\partial \bar z_j}}\right) < C (\delta_{i\bar j}).
\end{cases}
\label{eqn:SL26_2}
\end{align}
In order to show the metric $\tilde{g}$ is Euclidean, it suffices to show that $u$ is a global quadratic polynomial.
Without loss of generality, we may assume that $u(0) = D u(0) = 0$.
For every positive integer $k$, we can define a function $u^{(k)}$ in the unit ball by
\begin{align*}
u^{(k)} (z) = {u(kz)\over {k^2}}.
\end{align*}
Clearly, $u^{(k)}$ satisfies (\ref{eqn:SL26_2}). Note that $\norm{u^{(k)}}{C^2}$ is uniformly bounded,
in the unit ball $B(0,1)$. By standard Evans-Krylov theorem, there exists a
uniform constant $C$ such that
\begin{align*}
[D^2 u^{(k)}]_{C^\alpha(B(0,\frac{1}{2}))} \leq C
\end{align*}
for every $k$. Putting back the scaling factor, the above inequality is equivalent to
\begin{align*}
[D^2 u]_{C^\alpha(B(0, \frac{k}{2}))} \leq C k^{-\alpha},\qquad \forall \quad k=1,2,\cdots.
\end{align*}
Let $k\rightarrow \infty$, we have $[D^2 u]_{C^\alpha(\ensuremath{\mathbb{C}}^n)} = 0$. Therefore, $D^2 u$ is a constant matrix,
$u$ is a quadratic polynomial. So we finish the proof.
\end{proof}
\begin{proposition}[\textbf{Ball containing relationship implies regularity improvement}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$,
notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}).
Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$.
Moreover, we assume
\begin{align}
\Omega'' \subset B_{g(t)}(x_0,r) \subset \Omega' \label{eqn:MC09_1}
\end{align}
for every $0\leq t \leq t_0$.
Then the following estimates hold.
\begin{itemize}
\item In the geodesic ball $B_{g(t)}(x_0,r)$, we have
\begin{align}
\frac{1}{C} \omega_0 \leq \omega_t \leq C \omega_0
\label{eqn:SK23_6}
\end{align}
for some constant $C=C(n,B,k,\delta,r_0,r,t)$.
\item In the geodesic ball $B_{g(t)}(x_0,r-\xi)$, we have
\begin{align}
|Rm|(x,t) \xi^2 \leq C \label{eqn:SK23_5}
\end{align}
for each small $\xi$ and some constant $C=C(n,B,k,\delta,r_0,r,t_0)$.
\end{itemize}
\label{prn:HE11_2}
\end{proposition}
\begin{proof}
Note that Perelman's strong version of pseudolocality theorem, i.e., Theorem 10.3 of~\cite{Pe1}, can be modified and applied here.
In fact, the almost Euclidean volume ratio condition in that theorem can be replaced by $\kappa$-noncollapsing condition.
Since one has injectivity radius estimate when curvature and volume ratio bounds are available,
thanks to the work of Cheeger, Gromov and Taylor, in~\cite{ChGrTa}.
By shrinking the ball to some fixed smaller size, one can get back the condition of almost Euclidean volume ratio.
Up to a covering argument, we can apply this strong version pseudolocality theorem to show that
$|Rm|$ is uniformly bounded on $\Omega' \times [0,\eta]$ for some positive $\eta=\eta(n,\kappa,\delta)$.
Then (\ref{eqn:SK23_6}) and (\ref{eqn:SK23_5}) follows trivially. For this reason, we can assume $t_0>\eta$.
We first prove estimate (\ref{eqn:SK23_6}). Due to Fubini-Study metrics' approximation, Lemma~\ref{lma:SK25_1}, it is clear that one can regard
$\omega_0$ and $\tilde{\omega}_0^{(k)}$ as the same metric on $\Omega'$. Therefore, it follows from the combination of Lemma~\ref{lma:J02_1}
and Lemma~\ref{lma:SK23_1} that $F^{(k)}$ is bounded from above, which implies $\Lambda_{\omega_t} \omega_0 \leq C$.
Recall that the volume element $\omega_0^n$ and $\omega_t^n$ are uniformly equivalent, due to the uniform bound of $|R|+|\lambda|$ and the evolution equation
\begin{align*}
\frac{\partial}{\partial t} \log \omega_t^n= n\lambda-R.
\end{align*}
Consequently, (\ref{eqn:SK23_6}) follows. We remind the readers that condition (\ref{eqn:MC09_1}) is used in the above discussion.
Then we proceed to prove inequality (\ref{eqn:SK23_5}). Fix $L$ very large.
If (\ref{eqn:SK23_5}) does not hold uniformly, then we can find some space-time point $(y_0, s_0)$ such that $y_0 \in B_{g(s_0)}(x_0, r-\xi)$ and $Q_0 \triangleq |Rm|(y_0, s_0)>100L^2 \xi^{-2}$ is very large.
Set $\rho_0 \triangleq d_{g(s_0)}(y_0, x_0)$. On one hand, $\rho_0<r-\xi$ by the choice of $(y_0, s_0)$. On the other hand, $s_0>\eta$ for some uniform $\eta$ due to the application of Perelman's pseudo-locality,
as discussed above. Search whether there is a point $(x,t)$ satisfying
\begin{align*}
|Rm|(x,t)>4Q_0, \quad x \in B_{g(t)} \left(x_0, \rho_0 + LQ_0^{-\frac{1}{2}} \right), \; t \in \left[t_0-Q_0^{-1}, t_0 \right].
\end{align*}
If there exists such a point, we denote it by $(y_1, s_1)$ and continue the above searching.
We can find $(y_k, s_k)$ by induction. Actually, if $(y_{k-1}, s_{k-1})$ is defined, then we denote $|Rm|(y_{k-1}, s_{k-1})$ by $Q_{k-1}$, denote $d_{g(s_{k-1})}(x_0, y_{k-1})$ by $\rho_{k-1}$ and search point $(x,t)$ satisfying
\begin{align*}
|Rm|(x,t)>4Q_{k-1}, \quad x \in B_{g(t)} \left(x_0, \rho_{k-1}+ LQ_{k-1}^{-\frac{1}{2}} \right), \; t \in \left[t_{k-1}-Q_{k-1}^{-1}, t_{k-1} \right].
\end{align*}
If there is no such point, we stop the process.
Otherwise, we denote such a point by $(y_k, s_k)$ and continue the process. Clearly, we have
\begin{align*}
& Q_k=4^{k}Q_0 >100L^2 \xi^{-2}, \\
&\rho_k \leq \rho_0 + L \left(Q_0^{-\frac{1}{2}} + \cdots Q_{k-1}^{-\frac{1}{2}} \right)<\rho_0 + 4LQ_0^{-\frac{1}{2}}< r-0.5\xi, \\
&|s_0-s_k|=s_0-s_k \leq Q_0^{-1} + Q_{1}^{-1} + \cdots Q_{k-1}^{-1}< 2Q_0^{-1}<\frac{\xi^2}{50 L^2}<<\eta.
\end{align*}
Since the process happens in a compact space-time domain with bounded geometry, it must stop after finite steps. Let $k$ be the last $(y_k, s_k)$.
We denote it by $(y,s)$ and set $Q=|Rm|(y,s)$ and $\rho=d_{g(s)}(y, x_0)$. Then we have
\begin{align}
\begin{cases}
&Q>100 L^2 \xi^{-2}, \\
&\rho< r-0.5 \xi,\\
&s> 0.5 \eta, \\
&|Rm|(x,t)<4Q, \quad \forall \; x \in B_{g(t)}(x_0, \rho+LQ^{-\frac{1}{2}}), \quad t \in \left[s-Q^{-1}, s \right].
\end{cases}
\label{eqn:MC09_2}
\end{align}
By its choice, we have $d_{g(s)}(x_0,y)=\rho$. We observe that $y$ will stay in $B_{g(t)}(x_0, \rho+2Q^{-\frac{1}{2}})$ whenever $t \in [s-\frac{1}{5nQ}, s]$.
This is an application of Lemma 8.3 of Perelman~\cite{Pe1}, or section 17 of Hamilton~\cite{Ha95f}.
Actually, let $\theta_0$ be the largest positive number such that $y$ fails to locate in $B_{g(t-\theta_0 Q^{-1})}(x_0, \rho+ 2Q^{-\frac{1}{2}})$.
Then for each $t \in [s-\theta_0 Q^{-1}, s]$, triangle inequality implies that $B_{g(t)}(y, Q^{-\frac{1}{2}}) \subset B_{g(t)}(x_0, \rho+3Q^{-\frac{1}{2}})$. Conseqeuently, we have
\begin{align*}
&|Rm|(x,t) \leq 4Q, \quad \forall \; x \in B_{g(t)}(y, Q^{-\frac{1}{2}}), \\
&|Rm|(x_0,t) \leq 4Q, \quad \forall \; x \in B_{g(t)}(x_0, Q^{-\frac{1}{2}}).
\end{align*}
It follows from Lemma 8.3 (b) of Perelman~\cite{Pe1} that
\begin{align*}
\frac{d}{dt} d(x_0,y) \geq -10 n Q^{\frac{1}{2}} \quad \Rightarrow \quad d_{g(s)}(x_0,y)-d_{g(s-\theta_0 Q^{-1})}(x_0,y) \geq -10 n Q^{\frac{1}{2}} \cdot \theta_0 Q^{-1}.
\end{align*}
According to the choice of $\theta_0$, the left hand side of the second inquality is $-2Q^{-\frac{1}{2}}$. It follows that $\theta_0 \geq \frac{1}{5n}$.
Now we know that $y$ stays in $B_{g(t)}(x_0, \rho+2Q^{-\frac{1}{2}})$ for each $t \in [s-\frac{1}{5nQ}, s]$.
In view of (\ref{eqn:MC09_2}) and the fact $L>>1$, the triangle inequality implies that
\begin{align*}
|Rm|(x,t) < 4Q, \quad \forall x \in B_{g(t)}(y, 0.5LQ^{-\frac{1}{2}}), \quad t \in \left[s-\frac{1}{5nQ}, s \right].
\end{align*}
Let $\tilde{g}(t)=Qg(Q^{-1} t + s)$. We have
\begin{align*}
\begin{cases}
&|\widetilde{Rm}|(y,0)=1,\\
&|\widetilde{Rm}|(x,t) <4, \quad \forall x \in B_{\tilde{g}(t)}(y, 0.5 L), \quad t \in \left[-\frac{1}{5n}, 0 \right].
\end{cases}
\end{align*}
Note that $\left[-\frac{1}{5n}, 0 \right]$ is a fixed time period. The application of Perelman's pseudo-locality guarantees the existence of such a time period(c.f. (\ref{eqn:MC09_2})).
Now let $L \to \infty$, we can use the compactness theorem of Hamilton~\cite{Ha95} to obtain a limit Ricci flow solution, which is non-flat, K\"ahler Ricci-flat and non-collapsed on all scales.
We remark that the discussion above is nothing but repeating the argument of Claim 1 and Claim 2 in the proof of Perelman's pseudo-locality theorem, i.e, Theorem 10.1 in~\cite{Pe1}.
Similar argument was also used in the distance estimate of the work of Tian and the second named author~\cite{TW2}.
Note that $B_{g(s)}(y, 0.5LQ^{-\frac{1}{2}}) \subset B_{g(s)}(x_0, r-0.5\xi) \subset B_{g(s)}(x_0,r) \subset \Omega'=B_{g(0)}(x_0, 1-\delta)$.
Therefore, by the same scale blowup at $(y,0)$, we obtain nothing but $\ensuremath{\mathbb{C}}^n$. Recall we have (\ref{eqn:SK23_6}), so we obtain a nontrivial K\"ahler Ricci flat metric
$\tilde{g}_{i\bar{j}}$ on $\ensuremath{\mathbb{C}}^n$ such that (\ref{eqn:SL26_1}) holds for some $C$.
This contradicts Lemma~\ref{lma:HE11_1}.
\end{proof}
The rough estimate (\ref{eqn:SK23_6}) and (\ref{eqn:SK23_5}) can be improved when $|R|+|\lambda|$ is very small.
When curvature tensor is bounded in the space-time, one can estimate the Ricci curvature in terms of scalar curvature.
Let $|R|+|\lambda|$ tend to zero, we see that the Ricci curvature tends to zero at the space-time where $|Rm|$ is bounded. By adjusting $\xi$ if necessary, we obtain that in the limit, $B_{g(t)}(x_0,(1-\xi)r)$ is isometric to
$B_{g(0)}(x_0, (1-\xi)r)$ for every $0<t<t_0$. By adjusting $\xi$ and applying Perelman's pseudolocality theorem, we see the convergence at time $t=t_0$ is also smooth since curvature derivatives are all bounded in
the ball $B_{g(t_0)}(x_0,(1-\xi)r)$ at time $t_0$.
\begin{proposition}[\textbf{Volume element derivative small implies ball containing relationship}]
For every $r_0, T$ and small $\xi$, there exists an $\epsilon$ with the following property.
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$, notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}). Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$.
If $\displaystyle \sup_{\mathcal{M}} (|R| + |\lambda|) <\epsilon$, then for every $t \in [0,T]$ we have
\begin{align*}
&\Omega'' \subset B_{g(t)}\left(x_0, \left(1-\frac{3}{2}\delta \right) r_0 \right) \subset \Omega', \\
&\left(1- \xi \right)\omega(0) \leq \omega(t) \leq \left(1 + \xi \right) \omega(0), \quad \textrm{in} \; \Omega''.
\end{align*}
\label{prn:HE11_3}
\end{proposition}
\begin{proof}
If the statement was wrong, we can find a tuple $(n,B,\delta,r_0,T)$ and $\epsilon_i \to 0$ such that the property does not hold for every $\epsilon_i \to 0$.
Without loss of generality, we can assume $r_0=1$.
For each $\epsilon_i$, let $t_i \in [0,T]$ be the critical time of a flow $g_i(t)$ such that the properties hold on $[0,t_i]$.
In other words, for every $t \in [0, t_i]$, we have
\begin{align}
&\Omega_i'' \subset B_{g_i(t)}\left(x_i, 1-\frac{3}{2}\delta \right) \subset \Omega_i', \label{eqn:GE08_2}\\
&\left(1- \xi \right)\omega_i(0) \leq \omega_i(t) \leq \left(1 + \xi \right) \omega_i(0), \quad \textrm{in} \; \Omega_i''. \label{eqn:GE08_3}
\end{align}
However, for each time $t>t_i$, at least one of the above relations fails to hold. Related to (\ref{eqn:SL26_3}), here we set
\begin{align*}
\Omega_i \triangleq B_{g_i(0)}(x_i, 1), \quad \Omega_i' \triangleq B_{g_i(0)}(x_i, 1-\delta),
\quad \Omega_i'' \triangleq B_{g_i(0)}(x_i, 1-2\delta).
\end{align*}
We shall show that $t_i$ cannot locate in $[0, T]$ for large $i$ and therefore obtain a contradiction.
Note that $|Rm|_{g_i(0)} \leq 1$ at time $t=0$ in the ball $B_{g_i(0)}(x_i , 1)$.
By the strong version of Perelman's pseudolocality theorem, i.e., Theorem 10.3 of~\cite{Pe1}, one can find a uniform small constant $\eta$ such that
\begin{align}
|Rm|_{g_i}(x,t) \leq \frac{\xi}{100 n^2} \eta^{-2}, \quad \forall \; x \in B_{g_i(0)}(x_i, 1-\eta), \; t \in [0, \eta^2].
\label{eqn:GE08_1}
\end{align}
The existence of $\eta$ can be obtained by a contradiction blowup argument.
Since metrics evolve by $-Ric+\lambda g$, it follows from (\ref{eqn:GE08_1}) and the choice of $t_i$ that $\eta^2 \leq t_i \leq T$.
Recall that we have the relationship (\ref{eqn:GE08_2}) by the choice of $t_i$.
Therefore, Proposition~\ref{prn:HE11_2} can be applied to obtain a unfiorm $C$, independent of $i$, such that
\begin{align}
\frac{1}{C} g_i(0) \leq g_i(t_i) \leq Cg_i(0)
\label{eqn:GE08_4}
\end{align}
in the ball $B_{g_i(t_i)}(x_i, 1-\frac{3\delta}{2})$. Furthermore, the inequality (\ref{eqn:SK23_5}) in Proposition~\ref{prn:HE11_2} yields that
\begin{align*}
|Rm|_{g_i}(x,t) \leq \frac{C}{\psi^2}, \quad x \in B_{g(t)} \left(x_i, 1-\frac{3}{2} \delta -\psi \right), \; 0 \leq t \leq t_i,
\end{align*}
where $\psi$ is a small constant $\psi <<\delta$, to be determined. Note that we are in a setting where each geodesic ball's volume ratio is bounded from below, due to the bounds in (\ref{eqn:SK20_1}).
Consequently, injectivity radius has a lower bound(c.f.~\cite{ChGrTa}), by shrinking the ball if necessary.
Therefore, we can apply Theorem 3.2 of~\cite{Wa2} to obtain
\begin{align}
\sup_{\eta^2 \leq t \leq t_i, d_{g_i(t)}(x, x_i) \leq 1-\frac{3}{2} \delta -2\psi} |Ric|_{g_i}(x, t) \to 0, \quad \textrm{as} \; i \to \infty,
\label{eqn:GE08_5}
\end{align}
where $\eta$ is the constant in (\ref{eqn:GE08_1}). Alternatively, one can apply Lemma 2.1 of~\cite{CW5} to obtain the above estimate, with the fact that geodesic balls at different times can be compared due to
the Riemannian curvature bound and the evolution equation of the Ricci flow: the metrics evolve by $-Ric+\lambda g$.
Since $t_i$ is uniformly bounded by $T$, the above equation implies (up to a maximum principle type argument of the first violating time if necessary)
that
\begin{align}
B_{g_i(\eta^2)} \left( x_i, 1-\frac{3}{2}\delta -5\psi \right) \subset B_{g_i(t)} \left( x_i, 1-\frac{3}{2}\delta -4\psi \right) \subset B_{g_i(\eta^2)} \left( x_i, 1-\frac{3}{2}\delta -3\psi \right), \quad \forall \; t \in [\eta^2, t_i).
\label{eqn:GE08_6}
\end{align}
Combining the above relationship with (\ref{eqn:GE08_5}), we obtain that
\begin{align}
\sup_{\eta^2 \leq t \leq t_i, d_{g_i(\eta^2)}(x, x_i) \leq 1-\frac{3}{2} \delta-3\psi } |Ric|_{g_i}(x, t) \to 0, \quad \textrm{as} \; i \to \infty.
\label{eqn:GE08_7}
\end{align}
By (\ref{eqn:GE08_1}) and $|R|+|\lambda| \to 0$, we see the metric at $g_i(0)$ and $g_i(\eta^2)$ are almost isometric to each other on the ball $B_{g_i(0)}(x_i, 1-\frac{3}{2}\delta)$.
Consequently, we have
\begin{align*}
\Omega_i''=B_{g_i(0)}(x_i, 1-2\delta) \subset B_{g_i(\eta^2)} \left( x_i, 1-\frac{7}{4}\delta \right) \subset B_{g_i(t_i)} \left( x_i, 1-\frac{7}{4}\delta +\psi \right) \Subset B_{g_i(t_i)} \left( x_i, 1-\frac{3}{2}\delta \right),
\end{align*}
where $\Subset$ means ``compactly contained". We claim that we also have
\begin{align*}
B_{g_i(t_i)} \left( x_i, 1-\frac{3}{2}\delta \right) \Subset \Omega_i'.
\end{align*}
For otherwise, by the choice of $t_i$, the boundary of $B_{g_i(t_i)} \left( x_i, 1-\frac{3}{2}\delta \right)$ touches the boundary of $\Omega_i'$ at time $t_i$.
Therefore, we can find a point $y_i$ satisfying
\begin{align*}
d_{g_i(t_i)}(x_i, y_i)=1-\frac{3}{2}\delta, \quad d_{g_i(0)}(x_i, y_i)=1-\delta.
\end{align*}
Let $\gamma_i$ be a shortest unit-speed geodesic connecting $x_i$ and $y_i$, with respect to the metric $g_i(t_i)$.
Let $\gamma_i(0)=x_i$ and $\gamma_i(1-\frac{3}{2}\delta)=y_i$. By previous estimates, we see that
\begin{align*}
\gamma_i \left(1-\frac{3}{2}\delta -100\psi \right) \subset B_{g_i(0)} \left(x_i, 1-\frac{3}{2}\delta -50\psi \right).
\end{align*}
Let $\alpha_i$ be the part of $\gamma_i$, connecting $x_i=\gamma_i(0)$ and $\gamma_i(1-\frac{3}{2}\delta -100\psi )$.
Let $\beta_i$ be the remainded part of $\gamma_i$, i.e., the part connecting $\gamma_i(1-\frac{3}{2}\delta -100\psi )$ and $y_i=\gamma_i(1-\frac{3}{2}\delta)$.
Using $|\cdot|$ to denote the length of curves. It is clear that $|\beta_i|_{g_i(t_i)} =100\psi$.
Note that $\alpha_i$ locates in $B_{g_i(t_i)}(x_i, 1-\frac{3}{2}\delta -100\psi)$.
It follows from (\ref{eqn:GE08_6}), (\ref{eqn:GE08_7}) and (\ref{eqn:GE08_1}) that
\begin{align*}
\sup_{\alpha_i \times [\eta^2, t_i]} |Ric|(x, t) \to 0, \quad \textrm{as} \; i \to \infty; \quad \sup_{\alpha_i \times [0, \eta^2]} |Rm|(x, t) \leq \frac{\xi}{100n^2} \eta^{-2}.
\end{align*}
Together with $|R|+|\lambda| \to 0$ as $i \to \infty$, we can compare the length of $\alpha_i$ at time $t=t_i$ and $t=0$.
\begin{align*}
|\alpha_i|_{g_i(t_i)}=1-\frac{3}{2}\delta -100\psi, \quad |\alpha_i|_{g_i(0)} \leq 1-\frac{3}{2}\delta.
\end{align*}
However, since $d_{g_i(0)}(x_i, y_i)=1-\delta$, we have
\begin{align*}
1-\delta \leq |\gamma_i|_{g_i(0)}=|\alpha_i|_{g_i(0)} + |\beta_i|_{g_i(0)} \leq |\beta_i|_{g_i(0)} + 1-\frac{3}{2}\delta.
\end{align*}
It follows that $ |\beta_i|_{g_i(0)} \geq \frac{1}{2}\delta$. Recall that $|\beta_i|_{g_i(t_i)} =100\psi$.
Therefore, by mean value theorem, we must have
\begin{align*}
\frac{\sqrt{\langle V, V\rangle_{g_i(0)}}}{ \sqrt{\langle V, V \rangle_{g_i(t_i)}}} \geq \frac{\frac{1}{2}\delta}{100\psi}=\frac{\delta}{200\psi}.
\end{align*}
at some point $z_i \in \beta_i$, where $V$ is the unit tangent vector (with respect to $g_i(t_i)$) of $\beta_i$ at $z_i$.
Since $z_i \in B_{g_i(t_i)}(x_i, 1-\frac{3}{2}\delta)$, one can apply (\ref{eqn:GE08_4}) to
bound the left hand side of the above inequality by $\sqrt{C}$, where $C$ is the constant in (\ref{eqn:GE08_4}).
It follows that $ C \geq \frac{\delta^2}{40000\psi^2}$, which is impossible if we choose $\psi$ small enough. Therefore, for $i$ large, we must have
\begin{align*}
\Omega_i '' \Subset B_{g_i(t_i)} \left( x_i, 1-\frac{3}{2}\delta \right) \Subset \Omega_i'.
\end{align*}
Then we can apply (\ref{eqn:GE08_1}), (\ref{eqn:GE08_5}) and the fact that $|R|+|\lambda| \to 0$ to obtain that
\begin{align*}
\left(1- \frac{\xi}{100} \right)\omega_i(0) \leq \omega_i(t) \leq \left(1 + \frac{\xi}{100} \right) \omega_i(0), \quad \textrm{in} \; \Omega_i''= B_{g_i(0)}(x_i, 1-2\delta),
\end{align*}
whenever $i$ large enough. This means that for large $i$, we have both (\ref{eqn:GE08_2}) and (\ref{eqn:GE08_3}) hold for a short while beyond the time $t_i$.
This contradicts to the choice of time $t_i$.
\end{proof}
Combine Proposition~\ref{prn:HE11_2} and Proposition~\ref{prn:HE11_3}, by further applying the argument in Proposition~\ref{prn:HE11_2},
the following theorem is clear now.
\begin{theorem}[\textbf{Rough long-time pseudolocality theorem for polarized K\"ahler Ricci flow}]
For every group of numbers $\delta, \xi, r_0,T$, there exists an $\epsilon=\epsilon(n,B,\delta, \xi, r_0,T)$
with the following properties.
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$.
Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$, where $\Omega=B_{g(0)}(x_0,r_0)$.
If $\displaystyle \sup_{\mathcal{M}} (|R| + |\lambda|) <\epsilon$, then for every $t \in [0,T]$ we have
\begin{align}
& B_{g(t)}(x_0, (1-2\delta)r_0) \subset \Omega, \\
& |Rm|(\cdot, t) \leq 2r_0^{-2}, \quad \textrm{in} \; B_{g(t)}(x_0, (1-2\delta)r_0), \\
& \left(1-\xi \right) g(0) \leq g(t) \leq \left(1 +\xi \right) g(0), \quad \textrm{in}
\; B_{g(t)}(x_0, (1-2\delta)r_0).
\end{align}
\label{thm:K8_1}
\end{theorem}
\subsection{Motivation and definition}
In previous subsection, we see that the assumption (\ref{eqn:SK23_1}) helps a lot to relate different time slices of
the K\"ahler Ricci flow solution. However, why is this assumption reasonable? This question will be answered in this
subsection.
\begin{proposition}[\textbf{Weak continuity of Bergman function}]
There is a big integer constant $k_0=k_0(n,A)$ and small constant $\epsilon=\epsilon(n,A)$
with the following property.
Suppose $(M,g,J,L,h)$ is a polarized K\"ahler manifold, taken out from a polarized K\"ahler Ricci flow
in $\mathscr{K}(n,A)$ as a central time slice. In particular, we have
\begin{align}
Osc_{M} \dot{\varphi} + C_S(M) + |\lambda| \leq B,
\label{eqn:SK27_1}
\end{align}
where $B=B(n,A)$.
If $\mathbf{cr}(M) \geq 1$, then
\begin{align}
\sup_{1 \leq k \leq k_0} \mathbf{b}^{(k)}(x) > -k_0
\label{eqn:SK27_2}
\end{align}
whenever $d_{PGH}((M,x,g), (\tilde{M}, \tilde{x}, \tilde{g}))<\epsilon$ for some space
$(\tilde{M}, \tilde{x}, \tilde{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{prn:SK27_1}
\end{proposition}
\begin{proof}
Note that the model moduli space $\widetilde{\mathscr{KS}}(n,\kappa)$ has compactness
under the pointed Gromov-Hausdorff topology. This compactness will be essentially used in
the following argument.
Suppose the statement was wrong, then there is a sequence of polarized K\"ahler manifolds,
whose underlying K\"ahler manifolds converge to
$(\bar{M},\bar{x},\bar{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$,
satisfying the estimate (\ref{eqn:SK27_1}) and violating (\ref{eqn:SK27_2}) for $k_i \to \infty$.
To be explicit, we have
\begin{align}
(M_i,x_i,g_i) \longright{G.H.} (\bar{M},\bar{x},\bar{g}); \qquad
\sup_{1 \leq j \leq k_i} \mathbf{b}^{(j)}(x_i) \to -\infty, \quad k_i \to \infty. \label{eqn:K8_4}
\end{align}
Then we shall use the argument of the proof of Theorem 3.2 of~\cite{DS} by Donaldson-Sun to find positive integer $q=q(\bar{x})$, and real numbers $r=r(\bar{x})$, $C=C(\bar{x})$
such that
\begin{align}
\inf_{y \in B(x_i, r)} \mathbf{b}^{(q)}(y) \geq -C.
\label{eqn:GE09_1}
\end{align}
Note that the proof of Theorem 3.2~\cite{DS} is based on a blowup argument.
The essential ingredients there are the convergence theory, the H\"{o}mander's estimate, and the fact that each tangent space in the limit space is a good metric cone.
By ``good" we mean the singular set of the metric cone has Hausdorff codimension strictly greater than $2$.
It is important to observe that whether the limit space $\bar{M}$ is compact or not does not affect the argument.
Basically, this is because of the local property of the H\"{o}mander's estimate.
Actually, no matter whether $\bar{M}$ is compact or not, every tangent space of a point on $\bar{M}$ must be non-compact.
The contradiction is obtained from the convergence to the good tangent metric cone.
With the argument of Theorem 3.2 of~\cite{DS} in mind, we now check the conditions available to us in the current case.
Firstly, the canonical radius assumption makes sure that the topology of the convergence can be improved to the
$\hat{C}^4$-Cheeger-Gromov topology.
Secondly, by the uniform bound of Sobolev constant and $\norm{\dot{\varphi}}{C^0}$, the general H\"ormander's estimate(c.f. section 3 of~\cite{CW4} and section 5 of~\cite{Wa1} for this particular case) can be applied.
Thirdly, we know each tangent space at $\bar{x}$ is a good metric cone, by Theorem~\ref{thm:HE08_1} since $(\bar{M},\bar{x},\bar{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
Therefore, we can use a contradiction blowup argument, like that in Theorem 3.2 of~\cite{DS}, to obtain (\ref{eqn:GE09_1}).
Consequently, we have
\begin{align*}
\mathbf{b}^{(q)}(x_i) \geq -C, \quad \Rightarrow \quad \sup_{j \leq k_i} \mathbf{b}^{(j)} (x_i) \geq -C,
\end{align*}
which contradicts (\ref{eqn:K8_4}), the assumption.
\end{proof}
Proposition~\ref{prn:SK27_1} means that the Bergman function has a weak continuity under the Cheeger-Gromov
convergence if the limit space is the model space. Inspired by this property, we can define the polarized canonical radius as follows.
\begin{definition}
Suppose $\left( M,g,J,L,h\right)$ is a polarized K\"ahler manifold satisfying (\ref{eqn:SK27_1}), $x \in M$. We say the polarized canonical radius of $x$ is not less than $1$ if
\begin{itemize}
\item $\mathbf{cr}(x) \geq 1$.
\item $\displaystyle \sup_{1 \leq j \leq 2k_0}\mathbf{b}^{(j)}(x) \geq -2k_0$.
\end{itemize}
For every $r=\frac{1}{j}, j \in \ensuremath{\mathbb{Z}}^{+}$, we say the polarized canonical radius of $x$ is not less than $r$ if the rescaled polarized manifold $\left( M, j^2 g, J, L^j, h^j\right)$
has polarized canonical radius at least $1$ at the point $x$.
Fix $x$, let $\mathbf{pcr}(x)$ be the supreme of all the $r$ with the above property and call it
as the polarized canonical radius of $x$.
\label{dfn:SK27_1}
\end{definition}
We can define the polarized canonical radius of a manifold as the infimum of the polarized canonical radii of all points in that manifold.
Similarly, we can define the polarized canonical radius of time slices of a flow.
Note that from the above definition, $\mathbf{pcr}$ is always the reciprocal of a positive integer.
It could not be zero because of (\ref{eqn:MC18_1}) in the proof of Lemma~\ref{lma:SK25_1} and the fact that every compact smooth manifold has bounded geometry and positive $\mathbf{cr}$.
Under this terminology, the continuity of Bergman function implies the following corollary.
\begin{corollary}[\textbf{Weak equivalence of $\mathbf{cr}$ and $\mathbf{pcr}$}]
There is a small constant $\epsilon=\epsilon(n,B,\kappa)$ with the following property.
Suppose $(M,g,J,L,h)$ is a polarized K\"ahler manifold satisfying (\ref{eqn:SK27_1}) and
$\mathbf{cr}(M) \geq 1$. Then
\begin{align}
\mathbf{pcr}(x) \geq 1
\label{eqn:SK27_4}
\end{align}
whenever $d_{PGH}((M,x,g), (\tilde{M}, \tilde{x}, \tilde{g}))<\epsilon$ for some space
$(\tilde{M}, \tilde{x}, \tilde{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{cly:SK27_1}
\end{corollary}
\subsection{K\"ahler Ricci flow with lower bound of polarized canonical radius}
Suppose the polarized canonical radius is uniformly bounded from below, then the convergence theory is much better
than that in section 3. This is basically because of the rough long-time pseudolocality theorem, Theorem~\ref{thm:K8_1}.
\begin{proposition}[\textbf{Improving regularity in forward time direction}]
For every $r_0>0$, $r \in (0,r_0)$ and $T_0>0$, there is an $\epsilon=\epsilon(n,A,r_0,r,T_0)$ with the following properties.
If $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK20_1}) and
\begin{align}
\mathbf{pcr}(\mathcal{M}^{t}) \geq r_0, \quad \forall \; t \in [0,T_0],
\label{eqn:SL04_1}
\end{align}
then \begin{align}
\mathcal{F}_{r}(M, 0) \subset \bigcap_{0\leq t \leq T_0} \mathcal{F}_{\frac{r}{K}}(M, t) \label{eqn:SC19_3}
\end{align}
whenever $\displaystyle \sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
Here $K$ is the constant in Proposition~\ref{prn:SC24_1}.
\label{prn:SC09_3}
\end{proposition}
\begin{proof}
It follows directly from Theorem~\ref{thm:K8_1}, the long time pseudolocality theorem for polarized K\"ahler Ricci flow with partial-$C^0$-estimate.
\end{proof}
\begin{proposition}[\textbf{Improving regularity in backward time direction}]
For every $r_0>0$, $r \in (0,r_0)$ and $T_0>0$, there is an $\epsilon=\epsilon(n,A,r_0,r,T_0)$ with the following properties.
If $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK20_1}) and (\ref{eqn:SL04_1}), then \begin{align}
\bigcup_{0 \leq t \leq T_0} \mathcal{F}_{r}(M, t) \subset \mathcal{F}_{\frac{r}{K}}(M,0) \label{eqn:SK27_6}
\end{align}
whenever $\displaystyle \sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
\label{prn:SK27_3}
\end{proposition}
\begin{proof}
At time $0$, $\mathcal{F}_r(M,0) \subset \mathcal{F}_{\frac{r}{K}}(M,0)$ trivially. Suppose $t_0>0$ is the first time such that
(\ref{eqn:SK27_6}) start to fail. It suffices to show that $t_0 >T_0$ whenever $\epsilon$ is small enough. Otherwise, at time $t_0 \in (0,T_0]$,
we can find a point $x_0 \in (\partial \mathcal{F}_{\frac{r}{K}}(M,0)) \cap (\partial \mathcal{F}_{r}(M,t_0))$. In other words, we have
\begin{align*}
\mathbf{cvr}(x_0,0)=\frac{r}{K}, \quad \mathbf{cvr}(x_0,t_0)=r.
\end{align*}
In particular, we have
\begin{align}
\left|B_{g(0)}\left(x_0,\frac{r}{K} \right) \right|_0 = \left(1-\delta_0 \right) \omega_{2n} \left(\frac{r}{K} \right)^{2n}. \label{eqn:SL16_1}
\end{align}
Let $\xi$ be a small number which will be fixed later. Let $\Omega_{\xi}(x_0,t_0)$ be the subset of unit sphere of tangent space of $T_{x_0}(M, g(t_0))$ such that every geodesic (under metric $g(t_0)$) emanating from $x_0$ along the direction in $\Omega_{\xi}(x_0,t_0)$ does not hit points in $\mathcal{D}_{\xi}(M,0)$ before
distance $\frac{r}{K}$. By canonical radius assumption, $|Rm|_{g(t_0)}$ is uniformly bounded in $B_{g(t_0)}(x_0,\frac{r}{K})$(See Figure~\ref{figure:backpseudo} for intuition).
By long-time pseudolocality theorem (c.f. Proposition~\ref{prn:SK27_3}), $B_{g(t_0)}(x_0,\frac{r}{K^3})$ has empty intersection with $\mathcal{D}_{\xi}(M,0)$
when $\xi<<\frac{r}{K^3}$. Note that every geodesic (emanating from $x_0$) entering $\mathcal{D}_{\xi}(M,0)$ must hit $\partial \mathcal{D}_{\xi}(M,0)$ first,
where $\mathbf{cvr}(\cdot, 0)=\xi$. So every point in $\partial \mathcal{D}_{\xi}(M,0)$ will be uniformly regular at time $t_0$, in light of the long-time pseudolocality.
At time $t_0$, observing from $x_0$, the set which stays behind $\partial \mathcal{D}_{\xi}(M,0)$ must have small measure.
Since $B_{g(t_0)}(x_0,\frac{r}{K})$ has uniformly bounded curvature,
it is clear that $\Omega_{\xi}(x_0,t_0)$ is an almost full measure subset of $S^{2n-1}$. Actually, we have
\begin{align*}
|\Omega_{\xi}(x_0,t_0)| \geq 2n\omega_{2n} \cdot \left(1-C \xi^{2p_0} \right)
\end{align*}
whenever $\epsilon$ is sufficiently small. On the other hand, we see that every geodesic (under metric $g(t_0)$) emanating from $\Omega_{\xi}(x_0,t_0)$ is almost geodesic at time $t=0$ (under metric $g(0)$), when $\epsilon$ small enough. Therefore,
$|B_{g(0)}(x_0,\frac{r}{K})|_{0}$ is almost not less than $|B_{g(t_0)}(x_0,\frac{r}{K})|_{t_0}$. Note that the volume ratio of
$B_{g(t_0)}(x_0,\frac{r}{K})$ is at least $(1-\frac{\delta_0}{100})\omega_{2n}$.
Suppose we choose $\xi$ small (according to $\delta_0$) and $\epsilon$ very small (based on $\xi,\delta_0,A,T_0$), we obtain
\begin{align*}
\left|B_{g(0)}\left(x_0,\frac{r}{K} \right) \right|_0 \geq \left(1-\frac{\delta_0}{2} \right) \omega_{2n} \left(\frac{r}{K} \right)^{2n},
\end{align*}
which contradicts (\ref{eqn:SL16_1}).
\end{proof}
\begin{figure}
\begin{center}
\psfrag{0}[c][c]{$0$}
\psfrag{t}[c][c]{$t$}
\psfrag{x0}[c][c]{$x_0$}
\psfrag{M1}[c][c]{$(M,x_0,g(t_0))$}
\psfrag{M2}[c][c]{$(M,x_0,g(0))$}
\psfrag{B1}[c][c]{$yellow: B_{g(t_0)}(x_0,r)$}
\psfrag{B2}[c][c]{$blue: B_{g(t_0)}(x_0,\frac{r}{K})$}
\psfrag{B3}[c][c]{$red: B_{g(t_0)}(x_0,\frac{r}{K^3})$}
\psfrag{B4}[c][c]{$green: B_{g(0)}(x_0,r)$}
\psfrag{B5}[c][c]{$black: \mathcal{D}_{\xi}(M,0)$}
\includegraphics[width=0.5 \columnwidth]{backpseudo}
\caption{Find a geodesic ball with almost Euclidean volume ratio}
\label{figure:backpseudo}
\end{center}
\end{figure}
\begin{definition}
Let $\mathscr{K}(n,A)$ be the collection of polarized K\"ahler Ricci flows satisfying (\ref{eqn:SK20_1}).
For every $r \in (0,1]$, define
\begin{align*}
\mathscr{K}(n,A;r)
\triangleq \left\{\mathcal{LM} \left| \mathcal{LM} \in \mathscr{K}(n,A), \mathbf{pcr}(M \times [-1,1]) \geq r \right. \right\}.
\end{align*}
\label{dfn:SC23_1}
\end{definition}
Clearly, $\mathscr{K}(n,A) \supset \mathscr{K}(n,A;r_1) \supset \mathscr{K}(n,A;r_2)$ whenever $1\geq r_2 >r_1>0$.
Since every polarized K\"ahler Ricci flow $\mathcal{LM} \in \mathscr{K}(n,A)$ has a smooth compact underlying manifold, we see $\mathcal{LM} \in \mathscr{K}(n,A;r)$
for some very small $r$, which depends on $\mathcal{LM}$. Therefore, it is clear that
\begin{align*}
\bigcup_{0<r<1} \mathscr{K}(n,A;r) = \mathscr{K}(n,A).
\end{align*}
Fix $r>0$, we shall first make clear the structure of $\mathscr{K}(n,A;r)$ under the help of polarized canonical radius.
Then we show that the canonical radius can actually been bounded a priori. In other words, there exists a uniform small constant $\hslash$ (Planck scale)
such that
\begin{align*}
\mathscr{K}(n,A) = \mathscr{K}(n,A;\hslash),
\end{align*}
which will be proved in Theorem~\ref{thm:SC28_1}.
\begin{proposition}[\textbf{Limit space-time with static regular part}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies the following properties.
\begin{itemize}
\item $\mathbf{pcr}(M_i \times [-T_i,T_i]) \geq r_0$ for each $i$.
\item $\displaystyle \lim_{i \to \infty} \sup_{\mathcal{M}_i} (|R|+|\lambda|)=0$.
\end{itemize}
Suppose $x_i \in M_i$ and
$\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i,0)>0$, then
\begin{align}
(M_i,x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g}). \label{eqn:MC05_10}
\end{align}
Moreover, we have
\begin{align}
(M_i,x_i, g_i(t)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g}) \label{eqn:MC05_11}
\end{align}
for every $t \in (-\bar{T}, \bar{T})$, where $\displaystyle \bar{T}=\lim_{i \to \infty} T_i>0$.
In particular, the limit space does not depend on time.
\label{prn:SL04_1}
\end{proposition}
\begin{proof}
It follows from the combination of Proposition~\ref{prn:SC09_3} and Proposition~\ref{prn:SK27_3} that the limit
space does not depend on time. From the definition of canonical radius, the convergence locate in
$\hat{C}^4$-topology for each time. However, this can be improved to $\hat{C}^{\infty}$-topology.
Actually, if $\bar{y}$ is a regular point of $\bar{M}$(c.f. the definition in Theorem~\ref{thm:HE11_1}), then we can find
$y_i \in M_i$ such that $y_i \to \bar{y}$ and $\mathbf{cvr}(y_i,0) \geq \eta$ uniformly for some $\eta>0$, in light of equation (\ref{eqn:HA11_3}) and the proof of Theorem~\ref{thm:HE11_1}.
It follows from Proposition~\ref{prn:SK27_3} that $\inf_{t \in [-1,0]} \mathbf{cvr}(y_i, t) \geq K^{-1} \eta$ for all large $i$.
By second property, or regularity estimate of canonical radius (c.f. Definition~\ref{dfn:SC02_1}), we know
\begin{align*}
|Rm|(z,t) \leq CK^{-4} \eta^{-2}, \quad \forall \; z \in B_{g_i(t)}(y_i, K^{-2}\eta), \; t \in [-1, 0].
\end{align*}
Note that $R \to 0$, which implies $|Ric| \to 0$ when we have $|Rm|$-bound in a bigger ball(c.f. the $|Ric| \leq \sqrt{|Rm||R|}$-type estimate in~\cite{Wa2}). In particular, we have
$B_{g_i(0)}(y_i, 0.1 K^{-2} \eta) \subset B_{g_i(t)}(y_i, K^{-2} \eta)$ for all $t \in [-0.5, 0]$. Hence, we obtain
\begin{align}
|Rm|(z,t) \leq CK^{-4} \eta^{-2}, \quad \forall \; z \in B_{g_i(0)}(y_i, 0.1 K^{-2}\eta), \; t \in [-0.5, 0]. \label{eqn:MC05_5}
\end{align}
Then we can apply Shi's estimate to obtain that $|\nabla^k Rm| \leq C_k$ on $B_{g_i(0)}(y_i, 0.01 K^{-2}\eta)$ for each positive integer $k$.
This is enough to set up a uniform sized harmonic coordinate chart around $y_i$ (with respect to metric $g_i(0)$) and all the metric tensor and its derivatives are uniformly bounded(c.f. Hamilton~\cite{Ha95}).
Clearly, the convergence around $\bar{y}$ happens in the $C^{\infty}$-topology. Since $\bar{y}$ is an arbitrary regular point, we see that the convergence to $\bar{M}$ is in $\hat{C}^{\infty}$-Cheeger-Gromov topology.
Note that we currently do not know whether $\bar{M}$ locates in the model space $\widetilde{\mathscr{KS}}(n,\kappa)$. However, we do know that
$\bar{M}=\mathcal{R}(\bar{M}) \cup \mathcal{S}(\bar{M})$. The regular part is a smooth Ricci-flat manifold, due to the smooth convergence and $|Ric| \to 0$ on regular part.
The singular part satisfies the Minkowski dimension bound(c.f. (\ref{eqn:SB13_4}) in Theorem~\ref{thm:HE11_1}):
\begin{align}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-2p_0< 2n-4+\frac{2}{1000n}<2n-4+\frac{2}{2n-1}, \label{eqn:MC05_1}
\end{align}
where we used the choice of $p_0$, which is discussed above the equation (\ref{eqn:SC17_11}).
Recall that each $(M_i, g_i(0))$ has uniform Sobolev constant (c.f. equation (\ref{eqn:SK20_1})) and uniform volume doubling condition(c.f. Q. Zhang~\cite{Zhq3} and Chen-Wang~\cite{CW5}).
Therefore, there is a uniform local $L^2$-Poincar\'e constant. All these estimates bypass to the limit space $(\bar{M}, \bar{x}, \bar{g})$.
There exists a good heat semigroup theory on $\bar{M}$. By the high codimension(c.f. inequality (\ref{eqn:MC05_1})) of $\mathcal{S}(\bar{M})$, we know that for every bounded heat solution $u$ on $\bar{M}$,
the gradient function $|\nabla u|\in N_{loc}^{1,2}(\bar{M})$ and it is a heat subsolution(c.f. Lemma~\ref{lma:HH18_2} and Remark~\ref{rmk:GD11_1}).
Consequently, the technique used in the proof of the Cheeger-Gromoll splitting, i.e., Claim~\ref{clm:GD11_1} in Lemma~\ref{lma:HE04_1},
can be applied here. Basically, every function $f \in N_0^{1,2}(\hat{Y})$ has a better version $\displaystyle \tilde{f}=\lim_{t \to 0^{+}} e^{-\Delta t}(f)$,
whose point-wise gradient(even on $\mathcal{S}(\bar{M})$, understood as weak upper gradient, c.f. Definition~\ref{dfn:HD18_1}, or Cheeger~\cite{Cheeger99}) can be bounded by the $L^{\infty}$-norm of the gradient.
We remark that such technique should be standard in the study of general metric measure spaces.
For example, it was used in the discussion in section 4.1.3 of Gigli's work~\cite{Gigli13}.
Actually, the Lipshitz function approximation Lemma, i.e., Lemma 10.7 of Cheeger's work in 1990's~\cite{Cheeger99}, can be regarded as an predecessor of the technique mentioned above.
For the convenience of the readers who are not familiar with the singular space, we also provide a detailed alternative proof as follows, which only uses the Ricci flow, heat equation, Proposition~\ref{prn:SK27_3}
and the canonical radius assumption.
\begin{claim}[\textbf{Good version of Lipshitz function}]
Every bounded function $f \in N_0^{1,2}(\bar{M})$ with finite $\norm{\nabla f}{L^{\infty}(\bar{M})}$ and Lipschitz on $\mathcal{R}(\bar{M})$
has a good version $\tilde{f}$ such that
\begin{align}
&f(x)=\tilde{f}(x), \quad \forall \; x \in \mathcal{R}(\bar{M}), \label{eqn:MC05_8} \\
&\sup_{\bar{M}} |\nabla \tilde{f}| \leq \norm{\nabla f}{L^{\infty}(\bar{M})}, \label{eqn:MC05_9}
\end{align}
where the inequality (\ref{eqn:MC05_9}) can be understood as
\begin{align*}
|\tilde{f}(x)-\tilde{f}(y)| \leq \norm{\nabla f}{L^{\infty}(\bar{M})} \cdot d(x, y), \quad \forall \; x, y \in \bar{M}.
\end{align*}
\label{clm:MC05_1}
\end{claim}
This is a flow property, so we assume $\lambda=0$ without loss of generality.
For simplicity of notation, we also assume that $\norm{\nabla f}{L^{\infty}(\bar{M})}=1$ and the support of $f$ is contained in $B(\bar{x}, 1)$.
Note that these assumptions can always be achieved up to rescaling argument.
Let $\chi_{\epsilon}=\phi(\frac{d(x, \mathcal{S})}{\epsilon})$ be the cutoff function where $\phi$ is a smooth cutoff function such that
$\phi \equiv 1$ on $(2,\infty)$ and $\phi \equiv 0$ on $(-\infty, 1)$ and $|\nabla \phi| \leq 2$, $\phi' \leq 0$.
Then $\chi_{\epsilon} f$ is a Lipschitz function with compact support.
By the smooth convergence away from singularity(c.f. Proposition~\ref{prn:SK27_3} and the discussion around inequality (\ref{eqn:MC05_5})),
we can regard $\chi_{\epsilon} f$ as a Lipshitz function on $(M_i, g_i(-\delta))$, denoted by $f_{\epsilon, i}$, where $\delta=\epsilon^{\frac{1}{n}}$.
Starting from $f_{\epsilon,i}$, we solve the heat equation until time $t=0$
and obtain a function $h_{\epsilon,i}=f_{\epsilon,i}(0)$, together with the metric evolving by the Ricci flow. Then we have
\begin{align*}
h_{\epsilon, i}(x)=\int_{M_i} w(x,y,-\delta) f_{\epsilon, i}(y) dv_y, \quad \forall \; x \in M_i,
\end{align*}
where $w$ is the fundamental solution of $\square^{*} w=(\partial_{\tau}-\Delta +R) w=0$. Recall that $\int_{M_i} w dv \equiv 1$ and $|f_{\epsilon, i}| \leq C$ uniformly, we have
\begin{align}
|h_{\epsilon, i}|(x)= \left| \int_{M_i} w(x,y,-\delta) f_{\epsilon, i}(y) dv_y \right| \leq \sup_{M_i} |f_{\epsilon, i}| \int_{M_i} w(x,y, -\delta) dv_y = \sup_{M_i} |f_{\epsilon, i}| \leq C.
\label{eqn:MC05_4}
\end{align}
Direct calculation shows that
\begin{align*}
\square |\nabla f_{\epsilon, i}|^2=\left( \partial_t -\Delta \right) |\nabla f_{\epsilon,i}|^2=-2|\nabla \nabla f_{\epsilon, i}|^2 \leq 0.
\end{align*}
It follows that
\begin{align*}
|\nabla h_{\epsilon,i}|^2(x) - \int_{M_i} w(x,y,-\delta) |\nabla f_{\epsilon, i}|^2 (y) dv_y=-2 \int_{-\delta}^{0} \int_{M_i} w(x,y,t) |\nabla \nabla f_{\epsilon, i}|^2 dv_y dt \leq 0.
\end{align*}
Consequently, we have
\begin{align}
|\nabla h_{\epsilon,i}|^2(x) &\leq \int_{M_i} w(x,y,-\delta) |\nabla f_{\epsilon, i}|^2 (y) dv_y \notag\\
&=\int_{\Omega_i \backslash A_i} w(x,y,-\delta) |\nabla f_{\epsilon, i}|^2 (y) dv_y + \int_{A_i} w(x,y,-\delta) |\nabla f_{\epsilon, i}|^2 (y) dv_y, \label{eqn:MC05_2}
\end{align}
where $A_i$ is the set where the pull back of $\chi_{\epsilon}$ achieves values in $(0,1)$, $\Omega_i$ is the support of the pull back function $f_i$. Note that $|\nabla f_{\epsilon, i}|(x) \leq 1+\xi$ for arbitrary small, but fixed $\xi$,
whenever $i$ is large enough and $x \in \Omega_i \backslash A_i$. On $A_i$, we have $|\nabla f_{\epsilon, i}| \leq C \epsilon^{-1}$ for some universal constant $C$. Note that $\Omega_i \subset B_{g_i(0)}(x_i, 1)$, the canonical assumption then
implies the density estimate $ |A_i| \leq C \epsilon^{2p_0}$. Recall that we have the heat kernel(hence conjugate heat kernel estimate) estimate $w(x,y,-\delta) \leq C \delta^{-n}$ for some universal constant $C$.
Plugging these inequalities into (\ref{eqn:MC05_2}), we obtain
\begin{align*}
|\nabla h_{\epsilon,i}|^2(x) &\leq \int_{\Omega_i \backslash A_i} w(x,y,-\delta) |\nabla f_{\epsilon, i}|^2 (y) dv_y + \int_{A_i} w(x,y,-\delta) |\nabla f_{\epsilon, i}|^2 (y) dv_y\\
&\leq (1+\xi) \int_{\Omega_i \backslash A_i} w(x,y,-\delta) dv_y + C\epsilon^{-2} \cdot C \delta^{-n} \cdot |A_i|\\
&\leq (1+\xi) \int_{M_i} w(x,y,-\delta) dv_y + C \epsilon^{2p_0-2} \delta^{-n}\\
&\leq (1+\xi) + C \epsilon^{2p_0-2} \delta^{-n} \leq 1+\xi + C \epsilon^{2-\frac{1}{500n}} \delta^{-n},
\end{align*}
where we used the fact $\epsilon<1$ and inequality (\ref{eqn:MC05_1}) in the last step.
Recall that $\delta=\epsilon^{\frac{1}{n}}$ and let $\xi=\epsilon^{1-\frac{1}{500n}}<<\sqrt{\epsilon}$, we then have
\begin{align}
|\nabla h_{\epsilon,i}|^2(x) \leq1+ \sqrt{\epsilon}. \label{eqn:MC05_3}
\end{align}
Moreover, if $\bar{z}$ is a regular point of $\bar{M}$, i.e., $\bar{z} \in \mathcal{R}(\bar{M})$. Let $z_i \in M_i$ and $z_i \to \bar{z}$. Then we have
\begin{align*}
&\quad \left| h_{\epsilon, i}(z_i) -f_{\epsilon, i}(z_i) \right| \\
&=\left| \int_{M_i} w(z_i,y,-\delta) \left\{ f_{\epsilon, i}(y) - f_{\epsilon, i}(z_i) \right\} dv_y \right|\\
&\leq \int_{B_{g_i(0)}(z_i, \delta^{\frac{1}{4}})} w(z_i,y,-\delta) \left| f_{\epsilon, i}(y) - f_{\epsilon, i}(z_i) \right| dv_y + \int_{M_i \backslash B_{g_i(0)}(z_i, \delta^{\frac{1}{4}})} w(z_i,y,-\delta) \left| f_{\epsilon, i}(y) - f_{\epsilon, i}(z_i) \right| dv_y.
\end{align*}
Note that $\bar{z}$ is regular, we can assume that the regularity scale(for example, $\mathbf{cvr}$) of each $z_i$ is much larger than $\delta=\epsilon^{\frac{1}{n}}$, if we choose $\epsilon$ small enough.
Clearly, $B_{g_i(0)}(z_i, \delta^{\frac{1}{4}}) \cap A_i =\emptyset$, which implies the Lipschitz constant of $f_{\epsilon, i}$ on $B_{g_i(0)}(z_i, \delta^{\frac{1}{4}})$ is uniformly bounded by $C$.
Recall that $\int_{M_i} w dv \equiv 1$. Therefore, we have
\begin{align}
\left| h_{\epsilon, i}(z_i) -f_{\epsilon, i}(z_i) \right| \leq C \delta^{\frac{1}{4}} + C \int_{M_i \backslash B_{g_i(0)}(z_i, \delta^{\frac{1}{4}})} w(z_i,y,-\delta) dv_y. \label{eqn:MC05_6}
\end{align}
The last term of the above inequality is a small term which can be absorbed in $C\delta^{\frac{1}{4}}$.
Actually, let $\psi$ be a cutoff function such that $\psi \equiv 0$ on $B_{g_i(0)}(z_i, 0.5 \delta^{\frac{1}{4}})$, $\psi \equiv 1$ on $M_i \backslash B_{g_i(0)}(z_i, \delta^{\frac{1}{4}})$.
Moreover, $|\nabla \psi|^2 + |\Delta \psi| \leq C \delta^{-\frac{1}{2}}$. This can be done since $\delta^{\frac{1}{4}}$ is much less than the regularity scale of $z_i$.
Now we extend $\psi$ to be a function on space-time by letting $\psi(x,t)=\psi(x)$. Due to Proposition~\ref{prn:SK27_3} and the discussion before(c.f. inequality (\ref{eqn:MC05_5})), we obtain
$|\nabla \psi|^2 + |\Delta \psi| \leq C \delta^{-\frac{1}{2}}$ on $M_i \times [-\delta, 0]$. Consequently, we obtain
\begin{align*}
\frac{d}{dt} \int_{M_i} \psi(y, t) w(z_i,y,t) dv_y=\int_{M_i} ( w \square \psi - \psi \square^* w ) dv_y=- \int_{M_i} w \Delta \psi dv_y.
\end{align*}
As $w$ converges to the $\delta$-function at $z_i$ as $t$ approaches $0$, $\psi(z_i, 0)=0$, we have
\begin{align*}
0- \int_{M_i} \psi(y, -\delta) w(z_i,y,-\delta) dv_y =-\int_{-\delta}^{0} \int_{M_i} w \Delta \psi dv_y dt \geq -C \delta^{-\frac{1}{2}} \int_{-\delta}^{0} \int_{M_i} w dv_y dt =-C \delta^{\frac{1}{2}},
\end{align*}
which implies that
\begin{align*}
\int_{M_i \backslash B_{g_i(0)}(z_i, \delta^{\frac{1}{4}})} \psi(y, -\delta) w(z_i,y,-\delta) dv_y \leq \int_{M_i} \psi(y, -\delta) w(z_i,y,-\delta) dv_y \leq C \delta^{\frac{1}{2}}.
\end{align*}
Plugging the above inequality into (\ref{eqn:MC05_6}), and noticing that $\delta=\epsilon^{\frac{1}{n}}$, we obtain
\begin{align}
\left| h_{\epsilon, i}(z_i) -f_{\epsilon, i}(z_i) \right| \leq C \delta^{\frac{1}{4}} \leq C \epsilon^{\frac{1}{4n}}. \label{eqn:MC05_7}
\end{align}
It follows from the combination of (\ref{eqn:MC05_4}), (\ref{eqn:MC05_3}) and (\ref{eqn:MC05_7}) that there is a limit function $h_{\epsilon}$ on $\bar{M}$.
Let $\epsilon=2^{-i} \to 0$, up to a diagonal sequence argument, we can assume that $h_{2^{-i}, i}$ converges to a limit function $h$, which satisfies
\begin{align*}
&\sup_{\bar{M}} |h| \leq C, \quad \sup_{\bar{M}} |\nabla h| \leq 1= \norm{\nabla f}{L^{\infty}(\bar{M})}, \\
& h(x)=f(x), \quad \forall \; x \in \mathcal{R}(\bar{M}).
\end{align*}
In particular, $h$ is a good version of $f$. We finish the proof of Claim~\ref{clm:MC05_1}.
Based on Claim~\ref{clm:MC05_1}, the proof of (\ref{eqn:MC05_11}) follows from the standard technique used in the proof of Lemma~\ref{lma:HE04_1}.
Actually, for each $t \neq 0$, we already know that $(M_i, x_i, g_i(t))$ converges in the pointed Gromov-Hausdorff topology to some $(\bar{M}', \bar{x}', \bar{g}')$.
We only need to show that $\bar{M}'$ is isometric to $\bar{M}$.
By Proposition~\ref{prn:SK27_3} and the fact $|R|+|\lambda| \to 0$, we know that there is a natural identification map between $\mathcal{R}(\bar{M})$ and $\mathcal{R}(\bar{M}')$, which contain a common point $\bar{x}$.
In the following discussion, we shall show that this identification map can be extended to an isometry between $\bar{M}$ and $\bar{M}'$.
Let $\bar{y}, \bar{z}$ be two regular points of $\bar{M}$. Clearly, $\bar{y}$ and $\bar{z}$ can also be regarded as regular points on $\bar{M}'$. We omit the identification map for the simplicity of notations.
Suppose $d_{\bar{g}}(\bar{y}, \bar{z})=D>0$. We can construct a function $\chi$ on $\bar{M}'$ as follows
\begin{align}
\chi(x) =
\begin{cases}
\max\{ D-d_{\bar{g}}(x, \bar{y}), 0\}, \quad \; &\textrm{if} \quad x \in \mathcal{R}(\bar{M}'), \\
0, \quad &\textrm{if} \quad x \in \mathcal{S}(\bar{M}').
\end{cases}
\label{eqn:MC07_1}
\end{align}
Fix point $x \in \mathcal{R}(\bar{M}') \backslash B_{\bar{g}'}(\bar{y}, 3D)$, every smooth curve connecting $x$ and $\bar{y}$ has length as least $3D$.
In light of inequality (\ref{eqn:HE11_1}) in Theorem~\ref{thm:HE11_1}(applying to both $\bar{g}$ and $\bar{g}'$), we know that
\begin{align*}
\min\{d_{\bar{g}}(x, \bar{y}), d_{\bar{g}'}(x, \bar{y}) \} \geq D,
\end{align*}
which implies that $\chi(x)=0$ by definition equation (\ref{eqn:MC07_1}). We remark that inequality (\ref{eqn:HE11_1}) together with the high codimension of $\mathcal{S}$ implies that
$\chi \in N_0^{1,2}(\bar{M}')$ and we have
\begin{align*}
\norm{\nabla \chi}{L^{\infty}(\bar{M}')}=\norm{\nabla \chi}{L^{\infty}(\mathcal{R}(\bar{M}'))}=\norm{\nabla \chi}{L^{\infty}(\mathcal{R}(\bar{M}))}=1.
\end{align*}
Then we can apply Claim~\ref{clm:MC05_1} to obtain a good version $\tilde{\chi}$ of $\chi$. In particular, we have
\begin{align}
d_{\bar{g}}(\bar{y}, \bar{z})=D=\left| \chi(\bar{y}) - \chi(\bar{z}) \right|=\left| \tilde{\chi}(\bar{y}) - \tilde{\chi}(\bar{z}) \right| \leq d_{\bar{g}'}(\bar{y}, \bar{z}) \norm{\nabla \chi}{L^{\infty}(\bar{M}')} \leq d_{\bar{g}'}(\bar{y}, \bar{z}).
\label{eqn:MC07_2}
\end{align}
Similarly, by reversing the role of $\bar{g}'$ and $\bar{g}$ when we choose the test function, we obtain that
\begin{align}
d_{\bar{g}'}(\bar{y}, \bar{z}) \leq d_{\bar{g}}(\bar{y}, \bar{z}).
\label{eqn:MC07_3}
\end{align}
By the arbitrary choice of $\bar{y}, \bar{z}$, we know the identity map between $\mathcal{R}(\bar{M})$ and $\mathcal{R}(\bar{M}')$ is an isometry map by (\ref{eqn:MC07_2}) and (\ref{eqn:MC07_3}).
Since $\mathcal{R}(\bar{M})$ is dense in $\bar{M}$, $\mathcal{R}(\bar{M}')$ is dense in $\bar{M}'$, we obtain $\bar{M}$ and $\bar{M}'$ are isometric to each other by taking metric completion.
Consequently, (\ref{eqn:MC05_11}) follows from (\ref{eqn:MC05_10}).
\end{proof}
In Proposition~\ref{prn:SL04_1},
we show that the limit flow exists and is static in the regular part, whenever we have $|R|+|\lambda| \to 0$.
It is possible that the limit points in the singular part $\mathcal{S}$ are moving as time evolves. However, this possibility
will be ruled out finally(c.f. Proposition~\ref{prn:HE15_1}).
\subsubsection{Tangent structure of the limit space}
In this subsection, we shall show that the tangent space of each point in the limit space has a metric cone structure, provided polarized canonical radius is uniformly bounded below. Basically, the cone structure is induced from
the localized $W$-functional's monotonicity. Up to a parabolic rescaling, we can assume $\lambda=0$ without loss
of generality.
\begin{proposition}[\textbf{Local $W$-functional}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;r_0)$ and
$\sup_{\mathcal{M}_i} (|R|+|\lambda|) \to 0$.
Let $u_i$ be the fundamental solution of the
backward heat equation $\left[-\frac{\partial}{\partial t} - \triangle + R \right] u_i =0$ based at the space-time point $(x_i, 0)$. Then $u_i$ converges to
a limit positive solution $\bar{u}$ on $\mathcal{R} \times (-1, 0]$, i.e.,
\begin{align*}
\left[-\frac{\partial}{\partial t} - \Delta + R \right] \bar{u}=0.
\end{align*}
Moreover, we have
\begin{align}
\iint_{\mathcal{R} \times (-1, 0]} 2|t| \left|Ric+\nabla \nabla \bar{f} +\frac{\bar{g}}{2t} \right|^2 \bar{u} dv_{\bar{g}} dt \leq C,
\label{eqn:SC16_4}
\end{align}
where $C=C(n,A)$, $\bar{u}=(4\pi |t|)^{-n} e^{-\bar{f}}$.
\label{prn:SC16_1}
\end{proposition}
\begin{proof}
This is a flow property and has nothing to do with polarization. So we can assume $\lambda=0$ for simplicity.
Fix $r>0$. Choose a point $\bar{y} \in \mathcal{R}_{r}$ and a time $\bar{t}<0$. Without loss of generality, we assume that
there is a sequence of points $(y_i, t_i)$ converging to $(\bar{y}, \bar{t})$. Note that $d_{g_i(0)}(y_i, x_i)$ is uniformly bounded. It is not hard to see that $u_i$ is uniformly bounded around $(y_i, t_i)$.
Actually, let $w_i$ be the heat equation $\square w_i=(\partial_t -\Delta) w_i=0$, starting from a $\delta$-function at $(y_i, t_i)$. Then by the heat kernel estimate of Cao-Zhang(c.f.~\cite{CaoZhang}), we obtain the
on-diagonal bound $\frac{1}{C} |t_i|^{-n}<w_i(y_i, 0)<C|t_i|^{-n}$ for some uniform constant $C$. Then the gradient estimate of Cao-Hamilton-Zhang(c.f.~\cite{Zhq2},~\cite{CaHa}) and the fact $d_{g_i(0)}(y_i, x_i)<C$ implies that $|\log w_{i}(x_i, 0)|$
is uniformly bounded. Note that $w_i(x_i, 0)=u_i(y_i, t_i)$ since the integral $\int_M u_i v_i d\mu$ does not depend on time. Therefore, we have
\begin{align}
\frac{1}{C} \leq u_i(y_i, t_i)=w_i(x_i, 0) \leq C, \label{eqn:MC13_3}
\end{align}
where $C$ depends on $|t_i|$ and $d_{g_i(0)}(y_i, x_i)$. It clearly works uniformly for a fixed-sized space-time neighborhood of $(y_i, t_i)$, where curvatures are uniformly bounded.
Then standard regularity argument from heat equation shows that all derivatives of $u_i$ are uniformly bounded around $(y_i, t_i)$. Therefore, there is a limit positive solution $\bar{u}$ around $(\bar{y}, \bar{t})$. By the arbitrary choice of $r,\bar{y}, \bar{t}$. It is clear
that there is a smooth heat solution $\bar{u}$ defined on $\mathcal{R} \times (-1, 0)$.
By Perelman's calculation, for each flow $g_i$, we have
\begin{align}
\int_{-1}^{0}\int_{M_i} 2|t| \left|Ric_{g_i}+\nabla \nabla f_i +\frac{g_i}{2t} \right|^2 u_i dv_{g_i} dt=-\mu(M_i, g_i(t_i), 1) \leq C,
\label{eqn:SK17_1}
\end{align}
since Sobolev constant is uniformly bounded. By passing to limit, (\ref{eqn:SC16_4}) follows.
\end{proof}
\begin{theorem}[\textbf{Tangent cone structure}]
Suppose $\mathcal{LM}_i$ is a sequence of polarized K\"ahler Ricci flow solutions in $\mathscr{K}(n,A;r_0)$,
$x_i \in M_i$. Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of
$(M_i,x_i,g_i(0))$, $\bar{y}$ be an arbitrary point of $\bar{M}$.
Then every tangent space of $\bar{y}$ is an irreducible metric cone.
\label{thm:SC09_1}
\end{theorem}
\begin{proof}
Suppose $\hat{Y}$ is a tangent space of $\bar{M}$ at the point $\bar{y}$, i.e., there are scales $r_k\to 0$ such that
\begin{align}
(\hat{Y}, \hat{y}, \hat{g})=\lim_{k \to \infty} (\bar{M}, \bar{y}, \bar{g}_k) \label{eqn:MC12_1}
\end{align}
where $\bar{g}_k= r_k^{-2}\bar{g}$. By taking subsequence if necessary, we can assume $ (\hat{Y}, \hat{y}, \hat{g})$
as the limit space of $(M_{i_k}, y_{i_k}, \tilde{g}_{i_k})$ where $\tilde{g}_{i_k}=r_i^{-2}g_{i_k}(0)$.
Denote the regular part of $\hat{Y}$ by $\mathcal{R}(\hat{Y})$.
Then on the space-time
$\mathcal{R}(\hat{Y}) \times (-\infty, 0]$, there is
a smooth limit backward heat solution $\hat{u}$.
Recall that $\hat{u}$ is positive by Proposition~\ref{prn:SC16_1}.
For every compact subset $K \subset \mathcal{R}(\hat{Y})$ and positive number $H$, it follows from Cheeger-Gromov convergence and
the estimate (\ref{eqn:SK17_1}) that
\begin{align*}
\iint_{K \times [-H, 0]} 2|t| \left| Ric + \nabla \nabla \hat{f} + \frac{\hat{g}}{2t}\right|^2 \hat{u} dv dt =0.
\end{align*}
Note the scaling invariance of $\hat{u} dv$ and $|t| \left| Ric + \nabla \nabla \hat{f} + \frac{\hat{g}}{2t}\right|^2 dt$. Actually, if the above equality fails for some $K$ and $H$, then by definition of tangent space and
the integral accumulation, we shall obtain the left hand side of (\ref{eqn:SK17_1}) is infinity and obtain a contradiction.
Then by the arbitrary choice of $K$ and $H$, we arrive
\begin{align*}
\iint_{\mathcal{R}(\hat{Y}) \times (-\infty, 0]} 2|t| \left| Ric + \nabla \nabla \hat{f} + \frac{\hat{g}}{2t}\right|^2 \hat{u} dv dt =0.
\end{align*}
Note that $\mathcal{R}(\hat{Y})$ is Ricci flat. So there is a smooth function $\hat{f}$ defined on $\mathcal{R}(\hat{Y}) \times (-\infty, 0]$ such that
\begin{align}
\nabla \nabla \hat{f} + \frac{\hat{g}}{2t} \equiv 0.
\label{eqn:SL13_1}
\end{align}
The above equation means that $\nabla \hat{f}$ is a conformal Killing vector field, when restricted on each time slice $t<0$.
It follows from the work of Cheeger-Colding(c.f.~\cite{CCWarp}) that there is a local cone structure around each regular point.
We shall show that a global cone structure can be obtained due to the high co-dimension of the singular set $\mathcal{S}$
and the Killing property arised from (\ref{eqn:SL13_1}).
The basic techniques we shall use in our proof is very similar to that in the proof of Lemma~\ref{lma:HE04_1}, Lemma~\ref{lma:HE04_2} and Proposition~\ref{prn:SL04_1}.
Let's first list the excellent properties of $\hat{f}$. Recall that $\hat{f}$ satisfies the following differential equation on $\mathcal{R} \times (-\infty, 0)$ from the limit process.
\begin{align}
\hat{f}_{t}=-\Delta \hat{f} + |\nabla \hat{f}|^2-R -\frac{n}{t}=|\nabla \hat{f}|^2. \label{eqn:MC12_6}
\end{align}
On the other hand, it follows from (\ref{eqn:SL13_1}) that
\begin{align*}
\nabla \left(t|\nabla \hat{f}|^2 +\hat{f} \right)=2t Hess_{\hat{f}}(\nabla \hat{f}, \cdot) + \nabla \hat{f}= -\nabla \hat{f} + \nabla \hat{f} \equiv 0.
\end{align*}
So we have $t|\nabla \hat{f}|^2 +\hat{f}=C(t)$, whose time derivatives calculation yields that
\begin{align*}
C'(t)&=|\nabla \hat{f}|^2 + 2t \left \langle \nabla \hat{f}, \nabla \hat{f}_t \right \rangle + \hat{f}_t
=2|\nabla \hat{f}|^2 +2t\left\langle \nabla \hat{f}, \nabla |\nabla \hat{f}|^2 \right \rangle\\
&=2|\nabla \hat{f}|^2 +4t Hess_{\hat{f}} \left(\nabla \hat{f}, \nabla \hat{f} \right)=2|\nabla \hat{f}|^2-2|\nabla \hat{f}|^2=0,
\end{align*}
where we repeatedly used (\ref{eqn:MC12_6}) and (\ref{eqn:SL13_1}). Therefore, $t|\nabla \hat{f}|^2 +\hat{f} \equiv C$ on $\mathcal{R} \times (-\infty, 0)$.
Replacing $\hat{f}$ by $\hat{f}+C$ if necessary, we can assume that $t|\nabla \hat{f}|^2 + \hat{f} \equiv 0$, which implies that
\begin{align}
\left(t\hat{f} \right)_{t}=\hat{f}+ t \hat{f}_t=\hat{f}+t|\nabla \hat{f}|^2 \equiv 0. \label{eqn:MC12_5}
\end{align}
Consequently, we have
\begin{align}
&\hat{f}(x, t)= \frac{-1}{t} \hat{f}(x, -1), \quad \forall \; x \in \mathcal{R}(\hat{Y}). \label{eqn:MC12_7}\\
& \left|\nabla \sqrt{\hat{f}(x, t)} \right|=\frac{1}{2|t|}. \label{eqn:MC15_2}
\end{align}
We remark that the above discussion is nothing but the application of general property of gradient shrinking solitons(c.f. Chapter 4 of Chow-Lu-Ni~\cite{CLN}), in the special case that $Ric \equiv 0$.
Intuitively, a space which is both Ricci-flat and is a gradient shrinking soliton must be a metric cone.
This can be easily proved if the underlying space is smooth.
In our current situation, due to high codimension of $\mathcal{S}$, the cone structure can be established using the technique developed in section 2.
Suppose $\hat{Y}$ is a metric cone based at $\hat{y}$, then we should have
\begin{align}
\hat{f}=\frac{d^2}{4|t|}, \label{eqn:MC17_7}
\end{align}
where $d$ is the distance to the origin. This will be confirmed in the following discussion.
The cone structure of $\hat{Y}$ will be established together with equality (\ref{eqn:MC17_7}).
The basic idea to prove (\ref{eqn:MC17_7}) is to compare the level sets of $\hat{f}$ with geodesic balls, with more and more preciseness.
Note that similar ideas to estimate distance will be essentially used in section 5.3(c.f. Lemma~\ref{lma:HA15_1}).
We remark that our proof could be much simpler if we use Lemma~\ref{lma:SK27_4}, which is independent(c.f. Remark~\ref{rmk:MC07_1}).
For example, the application of Lemma~\ref{lma:SK27_4} directly implies that $\hat{f}$ must achieve minimum only at base point $\hat{y}$(see step 3 below),
since $\hat{f}$ is a strictly convex function in regular part $\cal R$ and can be regarded as a continuous function on $\hat{Y}$(c.f. setp 1 below).
Here we want to give a self-contained proof, using only the good property of $\hat{f}$ to improve the regularity of $\hat{Y}$.
We divide the proof of (\ref{eqn:MC17_7}) into four steps.
\textit{Step 1. $\hat{f}$ is a nonngative, continuous, proper function which achieves minimum value $0$ at $\hat{y}$. }
Let us focus our attention on time slice $t=-1$ for a while. Denote $\hat{f}(x, -1)$ by $\hat{f}(x)$ for simplicity of notation.
It is not hard to observe that $\hat{f}(x)$ is weakly proper. In other words, we have
\begin{align}
\lim_{\mathcal{R}(\hat{Y}) \ni x \to \infty} \hat{f}(x)=\infty. \label{eqn:MC17_4}
\end{align}
For otherwise, we can find a sequence of points $z_i \in \mathcal{R}(\hat{Y})$ such that
$d(z_i, \hat{y}) \to \infty$ and $\hat{f}(z_i) \leq D$ for some positive number $D$.
Note that $\hat{f}$ is uniformly bounded from below in the ball $B(z_i, 1)$.
Actually, for every smooth point $x \in B(z_i, 1)$, we can find a smooth curve $\gamma$ connecting $x$ to $z_i$ such that $|\gamma| \leq 3 d(x,z_i)$.
This is an application of inequality (\ref{eqn:HE11_1}) in Theorem~\ref{thm:HE11_1}. Note that the caonical radius is very large in the current situation.
Parametrize $\gamma$ by arc length and let $\gamma(0)=z_i$ and $\gamma(L)=x$. Then $|\gamma|=L \leq 3$.
Along the curve $\gamma$, by (\ref{eqn:MC15_2}), we have
\begin{align*}
\frac{d}{ds} \sqrt{\hat{f}}(\gamma(s))=\left \langle \nabla \sqrt{\hat{f}}, \dot{\gamma}(s) \right \rangle \leq \left|\nabla \sqrt{\hat{f}} \right| = \frac{1}{2}.
\end{align*}
Integration of the above inequality implies that
\begin{align}
\hat{f}(x)=\hat{f}(\gamma(L)) \leq \left( \frac{1}{2}L + \hat{f}(\gamma(0)) \right)^2= \left( \frac{1}{2}L + \hat{f}(z_i) \right)^2 \leq (1.5+D)^2 \leq 2(1+D)^2.
\label{eqn:MC13_7}
\end{align}
The above inequality holds for every regular point $x \in B(z_i,1)$. In particular, we know $\int_{B(z_i, 1)} e^{-\hat{f}} dv$ is uniformly bounded from below by some $C^{-1}$.
Consequently, we have
\begin{align*}
\int_{B(z_i,1)} \hat{u} dv = (4\pi)^{-n} \int_{B(z_i,1)} e^{-\hat{f}} dv \geq \frac{1}{C},
\end{align*}
for some uniform constant $C$ depending on $\kappa$ and $D$. Up to reselecting a subsequence if necessary, we can assume that all $B(z_i,1)$ are disjoint to each other. Then we have
\begin{align*}
C \geq \sum_{i=1}^{\infty} \int_{B(z_i, 1)} \hat{u} dv \geq \infty,
\end{align*}
which is impossible. This contradiction establishes the proof of (\ref{eqn:MC17_4}).
Note that in the above disccusion, we already know that the function $\hat{f}$ is bounded on $B \cap \mathcal{R}(\hat{Y})$ for each fixed geodesic ball $B$, by the application of the proof of (\ref{eqn:MC13_7}).
Consequently, we have uniform gradient estimate of $\hat{f}$ in $B \cap \mathcal{R}(\hat{Y})$ by (\ref{eqn:MC12_5}), since $t=-1$.
The locally Lipschitz condition guarantees that $\hat{f}$ can be extended as a continuous function on whole $\hat{Y}$.
Actually, let $\bar{z}$ be a singular point on $\hat{Y}$. Suppose $a_k$ and $b_k$ are two sequences of regular points in $\mathcal{R}(\hat{Y})$ converging to $\bar{z}$.
Clearly, $d(a_k, b_k) \to 0$. By inequality (\ref{eqn:HE11_1}) in Theorem~\ref{thm:HE11_1}, we can find a smooth curve $\gamma_k \subset \mathcal{R}(\hat{Y})$ connecting
$a_k, b_k$ such that $|\gamma_k|<3d(a_k, b_k) \to 0$. The bound of $|\nabla \hat{f}|$ then implies that $|\hat{f}(a_k) -\hat{f}(b_k)| \to 0$. So we can define
$\displaystyle \hat{f}(\bar{z}) \triangleq \lim_{y \to \bar{z}, y \in \mathcal{R}(\hat{Y})} \hat{f}(y)$ without ambiguity(c.f. Proposition~\ref{prn:HD16_1} for similar discussion).
Therefore, from now on we can regard $\hat{f}$ as a continuous function on $\hat{Y}$, rather than only on $\mathcal{R}(\hat{Y})$.
Clearly, the previous discussion implies that $\hat{f}$ is proper. Namely, we have
\begin{align}
\lim_{\hat{Y} \ni x \to \infty} \hat{f}(x)=\infty. \label{eqn:MC17_5}
\end{align}
Consequently, the minimum value of $\hat{f}$ can be achieved at some point $\hat{z}$.
The above discussion can be trivially extended for the function $\hat{f}(\cdot, t)$ for each $t \in (-\infty, 0)$.
So we know $\hat{f}(\cdot, t)$ is a continuous proper function, which achieves minimum value at $\hat{z}$ also, by (\ref{eqn:MC12_7}).
Furthermore, it is also clear that (\ref{eqn:MC12_7}) and the first part of (\ref{eqn:MC12_5}) can be extended to hold on whole $\hat{Y} \times (-\infty, 0)$.
Then we observe that
\begin{align}
\hat{f}(\hat{y}, t)=\min_{x \in \hat{Y}} \hat{f}(x, t) =0, \quad \forall \; t \in (-\infty, 0). \label{eqn:MC17_2}
\end{align}
Actually, following the discussion around inequality (\ref{eqn:MC13_3}), we can use the on-diagonal estimate of Cao-Zhang and the gradient estimate of Cao-Hamilton-Zhang
to obtain that
\begin{align*}
(4\pi |t|)^{-n} e^{-\hat{f}(x,t)-C}=\hat{u}(x, t) \geq \frac{1}{C}|t|^{-n}, \quad \forall \; x \in B\left(\hat{y}, \sqrt{|t|} \right) \cap \mathcal{R}(\hat{Y}),
\end{align*}
where we used the fact that we adjusted $\hat{f}$ globally by adding a constant to obtain (\ref{eqn:MC12_5}). By the continuity of $\hat{f}$, the above inequality implies that
\begin{align*}
\frac{\hat{f}(\hat{y}, -1)}{|t|}=\hat{f}(\hat{y}, t) \leq C, \quad \Rightarrow \quad \hat{f}(\hat{y}, -1) \leq C|t|, \quad \forall \; t \in (-\infty, 0).
\end{align*}
This forces that $\hat{f}(\hat{y}, -1)=0$. Recall that $\hat{f}$ is a nonnegative function by (\ref{eqn:MC12_5}), so we obtain $\displaystyle \min_{x \in \hat{Y}} \hat{f}(x, -1)=0$.
Then (\ref{eqn:MC17_2}) follows from the extended version of (\ref{eqn:MC12_7}). So we finish Step 1.
\textit{Step 2. Unit level set of $\hat{f}$ is comparable with unite geodesic ball centered at $\hat{y}$.}
For each nonnegative number $a$, we define $\Omega_a \triangleq \{x \in \hat{Y} | \hat{f}(x, -1) \leq a^2 \}$.
According to this definition, we immediately know that $\hat{y} \in \Omega_0$. Furthermore, by (\ref{eqn:MC12_7}), it is clear that
\begin{align*}
\Omega_a = \{x \in \hat{Y} | \hat{f}(x, t) \leq |t|^{-1}a^2 \}, \quad \forall \; t \in (-\infty, 0).
\end{align*}
Note that $\Omega_1$ is bounded by the properness of $\hat{f}(\cdot)=\hat{f}(\cdot, -1)$.
For simplicity, we assume that $ \Omega_1 \subset B(\hat{y}, 0.5H)$ for some $H>0$.
On the other hand, applying the gradient estimate of $\sqrt{\hat{f}}$, i.e., (\ref{eqn:MC15_2}), and the smooth curve length estimate (\ref{eqn:HE11_1}), we have
\begin{align*}
\sqrt{\hat{f}(x)} \leq \sqrt{\hat{f}(\hat{y})} + \frac{1}{2} \cdot 4 \cdot H \leq 2H, \quad \forall \; x \in B(\hat{y}, H)
\end{align*}
which means that $B(\hat{y}, H) \subset \Omega_{2H}$. Let $D=2H$, we have the following relationships in short:
\begin{align}
\Omega_1 \subset B(\hat{y}, 0.25D) \subset B(\hat{y}, 0.5D) \subset \Omega_{D}. \label{eqn:MC16_5}
\end{align}
Euqation (\ref{eqn:MC16_5}) can be regarded as the first step to improve (\ref{eqn:MC17_5}) and (\ref{eqn:MC17_2}).
In order to obtain the estimates of general level sets of $\hat{f}$, we need to use the conformal Killing equation (\ref{eqn:SL13_1}).
We observe that the space-time vector field $(-\nabla \hat{f}, \frac{\partial}{\partial t})=(-\frac{\hat{r}}{2} \frac{\partial}{\partial \hat{r}}, \frac{\partial}{\partial t})=(-0.5 \hat{r}\partial_{\hat{r}}, \partial_t)$, as the ``lift"
of the conformal Killing vector field $-\nabla \hat{f}$(c.f. (\ref{eqn:SL13_1})), has many excellent properties. First, direct calculation(c.f. (\ref{eqn:MC12_6})) shows that
\begin{align}
\frac{d}{dt} \hat{f}=\hat{f}_t -|\nabla \hat{f}|^2 \equiv 0 \label{eqn:MC16_3}
\end{align}
along the integral curve of this space-time vector field. Second, it follows from (\ref{eqn:SL13_1}) that
\begin{align}
\mathcal{L}_{(-\nabla \hat{f}, \frac{\partial}{\partial t})} \left\{ |t|\hat{g} \right\}=0. \label{eqn:MC16_4}
\end{align}
Now we can regard $\hat{Y} \times (-\infty, 0)$ as a Riemannian manifold, equipped with metric $|t|\hat{g}(t) + dt^2$(c.f. section 6 of Perelman~\cite{Pe1}).
Then $(-\nabla \hat{f}, \frac{\partial}{\partial t})$ is really a Killing vector field.
\textit{Step 3. $\hat{f}$ and $d(\hat{y}, \cdot)$ have the same unique minimum value point $\hat{y}$.}
In other words, the infimum of $\hat{f}$ must be $0$ and it is only achieved at base point $\hat{y}$.
We shall use Killing vector field to generate quasi-isometric diffeomorphisms.
Then an application of the technique, i.e., bounding distance by choosing good Lipshitz functions, used in the proof Lemma~\ref{lma:HE04_1}, Lemma~\ref{lma:HE04_2} and Proposition~\ref{prn:SL04_1} will imply the diameter bound for general
level sets $\Omega_a$. For small $a$, we shall show that $\diam \Omega_a$ is also small. Then $\Omega_0$ has diameter $0$ and consists of only one point $\hat{y}$, which is of course the unique minimum point of $d(\hat{y}, \cdot)$.
Actually, if one only want to show $\Omega_0 =\{\hat{y}\}$, then there is a shortcut by using the uniform convexity of $\hat{f}$ on $\mathcal{R}(\hat{Y})$ (i.e. equation (\ref{eqn:SL13_1})), the homogeneity of $\hat{f}$
in time direction(i.e., equation (\ref{eqn:MC12_7})), the fact $\int_{\hat{Y}} e^{-\hat{f}} dv<C$ and the application of Lemma~\ref{lma:SK27_4}.
We leave the details to interested readers. In the following paragraph, we shall show $\Omega_0=\{\hat{y}\}$ together with the construction of the cone structure.
Killing vector property together with high codimension of $\mathcal{S}$ implies metric product rigidity, as we have done in Lemma~\ref{lma:HE04_1}, Lemma~\ref{lma:HE04_2}.
We repeat the dicussion here again for the convenience of the readers.
Fix each positive integer $k$. We claim that there is a bounded closed set $E_k \subset \hat{Y}$ satisfying $\dim_{\mathcal{M}}(E_k)<2n-2$.
Furthermore, for each $t \in [-2^{-k}, -2^{-k}]$, we have a family of smooth diffeomorphism $\varphi_{k,t}$
from $\Omega_{D} \backslash E_{k}$ to $\Omega_{\sqrt{2^k|t|}D} \backslash E_{k}$ with
\begin{align}
\varphi_k^*(\hat{g})(z)=2^k|t| \hat{g}(z), \quad \forall \; t \in [-2^{-k}, -2^{-k}], \; z \in \Omega_{\sqrt{2^k|t|}D} \backslash E_{k}. \label{eqn:MC17_6}
\end{align}
The set $E_k$ can be constructed similarly as the set $E_k$ in the proof of Claim~\ref{clm:MA20_1}.
Now the Killing vector field $\nabla b^{+}$ is replaced by the space-time ``Killing"(c.f. (\ref{eqn:MC16_4})) vector field $(-\nabla \hat{f}, \partial_t)$.
Let's describe more details about the construction of $E_k$. Actually, fixing a small positive number $\xi$, we define the set $E_{k,\xi}^{-}$ to be
\begin{align}
\{x \in \Omega_D | \textrm{flow line of} \; (-\nabla \hat{f}, \partial_t) \; \textrm{passing through} \; (x,-2^{-k}) \; \textrm{hits} \; \mathcal{D}_{\xi} \; \textrm{at some} \; t \in (-2^{k}, -2^{-k})\}.
\label{eqn:MC17_1}
\end{align}
The minus sign in $E_{k,\xi}^{-}$ indicates that we are flowing backward along the space-time integral curve of $(-\nabla \hat{f}, \partial_t)$, since $-2^{-k}>t$ for each $t \in (-2^{k}, -2^{-k})$.
Note that the intersection point to $\mathcal{D}_{\xi}$ locates in a uiformly bounded set.
This can be simply proved as follows.
Let $(y, -\tau)$ be the first point on $\mathcal{D}_{\xi}$.
By (\ref{eqn:MC16_3}) and (\ref{eqn:MC12_7}), we have
\begin{align*}
\frac{\hat{f}(y, -1)}{\tau}=\hat{f}(y, -\tau)=\hat{f}(x, -2^{-k})=2^k \hat{f}(x, -1) \leq2^k \cdot D^2, \quad \Rightarrow \quad \hat{f}(y, -1) \leq 2^k \tau D^2 \leq 4^k D^2,
\end{align*}
which means that $y \in \Omega_{2^k D}$, a uniformly bounded set by the properness of $\hat{f}$.
By high Minkowski codimension of $\mathcal{S}$ and the application of the Killing condiiton (\ref{eqn:MC16_4}), similar argument for (\ref{eqn:MA20_3}) in Claim~\ref{clm:MA20_1} implies that
\begin{align*}
|E_{k,\xi}^{-}| \leq C \xi^{2p_0-1-\epsilon},
\end{align*}
where $p_0$ is the constant appeared in (\ref{eqn:MC05_1}), i.e., $\dim_{\mathcal{M}} \mathcal{S}<2n-2p_0$, $C$ may depends on $\epsilon$ also.
Let $\xi_i \to 0$ and define $E_k^{-}=\cap_{i=1}^{\infty} E_{k, \xi}^{-}$. We obtain a measure-zero closed set $E_{k}^{-}$. Moreover, same as (\ref{eqn:MC14_0}) in the proof of Claim~\ref{clm:MA20_1},
the $\xi$-neighborhood of $E_{k}^{-}$ is contained in $E_{k, C\xi}^{-}$
for some uniform constant $C$. Then the above volume estimate implies that $\dim_{\mathcal{M}} E_{k}^{-} \leq 2n-2p_0+1<2n-2$.
Now we reverse the direction. Similar to the definition of $E_{k,\xi}^{-}$ in (\ref{eqn:MC17_1}), we can define $E_{k,\xi}^{+}$ as follows:
\begin{align*}
\{x \in \Omega_{2^k D} | \textrm{flow line of} \; (-\nabla \hat{f}, \partial_t) \; \textrm{passing through} \; (x,-2^{k}) \; \textrm{hits} \; \mathcal{D}_{\xi} \; \textrm{at some} \; t \in (-2^{k}, -2^{-k})\}.
\end{align*}
Clearly, the plus sign in $E_{k,\xi}^{+}$ indicates that we are flowing forward along the space-time integral curve of $(-\nabla \hat{f}, \partial_t)$, since $t>-2^{k}$ for each $t \in (-2^k, -2^{-k})$.
Suppose we start from $(x, -2^{k})$ outside $\mathcal{D}_{\xi}$ and the flow line of $(-\nabla \hat{f}, \partial_t)$ enters $\mathcal{D}_{\xi}$ at some $(y, -\tau)$.
We know $\hat{f}(y, -\tau)=\hat{f}(x, -2^{k})$ since the flow preserves $\hat{f}$-value. Then we have
\begin{align*}
\hat{f}(y, -1)=\tau \hat{f}(y, -\tau)=\tau \hat{f}(x, -2^k)=2^{-k}\tau \hat{f}(x, -1) \leq \hat{f}(x, -1) \leq 4^k D^2.
\end{align*}
Consequently, $y \in \Omega_{2^k D}$. Therefore, the forward flow is also restricted in a bounded domain when we start from a point $(x, -2^{k})$ satisfying $x \in \Omega_{2^k D}$.
Applying high codimension of $\mathcal{S}$ and Killing condition again, we know $E_{k,\xi}^{+}$ has volume bounded by $C \xi^{2p_0-1-\epsilon}$. Let $\xi_i \to 0$ and
set $E_{k}^{+}$ to be $\cap_{i=1}^{\infty} E_{k, \xi}^{+}$. We know $E_{k}^{+}$ is a bounded closed set satisfying $\dim_{\mathcal{M}} E_{k}^{+}<2n-2$.
Now we define
\begin{align}
E_k \triangleq E_{k}^{+} \cup E_{k}^{-}. \label{eqn:MC17_3}
\end{align}
Then each $E_k$ is a closed bounded set satisfying $\dim_{\mathcal{M}} E_k<2n-2$.
According to their definitions and the above discussion, we know that there is a family of diffeomorphism $\varphi_{k,t}$, parametrized by $t \in [-2^{k}, -2^{-k}]$, from $\Omega_{D} \backslash E_k$ to $\Omega_{\sqrt{2^k|t|} D} \backslash E_k$, generated by the integral curve of $(-\nabla \hat{f}, \partial_t)$.
It is clear that (\ref{eqn:MC17_6}) follows from the integration of (\ref{eqn:MC16_4}).
The above argument is almost the same as that in the proof of Claim~\ref{clm:MA20_1} in Lemma~\ref{lma:HE04_1}.
In particular, the argument for the proof of equation (\ref{eqn:MC14_4}) is more or less repeated here.
We remind the readers that weak convexity of $\mathcal{R}$ is not used in the proof of equation (\ref{eqn:MC14_4}). Only the high codimension of $\mathcal{S}$ and the Killing vector properties are used.
Now we are ready to use the existence of the diffeomrophism(c.f. discussion around (\ref{eqn:MC17_6})) $\varphi_{k, -2^{k}}: \Omega_D \backslash E_k \to \Omega_{2^kD} \backslash E_k$ to relate the estimate of general $\Omega_a$ to (\ref{eqn:MC16_5}).
We are particularly interested in the sets $\Omega_a$ for small $a$'s. Without loss of generality, let $a=2^{-k}$. Fix some points $x,y \in \Omega_{2^{-k}} \backslash E_k$. Denote $\rho=d(x,y)$.
Similar to (\ref{eqn:MC07_1}) in the proof of Proposition~\ref{prn:SL04_1}, we choose a function
\begin{align}
\tilde{\chi} \triangleq \max \{\rho-d(\cdot, x), 0\}.
\label{eqn:MC12_3}
\end{align}
Note that $x,y \in \Omega_{2^{-k}} \subset \Omega_1 \subset B(\hat{y}, 0.25D)$ by (\ref{eqn:MC16_5}), which forces that $\rho=d(x,y)<0.25D$.
Also by (\ref{eqn:MC16_5}), we know that $\tilde{\chi}$ is supported in $B(x,\rho) \subset B(x, 0.25D) \subset B(\hat{y}, 0.5D) \subset \Omega_{D}$.
Let $\varphi$ be the diffeomorphism generated by integrating $(-\nabla \hat{f}, \partial_t)$ from $t=-2^{-k}$ to $t=-2^{k}$.
In other words, $\varphi=\varphi_{k, -2^{-k}}$. Using $\varphi$, we can push forward the function $\tilde{\chi}$ to obtain
\begin{align*}
\varphi_*(\tilde{\chi})(z) \triangleq \tilde{\chi} (\varphi^{-1} (z)), \quad \forall \; z \in \Omega_{2^kD} \backslash E_k.
\end{align*}
Clearly, $\varphi_*(\tilde{\chi})$ is supported on $\Omega_{2^k D} \backslash E_k$ with $\norm{\nabla \varphi_*(\tilde{\chi})}{L^{\infty}(\hat{Y})} \leq 2^{-k} \norm{\nabla\tilde{\chi}}{L^{\infty}(\hat{Y})}=2^{-k}$, in light of
(\ref{eqn:MC17_6}) and $t=-2^{k}$.
By the high codimension of $E_k$, we know that $\varphi_*(\tilde{\chi})$ is an $N_0^{1,2}$-function,
which has a good version such that $\sup_{\hat{Y}} |\nabla \varphi_*(\tilde{\chi})| \leq \norm{\nabla \varphi_*(\tilde{\chi})}{L^{\infty}(\hat{Y})}$, due to the high codimension of $\mathcal{S}$(c.f. Claim~\ref{clm:MC05_1}).
For simplicity of notation, we still denote the new version of $\tilde{\chi}$ by $\tilde{\chi}$. Note that the values of $\tilde{\chi}(x)$ and $\tilde{\chi}(y)$ are independent of the different versions, since $x,y$ are away from $E_k$.
Recall that $x, y \in \Omega_{2^{-k}}$. Integration of (\ref{eqn:MC16_3}) implies that
\begin{align*}
\hat{f}(x, -2^{k})=2^{-k} \hat{f}(x, -1)=4^{-k} \hat{f}(x, 2^{-k}) \leq 4^k \cdot 4^{-k} =1.
\end{align*}
Therefore, $\varphi(x) \in \Omega_1$. Similarly, we also know $\varphi(y) \in \Omega_1$.
Combining the previous inequalities and use (\ref{eqn:MC16_5}) again, we obtain that
\begin{align*}
0.5D \geq d(\varphi(x), \varphi(y)) \geq \frac{|\varphi_*(\tilde{\chi})( \varphi(x))-\varphi_*(\tilde{\chi})(\varphi(y))|}{\sup_{\hat{Y}} |\nabla \varphi_*(\tilde{\chi})|}
\geq \frac{|\varphi_*(\tilde{\chi})(\varphi(x))-\varphi_*(\tilde{\chi})(\varphi(y))|}{\norm{\nabla \varphi_*(\tilde{\chi})}{L^{\infty}(\hat{Y})}} \geq \frac{|\tilde{\chi}(x)-\tilde{\chi}(y)|}{2^{-k}}.
\end{align*}
Recall that $\tilde{\chi}(x)=\rho$ and $\tilde{\chi}(y)=0$ by (\ref{eqn:MC12_3}).
It follows from the above inequality that
\begin{align}
\rho=d(x,y) \leq 0.5D \cdot 2^{-k}=2^{-1-k} D, \label{eqn:MC12_4}
\end{align}
which is independent of the choice of $x,y \in \Omega_{2^{-k}} \backslash E_k$. Recall that $\Omega_{2^{-k}} \backslash E_k$ is dense in $\Omega_{2^{-k}}$. So we have
\begin{align*}
\diam \Omega_{2^{-k}}=\diam \{\Omega_{2^{-k}} \backslash E_k\} \leq 2^{-1-k}D.
\end{align*}
Consequently, $\displaystyle \lim_{k \to \infty} \diam(\Omega_{2^{-k}})=0$.
Since $\Omega_0=\bigcap_{1 \leq k <\infty} \overline{\Omega_{2^{-k}}}$, we know that $\Omega_0$ consists of only one point $\{\hat{y}\}$.
\textit{Step 4. The level sets of $\hat{f}$ coincide the geodesic balls centered at $\hat{y}$.}
Define
\begin{align}
\hat{r}(x) \triangleq \sqrt{4\hat{f}(x, -1)}=\sqrt{4\hat{f}(x)}, \quad d(x) \triangleq d(x, \hat{y}). \label{eqn:MC16_1}
\end{align}
Recall that in the standard Euclidean case, $\hat{f}=\frac{d^2}{4}$ and $\hat{r}=d$.
Our destination (\ref{eqn:MC17_7}) is equivalent to the equation $\hat{r}-d \equiv 0$.
Clearly, we have $ |\nabla \hat{r}|=\frac{|\nabla \hat{f}|}{\sqrt{\hat{f}}}=1$.
Recall that (c.f. (\ref{eqn:MC17_3})) each $E_k$ is a bounded closed set with $\dim_{\mathcal{M}} E_k<2n-2$. Let $E=\cup_{k=1}^{\infty} E_k$.
Then it is clear that $E$ is measure-zero and $\hat{Y} \backslash E$ is dense in $\hat{Y}$.
Note that $\hat{Y} \backslash E$ has a cone structure, as every point $x \in \hat{Y} \backslash E$ can be flowed to $\hat{y}$ along the integral curve of $\nabla \hat{f}=\frac{1}{2}\hat{r}\partial_{\hat{r}}$
without hitting singularities(c.f. Section 1 of~\cite{CCWarp}).
Let $x \in \mathcal{R}(\hat{Y})$ and $a=\hat{r}(x)>0$, we can find $x_k \in \mathcal{R}(\hat{Y}) \backslash E$ approaching $y$.
Every point $x_k$ can be flowed to a point nearby $\hat{y}$. So we obtain
\begin{align}
d(x)=d(x, \hat{y}) \leq \lim_{k \to \infty} d(x_k, \hat{y}) \leq\lim_{k \to \infty} \hat{r}(x_k) \leq \hat{r}(x). \label{eqn:MC16_7}
\end{align}
On the other hand, we can construct a function $\chi$ as
\begin{align*}
\chi(x) \triangleq \max\{a-\hat{r}(x), 0\},
\end{align*}
which is supported on a bounded set $\Omega_{0.5 a}$. Clearly, $\chi$ is Lipshitz. By the high codimeision of $\mathcal{S}$, by replacing $\chi$ with a new version if necessary, we can assume
$\sup_{\hat{Y}} |\nabla \chi| \leq \norm{\nabla \chi}{L^{\infty}(\hat{Y})} \leq 1$. Note the values at $x$ adn $y$ does not depend on the choice of versions since they are regular points.
Therefore, we have
\begin{align*}
d(x, y) \geq \frac{|\chi(x)-\chi(y)|}{\sup_{\hat{Y}} |\nabla \chi|} \geq = \frac{|\chi(x)-\chi(y)|}{\norm{\nabla \chi}{L^{\infty}(\hat{Y})}}=|\chi(x)-\chi(y)|=|\chi(y)|\geq a-|\hat{r}(y)|,
\end{align*}
for every $y \in \mathcal{R}(\hat{Y})$. Let $y$ approach $\hat{y}$ in $\mathcal{R}(\hat{Y}) \backslash E$, we obtain $d(x) \geq a=\hat{r}(x)$, which together with (\ref{eqn:MC16_7}) yields that
\begin{align}
d(x)=\hat{r}(x) \label{eqn:MC16_8}
\end{align}
for arbitrary $x \in \mathcal{R}(\hat{Y}) \backslash \{\hat{y}\}$. Since both $d$ and $\hat{r}$ are uniformly Lipshitz, the equation (\ref{eqn:MC16_8}) holds for every $y \in \hat{Y}$ by continuity and density reason.
In particular, the relationship (\ref{eqn:MC16_5}) can be improved to the following one:
\begin{align*}
\Omega_a =B(\hat{y}, 2a), \quad \forall \; a \geq 0.
\end{align*}
This confirms our expectation. Clearly, (\ref{eqn:MC17_7}) follows from the combination of (\ref{eqn:MC16_1}) and the extended version of (\ref{eqn:MC16_8}).
The proof of (\ref{eqn:MC17_7}) is complete.
From the discussion in Step 4 of the proof of (\ref{eqn:MC17_7}), we already know that $\hat{Y} \backslash \{E \cup \{\hat{y}\}\}$ has a local cone structure, which induces the global cone structure of $\hat{Y}$ by taking completion.
In view of (\ref{eqn:HE11_1}) in Proposition~\ref{thm:HE11_1}, we know $\mathcal{R}(\hat{Y})$ is path connected. Therefore, the cone $\hat{Y}$ is irreducible, i.e., $\hat{Y} \backslash \{\hat{y}\}$ is path connected.
Therefore, we obtain the global cone structure from the local cone structure, due to the high co-dimension of the singular set $\mathcal{S}$ and the Killing property arised from (\ref{eqn:SL13_1}), as we claimed.
\end{proof}
\subsubsection{Improved estimates in $\mathscr{K}(n,A;r_0)$}
In this subsection, we shall improve the limit space structure by the fact that every tangent space is a metric cone.
For simplicity, we assume $r_0=1$ if we do not mention otherwise.
\begin{proposition}[\textbf{Improvement of codimension estimate of $\mathcal{S}$}]
Suppose $\mathcal{LM}_i$ is a sequence of polarized K\"ahler Ricci flow solutions in $\mathscr{K}(n,A;1)$,
$x_i \in M_i$. Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i,x_i,g_i(0))$.
Let $\mathcal{S}$ be the singular part of $\bar{M}$. Then
\begin{align}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-2p_0, \quad
\dim_{\mathcal{H}}\mathcal{S} \leq 2n-4. \label{eqn:SL15_1}
\end{align}
\label{prn:SL15_1}
\end{proposition}
\begin{proof}
The Minkowski dimension estimate follows from Theorem~\ref{thm:HE11_1}.
Recall that we are in a situation where canonical radius is uniformly bounded from below.
Therefore, there is a gap between local behavior of singular point and regular point.
In particular, if one tangent space is Euclidean space, then the base point has a neighborhood with smooth manifold structure.
This follows from the volume convergence(c.f. Proposition~\ref{prn:SC12_1}) and the \textit{regularity estimate} in the definition of canonical radius(c.f. Definition~\ref{dfn:SC02_1}).
One can find the detailed argument in the proof of Proposition~\ref{prn:HA08_1}, where only polarized canonical radius lower bound is used.
Note that each iterated tangent space(away from vertex) is also a tangent space, and henceforce a tangent cone with more splitting directions.
Consequently, we can use induction to show that every tangent cone's singularity has an integer Hausdorff dimension(c.f.~\cite{CC1}).
However, the Minkowski dimension of singularity is
at most $2n-2p_0$. This forces that every tangent cone's singularity has Hausdorff dimension $2n-4$ at most,
which in turn implies $\dim_{\mathcal{H}} \mathcal{S}\leq 2n-4$.
\end{proof}
After we set up the tangent cone structure, we can improve Proposition~\ref{prn:SB27_1}.
\begin{proposition}[\textbf{Improvement of regular curve estimate}]
Same conditions as in Proposition~\ref{prn:SL15_1}.
For every two points $x,y \in \mathcal{R}$ and every small positive number $\epsilon>0$, there exists a rectifiable curve connecting $x,y$ such that
\begin{itemize}
\item $\gamma$ locates in $\mathcal{R}$.
\item $|\gamma| \leq (1+\epsilon)d(x,y)$.
\end{itemize}
\label{prn:SC17_4}
\end{proposition}
\begin{proof}
The proof is very similar to the proof of Proposition~\ref{prn:SB27_1}. The basic idea is to use the tangent cone structure, i.e., Theorem~\ref{thm:SC09_1} to improve Proposition~\ref{prn:SB27_1}.
\textit{First, every point in $\bar{M}$ has a cone-like neighborhood.}
To be more precise, fix $\epsilon>0$, for every point $z \in \bar{M}$, there is a radius $r_z$, depending on $z$ and $\epsilon$, with the following property:
\textit{For every point $v \in B(z, r_z)$, one can find a curve $\alpha$ such that}
\begin{itemize}
\item Initial point of $\alpha$ locates in $B(z, \epsilon d(v,z))$, end point of $\alpha$ locates in $B(v, \epsilon d(v,z))$.
\item $\alpha \subset \mathcal{R}$, \; $|\alpha|<(1+\epsilon) d(v,z)$.
\end{itemize}
The existence of $r_z$ can be obtained by application of Theorem~\ref{thm:SC09_1} and a contradiction blowup argument.
Actually, if for some $z$ such $r_z$ does not exist, we can find $v_i \to z$ such that corresponding $\alpha_i$ does not exist.
Blowup by $d^{-2}(v_i, z)$, we obtain a tangent cone $M_{\infty}$ with vertex $z_{\infty}$ and a point $v_{\infty}$ on the unit sphere of the cone.
By the density of regular part in the tangent cone, we have a regular point $\tilde{v}_{\infty} \in B(v_{\infty}, 0.5 \epsilon)$.
The cone structure guarantees that the shortest geodesic connecting $\tilde{v}_{\infty}$ to $z_{\infty}$, which we denote by $\overline{z_{\infty} \tilde{v}_{\infty}}$, has regular interior(c.f. (\ref{eqn:MC16_8})).
Denote the intersection of $\overline{z_{\infty} \tilde{v}_{\infty}}$ and $M_{\infty} \backslash B(z_{\infty}, 0.5 \epsilon)$ by $\alpha_{\infty}$.
Then $\alpha_{\infty}$ is a compact curve and locates in the regular part of $M_{\infty}$.
By the uniform convergence around $\alpha_{\infty}$, we obtain a curve $\alpha_i$ with the desired property before we arrive limit.
Contradiction.
\textit{Second, we can find a good covering of each shortest geodesic by cone-like neighborhoods.}
Fix any two points $x,y \in \mathcal{R}$.
Let $\beta$ be a shortest geodesic connecting $x,y$.
Since $\bigcup_{z \in \beta} B(z, \frac{1}{4} r_z)$ is a cover of a compact curve $\beta$, we can find a finite covering.
Starting from this finite covering, by deleting redundant extra balls from $x$ to $y$(e.g., using the ``greedy algorithm"), we obtain a covering $\cup_{i=1}^N B(z_i, \frac{1}{4}r_{z_i})$ with the following properties.
\begin{itemize}
\item $z_i$'s are ordered by their distance to $x$.
\item Each point on $\beta$ locates in at most two balls. If a point on $\beta$ is contained in two balls, then these two balls must be ``adjacent". In other words, if $z \in \beta \cap B(z_k, \frac{1}{4}r_k) \cap B(z_l, \frac{1}{4}r_l)$, then $|k-l|=1$.
\item Every pair of ``adjacent" balls have nonempty intersection, i.e., if $|k-l|=1$, then $B(z_k, \frac{1}{4}r_k) \cap B(z_l, \frac{1}{4}r_l) \neq \emptyset$.
\end{itemize}
\textit{Third, based on the good covering, one can construct approximation curve. }
Now we have a covering of $\beta$ by $\cup_{k=0}^{N} B(z_k, \frac{1}{4}r_k)$ with the property mentioned in the second step.
Without loss of generalirty, we further assume $z_0=x, z_N=y$. For each $0 \leq k \leq N-1$, let $\beta_k$ be the part of $\beta$ connecting $z_{k}$ and $z_{k+1}$,
let $d_k$ be the length of $\beta_k$. Then we have
\begin{align*}
d_k=d(z_k, z_{k+1}) < \frac{1}{4}r_k + \frac{1}{4}r_{k+1} \leq \frac{1}{2} \max \{ r_k, r_{k+1}\}.
\end{align*}
Hence either $z_{k+1}$ locates in the cone-like neighborhood of $z_k$, or $z_k$ locates in the cone-like neighborhood of $z_{k+1}$.
No matter what case happens, we can find an approximation curve $\alpha_k \subset \mathcal{R}$, whose two ends locate in the $\epsilon d_k$ neighborhood of $z_{k}$ and $z_{k+1}$, satisfying
$|\alpha_k|<(1+\epsilon) d_k$. According to this choice, the end point of $\alpha_{k-1}$ and the initial point of $\alpha_k$ have distance bounded by $\epsilon (d_k+d_{k-1})$, whenever $1 \leq k \leq N-1$.
So they can be connected by a curve $\gamma_k \subset \mathcal{R}$ with $|\gamma_k| \leq 3\epsilon (d_k+d_{k-1})$, due to Proposition~\ref{prn:SB27_1}.
For the boundary case, it is not hard to see that $z_0=x$ can be connected to the initial point of $\alpha_0$ by $\gamma_0 \subset \mathcal{R}$ and $|\gamma_0|<3\epsilon d_0$.
Similarly, $z_N=y$ can be connected to the end point of $\alpha_{N-1}$ by $\gamma_N \subset \mathcal{R}$ and $|\gamma_N|<3\epsilon d_{N-1}$.
Concatenating all the curves $\alpha_k$ and $\gamma_k$, we obtain a curve $\gamma \subset \mathcal{R}$ connecting $x,y$ and satisfying(c.f. Figure~\ref{figure:gcovering} for the case $N=2$)
\begin{align*}
|\gamma| &=\sum_{k=0}^{N-1} |\alpha_k| + \sum_{k=0}^{N} |\gamma_k| \leq \left( \sum_{k=0}^{N-1} (1+\epsilon) d_k \right) +\left( 3\epsilon d_0 +\sum_{k=0}^{N-2} 3\epsilon (d_k+d_{k+1}) + 3\epsilon d_{N-1} \right) \\
&=(1+\epsilon) \sum_{k=0}^{N-1} d_k + 6 \epsilon \sum_{k=0}^{N-1} d_k =(1+7\epsilon) |\beta|=(1+7 \epsilon) d(x,y).
\end{align*}
Replacing $\epsilon$ by $0.1 \epsilon$ at the beginning, we then find a curve $\gamma$ satisfying the requirement.
\end{proof}
\begin{figure}
\begin{center}
\psfrag{B0}[c][c]{$B(z_0,\frac14 r_0)$}
\psfrag{B1}[c][c]{$B(z_1,\frac14 r_1)$}
\psfrag{B2}[c][c]{$B(z_2,\frac14 r_2)$}
\psfrag{z0}[c][c]{$z_0=x$}
\psfrag{z1}[c][c]{$z_1$}
\psfrag{z2}[c][c]{$z_2=y$}
\psfrag{g0}[c][c]{$\color{blue}{\gamma_0}$}
\psfrag{g1}[c][c]{$\color{blue}{\gamma_1}$}
\psfrag{g2}[c][c]{$\color{blue}{\gamma_2}$}
\psfrag{a0}[c][c]{$\color{red}{\alpha_0}$}
\psfrag{a1}[c][c]{$\color{red}{\alpha_1}$}
\includegraphics[width=0.5 \columnwidth]{gcovering}
\caption{Construction of approximation curve $\gamma$}
\label{figure:gcovering}
\end{center}
\end{figure}
\begin{lemma}[\textbf{Rough estimate of reduced distance}]
There is an $\epsilon=\epsilon(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A;1)$,
$x,y\in M$ and $r=d_0(x,y)<1$. Suppose $y \in \mathcal{F}_{\frac{\epsilon_b}{2} r}(M,0)$. Then we have
\begin{align}
l((x,0),(y,-r^2))< 100 \label{eqn:SL25_3}
\end{align}
whenever $\sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
\label{lma:SL13_1}
\end{lemma}
\begin{proof}
Let $y_0=y$.
According to the construction in Proposition~\ref{prn:SL24_1}, there exists a point
$y_1 \in \partial B_{g(0)}(x,\frac{r}{2}) \cap \mathcal{F}_{\frac{\epsilon_b r}{4}}(M,0)$ and a curve
$\gamma_1 \subset \mathcal{F}_{\frac{\epsilon_b^2}{8} r}(M,0)$ connecting $y_0,y_1$, with length less than
$\frac{9}{2}r$.
Suppose $|R|+|\lambda|$ is small enough, then
$\displaystyle \gamma_1 \subset \bigcap_{-r^2 \leq t \leq 0} \mathcal{F}_{\frac{\epsilon_b^2 r}{16}}(M,t)$.
So $\gamma_1$ can be lifted as a space-time curve connecting $(y_1, -\frac{r^2}{4})$
and $(y_0, -r^2)$. Reparameterizing $\gamma_1$ by $\tau$, after a proper adjustment, we have
\begin{align*}
\int_{\frac{r^2}{4}}^{r^2} \sqrt{\tau} |\dot{\gamma}_1|_{g(-\tau)}^2 d\tau < 100 r.
\end{align*}
Following the same procedure, we can find $\gamma_2$ connecting $y_1$ to
$y_2 \in \partial B_{g(0)}(x,\frac{r}{4}) \cap \mathcal{F}_{\frac{\epsilon_b r}{8}}(M,0)$ with
$\displaystyle \gamma_2 \subset \bigcap_{-\frac{r^2}{4} \leq t \leq 0} \mathcal{F}_{\frac{\epsilon_b^2 r}{32}}(M,t)$.
By a proper reparameterization of $\tau$, we can regard $\gamma_2$ as a space-time curve connecting
$(y_1,-\frac{r^2}{4})$ and $(y_2, -\frac{r^2}{16})$, and it satisfies the estimate
\begin{align*}
\int_{\frac{r^2}{16}}^{\frac{r^2}{4}} \sqrt{\tau} |\dot{\gamma}_2|_{g(-\tau)}^2 d\tau < 100 \cdot \frac{r}{2}.
\end{align*}
Note that there is no need to choose a new $\epsilon$ because of the rescaling property of $|R|+|{\lambda}|$.
Repeating this process, we can find curve $\gamma_k$ connecting $(y_k, -\frac{r^2}{4^k})$
and $(y_{k+1}, -\frac{r^2}{4^{k+1}})$. Concatenating all $\gamma_k$'s together, we obtain a space-time curve
$\gamma$ connecting $(x,0)$ and $(y, -r^2)$ such that
\begin{align*}
\int_0^{r^2} \sqrt{\tau} |\dot{\gamma}|_{g(-\tau)}^2 d\tau < 100 \sum_{k=0}^{\infty} \frac{r}{2^k}=200r.
\end{align*}
It follows that
\begin{align*}
l((x,0), (y,-r^2)) < \frac{200r}{2\sqrt{r^2}}=100.
\end{align*}
\end{proof}
\begin{lemma}[\textbf{Most shortest reduced geodesics avoid high curvature part}]
For every group of numbers $0<\xi<\eta<1<H$, there is a big constant $C=C(n,A,\eta,H)$
and a small constant $\epsilon=\epsilon(n,A,H,\eta,\xi)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A;1)$, $x\in \mathcal{F}_{\eta}(M,0)$.
Let $\Omega_{\xi}$ be the collection of points $z \in M$ such that there exists a
shortest reduced geodesic $\boldsymbol{\beta}$ connecting $(x,0)$ and $(z,-1)$ satisfying
\begin{align}
\beta \cap \mathcal{D}_{\xi}(M,0) \neq \emptyset.
\label{eqn:SL24_4}
\end{align}
Then
\begin{align}
|B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap \Omega_{\xi}| < C \xi^{2p_0-1}
\label{eqn:SL24_5}
\end{align}
whenever $\sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
\label{lma:SL14_1}
\end{lemma}
\begin{proof}
This is a flow property, we assume $\lambda=0$ without loss of generality.
From the argument in Lemma~\ref{lma:SL13_1}, it is not hard to obtain the following bound
\begin{align}
l((x,0),(z,-1)) < C, \quad \forall \; z \in B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0), \label{eqn:SL24_6}
\end{align}
where $C=C(\eta,H)$.
Suppose $z \in B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0)$, $\boldsymbol{\beta}$ is a shortest reduced geodesic connecting $(x,0)$ and $(z,-1)$. Let $\beta$ be the corresponding space curve.
Note that $x$ and $z$, the two end points of $\beta$, locate outside of $ \mathcal{D}_{\xi}(M,0)$.
Therefore if $z \in \Omega_{\xi}$, then (\ref{eqn:SL24_4}) is satisfied. In other words, the shortest reduced geodesic connecting $(x, 0)$ and $(z, -1)$ cannot avoid the ``high curvature" part $\mathcal{D}_{\xi}(M,0)$.
By continuity(c.f. Appendix~\ref{app:B}), we have
\begin{align*}
\beta \cap \partial \mathcal{F}_{\xi}(M,0) =\beta \cap \partial \mathcal{D}_{\xi}(M,0) \neq \emptyset.
\end{align*}
Let $\tau_a$ be the first time $\boldsymbol{\beta}$ escape from $\mathcal{F}_{K^{-1}\eta}$, $\tau_b$ be the last time such that $\boldsymbol{\beta}(\tau)$ re-enter $\mathcal{F}_{K^{-1}\eta}$, when we move along backward time direction.
Here $K$ is the constant defined in Proposition~\ref{prn:SC24_1}. To be more precise, we define
\begin{align*}
&\tau_a \triangleq \sup \left\{ \tau \left| \beta(s) \in \mathcal{F}_{K^{-1}\eta}, \quad \forall \; s \in (0,\tau) \right. \right\}, \\
&\tau_b \triangleq \inf \left\{ \tau \left| \beta(s) \in \mathcal{F}_{K^{-1}\eta}, \quad \forall \; s \in (\tau, 1) \right. \right\}.
\end{align*}
By the choice of $x$ and $z$, it is clear that $0<\tau_a<\tau_b<1$. We can further estimate $\tau_a$ and $\tau_b$ uniformly.
Actually, since $l$ is achieved by $\boldsymbol{\beta}$ and is bounded by $C$, it follows from the definition of $l$(c.f. equation (\ref{eqn:MA22_1}) and (\ref{eqn:MA22_2})) that
\begin{align}
\int_0^1 \sqrt{\tau} \left( R + |\dot{\beta}|^2\right)_{g(-\tau)} d\tau<C.
\label{eqn:MA22_5}
\end{align}
Note that $\beta(\tau) \in \mathcal{F}_{\eta}(M, 0)$ whenever $\tau \in (0, \tau_a) \cup (\tau_b, 1)$. In view of Proposition~\ref{prn:SL04_1},
we have the metric equivalence $0.5 g(x, 0)<g(x, -\tau)<2g(x, 0)$
for all $x \in \mathcal{F}_{K^{-1}\eta}(M, 0)$ and $\tau \in (0,1)$. Recalling that $|R|$ is uniformly small. Then (\ref{eqn:MA22_5}) implies that
\begin{align*}
\int_0^{\tau_a} \sqrt{\tau} |\dot{\beta}|_{g(0)}^2 d\tau <C, \quad \int_{\tau_b}^1 \sqrt{\tau} |\dot{\beta}|_{g(0)}^2 d\tau <C.
\end{align*}
It follows from Proposition~\ref{prn:SC24_1} and the above inequality that
\begin{align}
&\frac{\eta}{C}<d_{g(0)}(x, \beta(\tau_a)) \leq \int_0^{\tau_a} |\dot{\beta}|_{g(0)} d\tau< \left( \int_0^{\tau_a} \sqrt{\tau} |\dot{\beta}|_{g(0)}^2 d\tau \right)^{\frac{1}{2}} \left( \int_0^{\tau_a} \frac{1}{\sqrt{\tau}} d\tau \right)^{\frac{1}{2}}<C\tau_a^{\frac{1}{4}}, \label{eqn:MA22_7}\\
&\frac{\eta}{C}<d_{g(0)}(\beta(\tau_b), z) \leq \int_{\tau_b}^{1} |\dot{\beta}|_{g(0)} d\tau< \left( \int_{\tau_b}^{1} \sqrt{\tau} |\dot{\beta}|_{g(0)}^2 d\tau \right)^{\frac{1}{2}} \left( \int_{\tau_b}^{1} \frac{1}{\sqrt{\tau}} d\tau \right)^{\frac{1}{2}}
<C\sqrt{1-\sqrt{\tau_b}},
\label{eqn:MA22_8}
\end{align}
where $C=C(\eta,H,K)$. Consequently, we have
\begin{align*}
\tau_a>\frac{\eta^4}{C}, \quad 1-\tau_b=\left(1-\sqrt{\tau_b} \right) \left(1+\sqrt{\tau_b} \right) \geq 1-\sqrt{\tau_b} \geq \frac{\eta^2}{C}.
\end{align*}
This means that $[\tau_a, \tau_b] \subset \left[\frac{\eta^4}{C}, 1-\frac{\eta^2}{C} \right]$.
Define $\bar{\tau}$ as
\begin{align}
\bar{\tau} \triangleq \max\{ \tau| \beta(\tau) \in \mathcal{D}_{\xi}(M, 0)\}. \label{eqn:MC23_1}
\end{align}
Clearly, we have $\beta(\bar{\tau}) \in \partial \mathcal{D}_{\xi}(M, 0)=\partial \mathcal{F}_{\xi}(M, 0)$.
Since $\xi<K^{-1}\eta$, we have $\bar{\tau} \in [\tau_a, \tau_b]$ for continuity reason. Consequently, we know
\begin{align}
\frac{\eta^4}{C}< \bar{\tau} < 1-\frac{\eta^2}{C},
\label{eqn:MA22_6}
\end{align}
for some $C=C(\eta, H, K)$, whenever $\xi<K^{-1} \eta$ and $|R|$ very small.
Beyond the estimate of $\bar{\tau}$, there are more estimates around $\boldsymbol{\beta}(\bar{\tau})$.
In light of the choice of $\bar{\tau}$, we have $\beta(\tau) \in \mathcal{F}_{\xi}(M,0)$ for each $\tau \in [\bar{\tau}, 1]$.
By (\ref{eqn:SL24_6}), we have uniform rough bound of the reduced distance from $(x, 0)$ to $(z, -1)$.
Noting that $R$ may be negative and $|R|$ is very small, we have
\begin{align*}
\int_{\bar{\tau}}^{1} \sqrt{\tau} \left( R + |\dot{\beta}|^2\right)_{g(-\tau)} d\tau <1+ \int_0^1 \sqrt{\tau} \left( R + |\dot{\beta}|^2\right)_{g(-\tau)} d\tau<C(\eta,H).
\end{align*}
Following the route of (\ref{eqn:MA22_7}), noting that metrics $g(0)$, $g(-\tau)$ and $g(-1)$ are all uniformly equivalent on $\beta(\tau)$ whenever $\tau \in [\bar{\tau}, 1]$, we have
\begin{align*}
d_{g(0)}(z, \beta(\bar{\tau})) \leq \int_{\bar{\tau}}^{1} |\dot{\beta}|_{g(0)} d\tau \leq 2\int_{\bar{\tau}}^{1} |\dot{\beta}|_{g(-1)} d\tau < 2\left( \int_{\bar{\tau}}^{1} \sqrt{\tau} |\dot{\beta}|_{g(-1)}^2 d\tau \right)^{\frac{1}{2}} \left( \int_{\bar{\tau}}^{1} \frac{1}{\sqrt{\tau}} d\tau \right)^{\frac{1}{2}}
< C.
\end{align*}
Note that $d_{g(0)}(z,x)<H$. Triangle inequality then implies that
\begin{align}
d_{g(0)}(\beta(\bar{\tau}), x)<F
\label{eqn:MA22_9}
\end{align}
for some $F$ independent of $\xi$ when $|R|+|\lambda|$ small enough.
The purpose of this paragraph is to estimate $\dot{\beta}(\bar{\tau})$.
Recalling the reduced geodesic equation (\ref{eqn:MA22_3}):
\begin{align*}
\nabla_V V +\frac{V}{2\tau} + 2Ric(V, \cdot) + \frac{\nabla R}{2}=0,
\end{align*}
where $V=\dot{\beta}$. It follows that along the reduced geodesic $\boldsymbol{\beta}$, we have
\begin{align*}
& \frac{d}{d\tau} |\dot{\beta}|^2=2\langle \nabla_{\dot{\beta}} \dot{\beta}, \dot{\beta} \rangle +2Ric(\dot{\beta}, \dot{\beta})=-\frac{|\dot{\beta}|^2}{\tau} -2Ric(\dot{\beta}, \dot{\beta}) - \langle \nabla R, \dot{\beta} \rangle,\\
& \frac{d}{d\tau} \left\{\tau |\dot{\beta}|^2 \right\}=-\tau \left\{ 2Ric(\dot{\beta}, \dot{\beta}) +\langle \nabla R, \dot{\beta} \rangle \right\}.
\end{align*}
Note that $\beta(\tau) \in \mathcal{F}_{\xi}(M,0)$ for each $\tau \in [\bar{\tau}, 1]$. By Proposition~\ref{prn:SC09_3}, we can assume $\beta(\tau) \in \mathcal{F}_{K^{-1}\xi}(M,-\tau)$.
It follows that $|Ric|$ and $|\nabla R|$ uniformly small whenever $|R|$ globally very small. Therefore, the above equation implies
\begin{align}
\left| \frac{d}{d\tau} \left( \tau |\dot{\beta}|^2 +1\right) \right| <\theta \left( \tau |\dot{\beta}|^2 + 1\right) , \quad \forall \; \tau \in (\bar{\tau}, 1)
\label{eqn:MA22_10}
\end{align}
for some small constant $\theta$ depending on $\xi$ and $\sup_{\mathcal{M}} |R|$. Moreover, $\theta \to 0$ if $\sup_{\mathcal{M}} |R| \to 0$ and $\xi$ is fixed.
Integrating (\ref{eqn:MA22_10}) and using (\ref{eqn:MA22_6}), we obtain
\begin{align*}
\tau |\dot{\beta}|^2+1> e^{-\theta} \left( \bar{\tau} |\dot{\beta}|_{g(-\bar{\tau})}^2 +1\right).
\end{align*}
It follows that
\begin{align*}
\int_{\bar{\tau}}^{1} \sqrt{\tau}|\dot{\beta}|_{g(-\tau)}^2 d\tau &= \int_{\bar{\tau}}^{1} \frac{1}{\sqrt{\tau}} \tau |\dot{\beta}|_{g(-\tau)}^2 d\tau
\geq \int_{\bar{\tau}}^{1} \left\{ e^{-\theta} \left( \bar{\tau} |\dot{\beta}|_{g(-\bar{\tau})}^2 +1\right)-1 \right\} d\tau\\
&=e^{-\theta} |\dot{\beta}|_{g(-\bar{\tau})}^2 \bar{\tau} (1-\bar{\tau}) + (e^{-\theta}-1)(1-\bar{\tau}).
\end{align*}
In view of (\ref{eqn:MA22_5}) and the fact that $|R|$ is very small, we know the left hand side of the above inequality is bounded above by $C=C(\eta, H)$. Since $\theta$ is very small,
$\bar{\tau} \in \left[ \frac{\eta^4}{C}, 1-\frac{\eta^2}{C} \right]$ by (\ref{eqn:MA22_6}), the above inequality yields that
\begin{align}
\left|\dot{\beta}(\bar{\tau}) \right|_{g(-\bar{\tau})} < C,
\label{eqn:MA22_11}
\end{align}
where $C=C(\eta, H, K)$ is independent of $\xi$. Note that $\boldsymbol{\beta}(\tau)=(\beta(\tau), -\tau)$, the space-time tangent vector of $\boldsymbol{\beta}$ is $(\dot{\beta}, -1)$.
Intuitively, (\ref{eqn:MA22_11}) can be understood that the ``angle" between the space-time tangent and the space tangent form a positive ``angle" which is uniformly bounded below.
Note that the reduced volume element $(4\pi \tau)^{-n}e^{-l} dv$ is decreasing along $\boldsymbol{\beta}$.
Up to a perturbation, $\partial \mathcal{F}_{\xi}(M,0)$ can be regarded(c.f. Corollary~\ref{cor:GI26_1}) as a smooth hypersurface in $M$ satisfying
\begin{align}
\left| \partial \mathcal{F}_{\xi}(M,0) \cap B_{g(0)}(x,F) \right|_{\mathcal{H}^{2n-1}} \leq C\xi^{2p_0-1},
\label{eqn:MA21_1}
\end{align}
for some $C=C(n,A,F)$, $F=F(\eta, H)$ is the constant in (\ref{eqn:MA22_9}).
Consequently, $\partial \mathcal{F}_{\xi}(M,0) \times [-1, 0]$ can be regarded as a hypersurface in the space-time.
Recall that $\Omega_{\xi}$ is the collection of points $z \in M$ such that there exists a
shortest reduced geodesic $\boldsymbol{\beta}$ connecting $(x,0)$ and $(z,-1)$ satisfying (\ref{eqn:SL24_4}).
By reduced geodesic theory(c.f. Section 7 of~\cite{Pe1} and the corresponding sections in~\cite{KL} for more details),
the following results are known.
\begin{itemize}
\item[(a).] For every $z \in M$, $(z, -1)$ can be connected to $(x, 0)$ by a shortest reduced geodesic.
\item[(b).] For every $z \in M \backslash E$, $(z, -1)$ can be connected to $(x, 0)$ by a unique shortest reduced geodesic, where $E$ is a measure-zero set and is called the $\mathcal{L}$-cut-locus.
\end{itemize}
Therefore, we can define a projection map $\varphi$ as follows.
\begin{align}
\varphi: B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap \left\{ \Omega_{\xi} \backslash E \right\} &\mapsto \partial \mathcal{F}_{\xi}(M,0) \times [-1, 0], \notag\\
z &\mapsto \boldsymbol{\beta}(\bar{\tau}).
\label{eqn:MC23_2}
\end{align}
For simplicity, let $\Omega=B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap \left\{ \Omega_{\xi} \backslash E \right\}$.
The reduced distance bound (\ref{eqn:SL24_6}) and the entering-time bound (\ref{eqn:MA22_6}) implies that the reduced volume element $(4\pi \tau)^{-n}e^{-l} dv$ along $\boldsymbol{\beta}$ is uniformly equivalent to $dv$,
whenever $\tau \in [\bar{\tau}, 1]$. Since $(4\pi \tau)^{-n}e^{-l} dv$ is monotone along $\boldsymbol{\beta}$, we can regard $dv$ as almost monotone, up to multiplying a uniform constant $C$.
Therefore, we have
\begin{align*}
|\Omega|_{\mathcal{H}^{2n}}=\int_{\Omega}1 dv \leq C \int_{\Omega} e^{-l(z)} dv_z \leq \int_{\varphi(\Omega)} \bar{\tau}^{-n} e^{-l(y)} dv_y \leq C \int_{\varphi(\Omega)} dv_y,
\end{align*}
where $y=\varphi(x)$. Note that inequality (\ref{eqn:MA22_11}) can be regarded as an ``angle" bound, since $\dot{\boldsymbol{\beta}}=(\dot{\beta}, -1)$.
The uniform bound of $|\dot{\beta}|$ guarantees that $dv_y \leq C |d\sigma_y \wedge dt|$ where $d \sigma_y$ is the ``area" element of $\partial \mathcal{F}_{\xi}$.
Then we have
\begin{align*}
|\Omega|_{\mathcal{H}^{2n}}&\leq C \int_{\varphi(\Omega)} |d\sigma_y \wedge dt| \leq C\int_{\left\{\partial \mathcal{F}_{\xi}(M,0) \cap B_{g(0)}(x, F) \right\} \times [-1,0]} |d\sigma_y \wedge dt|\\
&=C \left| \left\{\partial \mathcal{F}_{\xi}(M,0) \cap B_{g(0)}(x, F) \right\} \times [-1, 0] \right|_{\mathcal{H}^{2n}}\\
&\leq C \left| \partial \mathcal{F}_{\xi}(M,0) \cap B_{g(0)}(x,F) \right|_{\mathcal{H}^{2n-1}},
\end{align*}
where we used the almost product structure of $\partial \mathcal{F}_{\xi} \times [-1, 0]$ in the last step.
Note that $C=C(\eta, H, K)=C(n,A,\eta,H)$ since $K$ is determined by $n,A$(c.f. Proposition~\ref{prn:SC24_1}).
Recall that $\Omega=B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap \left\{ \Omega_{\xi} \backslash E \right\}$.
Plugging (\ref{eqn:MA21_1}) into the above inequality, we obtain (\ref{eqn:SL24_5}).
\end{proof}
Note that in Lemma~\ref{lma:SL14_1}, for every point $z \in \left\{ B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \right\} \backslash \{\Omega_{\xi} \cup E\}$, there is
a unique shortest reduced geodesic connecting $(z,-1)$ to $(x, 0)$ and avoiding $\mathcal{D}_{\xi}(M,0)$.
If $z \in \left\{ B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap E \right\} \backslash \Omega_{\xi}$, then every shortest reduced geodesic connecting
$(z, -1)$ to $(x, 0)$ avoids $\mathcal{D}_{\xi}(M,0)$. However, we may not have uniqueness.
Now we pass Lemma~\ref{lma:SL14_1} to Cheeger-Gromov limit and have the following property.
\begin{lemma}[\textbf{Rough weak convexity by reduced geodesics}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;1)$ satisfies
\begin{align}
\lim_{i \to \infty} \left( \frac{1}{T_i} + \frac{1}{\Vol(M_i)} + \sup_{\mathcal{M}_i} (|R|+|\lambda|) \right)=0. \label{eqn:SL06_1}
\end{align}
Suppose $x_i \in M_i$.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i, x_i, g_i(0))$, $\mathcal{R}$ be the regular part of $\bar{M}$ and $\bar{x} \in \mathcal{R}$.
Suppose $\bar{t}<0$ is a fixed number. Then every $(\bar{z}, \bar{t})$ can be connected to $(\bar{x}, 0)$ by a smooth reduced geodesic, whenever $\bar{z}$ is away from a closed measure-zero set.
\label{lma:SK27_4}
\end{lemma}
\begin{proof}
Without loss of generality, let $\bar{t}=-1$ and $\lambda=0$. We use $\bar{g}(0)$ as the default metric on the limit space.
Since $\bar{x} \in \mathcal{R}$, it locates in $\mathcal{R}_{\eta_0}$ for some $\eta_0 \in (0,1)$, where we used the notation defined in equation (\ref{eqn:HA11_3}).
Fix $\eta \in (0, \eta_0)$. Let $E_{\eta,\xi}$ be the closure of the limit set of $B_{g_i(0)}\left(x_i, \eta^{-1} \right) \cap \mathcal{F}_{\eta}(M_i, 0) \cap \Omega_{\xi}(M_i)$, which
we denote by $E_{\eta, \xi}'(M_i)$ for simplicity. Suppose $\bar{z} \in E_{\eta,\xi}$ is the limit of some sequence $z_i \in E_{\eta,\xi}'(M_i)$.
Then it is easy to see that $\bar{z} \in \overline{B(\bar{x}, \eta^{-1})} \cap \mathcal{R}_{\eta}(\bar{M})$.
For each $i$, there is a shortest reduced geodesic $\boldsymbol{\beta}_i$ connecting $(x_i, 0)$ to $(z_i, -1)$ and passing through $\mathcal{D}_{\xi}(M_i, 0)$.
Let $\boldsymbol{\beta}$ be the limit of $\boldsymbol{\beta}_i$.
Note that $\boldsymbol{\beta}$ may pass through singularity. The largest $\tau$ such that $\boldsymbol{\beta}(\tau)$ comes out of $\mathcal{S}_{\xi}$(c.f. equation (\ref{eqn:HA11_4}) for notations)
is denoted by $\bar{\tau}$(c.f. equation (\ref{eqn:MC23_1})).
By (\ref{eqn:MA22_6}), i.e., $ \frac{\eta^4}{C}< \bar{\tau} < 1-\frac{\eta^2}{C}$, we know $\bar{\tau}$ is uniformly bounded away from $0$ and $1$. Moreover, $d(\bar{x}, \beta(\bar{\tau}))$ is uniformly
bounded by some constant $F$(c.f. inequality (\ref{eqn:MA22_9})), the value $|\dot{\beta}(\bar{\tau})|$ is uniformly bounded by inequality (\ref{eqn:MA22_11}).
By taking limit on $\bar{M}$, we see that for every point $\bar{z}$(no matter whether it is a limit of points in $E_{\eta,\xi}'(M_i)$), we can find a shortest reduced geodesic $\boldsymbol{\beta}$ connecting $(\bar{x}, 0)$ and $(\bar{z}, -1)$, with $\bar{\tau}$ satisfying (\ref{eqn:MA22_6})
and $\beta(\bar{\tau})$ locating in $B(\bar{x}, F)$ for some uniform constant $F$, and $|\dot{\beta}(\bar{\tau})|$ uniformly bounded by $C$. Note that both $C$ and $F$ are independent of $\xi$.
As a closure, $E_{\eta, \xi}$ is clearly a closed set. Note that $E_{\eta, \xi} \subset \overline{B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}} \subset B(\bar{x}, 2\eta^{-1}) \cap \overset{\circ}{\mathcal{R}}_{0.5\eta}$, which is an open smooth manifold.
Therefore, $E_{\eta, \xi}$ is measurable.
Suppose $\bar{z}_a, \bar{z}_b$ are two points in $\bar{E}_{\eta, \xi}$. Tracing their origin and use the shortest property, it is clear that $\boldsymbol{\beta}_a$ and $\boldsymbol{\beta}_b$ have no intersection
except $(\bar{x}, 0)$, where $\boldsymbol{\beta}_a$ is a shortest reduced geodesic connecting $(\bar{x}, 0)$ to $(\bar{z}_a, -1)$, $\boldsymbol{\beta}_b$ is a shortest reduced geodesic connecting $(\bar{x}, 0)$ to $(\bar{z}_b, -1)$.
Similar to (\ref{eqn:MC23_2}) in the proof of Lemma~\ref{lma:SL14_1}, we now define a mutli-value projection map $\tilde{\varphi}$ from $E_{\eta,\xi}$ to $\partial \mathcal{R}_{\xi}$ as follows:
\begin{align*}
\tilde{\varphi}: E_{\eta,\xi} &\mapsto \partial \mathcal{R}_{\xi} \times [-1, 0], \\
z &\mapsto \{\boldsymbol{\beta}(z), \; \boldsymbol{\beta} \; \textrm{is a shortest reduced geodesic connecting} \; (z, -1) \; \textrm{to} \; (\bar{x}, 0) \; \textrm{with} \; \beta \cap \mathcal{S}_{\xi} \neq \emptyset \}.
\end{align*}
Following the argument at the end of the proof of Lemma~\ref{lma:SL14_1}, we have
\begin{align*}
|E_{\eta,\xi}|_{\mathcal{H}^{2n}} = \int_{E_{\eta,\xi}} 1 dv \leq C \int_{E_{\eta,\xi}} e^{-l(z)} dv_{z} \leq \int_{\tilde{\varphi}(E_{\eta,\xi})} \bar{\tau}^{-n} e^{-l(y)} dv_y,
\end{align*}
where $(y, -\bar{\tau})=\boldsymbol{\beta}(\bar{\tau})$ for some $\boldsymbol{\beta}$ connecting $(\bar{x}, 0)$ to $(z, -1)$ satisfying $\beta \cap \mathcal{S}_{\xi} \neq \emptyset$.
Note that the last inequality holds even if $\tilde{\varphi}$ is multi-valued.
Starting from the above step, the remainder argument exactly follows from the proof of (\ref{eqn:SL24_5}). Consequently, we have
\begin{align}
|E_{\eta,\xi}| \leq C\xi^{2p_0-1} \label{eqn:MA23_2}
\end{align}
for some $C$ independent of $\xi$.
Note that $E_{\eta_1,\xi_1} \subset E_{\eta_2, \xi_2}$ whenever $0<\xi_1<\xi_2$, $0<\eta_2 \leq \eta_1$.
Then we define
\begin{align}
E_{\eta} \triangleq \bigcap_{\xi \in (0, \eta)} E_{\eta, \xi}.
\label{eqn:MA23_1}
\end{align}
In light of (\ref{eqn:MA23_2}), we see that $E_{\eta}$ is a closed subset of $\overline{B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}(\bar{M})}$ with measure zero.
Suppose
\begin{align*}
\bar{z} \in \left\{ B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}(\bar{M}) \right\} \backslash E_{\eta}=\bigcup_{\xi \in (0, \eta)} \left\{ \left\{B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}(\bar{M}) \right\} \backslash E_{\eta, \xi} \right\},
\end{align*}
then $\bar{z} \in B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}(\bar{M}) \backslash E_{\eta, \xi}$ for some $\xi \in (0,\eta)$.
By the smooth flow convergence on $\mathcal{F}_{\xi}(M_i, 0) \times [-1, 0]$(c.f. Proposition~\ref{prn:SL04_1}) and the definition of $E_{\eta, \xi}$, we obtain that
$(\bar{z}, -1)$ can be connected to $(\bar{x}, 0)$ by some shortest smooth reduced geodesic contained in $\mathcal{R}_{\xi}(\bar{M}) \times [-1,0]$.
Moreover, every smooth shortest reduced geodesic connecting $(\bar{x}, 0)$ and $(\bar{z}, -1)$ are uniformly $\xi$-regular.
To be more precise, every point $\bar{z} \in \left\{ B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}(\bar{M}) \right\} \backslash E_{\eta}$ satisfies the following property:
\textit{$(\bar{z}, -1)$ can be connected to $(\bar{x}, 0)$ by a shortest smooth reduced geodesic $\boldsymbol{\beta}$. In other words, for every other smooth reduced geodesic $\boldsymbol{\gamma}$ with the same ends, we have
$\mathcal{L}(\boldsymbol{\gamma}) \geq \mathcal{L}(\boldsymbol{\beta})$.}
Now we define
\begin{align}
E \triangleq \bigcup_{k\in \{1, 2, \cdots \}} E_{2^{-k}\eta_0} \backslash \left\{ B(\bar{x}, 2^{k-2} \eta_0^{-1}) \cap \overset{\circ}{\mathcal{R}}_{2^{-k+2} \eta_0}\right\}.
\label{eqn:MC25_1}
\end{align}
The $\eta_0$ above is some fixed positive number.
According to this definition, every regular point locates in finitely many closed sets $E_{2^{-k}\eta_0} \backslash \left\{ B(\bar{x}, 2^{k-2} \eta_0^{-1}) \cap \overset{\circ}{\mathcal{R}}_{2^{k-2} \eta_0}\right\}$.
The reason we choose to define $E$ in the way of (\ref{eqn:MC25_1}) is to obtain the closedness of $E \cup \mathcal{S}$. Note that if we simply define $E$ to be the union of all $E_{2^{-k}\eta_0}$, then $E \cup \mathcal{S}$ may not be closed set.
It is possible to obtain points in $E_{2^{-k} \eta_0}$ converging to a regular point. However, from the discussion in the above paragraph, it is clear that for every regular point, one can find a small closed ball regular neighborhood $\bar{B}$ where
every point (with time $t=-1$) can be connected to $(\bar{x}, 0)$ away from a closed set $E_{\bar{B}}=E_{\eta} \cap \bar{B}$, where $\eta$ depends on $\bar{B}$.
Taking a countable, locally finite cover of $\mathcal{R}$ by such $\bar{B}$'s and let $E'$ be the union of such $E_{\bar{B}}$.
Then $E'$ is measure zero and relatively closed in $\mathcal{R}$. The choice of $E$ in (\ref{eqn:MC25_1}) follows the same idea, with the covering of $\mathcal{R}$ being written down explicitly.
It follows from (\ref{eqn:MC25_1}) that $E$ is the union of countably many measure-zero sets.
Consequently, $E$ is measure-zero.
Fix arbitrary $\bar{z} \in \mathcal{R} \backslash E$.
Because $\bar{z} \in \mathcal{R}$, we see that $\bar{z} \in B(\bar{x}, \eta^{-1}) \cap \mathcal{R}_{\eta}(\bar{M})$ for some $\eta>0$.
Accordingly, we can find $k_0$ very large such that $\bar{z} \in B(\bar{x}, 2^{k_0}\eta_0^{-1}) \cap \mathcal{R}_{2^{-k_0}\eta_0}(\bar{M})$.
Now using $\bar{z} \notin E$ and the decomposition of $E$ in (\ref{eqn:MA23_1}), we have
\begin{align*}
\bar{z} \notin E_{2^{-k_0}\eta_0} \backslash \left\{ B(\bar{x}, 2^{k_0-2} \eta_0^{-1}) \cap \overset{\circ}{\mathcal{R}}_{2^{-k_0+2} \eta_0}\right\} \quad
&\Leftrightarrow \quad \bar{z} \in \left\{ B(\bar{x}, 2^{k_0-2} \eta_0^{-1}) \cap \overset{\circ}{\mathcal{R}}_{2^{-k_0+2} \eta_0}\right\} \backslash E_{2^{-k_0}\eta_0}, \\
&\Rightarrow \quad \bar{z} \in \left\{ B(\bar{x}, 2^{k_0} \eta_0^{-1}) \cap \mathcal{R}_{2^{-k_0} \eta_0}\right\} \backslash E_{2^{-k_0}\eta_0}.
\end{align*}
Then it follows from our discussion in the previous paragraph that $(\bar{z}, -1)$ can be connected to $(\bar{x}, 0)$ by a shortest smooth reduced geodesic in $\mathcal{R}(\bar{M}) \times [-1, 0]$.
It is not hard to see that $E \cup \mathcal{S}$ is a closed set, which will be proved in this paragraph.
Suppose $z_i$ is a sequence of points in $E$. Without loss of genearlity, we can assume
\begin{align*}
z_i \in E_{2^{-k}\eta_0} \backslash \left\{ B(\bar{x}, 2^{k-2} \eta_0^{-1}) \cap \mathcal{R}_{2^{k-2} \eta_0}\right\},
\end{align*}
where $k=k(i)$.
Let $z$ be a limit point of $z_i$. There are two possibilities(by taking subsequence if necessary):
\begin{itemize}
\item $z \in \mathcal{S}$.
\item $z \in \mathcal{R}$. Then $z \in \mathcal{R}_{2\eta} \cap B(\bar{x}, 0.5 \eta^{-1})$ for some $\eta>0$. Therefore, we can assume $z_i \in \mathcal{R}_{\eta} \cap B(\bar{x}, \eta^{-1})$ for large $i$.
This forces that $k(i)$ is uniformly bounded. By taking subsequence if necessary, we can assume that $z_i \in E_{2^{-k}\eta_0} \backslash \left\{ B(\bar{x}, 2^{k-2} \eta_0^{-1}) \cap \overset{\circ}{\mathcal{R}}_{2^{k-2} \eta_0}\right\}$ for a fixed $k$.
By closedness of each $E_{\eta}$, we see that $z \in E_{2^{-k}\eta_0} \backslash \left\{ B(\bar{x}, 2^{k-2} \eta_0^{-1}) \cap \overset{\circ}{\mathcal{R}}_{2^{k-2} \eta_0}\right\} \subset E$.
\end{itemize}
Therefore, we conclude that $z \in E \cup \mathcal{S}$. Note that $\mathcal{S}$ is a closed set and has measuzre($2n$-Hausdorff measure) zero.
Then we obtain $E \cup \mathcal{S}$ is a closed measure-zero set.
Clearly, away from the closed measure-zero set $E \cup \mathcal{S}$, every point $\bar{z} \in \bar{M}$ satisfies the following property:
\textit{$(\bar{z}, -1)$ can be connected to $(\bar{x}, 0)$ by a shortest smooth reduced geodesic.}
\end{proof}
\begin{remark}
Note that the devlopment from Lemma~\ref{lma:SL13_1} to Lemma~\ref{lma:SK27_4} is parallel, or independent to the development
from Proposition~\ref{prn:SC16_1} to Prposition~\ref{prn:SC17_4}.
Our key observation is that the limit space has weakly convex regular part, which essentialy arises from
the weak convexity of $\mathcal{R} \times [-1, 0]$ in terms of reduced geodesics. For the convenience of the readers who are not familiar with singular space theory, we also provide an alternative proof of Proposition~\ref{prn:SC17_4}
in Appendix~\ref{app:C}, based on the discussion from Lemma~\ref{lma:SL13_1} to Lemma~\ref{lma:SK27_4}.
Then Proposition~\ref{prn:SC17_4} follows directly from Proposition~\ref{prn:MC04_1}.
\label{rmk:MC07_1}
\end{remark}
By natural projection to the time slice $t=0$, we obtain the following property.
\begin{proposition}[\textbf{Weak convexity by Riemannian geodesics}]
Same conditions as in Lemma~\ref{lma:SK27_4}.
Then away from a measure-zero set, every point in $\mathcal{R}$ can be connected to $\bar{x}$ with a unique smooth shortest geodesic. Consequently, $\mathcal{R}$ is weakly convex.
\label{prn:SC30_1}
\end{proposition}
\begin{proof}
Fix $\bar{x} \in \mathcal{R}$ and let $E$ be the measure-zero set constructed in the proof of Lemma~\ref{lma:SK27_4}.
Therefore, $(\bar{y}, -1)$ can be connected to $(\bar{x}, 0)$ by a smooth shortest reduced geodesic $\boldsymbol{\beta}$, with space projection curve $\beta$, whenever $\bar{y} \in \mathcal{R} \backslash E$.
For our purpose of weak convexity, it suffices to show that each $\beta$ is a smooth shortest geodesic connecting $\bar{x}$ and $\bar{y}$.
Actually, it follows from reduced geodesic equation on Ricci-flat manifold (c.f. equations (\ref{eqn:SL25_6})) that $\displaystyle \mathcal{L}(\boldsymbol{\beta})=\frac{1}{2}|\beta|^2$, where $|\beta|$ is the length of $\beta$.
Since both $\bar{x}$ and $\bar{y}$ are regular, for each small $\epsilon>0$, we can find a smooth geodesic $\gamma$ such that $|\gamma|<d_0(\bar{x}, \bar{y})+\epsilon$, by Proposition~\ref{prn:SC17_4}.
Because the limit space-time is static, we can lift $\gamma$ to be a space-time curve $\boldsymbol{\gamma}$ such that $\mathcal{L}(\boldsymbol{\gamma})=\frac{|\gamma|^2}{2}$.
Using the shortest property of $\boldsymbol{\beta}$ and the construction of $\gamma$, we have
\begin{align*}
\frac{|\beta|^2}{2}=\mathcal{L}(\boldsymbol{\beta}) \leq \mathcal{L}(\boldsymbol{\gamma})=\frac{|\gamma|^2}{2}< \frac{(d_0(\bar{x},\bar{y}) +\epsilon)^2}{2},
\quad \Rightarrow \quad |\beta|<d_0(\bar{x}, \bar{y}) + \epsilon.
\end{align*}
Since $\epsilon$ can be chosen arbitrarily small, we have $|\beta| \leq d_0(\bar{x}, \bar{y})$, which means $|\beta|=d_0(\bar{x}, \bar{y})$ and $\beta$ is a shortest Riemannian geodesic.
By adjusting $E$ to a bigger measure zero set $E'$ if necessary, we obtain the uniqueness of geodesics from $\bar{y}$ to $\bar{x}$ for each $\bar{y} \in \mathcal{R} \backslash E'$.
This follows from standard Riemannian geometry argument since $E' \backslash E \subset \mathcal{R}$.
\end{proof}
By the correspondence between smooth Riemannian geodesic and smooth reduced geodesic(c.f. the discussion in Section~\ref{subsec:reduced}), it is clear (from the proof of Proposition~\ref{prn:SC30_1}) now that most smooth reduced geodesics obtained in Lemma~\ref{lma:SK27_4} are shortest among all smooth reduced geodesics.
Furthermore, the rough estimate in Lemma~\ref{lma:SL13_1} can be improved as the following proposition.
\begin{proposition}[\textbf{Continuity of reducecd distance}]
Same conditions as in Lemma~\ref{lma:SK27_4}.
Suppose $(y_i,t_i) \in \mathcal{M}_i$ converges to $(\bar{y}, \bar{t})$, which is regular and $\bar{t}<0$.
Then we have
\begin{align}
\lim_{i \to \infty} l((x_i,0), (y_i,t_i))=\frac{d_0^2(\bar{x}, \bar{y})}{4|\bar{t}|}=l((\bar{x},0), (\bar{y},\bar{t}))
\label{eqn:SK27_9}
\end{align}
where $l$ is Perelman's reduced distance. Therefore, reduced distance is continuous function under Cheeger-Gromov topology whenever $\bar{y}$ is regular.
\label{prn:SL14_1}
\end{proposition}
\begin{proof}
Without loss of generality, we assume $t_i \equiv -1$, $d_0(x_i,y_i) \equiv 1$.
We first show
\begin{align}
\lim_{i \to \infty} l((x_i,0), (y_i,t_i)) \leq \frac{1}{4}.
\label{eqn:SL14_5}
\end{align}
If $x_i$ are uniformly regular, then there is a limit smooth geodesic connecting $\bar{x}$ and $\bar{y}$, which can
be lifted to a smooth reduced geodesic connecting $(\bar{x},0)$ and $(\bar{y},-1)$ with reduced length $\frac{1}{4}$.
Then (\ref{eqn:SL14_5}) follows trivially. So we focus on the case when $\bar{x}$ is a singular point.
Choose a smooth point $\bar{z}$ very close to $\bar{x}$, say $\delta$-away from $\bar{x}$ under metric $\bar{g}(0)$.
From Lemma~\ref{lma:SL13_1}, the reduced length from $(x_i, 0)$ to $(z_i, -\delta^2)$ is uniformly less than $100$. So we have space-time curves $\boldsymbol{\alpha}_i$ connecting these two points such that
\begin{align*}
\int_0^{\delta^2} \sqrt{\tau} |\dot{\alpha}_i|^2 d\tau< 200\delta.
\end{align*}
Note that $(\bar{z},-\delta^2)$ and $(\bar{y}, -1)$ can be connected by a space-time curve $\boldsymbol{\beta}$ such that
\begin{align*}
\int_{\delta^2}^{1} \sqrt{\tau} |\dot{\beta}|^2 d\tau< \frac{1}{2} + 100\delta
\end{align*}
if $\delta$ is small enough.
So for large $i$, we have space-time curve $\boldsymbol{\beta}_i$ connecting $(z_i,-\delta^2)$ and $(y_i, -1)$ such that
\begin{align*}
\int_{\delta^2}^{1} \sqrt{\tau} |\dot{\beta_i}|^2 d\tau < \frac{1}{2} + 200\delta.
\end{align*}
Concatenating $\boldsymbol{\alpha}_i$ and $\boldsymbol{\beta}_i$ to obtain $\boldsymbol{\gamma}_i$ such that
\begin{align*}
\int_{\delta^2}^{1} \sqrt{\tau} |\dot{\gamma_i}|^2 d\tau < \frac{1}{2} + 400\delta,
\end{align*}
which implies $l((x_i,0),(y_i,-1)) < \frac{1}{4} + 200\delta$ for large $i$. Thus (\ref{eqn:SL14_5}) follows by letting $i \to \infty$ and $\delta \to 0$.
Then we show the equality holds. Otherwise, there exists a small $\epsilon$ such that
\begin{align*}
\lim_{i \to \infty} l((x_i,0), (y_i, -1)) < \frac{1}{4} -\epsilon.
\end{align*}
Note that $(y_i,-1)$ is uniformly regular. So we can find small $\delta$ such that
\begin{align*}
l((x_i,0), (z, -1-\delta^2)) < \frac{1}{4}-\frac{1}{2}\epsilon, \quad \forall \;
z \in B_{g(-1-\delta^2)}(y_i, \epsilon \delta).
\end{align*}
By Lemma~\ref{lma:SK27_4}, we obtain a point $(\bar{z}, -1-\delta^2)$, which can be connected to
$(\bar{x},0)$ by a smooth reduced geodesic, with reduced length smaller than $\frac{1}{4}-\frac{1}{2}\epsilon$.
Projecting this reduced geodesic to time zero slice, we obtain a curve connecting $\bar{x}$ and $\bar{z}$ with
\begin{align*}
d_0^2(\bar{x}, \bar{y}) < 4(1+\delta^2) \cdot \left(\frac{1}{4}-\frac{1}{2}\epsilon \right)=(1+\delta^2)(1- 2\epsilon)<1-\epsilon
\end{align*}
if we choose $\delta$ sufficiently small. This is impossible since $d_0(\bar{x},\bar{y})=1$. Therefore, we have
\begin{align*}
\lim_{i \to \infty} l((x_i,0), (y_i, -1)) = \frac{1}{4}.
\end{align*}
\end{proof}
Since singular set has measure zero, it is clear that
\begin{align}
\mathcal{V}((\bar{x},0), |\bar{t}|) \leq \lim_{i \to \infty} \mathcal{V}((x_i,0),|\bar{t}|),
\label{eqn:SK27_10}
\end{align}
where the ``$\lim$" of the right hand side of the above inequality should be understood as ``$\limsup$".
We shall improve the above inequality as equality.\\
\begin{lemma}[\textbf{Major part of reduced volume}]
For every positive $\eta$ and $H$, there exists an $\epsilon=\epsilon(n,A,\eta,H)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in \mathcal{F}_{\eta}(M,0)$. Then we have
\begin{align}
\left|\mathcal{V}((x,0),1)-(4\pi)^{-n}\int_{B_{g(0)}(x,H)} e^{-l}dv \right| \leq 2a(H), \label{eqn:SL18_1}
\end{align}
whenever $\sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
Here $a$ is a positive function defined as
\begin{align}
a(H) \triangleq (4\pi)^{-n} \int_{\{|\vec{w}|>\frac{H}{100}\} \subset \ensuremath{\mathbb{R}}^{2n}} e^{-\frac{|\vec{w}|^2}{4}} dw. \label{eqn:SL18_2}
\end{align}
\label{lma:SL18_1}
\end{lemma}
\begin{proof}
The line bundle structure is not used in the following proof. So up to a parabolic rescaling if necessary, we can assume $\lambda=0$.
For every $y \in M$, there is at least one shortest reduced geodesic $\boldsymbol{\gamma}$ connecting $(x,0)$ and $(y,-1)$.
By standard ODE theory, the limit $\displaystyle \lim_{\tau \to 0} \sqrt{\tau} \gamma'(\tau)$ is unique as a vector in $T_{x} M$,
which is called the reduced tangent vector of $\boldsymbol{\gamma}$.
Away from a measure-zero set, every $(y,-1)$ can be connected to $(x,0)$ by a
unique shortest reduced geodesic. For simplicity for our argument, we may assume this measure-zero set is empty, since measure-zero
set does not affect integral at all.
So there is a natural injective map from $M$ to $T_xM$, by mapping $y$ to the corresponding
reduced tangent vector $\vec{w}$.
We define
\begin{align*}
\Omega(H) \triangleq \{ y \in M | |\vec{w}|>H \}.
\end{align*}
It follows from the monotonicity of reduced element along reduced geodesic that
\begin{align*}
\int_{\Omega(H)} (4\pi)^{-n} e^{-l} dv \leq \int_{\{|\vec{w}| > H\} \subset \ensuremath{\mathbb{R}}^{2n}} (4\pi)^{-n} e^{-\frac{|\vec{w}|^2}{4}} dw.
\end{align*}
Choose $\xi<\eta$, with size to be determined.
Suppose $\boldsymbol{\gamma}$ is a reduced geodesic connecting $(x,0)$ to $(y,-1)$ for some $y \in M$.
It is clear that $\gamma(0)$ is in the interior part of $\mathcal{F}_{\xi}(M,0)$. Let $\tau$ to be the first time such that $\gamma(\tau)$ touches
the boundary of $\mathcal{F}_{\xi}(M,0)$. Then we see that $\boldsymbol{\gamma}([0,\tau])$ locates in a space-time domain
with uniformly bounded geometry, Ricci curvature very small. In particular, the reduced distance between $(x,0)$ and $\boldsymbol{\gamma}(\tau)$
is comparable to the length of $\vec{w}$, which is the reduced tangent vector of $\boldsymbol{\gamma}$ at $(x,0)$.
If $|\vec{w}|<H$, then we see that
\begin{align*}
\frac{H^2}{4} > \frac{|\vec{w}|^2}{4} \sim \frac{d_{g(0)}^2(x, \gamma(\tau))}{4\tau}> \frac{c_a^2 \eta^2}{100\tau},
\quad \Rightarrow \quad \tau > \frac{c_a^2 \eta^2}{25H^2}.
\end{align*}
Note that $\boldsymbol{\gamma}([0,\tau])$ is in a space-time region where Ricci curvature is almost flat, geometry is uniformly bounded.
So the lower bound of $\tau$ and the upper bounded of $|\vec{w}|$ imply an upper bound of $d_{g(0)}(x,\gamma(\tau))$.
Say $d_{g(0)}(x,\gamma(\tau))<H'$.
Around $\boldsymbol{\gamma}$, there is a natural projection (induced by reduced geodesic) from the space-time hypersurface
$\partial \mathcal{F}_{\xi}(M,0) \times [-1, -\frac{c_a^2 \eta^2}{25H^2}]$,
to the time slice $M \times \{-1\}$. At point $\boldsymbol{\gamma}(\tau)$, $\boldsymbol{\gamma}$ has space-time tangent vector $(\gamma', -1)$,
with $\tau|\gamma'(\tau)|^2$ is almost less than $\frac{H^2}{4}$. Together with the lower bound of $\tau$, we obtain an upper bound of $|\gamma'(\tau)|$.
Up to a constant depending on $H,\eta$, the volume element of $\partial \mathcal{F}_{\xi}(M,0) \times [-1,-\frac{c_a^2 \eta^2}{25H^2}]$
is comparable to the reduced volume element
$(4\pi \tau)^{-n} e^{-l}$ of $M$, around the point $\boldsymbol{\gamma}(\tau)$.
Note that the reduced volume element is monotone along each reduced geodesic. This implies that the projection map mentioned above
``almost" decreases weighted hypersurface volume element, if we equip $\{B(x,H') \cap \partial \mathcal{F}_{\xi}(M,0)\} \times [-1, -\frac{\eta^2}{4H^2}]$ with the natural
weighted volume element $e^{-l}|d \sigma \wedge dt|$. Let $\Omega_{\xi}'$ be the collection of all $y$'s such that $(y,-1)$ cannot be connected to $(x,0)$ by a shortest reduced geodesic $\boldsymbol{\gamma}$ which locates completely in $\mathcal{F}_{\xi}(M,0) \times [-1,0]$.
Then we have
\begin{align*}
\int_{\Omega_{\xi}'} e^{-l}(4\pi \tau)^{-n} dv &\leq C \int_{\frac{c_a^2 \eta^2}{25H^2}}^1 \int_{B(x,H') \cap \partial \mathcal{F}_{\xi}(M,0)} e^{-l} d\sigma d\tau
\leq C \int_{B(x,H') \cap \partial \mathcal{F}_{\xi}(M,0)} d\sigma
\leq C \xi^{2p_0-1},
\end{align*}
where $C=C(n,H,H',\eta)=C(n,H,\eta)$. By choosing $\xi$ small enough, we have
\begin{align}
\int_{\Omega_{\xi}'} e^{-l}(4\pi \tau)^{-n} dv \leq (4\pi)^n a(H). \label{eqn:SL18_3}
\end{align}
Note that
\begin{align*}
\Omega_{100 H} \cap B_{g(0)}(x,H) \subset \Omega_{\xi}', \qquad
M \backslash (\Omega_{\xi}' \cup B_{g(0)}(x,H)) \subset \Omega_{\frac{H}{100}}.
\end{align*}
Therefore, recalling the definition of reduced volume (\ref{eqn:MA22_4}), we have
\begin{align*}
&\quad (4\pi)^n \mathcal{V}((x,0),1)
=\int_{M} e^{-l}dv
=\int_{M \backslash (\Omega_{\xi}' \cup B_{g(0)}(x,H))} e^{-l}dv + \int_{\Omega_{\xi}'} e^{-l}dv + \int_{B_{g(0)}(x,H) \backslash \Omega_{\xi}'} e^{-l}dv\\
&\leq \int_{|\vec{w}|>\frac{H}{100}} e^{-\frac{|\vec{w}|^2}{4}} dw + \int_{\Omega_{\xi}'} e^{-l}dv + \int_{B_{g(0)}(x,H)} e^{-l}dv
\leq \int_{|\vec{w}|>\frac{H}{100}} e^{-\frac{|\vec{w}|^2}{4}} dw + C \xi^{2p_0-1} + \int_{B_{g(0)}(x,H)} e^{-l}dv\\
&\leq 2 (4\pi)^n a(H) + \int_{B_{g(0)}(x,H)} e^{-l}dv.
\end{align*}
Then (\ref{eqn:SL18_1}) follows from the above inequality directly.
\end{proof}
Lemma~\ref{lma:SL18_1} is related to Corollary 6.82 of~\cite{MT}.
\begin{lemma}[\textbf{Uniform continuity of reduced volume}]
Suppose $\mathcal{M}=\{(M, g(t)), -\tau \leq t \leq 0\}$ is an unnormalized K\"ahler Ricci flow solution.
Suppose $x,y$ are two points in $M$, $d=d_{g(0)}(x,y)$. Then we have
\begin{align}
|\mathcal{V}((x,0),\tau) - \mathcal{V}((y,0),\tau)|<(4n+1) (e^{\frac{d}{2}}-1).
\label{eqn:SL15_2}
\end{align}
In particular, the reduced volume changes uniformly continuously with respect to the base point.
\label{lma:SL14_2}
\end{lemma}
\begin{proof}
Recall the definition of reduced volume (\ref{eqn:MA22_4}):
\begin{align*}
\mathcal{V}((x,0),\tau)= (4\pi \tau)^{-n} \int_M e^{-l} dv.
\end{align*}
Let $x$ move along a unit speed Riemannian geodesic $\alpha$, with respect to the metric $g(0)$.
Let $x=\alpha(0)$, $s$ be parameter of $\alpha$, $\vec{u}=\alpha'$.
For simplicity of notation, we denote $\mathcal{V}((\alpha(s),0),\tau)$ by $\mathcal{V}_s$.
It can be calculated directly the first variation of $l$ is $\langle\vec{u}, \vec{w} \rangle$ where $\vec{w}$ is the tangent vector of the reduced geodesic at time $t=0$.
Therefore, we have
\begin{align*}
\left|\frac{d}{ds} \mathcal{V}((\alpha(s),0),\tau) \right|&=\left| (4\pi \tau)^{-n} \int_M \langle \vec{u}, \vec{w} \rangle e^{-l}dv \right|
\leq (4\pi \tau)^{-n} \int_M \frac{1+|\vec{w}|^2}{2} e^{-l}dv\\
&=\frac{1}{2}\mathcal{V} + \frac{1}{2} \int_{\ensuremath{\mathbb{R}}^{2n}} |\vec{w}|^2e^{-\frac{|\vec{w}|^2}{4}} J dw
\leq \frac{1}{2}\mathcal{V} + \frac{(4\pi)^{-n}}{2} \int_{\ensuremath{\mathbb{R}}^{2n}} |\vec{w}|^2e^{-\frac{|\vec{w}|^2}{4}}dw,
\end{align*}
where $J$ is the Jacobian determinant of the reduced exponential map, which is always not greater than $1$, due to Perelman's argument in Section 7 of \cite{Pe1}.
Plugging the identity
\begin{align*}
(4\pi)^{-n} \int_{\ensuremath{\mathbb{R}}^{2n}} |\vec{w}|^2 e^{-\frac{|\vec{w}|^2}{4}} dw = 4n
\end{align*}
into the above inequality implies $ \left|\frac{d}{ds} \mathcal{V} \right| \leq \frac{1}{2}\mathcal{V} +2n$,
which can be integrated as
\begin{align*}
(-\mathcal{V}_0+4n) (1-e^{-\frac{s}{2}})
\leq \mathcal{V}_s -\mathcal{V}_0
\leq (\mathcal{V}_0 + 4n) (e^{\frac{s}{2}}-1).
\end{align*}
Note that $0<\mathcal{V}_0\leq 1, s>0$. So we obtain
\begin{align*}
|\mathcal{V}_s -\mathcal{V}_0| \leq (4n+1) (e^{\frac{s}{2}}-1),
\end{align*}
which yields (\ref{eqn:SL15_2}) by letting $s=d$.
\end{proof}
The above argument clearly works for every Riemannian Ricci flow.
Note that the reduced volume is continuous for geodesic balls of each fixed scale under the Cheeger-Gromov convergence.
Combining this continuity together with the estimate in Lemma~\ref{lma:SL18_1} and Lemma~\ref{lma:SL14_2},
we can improve (\ref{eqn:SK27_10}) as an equality.
\begin{proposition}[\textbf{Continuity of reduced volume}]
Same conditions as in Lemma~\ref{lma:SK27_4}, $\bar{t}<0$ is a finite number.
Then we have
\begin{align}
\mathcal{V}((\bar{x},0), |\bar{t}|) = \lim_{i \to \infty} \mathcal{V}((x_i,0),|\bar{t}|).
\label{eqn:SL09_1}
\end{align}
Therefore, reduced volume is a continuous function under the Cheeger-Gromov convergence.
\label{prn:SL14_2}
\end{proposition}
Then we can study the gap property of the singularities.
\begin{proposition}[\textbf{Gap of local volume density}]
Same conditions as in Theorem~\ref{thm:SC09_1}.
Suppose $\bar{y} \in \mathcal{S}(\bar{M})$, then we have
\begin{align}
\mathrm{v}(\bar{y})=\lim_{r \to 0} \omega_{2n}^{-1} r^{-2n}|B(\bar{y},r)| \leq 1-2\delta_0.
\label{eqn:SC30_4}
\end{align}
\label{prn:SC17_7}
\end{proposition}
\begin{proof}
Due to the tangent cone structure(c.f. Theorem~\ref{thm:SL25_1}), we have
\begin{align}
\mathrm{v}(\bar{y})=\lim_{r \to 0} \omega_{2n}^{-1} r^{-2n}|B(\bar{y},r)|=\lim_{r \to 0} \mathcal{V}((\bar{y},0), r^2).
\label{eqn:SL14_6}
\end{align}
Let $y_i \to \bar{y}$ under the metric $g_i(0)$. By rearranging points and taking subsequences if necessary, we can assume $y_i$ has the
``local minimum" canonical volume radius $\rho_i$.
The rearrangement is a standard point-picking technique.
In fact, since $\bar{y}$ is a singular point, it is clear that $r_i=\mathbf{cvr}(y_i, 0) \to 0$.
Since everything is done at time slice $t=0$, we shall drop the time in the following argument.
Fix $L \geq 1$ and $i$, we search if $y_i$ is the point such that
\begin{align*}
\mathbf{cvr}(y) <0.5 \mathbf{cvr}(y_i), \quad \forall \; y \in B(y_i, Lr_i).
\end{align*}
If so, we stop. Otherwise, we can find a point $z \in B(y_i, Lr_i)$ such that $\mathbf{cvr}(z)<0.5 \mathbf{cvr}(y_i)$.
Denote such $z$ by $y_i^{(1)}$ and set $r_i^{(1)}=\mathbf{cvr} \left(y_i^{(1)} \right)$. We then repeat the previous process for $y_i^{(1)}$ and $r_i^{(1)}$.
To search points in the ball $B\left(y_i^{(1)}, Lr_i^{(1)} \right)$ with $\mathbf{cvr}<0.5 r_i^{(1)}$. If no such points exist, we stop. Otherwise, we find such a point
and denote it by $y_i^{(2)}$ and set $r_i^{(2)}=\mathbf{cvr} \left(y_i^{(2)} \right)$. Note this process happens in a compact set since
\begin{align*}
d\left( y_i^{(k)}, y_i\right)<L\left( r_i + r_i^{(1)} + \cdots + r_i^{(k)} \right)<2Lr_i.
\end{align*}
Each $\mathcal{LM}_i$ is smooth. Therefore, the process above must stop at some finite step $k$. Denote $z_i=y_i^{(k)}$ and $\rho_i=\mathbf{cvr}(z_i)$.
Then we have
\begin{align*}
\mathbf{cvr}(y)>0.5 \rho_i, \quad \forall \; y \in B(z_i, L\rho_i).
\end{align*}
Note that $L\rho_i \to 0$ as $i \to \infty$. Therefore, the limit of $z_i$ and the limit of $y_i$ are the same point $\bar{y}$.
Then we let $L \to \infty$ and take diagonal sequence, we obtain $z_i$ such that
\begin{align*}
\mathbf{cvr}(y)>0.5 \rho_i, \quad \forall \; y \in B(z_i, 2^i \rho_i); \quad \quad \lim_{i \to \infty} z_i=\bar{y}.
\end{align*}
Therefore, we can regard $z_i$ as the rearrangement of $y_i$, with the property that each $z_i$ achieve the ``local minimum" of $\mathbf{cvr}$.
By rescaling $\rho_i$ to $1$, we obtain new Ricci flows $\tilde{g}_i$.
Taking limit of $(M_i,y_i,\tilde{g}_i(0))$, we have a complete, Ricci flat
eternal Ricci flow solution.
It is not hard to see the limit space is not Euclidean. For otherwise, each geodesic ball's volume ratio, under metric $\tilde{g}_{\infty}(0)$, is exactly the Euclidean volume ratio $\omega_{2n}$.
Following from the volume convergence and the definition of the canonical volume radius, it is clear that the canonical volume radius of the rescaled flow is strictly greater than $1$ which contradicts to our assumption.
So it has normalized asymptotic volume ratio less than $1-2\delta_0$, according to Anderson's gap theorem. Then the infinity tangent cone
structure implies the asymptotic reduced volume is the same as the asymptotic reduced volume ratio. So it is
at most $1-2\delta_0$. Therefore, there exists a big constant $H$ such that
\begin{align*}
\mathcal{V}_{\tilde{g}_i}((y_i,0),H) < 1-2\delta_0.
\end{align*}
Note that $H\rho_i^2<r$ for each fixed $r$ and the corresponding large $i$. Recall the scaling invariant property of reduced volume, we can apply the reduced volume monotonicity to obtain
\begin{align*}
\mathcal{V}_{g_i}((y_i,0),r^2) \leq \mathcal{V}_{\tilde{g}_i}((y_i,0), \rho_i^{-2} r^2)
\leq \mathcal{V}_{\tilde{g}_i}((y_i,0),H) < 1-2\delta_0.
\end{align*}
The continuity of reduced volume (Proposition~\ref{prn:SL14_2}) then implies that
\begin{align*}
\mathcal{V}((\bar{y},0),r^2) \leq 1-2\delta_0
\end{align*}
for each $r>0$, which in turn yields
\begin{align}
\lim_{r \to 0} \mathcal{V}((\bar{y},0),r^2) \leq 1-2\delta_0.
\label{eqn:SL14_7}
\end{align}
Then (\ref{eqn:SC30_4}) follows from the combination of (\ref{eqn:SL14_6}) and (\ref{eqn:SL14_7}).
\end{proof}
\begin{theorem}[\textbf{Metric structure of a blowup limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;1)$ satisfies (\ref{eqn:SL06_1}), $x_i \in M_i$.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i, x_i, g_i(0))$. Then $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{thm:SC04_1}
\end{theorem}
\begin{proof}
We only need to check $\bar{M}$ satisfies all the 6 properties required in the definition of $\widetilde{\mathscr{KS}}(n,\kappa)$. In fact,
the 1st property is implied by Theorem~\ref{thm:HE11_1}.
The 2nd property follows from the fact that $\mathcal{R}$ is scalar flat and satisfies K\"ahler Ricci flow equation.
The 3rd property, weak convexity of $\mathcal{R}$ is shown in Proposition~\ref{prn:SC30_1}.
The 4th property, codimension estimate of singularity follows from Proposition~\ref{prn:SL15_1}.
The 5th property, gap estimate, follows from Proposition~\ref{prn:SC17_7}.
The 6th property, asymptotic volume ratio estimate can be obtained by the condition $Vol(M_i) \to \infty$, Sobolev constant uniformly bounded, and the volume convergence, Proposition~\ref{prn:SC12_1}.
So we have checked all the properties needed to define $\widetilde{\mathscr{KS}}(n,\kappa)$ are satisfied by $\bar{M}$.
In other words, $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\end{proof}
Since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, it is clear that $\mathbf{cr}(\bar{x})=\infty$. Therefore, we have $\mathbf{vr}(\bar{x})=\mathbf{cvr}(\bar{x})$
by definition.
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}. Let $\displaystyle \bar{r}=\lim_{i \to \infty} \mathbf{cr}(x_i)$.
Then we have
\begin{align}
\min\{\bar{r}, \mathbf{vr}(\bar{x})\} = \lim_{i \to \infty} \mathbf{cvr}(x_i).
\label{eqn:SC06_9}
\end{align}
\label{prn:SC06_5}
\end{proposition}
\begin{proof}
We divide the proof in three cases according to the value of $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}$. \\
\textit{Case 1. $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}=0$.}
Otherwise, there exists a positive number $\rho_0$ such that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) \geq \rho_0$. Therefore,
$\bar{x} \in \mathcal{R}_{\rho_0} \subset \mathcal{R}$, which in turn implies that $\mathbf{vr}(\bar{x})>0$.
Consequently, we have $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}>0$. Contradiction.
\textit{Case 2. $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}=\infty$.}
In this case, $\bar{r}=\infty$.
By the gap theorem in the space $\widetilde{\mathscr{KS}}(n,\kappa)$, we see that $\bar{M}$ is the Euclidean space $\ensuremath{\mathbb{C}}^{n}$.
Therefore, for each $H>0$, we have $\omega_{2n}^{-1}H^{-2n}|B(x_i,H)|$ converges to $1$, the normalized volume ratio of $\ensuremath{\mathbb{C}}^{n}$.
Since $\displaystyle \bar{r}=\lim_{i \to \infty} \mathbf{cr}(x_i)=\infty$, this means that $\mathbf{cvr}(x_i) \geq H$ for large $i$ by the volume convergence.
Since $H$ is chosen arbitrarily, we obtain $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i)=\infty$.\\
So the remainder case is that $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}$ is a finite positive number.
Two more subcases can be divided.\\
\textit{Case 3(a). $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}<\bar{r}$.}
Let $H=\mathbf{vr}(\bar{x})$, a finite number in this case. Clearly, $\bar{x}$ is a regular point and the normalized volume ratio of the ball $B(\bar{x},H)$
is $1-\delta_0$. Clearly, $B(\bar{x},H)$ cannot be a isometric to a Euclidean ball.
Therefore, by the rigidity of $\widetilde{\mathscr{KS}}(n,\kappa)$(c.f. Proposition~\ref{prn:SC17_3}), we see that
\begin{align*}
&\omega_{2n}^{-1}r^{-2n}|B(\bar{x},r)|>1-\delta_0, \quad \forall \; r \in (0,H), \\
&\omega_{2n}^{-1}r^{-2n}|B(\bar{x},r)|<1-\delta_0, \quad \forall \; r \in (H,\bar{r}).
\end{align*}
Then the volume convergence implies that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i)=H$. \\
\textit{Case 3(b). $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}=\bar{r}$.}
In this case, we see that the normalized volume ratio of $B(\bar{x},\bar{r})$ is at least $1-\delta_0$. Also, we see that $\bar{x}$ is a regular point.
Same argument as in the previous case, we see that
\begin{align*}
\omega_{2n}^{-1}r^{-2n}|B(\bar{x},r)|>1-\delta_0, \quad \forall \; r \in (0,\bar{r}).
\end{align*}
Therefore, for every fixed $r \in (0, \bar{r})$, the volume convergence implies that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) \geq r$. Consequently,
we have $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) \geq \bar{r}$ by the arbitrariness of $r$. On the other hand, the definition of $\mathbf{cvr}(x_i)$
implies that
\begin{align*}
\lim_{i \to \infty} \mathbf{cvr}(x_i) \leq \lim_{i \to \infty} \mathbf{cr}(x_i)=\bar{r}.
\end{align*}
Therefore, we obtain $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) =\bar{r}$.
\end{proof}
\begin{corollary}
Same conditions as in Theorem~\ref{thm:SC04_1}. Then for each $r \in (0,1)$, we have
\begin{align}
\mathcal{F}_{r}(\bar{M})=\mathcal{R}_{r}(\bar{M}).
\label{eqn:SC06_11}
\end{align}
In particular, for each $0<r<1<H<\infty$, we have
\begin{align*}
B(x_i,H) \cap \mathcal{F}_{r}(M_i) \longright{G.H.} B(\bar{x},H) \cap \mathcal{F}_{r}(\bar{M}).
\end{align*}
Moreover, this convergence can be improved to take place in $C^{\infty}$-topology, i.e.,
\begin{align}
B(x_i,H) \cap \mathcal{F}_{r}(M_i) \longright{C^{\infty}} B(\bar{x},H) \cap \mathcal{F}_{r}(\bar{M}). \label{eqn:SC06_12}
\end{align}
\label{cly:SC06_1}
\end{corollary}
\begin{corollary}
Same conditions as in Theorem~\ref{thm:SC04_1}, $0<H\leq 3$.
Then we have
\begin{align}
\lim_{i \to \infty} \int_{B(x_i,H)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy \leq H^{2n-2p_0}\mathbf{E}.
\label{eqn:SC06_13}
\end{align}
\label{cly:SC06_2}
\end{corollary}
\begin{proof}
Fix two positive scales $r_1,r_2$ such that $0<r_2<r_1<1$.
\begin{align}
\int_{B(x_i,H) \cap \mathcal{F}_{r_1}} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\leq r_1^{-2p_0} |B(x_i,H) \cap \mathcal{F}_{r_1}| \leq r_1^{-2p_0}|B(x_i,H)|.
\label{eqn:SC17_6}
\end{align}
Fix arbitrary $r \in (0,1)$, then we have
\begin{align*}
\lim_{i \to \infty} \int_{B(x_i,H) \cap (\mathcal{F}_{r_2} \backslash \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy
=\int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy.
\end{align*}
Note that
\begin{align*}
&\qquad \int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\\
&\leq \int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \min\{\mathbf{vr}, 1\}^{-2p_0} dy
<\int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \left\{1+\mathbf{vr}(y)^{-2p_0} \right\} dy\\
&<\int_{B(\bar{x},H)} \left\{1+\mathbf{vr}(y)^{-2p_0} \right\} dy
<|B(\bar{x},H)| + H^{2n-2p_0}E(n,\kappa,p_0).
\end{align*}
It follows that
\begin{align}
\lim_{i \to \infty} \int_{B(x_i,H) \cap (\mathcal{F}_{r_2} \backslash \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy
\leq |B(\bar{x},H)| + H^{2n-2p_0}E(n,\kappa,p_0).
\label{eqn:SC17_7}
\end{align}
Note that $\mathcal{S} \cap \overline{B(\bar{x},H)}$ is a compact set with Hausdorff dimension at most
$2n-4$, which is strictly less than $2n-2p_0$. By the definition of Hausdorff dimension, for every small number $\xi$,
we can find finite cover
$\cup_{j=1}^{N_{\xi}} B(\bar{y}_j, \rho_j)$ of $\mathcal{S} \cap \overline{B(\bar{x},H)}$ , such that $\displaystyle \sum_{j=1}^{N_{\xi}} |\rho_{j}|^{2n-2p_0} < \xi$.
By the finiteness of this cover, we can choose an $r_2$ very small such that $\cup_{j=1}^{N_{\xi}} B(\bar{y}_j, \rho_j)$
is a cover of $\mathcal{D}_{r_2} \cap \overline{B(\bar{x},H)}$.
Therefore, for large $i$, we have a finite cover $\cup_{j=1}^{N_{\xi}} B(y_{i,j}, \rho_j)$ of the set
$\mathcal{D}_{r_2}(M_i) \cap \overline{B(x_i,H)}$ such that $\displaystyle \sum_{j=1}^{N_i} |\rho_{i,j}|^{2n-2p_0} < \xi$.
Combining this with the canonical radius density estimate, we have
\begin{align}
\int_{B(x_i,H) \cap \mathcal{D}_{r_2}} \mathbf{vr}^{(1)}(y)^{-2p_0}dy
\leq \sum_{j=1}^{N_i} \int_{B(y_{i,j},\rho_{i,j})} \mathbf{vr}^{(\rho_{i,j})}(y)^{-2p_0} dy
\leq 2\mathbf{E} \sum_{j=1}^{N_i} |\rho_{i,j}|^{2n-2p_0} < 2\mathbf{E}\xi.
\label{eqn:SC17_8}
\end{align}
Putting (\ref{eqn:SC17_6}), (\ref{eqn:SC17_7}) and (\ref{eqn:SC17_8}) together, we have
\begin{align*}
&\qquad \int_{B(x_i,H)} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\\
&\leq \int_{B(x_i,H) \cap \mathcal{F}_{r_1}} \mathbf{vr}^{(1)}(y)^{-2p_0}dy
+\int_{B(x_i,H) \cap (\mathcal{F}_{r_2} \backslash \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0}dy
+\int_{B(x_i,H) \cap \mathcal{D}_{r_2}} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\\
&\leq r_1^{-2p_0}|B(x_i,H)|+ |B(\bar{x},H)| +H^{2n-2p_0} E(n,\kappa,p_0) +2\mathbf{E}\xi.
\end{align*}
Taking limit on both sides and then letting $\xi \to 0, r_1 \to 1$, we have
\begin{align*}
\lim_{i \to \infty} \int_{B(x_i,H)} \mathbf{vr}^{(1)}(y)^{-2p_0} &\leq 2|B(\bar{x},H)| + H^{2n-2p_0} E(n,\kappa,p_0)
\leq \left( 2 \omega_{2n} H^{2p_0} + E(n,\kappa,p_0) \right) H^{2n-2p_0}\\
&\leq \left( 2 \cdot 9^{p_0}\omega_{2n} + E(n,\kappa,p_0) \right) H^{2n-2p_0},
\end{align*}
where we used the fact that $H\leq 3$ in the last step. Then (\ref{eqn:SC06_13}) follows from the definition of $\mathbf{E}$.
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $1 \leq H <\infty$. Then we have
\begin{align}
\lim_{i \to \infty} \sup_{1 \leq \rho \leq H} \omega_{2n}^{-1}\rho^{-2n}|B(x_i,\rho)| <\kappa^{-1}, \label{eqn:SC06_4}
\end{align}
where $g_i(0)$ is the default metric. In particular, for every large $i$, the volume ratio estimate holds
on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_1}
\end{proposition}
\begin{proof}
We argue by contradiction. If (\ref{eqn:SC06_4}) were false, by taking subsequence if necessary, one can assume that there exists $\rho_i \in [1,H]$ such that
$ \omega_{2n}^{-1}\rho_i^{-2n}|B(x_i,\rho_i)| >\kappa^{-1}$.
Let $\bar{\rho}$ be the limit of $\rho_i$, then by the volume continuity in the Cheeger-Gromov convergence, we see that
\begin{align}
\omega_{2n}^{-1}\bar{\rho}^{-2n}|B(\bar{x},\bar{\rho})| \geq \kappa^{-1}.
\label{eqn:SC06_5}
\end{align}
However, since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, we know $ \omega_{2n}^{-1}\bar{\rho}^{-2n}|B(\bar{x},\bar{\rho})| \leq 1$,
which contradicts (\ref{eqn:SC06_5}).
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $0<H<\infty$. For every large $i$, the regularity estimate holds
on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_2}
\end{proposition}
\begin{proof}
If the statement were false, then by taking subsequence if necessary, we can assume there exists $\rho_i \in (0,H]$ such that the regularity estimates fail on
the scale $\rho_i$, i.e., the following two inequalities hold simultaneously.
\begin{align}
&\omega_{2n}^{-1}\rho_i^{-2n}|B(x_i,\rho_i)|>1-\delta_0, \label{eqn:SC06_6}\\
&\max_{0 \leq k \leq 5} \left\{\rho_i^{2+k} \sup_{B(x_i,\frac{1}{2}c_a \rho_i)} |\nabla^k Rm| \right\} > 4c_a^{-2}. \label{eqn:SC06_7}
\end{align}
Clearly, $\rho_i \in [1,H]$ by the fact $\mathbf{cr}(x_i,0) \geq 1$. Let $\bar{\rho}$ be the limit of $\rho_i$. Then we have
\begin{align}
\omega_{2n}^{-1}\bar{\rho}^{-2n}|B(\bar{x},\bar{\rho})| \geq 1-\delta_0. \label{eqn:SC06_8}
\end{align}
Since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, (\ref{eqn:SC06_8}) implies
\begin{align*}
\max_{0 \leq k \leq 5} \left\{\bar{\rho}^{2+k} \sup_{B(\bar{x}, c_a \bar{\rho})} |\nabla^k Rm| \right\} < c_a^{-2},
\end{align*}
which contradicts (\ref{eqn:SC06_7}) in light of the smooth convergence(c.f. Proposition~\ref{prn:SL04_1}).
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $1 \leq H \leq 2$. Then we have
\begin{align}
\lim_{i \to \infty} \sup_{1 \leq \rho \leq H} \rho^{2p_0-2n} \int_{B(x_i,\rho)} \mathbf{vr}^{(\rho)}(y)^{-2p_0}dy \leq \frac{3}{2}\mathbf{E}. \label{eqn:SC06_16}
\end{align}
In particular, for every large $i$, the density estimate holds on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_3}
\end{proposition}
\begin{proof}
Since $\mathbf{vr}^{(\rho)} \geq \mathbf{vr}^{(1)}$ whenever $\rho \geq 1$, in order to show (\ref{eqn:SC06_16}), it suffices to show
\begin{align}
\lim_{i \to \infty} \sup_{1 \leq \rho
\leq H} \rho^{2p_0-2n} \int_{B(x_i,\rho)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy \leq \frac{3}{2}\mathbf{E}.
\label{eqn:SL23_9}
\end{align}
We argue by contradiction. If (\ref{eqn:SL23_9}) were false, by taking subsequence if necessary, one can assume that there exists $\rho_i \in [1,H]$ such that
\begin{align*}
\rho_i^{2p_0-2n} \int_{B(x_i,\rho_i)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy > \frac{3}{2}\mathbf{E}, \quad
\Rightarrow \quad \int_{B(x_i,\rho_i)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy \geq \frac{3}{2}\mathbf{E} \rho_i^{2n-2p_0}.
\end{align*}
Let $\bar{\rho}$ be the limit of $\rho_i$. Fix $\epsilon$ arbitrary small positive number, then we have
\begin{align}
\int_{B(x_i,\bar{\rho}+\epsilon)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy> \frac{3}{2}\mathbf{E} \rho_i^{2n-2p_0}> \frac{5}{4} \mathbf{E} (\bar{\rho}+\epsilon)^{2n-2p_0}
\label{eqn:SC07_1}
\end{align}
for large $i$. Note that $\bar{\rho}+\epsilon<3$, so (\ref{eqn:SC07_1}) contradicts (\ref{eqn:SC06_13}).
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $0<H\leq 2$.
Then for every large $i$, the connectivity estimate holds on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_4}
\end{proposition}
\begin{proof}
By the canonical radius assumption, we know the connectivity estimate holds for every scale $\rho \in (0,1]$.
If the statement were false, then by taking subsequence if necessary, we can assume that for each $i$,
there is a scale $\rho_i \in [1,H]$ such that the connectivity estimate fails on the scale $\rho_i$.
In other words, $\mathcal{F}_{\frac{1}{50}c_b\rho_i} \cap B(x_i,\rho_i)$ is not $\frac{1}{2}\epsilon_b \rho_i$-regular-connected.
So there exist points $y_i,z_i \in \mathcal{F}_{\frac{1}{50}c_b\rho_i} \cap B(x_i,\rho_i)$ which cannot be connected by a curve
$\gamma \subset \mathcal{F}_{\frac{1}{2}\epsilon_b \rho_i}$ satisfying $|\gamma|\leq 2d(y_i,z_i)$.
By the canonical radius assumption, it is clear that $\rho_i \in [1,H]$, $d(y_i,z_i) \in [1,2H]$.
Let $\bar{\rho}$ be the limit of $\rho_i$, $\bar{y}$ and $\bar{z}$ be the limit of $y_i$ and $z_i$ respectively.
Clearly, we have $\bar{y}, \bar{z} \in \mathcal{R}_{\frac{1}{50}c_b \bar{\rho}} \subset \mathcal{F}_{\frac{1}{100}c_b \bar{\rho}}(\bar{M})$.
Since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, we can find a shortest geodesic
$\bar{\gamma}$ connecting $\bar{y}$ and $\bar{z}$ such that $\bar{\gamma} \subset \mathcal{F}_{\epsilon_b \bar{\rho}}$.
Note that the limit set of $\mathcal{F}_{\frac{1}{50}c_b \rho_i}\cap B(x_i,\rho_i)$ falls into $\mathcal{F}_{\frac{1}{100}c_b \bar{\rho}}$.
Moreover, this convergence takes place in the smooth topology(c.f.~Corollary~\ref{cly:SC06_1}).
So by deforming $\bar{\gamma}$ if necessary, we can construct a curve $\gamma_i$ which locates in $\mathcal{F}_{\frac{1}{2}\epsilon_b \rho_i}$
and $|\gamma_i| <\frac{3}{2}d(\bar{y},\bar{z})<3d(y_i,z_i)$. The existence of such a curve contradicts the choice of the points
$y_i$ and $z_i$.
\end{proof}
Combining Proposition~\ref{prn:SC06_1} to~\ref{prn:SC06_4}, we obtain a weak-semi-continuity of canonical radius.
\begin{theorem}[\textbf{Weak continuity of canonical radius}]
Same conditions as in Theorem~\ref{thm:SC04_1}. Then we have
$\displaystyle \lim_{i \to \infty} \mathbf{cr}(\mathcal{M}_i^{0})=\infty$.
\label{thm:SC03_1}
\end{theorem}
\begin{proof}
If the statement were wrong, then we can find a sequence of polarized K\"ahler Ricci flow solutions
$\mathcal{LM}_i \in \mathscr{K}(n,A;1)$ satisfying (\ref{eqn:SL06_1}) and
\begin{align}
\lim_{i \to \infty} \mathbf{cr}(\mathcal{M}_i^{0})=H<\infty.
\label{eqn:SC06_17}
\end{align}
For each $\mathcal{M}_i$, we can find a point $x_i$ such that $\mathbf{cr}(x_i,0) \leq \frac{3}{2} \mathbf{cr}(\mathcal{M}_i^{0})$ by definition. So we have
\begin{align}
\lim_{i \to \infty} \mathbf{cr}(x_i,0) \leq \frac{3}{2}H<\infty. \label{eqn:SC04_3}
\end{align}
Note that $T_i \to \infty$, so the first property holds trivially on the scale $2H$ for large $i$.
By Propositions listed before, we see that there exists an $N=N(H)$ such that for every $i>N$, we have volume ratio estimate, regularity estimate,
density estimate and connectivity estimate hold on each scale $\rho \in (0,2H]$. Therefore, by definition, we obtain that
$\displaystyle \lim_{i \to \infty} \mathbf{cr}(x_i,0) \geq 2H$, which contradicts (\ref{eqn:SC04_3}).
\end{proof}
\begin{corollary}[\textbf{Weak continuity of canonical volume radius}]
Same conditions as in Theorem~\ref{thm:SC04_1}. Then we have
$\displaystyle \mathbf{vr}(\bar{x})=\lim_{i \to \infty} \mathbf{cvr}(x_i)$.
\label{cly:SL27_1}
\end{corollary}
\begin{proof}
It follows from the combination of Proposition~\ref{prn:SC06_5} and Theorem~\ref{thm:SC03_1}.
\end{proof}
\begin{theorem}[\textbf{Weak continuity of polarized canonical radius}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;0.5)$ satisfies (\ref{eqn:SL06_1}).
Then $\mathbf{pcr}(\mathcal{M}_i^{0}) \geq 1$ for $i$ large enough.
\label{thm:SK27_2}
\end{theorem}
\begin{proof}
It follows from the combination of Theorem~\ref{thm:SC04_1}, Theorem~\ref{thm:SC03_1} (for the case $r_0=0.5$) and Corollary~\ref{cly:SK27_1}.
\end{proof}
\subsection{A priori bound of polarized canonical radius}
We shall use a maximum principle type argument to show that the polarized canonical radius cannot be too small.
The technique used in the following proof is inspired by the proof of Theorem 12.1 of~\cite{Pe1}.
\begin{proposition}[\textbf{A priori bound of $\textbf{pcr}$}]
There is a uniform integer constant $j_0=j_0(n,A)$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, then
\begin{align}
\mathbf{pcr}(\mathcal{M}^{t}) \geq \frac{1}{j_0}
\label{eqn:SB27_4}
\end{align}
for every $t \in [-1, 1]$.
\label{prn:SC28_2}
\end{proposition}
\begin{proof}
Suppose for some positive integer $j_0$, (\ref{eqn:SB27_4}) fails at time $t_0 \in [-1,1]$.
Then we check whether $\mathbf{pcr}(\mathcal{M}^t) \geq \frac{1}{2j_0}$ on the interval $[t_0-\frac{1}{2j_0}, t_0+\frac{1}{2j_0}]$.
If so, stop. Otherwise, choose $t_1$ to be such a time and continue to check if $\mathbf{pcr}(\mathcal{M}^t) \geq \frac{1}{4j_0}$
on the interval $[t_1-\frac{1}{4j_0}, t_1+\frac{1}{4j_0}]$. In each step, we shrink the scale to one half of the scale in the previous step.
Note this process will never escape the time interval $[-2,2]$ since
\begin{align*}
|t_k-t_0|<\frac{1}{j_0} \left( \frac{1}{2} + \frac{1}{4} + \cdot +\frac{1}{2^k}\right)<\frac{1}{j_0}<1, \quad |t_k|<|t_0|+1 \leq 2.
\end{align*}
By compactness of the underlying manifold,
it is clear that the process stops after finite steps. So we can find $t_k$ such that $\frac{1}{2^{k+1}j_0} \leq \mathbf{pcr}(\mathcal{M}^{t_k}) < \frac{1}{2^kj_0}$ and
$\mathbf{pcr}(\mathcal{M}^{t}) \geq \frac{1}{2^{k+1}j_0}$ for every
$t \in [t_k-\frac{1}{2^{k+1} j_0}, t_k + \frac{1}{2^{k+1} j_0}]$.
Translate the flow and rescale by constant $4^{k}j_0^2$,
we obtain a new polarized K\"ahler Ricci flow $\widetilde{\mathcal{LM}} \in \mathscr{K}(n,A)$ such that
\begin{align}
\begin{cases}
&\mathbf{pcr}(\widetilde{\mathcal{M}}^0) <1, \\
&\mathbf{pcr}(\widetilde{\mathcal{M}}^{t}) \geq \frac{1}{2}, \quad \forall \; t \in [-2^{k-1}j_0,2^{k-1}j_0], \\
&|R|+|\lambda|<\frac{A}{4^{k}j_0^2}<\frac{A}{j_0^2}, \quad \textrm{on} \; \widetilde{\mathcal{M}}, \\
&\frac{1}{T}+\frac{1}{\Vol(M)}<\frac{1}{2^{k-1}j_0} + \frac{A}{j_0^2}, \quad \textrm{on} \; \widetilde{\mathcal{M}}.
\end{cases}
\label{eqn:SL04_3}
\end{align}
In other words, $\widetilde{\mathcal{LM}} \in \mathscr{K}(n,A;0.5)$ and $|R|+|\lambda|+\frac{1}{T}+\frac{1}{\Vol(M)}$ very small.
Now we return to the main proof. If the statement fails, after adjusting, translating and rescaling, we can find a sequence of polarized
K\"ahler Ricci flow $\widetilde{\mathcal{LM}}_i \in \mathscr{K}(n,A;0.5)$ satisfying
\begin{align*}
\begin{cases}
&\mathbf{pcr}\left( \mathcal{M}_i^{0} \right) <1, \\
&\frac{1}{T_i}+\frac{1}{\Vol(M_i)}+\sup_{\widetilde{\mathcal{M}}_i} (|R|+|\lambda|) \to 0,
\end{cases}
\end{align*}
which contradicts Theorem~\ref{thm:SK27_2}.
\end{proof}
Let $\hslash=\frac{1}{j_0}$. Then we have the following fact.
\begin{theorem}[\textbf{Homogeneity on small scales}]
For some small positive number $\hslash=\hslash(n,A)$, we have
\begin{align}
\mathscr{K}(n,A)=\mathscr{K}\left(n,A;\hslash \right). \label{eqn:SL04_4}
\end{align}
\label{thm:SC28_1}
\end{theorem}
\section{Structure of polarized K\"ahler Ricci flows in $\mathscr{K}(n,A)$}
Because of Theorem~\ref{thm:SC28_1}, $\mathscr{K}(n,A)=\mathscr{K}\left(n,A;\hslash \right)$.
We do have a uniform lower bound for polarized canonical radius.
\subsection{Local metric structure, flow structure, and line bundle structure}
The purpose of this subsection is to set up estimates related to the local metric structure, flow structure and
line bundle structure of every flow in $\mathscr{K}(n,A)$. In particular, we shall prove
Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1}.
\begin{proposition}[\textbf{K\"ahler tangent cone}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ is a sequence of polarized K\"ahler Ricci flows.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i,x_i, g_i(0))$.
Then for each $\bar{y} \in \bar{M}$, every tangent space of $\bar{M}$ at $\bar{y}$ is an irreducible metric cone.
Moreover, this metric cone can be extended as an eternal, possibly singular Ricci flow solution.
\label{prn:HA07_2}
\end{proposition}
\begin{proof}
It follows from Theorem~\ref{thm:SC28_1} and Theorem~\ref{thm:SC09_1} that every tangent space is an irreducible metric cone. From the proof of Theorem~\ref{thm:SC09_1}, it is clear that the tangent cone can be extended as
an eternal, static Ricci flow solution.
\end{proof}
\begin{proposition}[\textbf{Regularity equivalence}]
Same conditions as in Proposition~\ref{prn:HA07_2}, $\bar{y} \in \bar{M}$. Then the following statements are equivalent.
\begin{enumerate}
\item One tangent space of $\bar{y}$ is $\ensuremath{\mathbb{C}}^n$.
\item Every tangent space of $\bar{y}$ is $\ensuremath{\mathbb{C}}^n$.
\item $\bar{y}$ has a neighborhood with $C^{4}$-manifold structure.
\item $\bar{y}$ has a neighborhood with $C^{\infty}$-manifold structure.
\item $\bar{y}$ has a neighborhood with $C^{\omega}$-manifold (real analytic manifold) structure.
\end{enumerate}
\label{prn:HA08_1}
\end{proposition}
\begin{proof}
It is obvious that $5 \Rightarrow 4 \Rightarrow 3 \Rightarrow 2 \Rightarrow 1$. So it suffices to show $1 \Rightarrow 5$
to close the circle. Suppose $\bar{y}$ has a tangent space which is isometric to $\ensuremath{\mathbb{C}}^n$. So we can find a sequence $r_k \to 0$ such that
\begin{align*}
(\bar{M}, \bar{y}, r_k^{-2}\bar{g}) \stackrel{G.H.}{\longrightarrow} (\ensuremath{\mathbb{C}}^n, 0, g_{Euc}).
\end{align*}
So for large $k$, the unit ball $B_{r_k^{-2}\bar{g}}(\bar{y},1)$ has volume ratio almost the Euclidean one.
Fix such a large $k$, we see that $B_{\bar{g}}(\bar{y}, r_k)$ has almost Euclidean volume ratio. It follows from volume convergence that $\mathbf{cvr}(y_i,0) \geq r_k$ for large $i$, where $y_i \in M_i$ and $y_i \to \bar{y}$ as
$(M_i,x_i,g_i(0))$ converges to $(\bar{M}, \bar{x}, \bar{g})$. By the regularity improving property of
canonical volume radius, there is a uniform small constant $c$ such that $B(y_i,cr_k)$ is diffeomorphic to the
same radius Euclidean ball in $\ensuremath{\mathbb{C}}^n$ and the metrics on $B(y_i, c r_k)$ is $C^2$-close to the Euclidean metric.
Then one can apply the backward pseudolocality(c.f. Theorem~\ref{thm:SL27_2}) to obtain higher order derivative estimate for the metrics. Therefore, $B(y_i, \frac{1}{2}c r_k)$ will converge in smooth topology to a limit smooth geodesic ball $B(\bar{y}, \frac{1}{2}cr_k)$. Moreover, it is clear that geometry is uniformly bounded in a space-time neighborhood containing
$B(y_i, \frac{1}{2} cr_k) \times [-c^2 r_k^2, 0]$, by shrinking $c$ if necessary. So we obtain a limit K\"ahler Ricci flow
solution on $B(\bar{y}, \frac{1}{4}c r_k) \times [-\frac{1}{4}c^2 r_k^2, 0]$.
It follows from the result of Kotschwar(c.f.~\cite{Kotsch}), that $B(\bar{y}, \frac{1}{4}cr_k)$ is actually an analytic manifold, which is the desired neighborhood of $\bar{y}$.
So we finish the proof of $1 \Rightarrow 5$ and close the circle.
\end{proof}
\begin{remark}
By Proposition~\ref{prn:HA08_1}, our initial non-classical definition of regularity is proved to be the same as the classical
one (c.f.~Remark~\ref{rmk:HA01_1}).
\label{rmk:HA07_1}
\end{remark}
\begin{proposition}[\textbf{Volume density gap}]
Same conditions as in Proposition~\ref{prn:HA07_2}, $\bar{y} \in \bar{M}$.
Then $\bar{y}$ is singular if and only if
\begin{align}
\limsup_{r \to 0} \frac{|B(\bar{y}, r)|}{\omega_{2n}r^{2n}} \leq 1-2\delta_0.
\label{eqn:HA07_3}
\end{align}
\label{prn:HA07_1}
\end{proposition}
\begin{proof}
If (\ref{eqn:HA07_3}) holds, then every tangent cone of $\bar{y}$ cannot be $\ensuremath{\mathbb{C}}^n$, so $\bar{y}$ is singular.
If $\bar{y}$ is singular, then every tangent space of $\bar{y}$ is an irreducible metric cone in the model space
$\widetilde{\mathscr{KS}}(n,\kappa)$ with vertex a singular point, it follows from the gap property of
$\widetilde{\mathscr{KS}}(n,\kappa)$ that asymptotic volume ratio of such a metric cone must be at most $1-2\delta_0$.
Then (\ref{eqn:HA07_3}) follows from the volume convergence and a scaling argument.
\end{proof}
\begin{proposition}[\textbf{Regular-Singular decomposition}]
Same conditions as in Proposition~\ref{prn:HA07_2}, $\bar{M}$ has the regular-singular decomposition
$\bar{M}=\mathcal{R} \cup \mathcal{S}$. Then the regular part $\mathcal{R}$ admits a natural K\"ahler structure $\bar{J}$.
The singular part $\mathcal{S}$ satisfies the estimate $\dim_{\mathcal{H}} \mathcal{S} \leq 2n-4$.
\label{prn:HA08_2}
\end{proposition}
\begin{proof}
The existence of $\bar{J}$ on $\mathcal{R}$ follows from smooth convergence,
due to the backward pseudolocality(c.f. Theorem~\ref{thm:SL27_2}) and Shi's estimate.
The Hausdorff dimension estimate of $\mathcal{S}$ follows from the
combination of Proposition~\ref{prn:SL15_1} and Theorem~\ref{thm:SC28_1}.
\end{proof}
Therefore, Theorem~\ref{thmin:SC24_1} follows from
the combinations from Proposition~\ref{prn:HA07_2} to Proposition~\ref{prn:HA08_2}.
Now we are going to discuss more delicate properties of the moduli space
$\widetilde{\mathscr{K}}(n,A)$.
\begin{proposition}[\textbf{Improve regularity in two time directions}]
There is a small positive constant $c=c(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$. Let $r_0=\min\{\mathbf{cvr}(x_0,0), 1\}$. Then we have
\begin{align*}
r^{2+k}|\nabla^k Rm|(x,t) \leq \frac{C_k}{c^{2+k}}, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{+}, \quad x \in B_{g(0)}(x_0, cr_0),
\quad t \in [-c^2 r^2, c^2 r^2],
\end{align*}
where $C_k$ is a constant depending on $n,A$ and $k$.
\label{prn:HA08_3}
\end{proposition}
\begin{proof}
Otherwise, there exists a fixed positive integer $k_0$ and a sequence of $c_i \to 0$ such that
\begin{align}
(c_i r_i)^{2+k_0}|\nabla^{k_0} Rm|(y_i,t_i) \to \infty
\label{eqn:HA08_4}
\end{align}
for some $y_i \in B_{g_i(0)}(x_i, r_i)$, $t_i \in [-c_i r_i^2, c_i r_i^2]$, where $r_i=\min\{\mathbf{cvr}(x_i,0), 1\}$.
Let
$\tilde{g}_i(t)=(c_i r_i)^{-2}g_i((c_ir_i)^2t+t_i)$.
Then we have $\mathbf{cvr}_{\tilde{g}_i}(y_i, 0)=(c_i r_i)^{-1} \to \infty$. Note that
$\mathbf{pcr}_{\tilde{g}_i}(y_i,0) \geq \min\{\hslash (c_i r_i)^{-1}, 1\} \geq 1$.
It is also clear that for the flows $\tilde{g}_i$, $|R|+|\lambda| \to 0$.
Therefore, Proposition~\ref{prn:SL04_1} can be applied to obtain
\begin{align}
(M_i, y_i, \tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{M},\hat{y},\hat{g}).
\label{eqn:HA08_5}
\end{align}
However, it follows from Theorem~\ref{thm:SC04_1} and Corollary~\ref{cly:SL27_1} that
\begin{align*}
(\hat{M},\hat{y},\hat{g}) \in \widetilde{\mathscr{KS}}(n,\kappa), \quad
\mathbf{cvr}(\hat{y})=\infty.
\end{align*}
In light of the gap property, Proposition~\ref{prn:SC17_1},
we know that $\hat{M}$ is isometric to $\ensuremath{\mathbb{C}}^n$.
So the convergence (\ref{eqn:HA08_5}) can be rewritten as
\begin{align*}
(M_i, y_i, \tilde{g}_i(0)) \stackrel{C^{\infty}}{\longrightarrow} (\ensuremath{\mathbb{C}}^n, 0, g_{Euc}).
\end{align*}
In particular, $|\nabla^{k_0} Rm|_{\tilde{g}_i}(y_i,0) \to 0$, which is the same as
\begin{align*}
(c_i r_i)^{2+k_0}|\nabla^{k_0} Rm|(y_i,t_i) \to 0.
\end{align*}
This contradicts the assumption (\ref{eqn:HA08_4}).
\end{proof}
Perelman's pseudolocality theorem says that an almost Euclidean domain cannot become very singular in a short time.
His almost Euclidean condition is explained as isoperimetric constant close to that of the Euclidean one. In our special setting,
we can reverse this theorem, i.e., an almost Euclidean domain cannot become very singular in the reverse time direction for a short time period.
\begin{theorem}[\textbf{Two-sided pseudolocality}]
There is a small positive constant $\xi=\xi(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$.
Let $\Omega=B_{g(0)}(x_0,r)$, $\Omega'=B_{g(0)}(x_0, \frac{r}{2})$ for some $0<r\leq 1$.
Suppose $\mathbf{I}(\Omega) \geq (1-\delta_0) \mathbf{I}(\ensuremath{\mathbb{C}}^n)$ at time $t=0$, then
\begin{align*}
(\xi r)^{2+k}|\nabla^k Rm|(x,t) \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0}, \quad x \in \Omega',
\quad t \in [-\xi^2 r^2, \xi^2 r^2],
\end{align*}
where $C_k$ is a constant depending on $n,A$ and $k$.
\label{thm:SL27_2}
\end{theorem}
\begin{proof}
Note that each geodesic ball contained in $\Omega$ has volume ratio at least $(1-\delta_0)\omega_{2n}$.
Then the theorem follows from directly from Proposition~\ref{prn:HA08_3}.
\end{proof}
After we obtain the bound of geometry, we can go further to study the evolution of potential functions.
\begin{theorem}[\textbf{Two-sided pseudolocality on the potential level}]
Same conditions as in Theorem~\ref{thm:SL27_2}.
Let $\omega_B$ be a smooth metric form in $2\pi c_1(M,J)$ and denote
$\omega_t$ by $\omega_B + \sqrt{-1} \partial \bar{\partial} \varphi(\cdot, t)$.
Suppose $\varphi(x_0,0)=0$ and $Osc_{\Omega} \varphi(\cdot, 0) \leq H$.
Let $\Omega''=B_{g(0)}(x_0,\frac{r}{4})$. Then we have
\begin{align}
(\xi r)^{-2+k}\norm{\varphi(\cdot, t)}{C^k(\Omega'', \omega_t)} \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0},
\quad t \in \left[-\frac{\xi^2}{2} r^2, \frac{\xi^2}{2} r^2 \right],
\label{eqn:HA05_2}
\end{align}
where $C_k$ depends on $k,n,A,\xi$ and $\frac{H}{r^2}$.
\label{thm:HA03_5}
\end{theorem}
\begin{proof}
Up to rescaling, we may assume $\xi r=1$.
Note that $\varphi$ and $\dot{\varphi}$ satisfy the equations
\begin{align*}
\begin{cases}
&\dot{\varphi}=\log \frac{\omega_t^n}{\omega_B^n}+\varphi +\dot{\varphi}(\cdot, 0), \\
&-\sqrt{-1} \partial \bar{\partial} \dot{\varphi}=Ric-\lambda g.
\end{cases}
\end{align*}
It follows from Theorem~\ref{thm:SL27_2} that geometry is uniformly bounded in $\Omega' \times [-\xi r^2, \xi r^2]$.
The trace form of the second equation in the above list is $-\Delta \dot{\varphi}=R-n\lambda$.
Therefore, the regularity theory
of Laplacian operator applies and we have uniform bound of $\norm{\dot{\varphi}}{C^k}$ in a neighborhood of
$\Omega'' \times [-\frac{\xi}{2}r^2, \frac{\xi}{2}r^2]$.
Up to a normalization, we can rewrite the first equation as
\begin{align*}
\log \frac{(\omega_t-\sqrt{-1}\partial \bar{\partial} \varphi)^n}{\omega_t^n}=\varphi - \dot{\varphi}+\dot{\varphi}(\cdot, 0).
\end{align*}
On $\Omega'$, the metric $g(0)$ and $g(t)$ are uniformly equivalent in each $C^k$-topology. So it is clear that
$\norm{\dot{\varphi}-\dot{\varphi}(\cdot, 0)}{C^k(\Omega')}$ are uniformly bounded, for each $k$, with respect to
metric $g(t)$.
Since all higher derivatives of curvature are uniformly bounded on $\Omega'$, (\ref{eqn:HA05_2})
follows from standard Monge-Ampere equation theory and bootstrapping argument.
\end{proof}
\begin{theorem}[\textbf{Improving regularity of potentials}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $\mathbf{cvr}(M,0)=r_0$.
Let $\omega_B$ be a smooth metric in $[\omega_0]$ such that
\begin{align}
\frac{1}{2}\omega_B \leq \omega_0 \leq 2\omega_B.
\label{eqn:HA06_1}
\end{align}
Let $\omega_0=\omega_B + \sqrt{-1} \partial \bar{\partial} \varphi$.
Suppose $\int_M \varphi \omega_0^n=0$ and $Osc_M \varphi \leq H$. Then we have
\begin{align}
\norm{\varphi}{C^k(M,\omega_B)} \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0},
\label{eqn:HA06_2}
\end{align}
where $C_k$ depends on $k,\omega_B, n,A,r_0$ and $H$.
\label{thm:HA06_1}
\end{theorem}
\begin{proof}
Since $\mathbf{cvr}(M,0)=r_0>0$, we see that all the possible $\omega_0$'s form a compact set under the smooth topology.
In other words, $\omega_0$ has uniformly bounded geometry in each regularity level. Fix a positive integer $k_0 \geq 4$.
Therefore, around each point $x \in M$, one can find a coordinate chart $\Omega$, with uniform size, such that
\begin{align*}
\omega_0=\omega_{Euc} + \sqrt{-1} \partial \bar{\partial} f, \quad
\norm{f}{C^{k_0}(\Omega, \omega_{Euc})} \leq 0.01.
\end{align*}
Note that in $\Omega$, the connection terms of the metric $\omega_0$ are pure derivatives $f_{i\bar{j}l}$,
which are uniformly bounded. Similarly, all derivatives of connection terms can be expressed as high order
pure derivatives of $f$. Therefore, up to order $k_0-3$, the derivatives of connections are uniformly bounded.
It is clear that the metric $\omega_0$ and $\omega_{Euc}$ are uniformly equivalent.
By the covariant derivatives' bounds $\norm{\varphi}{C^k(M, \omega_0)} \leq C_k$,
the bounds of connection derivatives yield that
\begin{align}
\norm{\varphi}{C^k(\Omega, \omega_{Euc})} \leq C_k, \quad \forall \; 0 \leq k \leq k_0-1.
\label{eqn:HA08_3}
\end{align}
In other words, we have uniform bound for every order pure derivatives of $\varphi$, up to order $k_0-1$.
Together with the choice assumption of $\Omega$, we have
\begin{align*}
\norm{f-\varphi}{C^k(\Omega, \omega_{Euc})} \leq C_k, \quad \forall \; 0 \leq k \leq k_0-1.
\end{align*}
Therefore, the connection derivatives of metric $\omega_B$ in $\Omega$ are uniformly bounded, up to order $k_0-4$.
Consequently, the pure derivative bound (\ref{eqn:HA08_3}) implies
\begin{align*}
\norm{\varphi}{C^k(\Omega, \omega_B)} \leq C_k, \quad \forall \; 0 \leq k \leq k_0-1,
\end{align*}
since $\omega_B$ is a fixed smooth, compact metric with every level of regularity. Clearly, the above constant $C_k$ depends on
$k,n,A,r_0,\omega_B$ and $H$. Recall that the size of $\Omega$ is uniformly bounded from below, $(M, \omega_B)$ is a compact
manifold. Consequently, a standard covering argument implies (\ref{eqn:HA06_2}) for each $k \leq k_0-1$.
In the end, we free $k_0$ and finish the proof.
\end{proof}
In Ricci-flat theory, a version of Anderson's gap theorem says that regularity can be improved in the center of a ball
if the volume ratio of the unit ball is very close to the Euclidean one.
In our special setting, this gap theorem has a reduced volume version.
\begin{theorem}[\textbf{Gap of reduced volume}]
There is a constant $\delta_0' \in (0, \delta_0]$ and a small constant $\eta$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$, $0<r \leq 1$.
If $\mathcal{V}((x_0,0),r^2) \geq 1-\delta_0'$, then we have
\begin{align}
\mathbf{cvr}(x_0,0) \geq \eta r.
\label{eqn:SL27_2}
\end{align}
\label{thm:SL27_3}
\end{theorem}
\begin{proof}
If $\lambda=0$, reduced volume is monotone. If $\lambda$ is bounded, then reduced volume is almost monotone.
A simple calculation shows that $\mathcal{V}((x_0,0), \rho^2) \geq 1-\delta_0$ for all $0<\rho \leq r^2$
whenever $\mathcal{V}((x_0,0),r^2) \geq 1-\delta_0'$ for some $0<r\leq 1$. Therefore, without loss of generality,
we may assume $\lambda=0$ and $\delta_0'=\delta_0$ in the proof.
If the statement was wrong, there exists a sequence of $\eta_i \to 0$, $0<r_i \leq 1$ and $x_i \in M_i$, and corresponding K\"ahler Ricci flows satisfying
\begin{align*}
\begin{cases}
&\mathcal{V}((x_i,0),r_i^2) \geq 1-\delta_0, \\
&\mathbf{cvr}(x_i,0) <\eta_i r_i.
\end{cases}
\end{align*}
By the monotonicity of reduced volume, we have
\begin{align*}
\begin{cases}
&\mathcal{V}((x_i,0),H\eta_i^2 r_i^2) \geq 1-\delta_0, \\
&\mathbf{cvr}(x_i,0) <\eta_i r_i,
\end{cases}
\end{align*}
for each fixed $H$ and large $i$.
Let $\tilde{g}_i(t)=(\eta_i r_i)^{-2}g((\eta_i r_i)^2t)$. It is clear that
\begin{align}
\mathbf{cvr}_{\tilde{g}_i}(x_i,0)=1. \label{eqn:HA09_1}
\end{align}
The canonical radius of $\tilde{g}_i$ tends to infinity, $|R|+|\lambda| \to 0$.
Similar to the proof of Proposition~\ref{prn:HA08_3}, we have the convergence:
\begin{align*}
(M_i, x_i, \tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{M}, \hat{x}, \hat{g}) \in \widetilde{\mathscr{KS}}(n,\kappa).
\end{align*}
The limit space $\hat{M}$ can be extended to a static eternal K\"ahler Ricci flow solution.
Moreover, Proposition~\ref{prn:SL14_2} can be applied here and guarantees the reduced volume convergence.
\begin{align*}
\mathcal{V}((\hat{x},0), H)=\lim_{i \to \infty} \mathcal{V}_{\tilde{g}_i}((x_i,0), H)
=\lim_{i \to \infty} \mathcal{V}((x_i,0), H(\eta_i r_i)^2) \geq 1-\delta_0.
\end{align*}
Note that $H$ is arbitrary. By the homogeneity
of reduced volume at infinity, Theorem~\ref{thm:SL25_1}, we see that
\begin{align*}
\mathrm{avr}(\hat{M})=\lim_{H \to \infty} \mathcal{V}((\bar{x},0),H) \geq 1-\delta_0 \geq 1-\delta_0.
\end{align*}
So Proposition~\ref{prn:SC17_1} applies to force $\hat{M}$ to be isometric to be $\ensuremath{\mathbb{C}}^n$.
In particular, $\mathbf{vr}(\hat{x})=\infty$.
It follows from Corollary~\ref{cly:SL27_1} that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}_{\tilde{g}_i}(x_i,0)=\infty$,
which contradicts (\ref{eqn:HA09_1}).
\end{proof}
According to Theorem~\ref{thm:SL27_3}, one can define a concept of reduced volume radius for the purpose of
improving regularity. Clearly, other regularity radius can also be defined. However, it seems all of them are equivalent.
For simplicity, we shall not compare all of them, but only prove an example case:
the equivalence of harmonic radius and canonical volume radius. The proof of other cases are verbatim.
\begin{proposition}[\textbf{Equivalence of regularity radii}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$.
Suppose $\max\{\mathbf{hr}(x,0), \mathbf{cvr}(x,0)\} \leq 1$,
then we have
\begin{align*}
\frac{1}{C} \mathbf{hr}(x,0) \leq \mathbf{cvr}(x,0) \leq C \mathbf{hr}(x,0)
\end{align*}
for some uniform constant $C=C(n,A)$.
\label{prn:HA09_2}
\end{proposition}
\begin{proof}
Clearly, $\mathbf{cvr}(x,0) \leq C \mathbf{hr}(x,0)$ follows from the $C^5$-regularity property of canonical volume radius. It suffices to show $\frac{1}{C} \mathbf{hr}(x,0) \leq \mathbf{cvr}(x,0)$. However, since $\mathbf{cr}(x,0) \geq \hslash$,
it is clear from definition that
\begin{align*}
\mathbf{cvr}(x,0) \geq \frac{1}{C} \min\{\mathbf{hr}(x,0), \hslash\}.
\end{align*}
If $\mathbf{hr}(x,0)\leq \hslash$, then we are done. Otherwise, we have $\hslash <\mathbf{hr}(x,0) \leq 1$. It follows that
\begin{align*}
\mathbf{cvr}(x,0) \geq \frac{1}{C}\hslash \geq \frac{\hslash}{C} \mathbf{hr}(x,0) \geq \frac{1}{C'} \mathbf{hr}(x,0).
\end{align*}
So we finish the proof.
\end{proof}
\begin{theorem}[\textbf{Improved density estimate}]
For arbitrary small $\epsilon$, arbitrary $0 \leq p<2$, there is a constant $\delta=\delta(n,A,p)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$. Then under the metric $g(0)$, we have
\begin{align}
\log \frac{\int_{B(x,r)} \mathbf{cvr}^{-2p}dv}{E(n,\kappa,p) r^{2n-2p} } < \epsilon
\label{eqn:SC13_2}
\end{align}
whenever $r<\delta$.
Here the number $E(n,\kappa,p)$ is defined in Proposition~\ref{prn:SB25_2}.
\label{thm:SC12_2}
\end{theorem}
\begin{proof}
We argue by contradiction.
Note that every blowup limit is in $\widetilde{\mathscr{KS}}(n,\kappa)$(c.f. Theorem~\ref{thm:SC04_1}).
Then a contradiction can be obtained by the weak continuity of $\mathbf{cvr}$(c.f. Corollary~\ref{cly:SL27_1})
if the statement of this theorem does not hold.
\end{proof}
Note that $E(n,\kappa,0)=\omega_{2n}$. So we are led to the volume ratio estimate immediately.
\begin{corollary}[\textbf{Volume-ratio estimate}]
For arbitrary small $\epsilon$, there is a constant $\delta=\delta(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$. Then under the metric $g(0)$, we have
\begin{align}
\log \frac{|B(x,r)|}{\omega_{2n} r^{2n}} < \epsilon
\label{eqn:SC13_1}
\end{align}
whenever $r<\delta$.
\label{cly:SC13_1}
\end{corollary}
In the K\"ahler Ricci flow setting, Corollary~\ref{cly:SC13_1} improves the volume ratio estimates in~\cite{Zhq3}
and~\cite{CW5} (c.f.~Remark 1.1 of \cite{CW5}). Note that the integral (\ref{eqn:SC13_2}) can be used to show that
for every $p \in (0,2)$, there is a $C=C(n,A,p)$ such that the $r$-neighborhood of $\mathcal{S}$ in a unit ball is
bounded by $C r^{2p}$(c.f. Theorem~\ref{thm:HE11_1}), where $\mathcal{S}$ is the singular part of a limit space.
By the definition of Minkowski dimension(c.f.~Definition~\ref{dfn:HE08_1}),
we can improve Proposition~\ref{prn:HA08_2} as follows.
\begin{corollary}[\textbf{Minkowski dimension of singular set}]
Same conditions as in Proposition~\ref{prn:HA07_2}, $\bar{M}$ has the regular-singular decomposition
$\bar{M}=\mathcal{R} \cup \mathcal{S}$. Then $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$.
\label{cly:HE25_2}
\end{corollary}
In \cite{Wa2}, the second author developed an estimate of the type $|Ric| \leq \sqrt{|Rm||R|}$, where $\sqrt{|Rm|}$ should be understood as the reciprocal of a regular scale. Due to the improving regularity property of canonical volume radius, it induces the estimate
$|Ric| \leq \frac{\sqrt{|R|}}{\mathbf{cvr}}$ pointwisely.
By the uniform bound of scalar curvature and Theorem~\ref{thm:SC12_2}, the following estimate is clear now.
\begin{corollary}[\textbf{Ricci curvature estimate}]
There is a constant $C=C(n,A,r_0)$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$, $0<r \leq r_0$, $0<p<2$.
Then under the metric $g(0)$, we have
\begin{align}
r^{2p-2n}\int_{B(x_0,r)} |Ric|^{2p} dv<C.
\label{eqn:SL27_3}
\end{align}
\label{cly:SL27_4}
\end{corollary}
Corollary~\ref{cly:SL27_4} localizes the $L^{2p}$-curvature estimate of \cite{TZZ2} in a weak sense, since
(\ref{eqn:SL27_3}) only holds for $p<2$. If $n=2$, (\ref{eqn:SL27_3}) also holds for $p=2$,
since the finiteness of singularity guarantees that one can choose good cutoff functions.
We believe that the same localization result hold for $p=2$ even if $n>2$.\\
We return to the canonical neighborhood theorems in the introduction,
Theorem~\ref{thmin:SC24_1}, Theorem~\ref{thmin:HC08_1} and Theorem~\ref{thmin:HC06_1}.
However, Theorem~\ref{thmin:HC06_1} is not completely local.
Actually, Theorem~\ref{thm:SL27_2} is enough to show
the local flow structure of $\mathscr{K}(n,A)$ can be approximated by $\mathscr{KS}(n,\kappa)$.
In light of its global properties, the proof of Theorem~\ref{thmin:HC06_1} is harder and is postponed to section 5.5.
On the other hand, Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1} are local.
We now close this subsection by proving Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1}.
\begin{proof}[Proof of Theorem~\ref{thmin:SC24_1}]
It follows from the combination of Proposition~\ref{prn:HA07_2}, Proposition~\ref{prn:HA08_1}, Proposition~\ref{prn:HA07_1},
and Proposition~\ref{prn:HA08_2}.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thmin:HC08_1}]
It follows from Theorem~\ref{thm:SC28_1}, Definition~\ref{dfn:SK27_1} and a scaling argument.
\end{proof}
\subsection{Local variety structure}
We focus on the variety structure of the limit space in this subsection.
We essentially follow the argument in~\cite{DS}, with slight modification.
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Let $(\bar{M}, \bar{x},\bar{g})$
be a pointed-Gromov-Hausdorff limit of $(M_i,x_i,g_i(0))$.
Since $\bar{M}$ may be non-compact, the limit line bundle $\bar{L}$ may have infinitely many orthogonal holomorphic sections.
Therefore, in general, we cannot expect to embed $\bar{M}$ into a projective space of finite dimension
by the complete linear system of $\bar{L}$. However, when
we focus our attention to the unit geodesic ball $B(\bar{x},1)$,
we can choose some holomorphic sections of $\bar{L}$, peaked around $\bar{x}$, to embed $B(\bar{x},1)$ into $\ensuremath{\mathbb{CP}}^N$ for a
finite $N$.
Actually, for every $\epsilon>0$, we can find an $\epsilon$-net of $B(\bar{x},2)$ such that every point in this net has canonical volume radius
at least $c_0 \epsilon$. For each point $y$ in this $\epsilon$-net, we have a peak section $s_y$, which is a holomorphic section such that
$\norm{s(y)}{}$ achieves the maximum among all unit $L^2$-norm holomorphic sections $s \in H^0(\bar{M},\bar{L})$. By the partial-$C^0$-estimate
argument(c.f.~\cite{CW4} for the flow case with weak convergence), we can assume that $\norm{s_y}{}^2$ is uniformly bounded below in $B(y,2\epsilon)$.
On the other hand, by the choice of $y$, $B(y, \eta \epsilon)$ has a smooth manifold structure for some $\eta=\eta(n)$.
Therefore, we can choose $n$ holomorphic sections of $\bar{L}^k$ such that these sections are the local deformation of $ z_1, z_2, \cdots, z_n$.
Here $k$ is a positive integer proportional to $\epsilon^{-2}$. Put these holomorphic sections together with $s_y^k$,
we obtain $(n+1)$-holomorphic sections of $\bar{L}^k$ based at the point $y$. Let $y$ run through all points in the $\epsilon$-net and collect all the holomorphic sections based at $y$, we obtain a set of holomorphic sections $\{s_i\}_{i=0}^N$ of $\bar{L}^{k}$. Let $\{\tilde{s}_i\}_{i=0}^N$
be the orthonormal basis of $span\{s_0,s_1, \cdots, s_N\}$. We define the Kodaira map $\iota$ as follows.
\begin{align*}
\iota: B(0,2) &\mapsto \ensuremath{\mathbb{CP}}^N, \\
x &\mapsto [\tilde{s}_0(x) : \tilde{s}_1(x): \cdots:\tilde{s}_N(x)].
\end{align*}
This map is well defined. In fact, for every $z \in B(\bar{x},1)$, we can find a point $y$ in the $\epsilon$-net and $z \in B(y,2\epsilon)$, then
$\norm{s_y}{}^2(z)>0$ by the partial-$C^0$-estimate. It forces that $\tilde{s}_j(z) \neq 0$ for some $j$.
Since $k$ is proportional to $\epsilon^{-2}$, we can just let $\epsilon=\frac{1}{\sqrt{k}}$ without loss of generality.
In the following argument, by saying ``raise the power of line bundle" from $k_1$ to $k_2$,
we simultaneously means the underlying $\epsilon$-net is
strengthened from a $\frac{1}{\sqrt{k_1}}$-net to a $\frac{1}{\sqrt{k_2}}$-net.
\begin{lemma}
Suppose $w \in \iota(B(\bar{x},1))$, then $ \iota^{-1}(w) \cap \overline{B(\bar{x},1)}$ is a finite set.
\label{lma:HB09_1}
\end{lemma}
\begin{proof}
Let $y \in \iota^{-1}(w) \cap \overline{B(\bar{x},1)}$. It is clear that $\iota^{-1}(w)$ is contained in a ball centered at $y$ with fixed radius, say $10\epsilon$.
Therefore, $\iota^{-1}(w)$ is a bounded, closed set and therefore compact. Let $F$ be a connected component of $\iota^{-1}(w)$. Then
$\iota(F)$ is a connected, compact subvariety of $\ensuremath{\mathbb{C}}^N$, and consequently is a point. Note that $\iota(F)$ is always a connected set no matter how do we raise the power of $\iota$. On the other hand, $\iota(F)$ will contain more than one point if $F$ is not a single point, after we raise power high enough.
These force that $F$ can only be a point. Since $\iota^{-1}(w) \cap \overline{B(\bar{x},1)}$ is compact, it must be union of finite points.
\end{proof}
Denote $\iota(\overline{B(\bar{x},1)})$ by $W$. Then $W$ is a compact set and locally can be extended as an analytic variety.
By dividing $W$ into different components, one can apply induction argument as that in~\cite{DS}.
Following verbatim the argument of Proposition 4.10, Lemma 4.11 of~\cite{DS}, one can show that $\iota$ is an injective, non-degenerate embedding
map on $B(\bar{x},1)$, by raising power of $\bar{L}$ if necessary. Furthermore, since being normal is a local property, one can
improve Lemma 4.12 of~\cite{DS} as follows.
\begin{lemma}
By raising power if necessary, $W$ is normal at the point $\iota(y)$ for every $y \in B(\bar{x}, \frac{1}{2})$.
\end{lemma}
Under the help of parabolic Schwarz lemma and heat flow localization technique(c.f. Section 4.1 and Proposition~\ref{prn:HB05_1}),
we can parallelly generalize Proposition 4.14 of~\cite{DS} as follows.
\begin{lemma}
Suppose $y \in B(\bar{x},\frac{1}{2})\cap \mathcal{S}$, then $\iota(y)$ is a singular point of $W$.
\end{lemma}
It follows from the proof of Proposition 4.15 of~\cite{DS} that there always exist a holomorphic form $\Theta$ on $\mathcal{R} \cap B(\bar{x},1)$ such that
$\int_{\mathcal{R} \cap B(\bar{x},1)} \Theta \wedge \bar{\Theta}<\infty$. This means that every singular point $y \in \iota(B(\bar{x},\frac{1}{2})) \cap W$ is
log-terminal.
Combining all the previous lemmas, we have the following structure theorem.
\begin{theorem}[\textbf{Analytic variety structure}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$,
$(\bar{M}, \bar{x},\bar{g})$ is a pointed Gromov-Hausdorff limit of $(M_i,x_i,g_i(0))$.
Then $\bar{M}$ is an analytic space with normal, log terminal singularities.
\label{thm:HE19_1}
\end{theorem}
\subsection{Distance estimates}
In this subsection, we shall develop the distance estimate along polarized K\"ahler Ricci flow in terms of the estimates
from line bundle.
\begin{lemma}
Suppose $(M,L)$ is a polarized K\"ahler manifold satisfying the following conditions
\begin{itemize}
\item $|B(x,r)| \geq \kappa \omega_{2n} r^{2n}, \quad \forall x \in M, 0<r<1$.
\item $|\mathbf{b}|\leq 2c_0$ where $\mathbf{b}$ is the Bergman function.
\item $\norm{\nabla S}{} \leq C_1$ for every $L^2$-unit section $S \in H^0(M,L)$.
\end{itemize}
For every positive number $a$, define $ \Omega(x,a)$ as the path-connected component containing $x$ of the set
\begin{align}
\left\{z \left| \norm{S}{}^2(z) \geq e^{-2a-2c_0},
\norm{S}{}^2(x)=e^{\mathbf{b}(x)}, \int_M \norm{S}{}^2dv=1 \right.\right\}.
\label{eqn:HA15_2}
\end{align}
Then we have
\begin{align}
B\left(x,r\right) \subset \Omega(x,a) \subset B(x, \rho)
\label{eqn:HA15_1}
\end{align}
for some $r=r(n,\kappa,c_0,C_1,a)$ and $\rho=\rho(n,\kappa,c_0,C_1,a)$.
\label{lma:HA15_1}
\end{lemma}
\begin{proof}
Define $r \triangleq \frac{1-e^{-a}}{C_1 e^{a+c_0}}$.
Recall that $\norm{S}{}(x) \geq e^{-c_0}$. By the gradient bound of $S$, it is clear that every point in
$B(x,r)$ satisfies $\norm{S}{} \geq e^{-a-c_0}$. In other words, we have
\begin{align*}
B(x,r) \subset \Omega(x,a).
\end{align*}
On the other hand, we can cover $\Omega(x,a)$ by finite balls $B(x_i, 2r)$ such that each $x_i \in \Omega(x,a)$
and different $B(x_i,r)$'s are disjoint to each other. Again, the gradient bound of $S$ implies that
$\norm{S}{} \geq e^{-2a-c_0}$ in each $B(x_i,r)$. Then we have
\begin{align*}
N \kappa \omega_{2n} r^{2n} \leq \sum_{i=1}^{N} |B(x_i,r)| \leq |\Omega(x, 2a)| \leq e^{4a+2c_0}.
\end{align*}
For every $z \in \Omega(x,a)$, we have
\begin{align*}
d(x,z) \leq 4Nr \leq \frac{4e^{4a+2c_0}}{\kappa \omega_{2n}r^{2n-1}}
=\frac{4e^{4a+2c_0}}{\kappa \omega_{2n}} \cdot \frac{C_1^{2n-1} e^{(2n-1)(a+c_0)}}{(1-e^{-a})^{2n-1}}.
\end{align*}
Let $\rho$ be the number on the right hand side of the above inequality. Then it is clear that
\begin{align*}
\Omega(x,a) \subset B(x,\rho).
\end{align*}
So we finish the proof.
\end{proof}
Lemma~\ref{lma:HA15_1} implies that the level sets of peak holomorphic sections are comparable to
geodesic balls. However, the norm of peak holomorphic section has stability under the K\"ahler Ricci flows
in $\mathscr{K}(n,A)$. Therefore, one can compare distances at different time slices in terms the values
of norms of a same holomorphic sections.
\begin{lemma}
There exists a small constant $\epsilon_0=\epsilon_0(n,A)$ such that
the following properties are satisfied.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, then we have
\begin{align}
B_{g(t_1)}(x,\epsilon_0) \subset B_{g(t_2)} \left(x,\epsilon_0^{-1} \right) \label{eqn:HA14_1}
\end{align}
whenever $t_1, t_2 \in [-1,1]$.
\label{lma:HA14_1}
\end{lemma}
\begin{proof}
Without loss of generality, we only need to show (\ref{eqn:HA14_1}) for time $t_1=0, t_2=1$.
Because of Theorem~\ref{thmin:HC08_1} and Moser iteration, we can assume $|\mathbf{b}| \leq 2c_0$ for some $c_0=c_0(n,A)$.
By Moser iteration technique, we can also assume $\norm{\nabla S}{} \leq C_1$ for every unit $L^2$-norm holomorphic
section of $L$(c.f. Lemma 5.1 of~\cite{Wa1} and Lemma 3.2 of~\cite{CW4}). Note that $e^{\mathbf{b}(x)}$ is the maximum value of $\norm{S}{}^2$ among all
unit $L^2$-norm holomorphic sections of $L$. So we can choose $\epsilon$ small enough such that
\begin{align*}
\norm{S}{}(z) \geq \frac{1}{2}e^{-c_0}, \quad \forall \; z \in B(x,2\epsilon)
\end{align*}
for some unit holomorphic section $S$. Note $\epsilon$ can be chosen uniformly, say $\epsilon=\frac{e^{-c_0}}{4C_1}$.
Fix $S$ and define
\begin{align*}
\Omega \triangleq \left\{z\left| \norm{S}{0}(z)\geq \frac{1}{2}e^{-c_0} \right. \right\}, \qquad
\tilde{\Omega} \triangleq \left\{z\left| \norm{S}{1}(z) \geq \frac{1}{4}e^{-c_0-A} \right. \right\}.
\end{align*}
Without loss of generality, we can assume both $\Omega$ and $\tilde{\Omega}$ are path-connected. Otherwise, just replace them by the corresponding path-connected part containing $z$.
It follows from definition that $B(x,\epsilon) \subset \Omega$.
In view of the volume element evolution equation,
it is also clear that $\Omega \subset \tilde{\Omega}$.
Note that $S$ is a unit section at time $t=0$.
At time $t=1$, its $L^2$-norm locates in $[e^{-2A}, e^{2A}]$. So we have
\begin{align*}
e^{2A} \geq \int_M \norm{S}{1}^2 dv>|\tilde{\Omega}|_{1} \frac{1}{16}e^{-2c_0-2A}, \quad
\Rightarrow \quad |\tilde{\Omega}|_{1}<16 e^{2c_0+4A}.
\end{align*}
Now we can follow the covering argument in the previous lemma to show a diameter bound of $\tilde{\Omega}$
under the metric $g(1)$. In fact, we can cover $\tilde{\Omega}$ by finite geodesic balls $B(x_i,2\epsilon)$ such that $x_i \in \tilde{\Omega}$
and all different $B(x_i,\epsilon)$'s are disjoint to each other. Clearly, each geodesic ball
$B(x_i,\epsilon)$
has volume at least $\kappa \omega_{2n} \epsilon^{2n}$, where $\kappa=\kappa(n,A)$. Let
$N$ be the number of balls, then
\begin{align*}
N \kappa \omega_{2n} \epsilon^{2n} \leq \sum_{i=1}^{N}|B(x_i,1)| \leq |\tilde{\Omega}| \leq 16e^{2c_0+4A}.
\end{align*}
Therefore, under metric $g(1)$, we obtain
\begin{align*}
\diam \Omega \leq \diam \tilde{\Omega} \leq 4N\epsilon \leq \frac{64e^{2c_0+4A}}{\kappa \omega_{2n}\epsilon^{2n-1}}.
\end{align*}
Recall that $B_{g(0)}(x,\epsilon) \subset \Omega \subset \tilde{\Omega}$, $\epsilon=\frac{e^{-c_0}}{4C_1}$.
So under metric $g(1)$, we have
\begin{align*}
\diam B_{g(0)}(x,\epsilon) \leq \diam \tilde{\Omega} \leq \frac{4^{2n+2}e^{(2n+3)c_0+4A}C_1^{2n-1}}{\kappa \omega_{2n}}.
\end{align*}
Define
\begin{align}
\epsilon_0 \triangleq
\min\left\{\frac{e^{-c_0}}{4C_1}, \frac{\kappa \omega_{2n}}{4^{2n+2}e^{(2n+3)c_0+4A}C_1^{2n-1}} \right\}.
\label{eqn:HA15_3}
\end{align}
Note that $\epsilon_0$ depends only on $n,A$.
Then we have $ \diam B_{g(0)}(x,\epsilon_0) \leq \epsilon_0^{-1}$, which implies
\begin{align*}
B_{g(0)}(x, \epsilon_0) \subset B_{g(1)}(x,\epsilon_0^{-1}).
\end{align*}
So we finish the proof.
\end{proof}
\begin{lemma}
For every $r$ small, there is a $\delta$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Suppose $|R|+|\lambda|<\delta$ on $M \times [-1,1]$,
then we have
\begin{align}
B_{g(t_1)}(x,\epsilon_0 r) \subset B_{g(t_2)}(x,\epsilon_0^{-1}r)
\label{eqn:HA15_4}
\end{align}
for every $t_1,t_2 \in [-1,1]$. Here $\epsilon_0$ is the constant in Lemma~\ref{lma:HA14_1}.
\label{lma:HA14_2}
\end{lemma}
\begin{proof}
We proceed by a contradiction argument.
Again, it suffices to show (\ref{eqn:HA15_4}) for $t_1=0$ and $t_2=1$. By adjusting $r$ if necessary, we can
also make a rescaling by integer factor. Up to rescaling, (\ref{eqn:HA15_4}) is the same as
\begin{align}
B_{g(0)}(x,\epsilon_0) \subset B_{g(r^{-2})}(x, \epsilon_0^{-1}). \label{eqn:HA15_5}
\end{align}
Suppose the statement of this lemma was wrong. Then there is an $r_0>0$ and a sequence of
points $x_i \in M_i$ such that
\begin{align*}
B_{g_i(0)}(x_i,\epsilon_0) \not\subset B_{g_i(r_0^{-2})}(x_i,\epsilon_0^{-1}).
\end{align*}
However, $|R|+|\lambda| \to 0$ in $C^0$-norm as $i \to \infty$. So we can take a limit
\begin{align*}
(M_i,x_i,g_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M},\bar{x},\bar{g}).
\end{align*}
As usual, we can find a regular point $\bar{z} \in \bar{M}$ near $\bar{x}$. Let $z_i \in M_i$ and $z_i \to \bar{z}$
as the above convergence happens. Then we can extend the above convergence to each time slice.
\begin{align*}
(M_i,z_i,g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M},\bar{z},\bar{g}), \quad \forall \; t \in [0,r_0^{-2}].
\end{align*}
Note that $\bar{x}$ may be a singular point of $\bar{M}$. So in the above convergence, we only have $x_i$
converges to $\bar{x}(t)$, which may depends on time $t$. Lemma~\ref{lma:HA14_1} guarantees that $\bar{x}(t)$
is not at infinity.
Note that $Osc_M \dot{\varphi}$ is scaling invariant and consequently uniformly bounded by condition inequality (\ref{eqn:SK20_1}).
Therefore $\dot{\varphi}$ converges to a limit bounded function which is harmonic function on $\mathcal{R}(\bar{M})$, the regular part of $\bar{M}$.
Such a function must be a constant by Corollary~\ref{cly:HE08_1}.
Actually, a bounded harmonic function on $\mathcal{R}(\bar{M})$ will automatically be a bounded Lipschitz function on $\bar{M}$, by Proposition~\ref{prn:HD16_1}.
Applying normalization condition, the limit function must be zero on $\bar{M} \times [0,r_0^{-2}]$. Therefore, the limit line bundle $\bar{L}$
admits a limit metric which does not evolve along time. Therefore, for a fixed holomorphic section $\bar{S}$ and a
fixed level value, the level sets of $\norm{\bar{S}}{}^2$ does not depend on time.
Choose $S_i$ be the peak section of $L_i$ at $x_i$, with respect to the metrics at time $t=0$. By the choice of
$\epsilon_0$, it is clear that $\norm{S_i}{}\geq \frac{1}{2}e^{-c_0}$ on the ball $B(x_i,\epsilon_0)$.
In other words, we have
\begin{align*}
B(x_i, \epsilon_0) \subset
\Omega_{i,t} \triangleq \left\{z\left| \norm{S_i}{t}(z)\geq \frac{1}{2}e^{-c_0} \right. \right\}.
\end{align*}
Without loss of generality, we can assume $\Omega_{i,t}$ is path connected.
Clearly, each $\Omega_{i,t}$ has uniformly bounded diameter, due to Lemma~\ref{lma:HA15_1}.
Let $\bar{\Omega}$ be the limit set of $\Omega_{i,0}$. Clearly, $\bar{z} \in \bar{\Omega}$.
Then the above discussion implies that $\bar{\Omega}$ is actually the limit set of each $\Omega_{i,t}$.
Let $\bar{y}$ be the limit point of $y_i$, which is a point in $B_{g_i(0)}(x_i, \epsilon_0)$ and start to escape
$B(x_i, \epsilon_0^{-1})$ at time $t_i$, which converges to $\bar{t}$. So we obtain
\begin{align}
\bar{y} \in B_{\bar{g}(0)}(\bar{x},\epsilon_0),
\quad
d(\bar{y}, \bar{x}(\bar{t}))=\epsilon_0^{-1}.
\label{eqn:HA15_6}
\end{align}
Since volume element of the underlying manifold and the line bundle metric are all almost static when time evolves,
it is easy to see that $y_i$ can never escape $\Omega_{i,t}$. So $\bar{y} \in \bar{\Omega}$.
Similarly, we know $\bar{x}(\bar{t}) \in \bar{\Omega}$. Therefore, at time $\bar{t}$, we have
\begin{align*}
d(\bar{y}, \bar{x}(\bar{t})) \leq \diam \bar{\Omega}.
\end{align*}
Note that the argument in the proof of Lemma~\ref{lma:HA15_1} holds for the polarized singular manifold
$(\bar{M}, \bar{L})$, due to the high codimension of $\mathcal{S}(M)$ and the gradient bound of each $S_i$.
Since $\int_{\bar{M}} \norm{\bar{S}}{}^2 dv$ is uniformly bounded from above by $1$, we can follow the proof of
Lemma~\ref{lma:HA15_1} to show that
\begin{align*}
\diam(\bar{\Omega}) \leq \rho(n,\kappa,c_0,C_1,\log 2)<\epsilon_0^{-1}
\end{align*}
by the choice of $\epsilon_0$ in (\ref{eqn:HA15_3}). Consequently, we have $d(\bar{y},\bar{x}(\bar{t}))<\epsilon_0^{-1}$,
which contradicts (\ref{eqn:HA15_6}).
\end{proof}
Based on Lemma~\ref{lma:HA14_2}, we can improve Proposition~\ref{prn:SL04_1}.
Namely, under the condition $|R|+|\lambda| \to 0$, the limit flow is static, even on the singular part.
Clearly, due to Theorem~\ref{thm:SC28_1}, we do not need the assumption of lower bound of polarized
canonical radius anymore.
\begin{proposition}[\textbf{Static limit space-time}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies
\begin{align*}
\lim_{i \to \infty} \sup_{\mathcal{M}_i} (|R|+|\lambda|)=0.
\end{align*}
Suppose $x_i \in M_i$. Then
\begin{align*}
(M_i,x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g}).
\end{align*}
Moreover, we have
\begin{align*}
(M_i,x_i, g_i(t)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g})
\end{align*}
for every $t \in (-\bar{T}, \bar{T})$, where $\displaystyle \bar{T}=\lim_{i \to \infty} T_i>0$.
In other words, the identity maps between different time slices converge to the limit identity map.
\label{prn:HE15_1}
\end{proposition}
As a direct application, we obtain the bubble structure of a given family of polarized K\"ahler Ricci flows.
\begin{theorem}[\textbf{Space-time structure of a bubble}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$, $t_i \in (-T_i, T_i)$, and $r_i \to 0$.
Suppose $\widetilde{\mathcal{M}}_i$ is the adjusting of $\mathcal{M}_i$ by shifting time $t_i$ to $0$ and then
rescaling the space-time by the factors $r_i^{-2}$, i.e., $\tilde{g}_i(t)=r_i^{-2}g(r_i^2 t + t_i)$.
Suppose $r_i^{-2} \max \{|t_i-T_i|, |t_i+T_i|\}=\infty$. Then we have
\begin{align*}
(M_i,x_i, \tilde{g}_i(t)) \longright{\hat{C}^{\infty}} (\hat{M}, \hat{x}, \hat{g})
\end{align*}
for each time $t \in (-\infty, \infty)$ with $\hat{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{thm:HE26_1}
\end{theorem}
Theorem~\ref{thm:HE26_1} means that the space-time structure of
$\hat{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$ is the model for the space-time structures around $(x_i, t_i)$,
up to proper rescaling. Therefore, Theorem~\ref{thm:HE26_1} is an improvement of Theorem~\ref{thm:SC04_1},
where we only concern the metric structure.
In view of Proposition~\ref{prn:HE15_1},
it is not hard to see that distance is a uniform continuous function of time in $\mathscr{K}(n,A)$.
\begin{theorem}[\textbf{Uniform continuity of distance function}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x,y \in M$.
Suppose $d_{g(0)}(x,y)<1$. Then for every small $\epsilon$, there is a
$\delta=\delta(n,A,\epsilon)$ such that
\begin{align*}
|d_{g(t)}(x,y)-d_{g(0)}(x,y)|<\epsilon
\end{align*}
whenever $|t|<\delta$.
\label{thm:HE15_1}
\end{theorem}
\begin{proof}
We argue by contradiction. Suppose the statement was wrong, we can find an $\bar{\epsilon}>0$ and a sequence of
flows violating the statement for time $|t_i| \to 0$. Around $x_i$,
in the ball $B_{g_i(0)}\left(x_i,\frac{\epsilon_0^2\bar{\epsilon}}{10} \right)$,
we can find $x_i'$ which are uniform regular at time $t=0$, where $\epsilon_0$
is the same constant in Lemma~\ref{lma:HA14_2} and Lemma~\ref{lma:HA14_1}.
By two-sided pseudolocality, Theorem~\ref{thm:SL27_2}, it is clear that $x_i'$ is also
uniform regular at time $t=t_i$. Similarly, we can choose $y_i'$.
By virtue of triangle inequality and Lemma~\ref{lma:HA14_2}, we obtain
\begin{align*}
& d_{g_i(0)}(x_i', y_i') -\frac{\epsilon_0^2 \bar{\epsilon}}{5}
\leq d_{g_i(0)}(x_i, y_i) \leq d_{g_i(0)}(x_i', y_i') + \frac{\epsilon_0^2 \bar{\epsilon}}{5},\\
& d_{g_i(t_i)}(x_i', y_i') -\frac{\bar{\epsilon}}{5}
\leq d_{g_i(t_i)}(x_i, y_i) \leq d_{g_i(t_i)}(x_i', y_i') + \frac{\bar{\epsilon}}{5}.
\end{align*}
By argument similar to that in Proposition~\ref{prn:SC17_4}, it is clear that
\begin{align*}
\lim_{i \to \infty} d_{g_i(t_i)}(x_i', y_i')= \lim_{i \to \infty} d_{g_i(0)}(x_i', y_i').
\end{align*}
Then it follows that
\begin{align*}
\lim_{i \to \infty} d_{g_i(0)}(x_i,y_i) -\frac{(1+\epsilon_0^2)}{5}\bar{\epsilon}
\leq
\lim_{i \to \infty} d_{g_i(t_i)}(x_i, y_i) \leq \lim_{i \to \infty} d_{g_i(0)}(x_i,y_i) + \frac{(1+\epsilon_0^2)}{5} \bar{\epsilon}.
\end{align*}
In particular, for large $i$, we have
\begin{align*}
\left| d_{g_i(0)}(x_i,y_i) - d_{g_i(t_i)}(x_i, y_i) \right|<\frac{(1+\epsilon_0^2)}{5} \bar{\epsilon}<\bar{\epsilon},
\end{align*}
which contradicts our assumption.
\end{proof}
\subsection{Volume estimate for high curvature neighborhood}
In this subsection, we shall develop the flow version of the volume estimate of
Donaldson and the first author(c.f.~\cite{CD1},~\cite{CD2}, see also~\cite{CN}).
\begin{proposition}[\textbf{K\"ahler cone complex splitting}]
Same conditions as in Proposition~\ref{prn:HA07_2}, $\bar{y} \in \bar{M}$. Suppose
$\hat{Y}$ is a tangent cone of $\bar{M}$ at $\bar{y}$, then there is a fixed nonnegative integer $k$
such that
\begin{align}
\hat{Y}=C(Z) \times \ensuremath{\mathbb{C}}^{n-k},
\label{eqn:HA05_1}
\end{align}
where $C(Z)$ is a metric cone without straight line. A point in $\bar{M}$ is regular if and only if one
of the tangent cone is $\ensuremath{\mathbb{C}}^{n}$.
\label{prn:SL26_1}
\end{proposition}
\begin{proof}
By definition of tangent cone, one can find a sequence of numbers $r_i \to 0$. Taking subsequence if necessary,
let $\tilde{g}_i(t)=r_i^{-2} g_i(r_i^2t)$, then we have
\begin{align*}
(M_i, y_i, \tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{Y}, \hat{y}, \hat{g}).
\end{align*}
By compactness, we see that $(\hat{Y},\hat{y},\hat{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
On the other hand, it is a metric cone, which is the tangent space of itself at the origin.
So $\hat{Y}$ has the decomposition (\ref{eqn:HA05_1}),
by Theorem~\ref{thm:HE08_1} or Lemma~\ref{lma:HD30_1}.
\end{proof}
\begin{proposition}[\textbf{K\"ahler tangent cone rigidity}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$.
Suppose $x_i \in M_i$ and $(\bar{M}, \bar{x}, \bar{g})$ is a limit space of $(M_i, x_i, g_i(0))$.
Let $\hat{Y}$ be
a tangent space of $\bar{M}$. Then $\hat{Y}$ satisfies the splitting (\ref{eqn:HA05_1}) for $k=2$ or $k=0$.
\label{prn:HA10_2}
\end{proposition}
\begin{proof}
Clearly, $k=0$ if and only if the base point is regular. So it suffices to show that for every singular tangent space we have
$k=2$. By Proposition~\ref{prn:SL26_1}, we only need to rule out the case $k \geq 3$. However, this follows from the rigidity of complex structure on the smooth annulus in $\ensuremath{\mathbb{C}}^k /\Gamma$, where $\Gamma$ is a finite group of holomorphic isometry
of $\ensuremath{\mathbb{C}}^k$, when $k \geq 3$. Note that $[\omega_i]=c_1(L_i)$, which is an integer class.
Therefore, the proof follows verbatim as that in~\cite{CD2}. Note that Ricci curvature uniformly bounded condition in~\cite{CD2} is basically used to guarantee the Cheeger-Gromov convergence.
In our case, the convergence can be obtained from Theorem~\ref{thm:SC28_1}.
\end{proof}
\begin{proposition}[\textbf{Existence of holomorphic slicing}]
Suppose $\hat{Y} \in \widetilde{\mathscr{KS}}(n,\kappa)$ is a metric cone satisfying the splitting (\ref{eqn:HA05_1}).
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$.
If $(M,x,g(0))$ is very close to $(\hat{Y}, \hat{y}, \hat{g})$, i.e., the pointed-Gromov-Hausdorff distance
\begin{align*}
d_{PGH}((M,x,g(0)), (\hat{Y}, \hat{y}, \hat{g}))<\epsilon
\end{align*}
for sufficiently small $\epsilon$, which depends on $n,A,\hat{Y}$, then there exists a holomorphic map
\begin{align*}
\Psi=(u_{k+1}, u_{k+2}, \cdots, u_n): B(x, 10) \mapsto \ensuremath{\mathbb{C}}^{n-k}
\end{align*}
satisfying
\begin{align}
& |\nabla \Psi| \leq C(n,A), \label{eqn:HA10_1}\\
& \sum_{k+1 \leq i,j\leq n}
\int_{B(x,10)} \left| \delta_{ij} - \langle \nabla u_i, \nabla u_j \rangle\right|dv
\leq \eta(n,A,\epsilon), \label{eqn:HA10_2}
\end{align}
where $\eta$ is a small number such that $\displaystyle \lim_{\epsilon \to 0} \eta=0$.
\label{prn:HA10_1}
\end{proposition}
\begin{proof}
It follows from the argument in~\cite{DS} that the constant section $1$ of the trivial bundle over $\hat{Y}$ can be
``pulled back" as a non-vanishing holomorphic section of $L$ over $B(x,10)$, up to a finite lifting of power of $L$.
Therefore, we can regard $L$ as a trivial bundle over $B(x,10)$ without loss of generality.
Let $S_0$ be the pull-back of the constant $1$ section. In particular, $S_0$ is a non-vanishing
holomorphic section on $B(x,10)$.
On $B(x,10)$, every holomorphic section $S$ of $L$ can be written as $S=u S_0$ for a holomorphic function $u$ and
$\norm{S}{h}^2= \snorm{u}{}^2 \norm{S_0}{h}^2$.
From the splitting (\ref{eqn:HA05_1}), there exist natural coordinate holomorphic functions
$\{z_j\}_{j=k+1}^{n}$ on $\hat{Y}$.
Same as~\cite{DS}, one can apply H\"ormander's estimate
to construct $\{S_j\}_{j=k+1}^{n}$, which are holomorphic sections of $L$.
Each $S_j$ can be regarded as an ``approximation" of $z_j$, although they have different base spaces.
Let $u_j=\frac{S_j}{S_0}$ for each $j \in \{k+1, \cdots, n\}$.
Then we can define a holomorphic map $\Psi$ from $B(x,10)$ to $\ensuremath{\mathbb{C}}^{n-k}$ as follows
\begin{align*}
\Psi(y) \triangleq \left( u_{k+1}, u_{k+2}, \cdots, u_n \right).
\end{align*}
Note that each $S_i$ is a holomorphic section of $L$ with $L^2$-norm bounded from two sides, according to its
construction. It is easy to check that $\norm{\nabla S_i}{h}^2$ satisfies a sub-elliptic equation. So there exists a uniform
bound $\norm{\nabla S_i}{h}^2 \leq C(n,A)$, which implies (\ref{eqn:HA10_1}) when restricted on $B(x,10)$.
Moreover, on $B(x,10)$, by smooth convergence, it is not hard to see that $ \langle \nabla u_i, \nabla u_j \rangle$ can pointwisely
approximate $\delta_{ij}$ away from singularities of $\hat{Y}$, in any accuracy level when $\epsilon \to 0$.
This approximation together with (\ref{eqn:HA10_1}) yields (\ref{eqn:HA10_2}).
\end{proof}
\begin{theorem}[\textbf{Weak monotonicity of curvature integral}]
There exists a small constant $\epsilon=\epsilon(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Suppose $x \in M$, $0<r \leq 1$.
Then under the metric $g(0)$, we have
\begin{align}
\sup_{B(x, \frac{1}{2}r)} |Rm| \leq r^{-2}
\label{eqn:SC13_3}
\end{align}
whenever $r^{4-2n}\int_{B(x,r)} |Rm|^2 dv \leq \epsilon$.
\label{thm:SC12_3}
\end{theorem}
\begin{proof}
Up to rescaling, we can assume $r=1$ without loss of generality. If the statement was wrong, we can find a sequence of
points $x_i \in M_i$ such that
\begin{align*}
\int_{B(x_i,1)} |Rm|^2 dv \to 0, \quad
\sup_{B(x_i,\frac{1}{2})} |Rm| \geq 1,
\end{align*}
where the default metric is $g_i(0)$, the time zero metric of a flow $g_i$, in the moduli space $\mathscr{K}(n,A)$.
By the smooth convergence at places when curvature uniformly bounded, it is clear that the above conditions imply that
\begin{align*}
\int_{B(x_i,1)} |Rm|^2 dv \to 0, \quad
\sup_{B(x_i,\frac{3}{4})} |Rm| \to \infty.
\end{align*}
Let $(\bar{M},\bar{x},\bar{g})$ be the limit space of $(M_i,x_i,g_i(0))$. Then $B(\bar{x}, \frac{3}{4})$ contains at least one
singularity $\bar{y}$. Without loss of generality, we can assume $\bar{x}$ is a singular point.
Note that $B(\bar{x}, \frac{1}{4})$ is a flat manifold away from singularities. So every tangent space
of $\bar{M}$ at $\bar{x}$ is a flat metric cone. Let $\hat{Y}$ be one of such a flat metric cone. By taking subsequence if necessary, we can assume
\begin{align*}
(M_i,x_i,\tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{Y}, \hat{x},\hat{g}),
\end{align*}
for some flow metrics $\tilde{g}_i$ satisfying $\displaystyle \tilde{g}_i(t)=r_i^{-2} g_i(r_i^2 t), \; r_i \to 0$. Since $\hat{Y}$ is
a flat metric cone, in light of Proposition~\ref{prn:HA10_2}, we have the splitting
\begin{align*}
\hat{Y}= (\ensuremath{\mathbb{C}}^2/\Gamma) \times \ensuremath{\mathbb{C}}^{n-2}.
\end{align*}
Let $(M,x,\tilde{g})$ be one of $(M_i,x_i,\tilde{g}_i(0))$ for some large $i$. Because of Proposition~\ref{prn:HA10_1}, we can construct a holomorphic map $\Psi:B(x,10) \to \ensuremath{\mathbb{C}}^{n-2}$
satisfying (\ref{eqn:HA10_1}) and (\ref{eqn:HA10_2}).
Then we can follow the slice argument as in~\cite{CCT} and~\cite{Cheeger03}.
Our argument will be simpler since our slice functions are holomorphic rather than harmonic.
Actually, for generic $\vec{z}=(z_{3}, z_{4}, \cdots, z_n)$ satisfying $|\vec{z}|<0.1$,
we know $\Psi^{-1}(\vec{z}) \cap B(x,5)$ is a complex surface with boundary.
Clearly, $\Psi^{-1}((S^3/\Gamma) \times \{\vec{z}\})$ is close to $(S^3/\Gamma) \times \vec{z}$,
if we regard $S^3/\Gamma$ as the unit sphere in $\ensuremath{\mathbb{C}}^2/\Gamma$. Deform the preimage a little bit if necessary, we can obtain
a $\partial \Omega$ which bounds a complex surface $\Omega$. By coarea formula and the bound of
$|\nabla \Psi|$, it is clear that for generic $\Omega$ obtained in this way, we have
\begin{align*}
\int_{\Omega} |Rm|^2 d\sigma\to 0.
\end{align*}
Consider the restriction of $TM$ on $\Omega$. Let $c_2$ be a form representing the second Chern class of the tangent bundle $TM$, obtained from the K\"ahler metric $\tilde{g}(0)$ from the classical way.
Let $\hat{c}_2$ be the corresponding differential character with value in $\ensuremath{\mathbb{R}}/\ensuremath{\mathbb{Z}}$. Since the pointwise norm of $c_2$ is bounded by $|Rm|^2$, it is clear that
\begin{align}
\hat{c}_2(\partial \Omega)=\int_{\Omega} c_2 \quad (mod \; \ensuremath{\mathbb{Z}}) \to 0. \label{eqn:HA10_3}
\end{align}
On the other hand, since $\partial \Omega$ converges to $S^3/\Gamma$, we have
\begin{align}
\hat{c}_2(\Omega) \to \frac{1}{|\Gamma|}. \label{eqn:HA10_4}
\end{align}
Therefore, the combination of (\ref{eqn:HA10_3}) and (\ref{eqn:HA10_4}) forces that $|\Gamma|=1$. This is impossible
since $|\Gamma| \geq 2$ by our assumption that $\hat{Y}$ is a singular metric cone.
\end{proof}
From now on to the end of this subsection, we use $g(0)$ as the default metric.
Similar to the definition in~\cite{CD1}, for any small $r$, let $\mathcal{Z}_r$ be the $r$-neighborhood of the points where $|Rm| > r^{-2}$.
Recall the definition equation (\ref{eqn:SL27_1}), we denote $\mathcal{F}_r$ as the collection of points
whose canonical volume radii are greater than $r$, $\mathcal{D}_r$ as the component of $\mathcal{F}_r$.
Under these notations, we have the following property.
\begin{proposition}[\textbf{Equivalence of singular neighborhoods}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $0<r<\hslash$.
Then at time zero, we have
\begin{align}
\mathcal{D}_{cr} \subset \mathcal{Z}_r \subset \mathcal{D}_{\frac{1}{c}r}
\label{eqn:HA11_1}
\end{align}
for some small constant $c=c(n,A)$.
\label{prn:HA11_1}
\end{proposition}
\begin{proof}
Let us first prove $\mathcal{D}_{cr} \subset \mathcal{Z}_r$. Suppose the statement was wrong, we can find a sequence $c_i \to 0$
and flows in $\mathscr{K}(n,A)$ such that $\mathcal{D}_{c_ir_i} \not\subset Z_{r_i}$ for some $r_i <\hslash$.
Choose $x_i \in \mathcal{D}_{c_i r_i} \cap \mathcal{Z}_{r_i}^{c}$. Let $\rho_i$ be the canonical volume radius of $x_i$. Rescale
the flow such that the canonical volume radius at $x_i$ becomes $1$. Take limit, we will obtain a smooth flat space in
$\widetilde{\mathscr{KS}}(n,\kappa)$, which is nothing but $\ensuremath{\mathbb{C}}^n$. Therefore, the canonical volume radii
of the base points $x_i$ should tend to infinity, which is a contradiction.
Then we prove $\mathcal{Z}_r \subset \mathcal{D}_{\frac{1}{c}r}$. Suppose $x \in \mathcal{Z}_r$, then $|Rm|(y) \geq r^{-2}$ for some $y \in B(x,r)$. By the regularity improving property of canonical volume radius, it is clear
that $\mathbf{cvr}(x) \leq \frac{2}{c_a} r$. In other words, $x \in \mathcal{D}_{\frac{2}{c_a}r}$.
\end{proof}
\begin{theorem}[\textbf{Volume estimates of high curvature neighborhood}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Under the metric $g(0)$, we have
\begin{align*}
|\mathcal{Z}_r| \leq C r^4,
\end{align*}
where $C$ depends on $n,A$ and the upper bound of $\int_M |Rm|^2 dv$.
\label{thm:HA11_1}
\end{theorem}
\begin{proof}
Because of Proposition~\ref{prn:HA11_1}, it suffices to show $|\mathcal{D}_{cr}| \leq C r^4$.
In light of Theorem~\ref{thm:SC12_3}, if $r^{4-2n}\int_{B(x,r)} |Rm|^2 dv<\epsilon$ for some
$r<\hslash$, then $x \in \mathcal{F}_{cr}$. In other words, if $x \in \mathcal{D}_{cr}(M,0)$, then it is forced that
\begin{align*}
r^{4-2n} \int_{B(x,r)} |Rm|^2 dv \geq \epsilon.
\end{align*}
Let $\bigcup_{i=1}^{N} B(x_i,2r)$ be a finite cover of $\mathcal{D}_{cr}$ such that
\begin{itemize}
\item $x_i \in \mathcal{D}_{cr}$.
\item $B(x_i,r)$ are disjoint to each other.
\end{itemize}
Then we can bound $N$ as follows.
\begin{align*}
N\epsilon r^{2n-4} \leq \sum_{i=1}^{N} \int_{B(x_i,r)} |Rm|^2 dv \leq \int_M |Rm|^2 dv \leq H.
\end{align*}
Consequently, we have
\begin{align*}
|\mathcal{D}_{cr}(M)| \leq \sum_{i=1}^{N} |B(x_i, 2r)| \leq \frac{H}{\epsilon} r^{4-2n} \kappa^{-1} \omega_{2n} (2r)^{2n}
\leq C r^4.
\end{align*}
Since both $\kappa$ and $\epsilon$ depends only on $n$ and $A$. It is clear that $C=C(n,A,H)$ where $H$
is the upper bound of $\int_M |Rm|^2 dv$.
\end{proof}
\begin{corollary}[\textbf{Volume estimates of singular neighborhood}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$.
Suppose $\int_{M_i} |Rm|^2 dv \leq H$ uniformly under the metric $g_i(0)$. Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit
space of $(M_i,x_i,g_i(0))$. Let $\mathcal{S}_r$ be the set defined in (\ref{eqn:HA11_4}), then we have
\begin{align*}
|\mathcal{S}_{r}| \leq C r^4
\end{align*}
for each small $r$ and some constant $C=C(n,A,H)$.
In particular, we have the estimate of Minkowski dimension of the singularity
\begin{align*}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4.
\end{align*}
\label{cly:HA11_1}
\end{corollary}
Following~\cite{CW4}, the space $\bar{M}=\mathcal{R} \cup \mathcal{S}$
is called a metric-normal $Q$-Fano variety if there exists
a homeomorphic map $\varphi: \bar{M} \to Z$ for some $Q$-Fano normal variety $Z$ such that $\varphi|_{\mathcal{R}}$
is a biholomorphic map. Moreover, $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$.
\begin{theorem}[\textbf{Limit structure}]
Suppose that $\mathcal{LM}_i \in \mathscr{K}(n,A)$.
Under the metric $g_i(0)$, suppose
\begin{align}
\Vol(M_i) + \int_{M_i} |Rm|^2 dv \leq H \label{eqn:HA11_2}
\end{align}
for some uniform constant $H$.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i,x_i,g_i(0))$.
Then $\bar{M}$ is a compact metric-normal $Q$-Fano variety.
\label{thm:SC13_4}
\end{theorem}
\begin{proof}
It follows from (\ref{eqn:HA11_2}) and the non-collapsing that $diam(M_i)$ is uniformly bounded.
So the limit space $\bar{M}$ is compact. Due to Theorem~\ref{thmin:HC08_1}, the partial $C^0$-estimate,
one can follow the argument in~\cite{DS} to show that $\bar{M}$ is a $Q$-Fano, normal variety.
The metric-normal property follows from Corollary~\ref{cly:HA11_1}.
\end{proof}
Based on the estimates developed in this subsection, we can easily prove Corollary~\ref{clyin:SC24_4}
and Corollary~\ref{clyin:SC24_5} in the introduction.
\begin{proof}[Proof of Corollary~\ref{clyin:SC24_4} and Corollary~\ref{clyin:SC24_5}]
It follows from the combination of Theorem~\ref{thm:SC13_4}, Corollary~\ref{cly:HA11_1} of this paper and
main results in~\cite{CW4}. Note that the line bundle metric choice in this paper is
equivalent to that in~\cite{CW4}, due to the bound of $\dot{\varphi}$.
\end{proof}
\subsection{Singular K\"ahler Ricci flow solution}
\label{sec:sflow}
In this subsection, we shall relate the different limit time slices, without the assumption of $|R|+|\lambda| \to 0$.
We shall further improve regularity, by estimates essentially arising from complex analysis of holomorphic sections.
We want to compare $\omega_t$, the K\"ahler Ricci flow metrics, and $\tilde{\omega}_t$, the evolving Bergman metrics.
We first show that $\tilde{\omega}_t$ is very stable when $t$ evolves.
\begin{lemma}
Suppose $G(t)$ is a family of $(N+1)\times (N+1)$ matrices parameterized by $t \in [-1,1]$. Suppose $G(0)=Id$, $\dot{G}(0)=B$.
Let $\lambda_0\leq \lambda_1 \leq \cdots \leq \lambda_N$ be the real eigenvalues of the Hermitian matrix $B+\bar{B}^{\tau}$.
If we regard $G$ as a holomorphic map from $\ensuremath{\mathbb{CP}}^N$ to $\ensuremath{\mathbb{CP}}^N$, then we have
\begin{align}
(\lambda_0-\lambda_N) \omega_{FS} \leq
\left. \frac{d}{dt}G(t)^*(\omega_{FS}) \right|_{t=0} \leq (\lambda_N-\lambda_0) \omega_{FS}.
\label{eqn:HB02_1}
\end{align}
\label{lma:HB02_1}
\end{lemma}
\begin{proof}
Let $\{z_i\}_{i=0}^N$ be the homogeneous coordinate of $\ensuremath{\mathbb{CP}}^N$. Let $G=G(t)$. Then we have
\begin{align*}
&\omega_{FS}=\sqrt{-1} \partial \bar{\partial} \log (|z_0|^2+|z_1|^2+\cdots |z_N|^2)
=\sqrt{-1} \left\{\frac{\partial z_i \wedge \bar{\partial} \bar{z}_i}{|z|^2}
+\frac{(z_i \bar{\partial} \bar{z}_i) \wedge (\bar{z}_j \partial z_j)}{|z|^4}\right\}, \\
&G^*(\omega_{FS})=\sqrt{-1} \partial \bar{\partial} \log (|\tilde{z}_0|^2+|\tilde{z}_1|^2+\cdots |\tilde{z}_N|^2)
=\sqrt{-1} \left\{\frac{\partial \tilde{z}_i \wedge \bar{\partial} \bar{\tilde{z}}_i}{|\tilde{z}|^2}
+\frac{(\tilde{z}_i \bar{\partial} \bar{\tilde{z}}_i) \wedge (\bar{\tilde{z}}_j \partial \tilde{z}_j)}{|\tilde{z}|^4}\right\},
\end{align*}
where $\tilde{z}_i=G_{ij}z_j$. Let $\{ w_1, \cdots, w_N\}$ be local coordinate.
At point $z$, the matrix of $\omega_{FS}$ is
\begin{align*}
E_0=J \left(\frac{Id}{|z|^2}-\frac{\bar{z}^{\tau}z}{|z|^4}\right) \bar{J}^{\tau}=J F_0 \bar{J}^{\tau},
\end{align*}
where $J$ is an $N \times (N+1)$ matrix which is the Jacobi matrix $\left(\D{z_j}{w_{\alpha}}\right)$.
The matrix of
$\omega_{G^*\omega_{FS}}$ is
\begin{align*}
E_t=J G\left(\frac{Id}{|\tilde{z}|^2}-\frac{\bar{\tilde{z}}^{\tau} \tilde{z}}{|\tilde{z}|^4}\right) \bar{G}^{\tau} \bar{J}^{\tau}
=JF_t \bar{J}^{\tau}.
\end{align*}
Clearly, we have
\begin{align*}
\left. \frac{d}{dt} F_t \right|_{t=0}&=B\left(\frac{Id}{|z|^2}-\frac{\bar{z}^{\tau}z}{|z|^4}\right)
+\left(\frac{Id}{|z|^2}-\frac{\bar{z}^{\tau}z}{|z|^4}\right) \bar{B}^{\tau}
-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} Id +\frac{2z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^6} \bar{z}^{\tau}z
-\frac{\bar{B}^{\tau}\bar{z}^{\tau}z+\bar{z}^{\tau}zB}{|z|^4}\\
&=\left\{ \frac{B+\bar{B}^{\tau}}{|z|^2} -\frac{(B+\bar{B}^{\tau})\bar{z}^{\tau}z + \bar{z}^{\tau}z(B+\bar{B}^{\tau})}{|z|^4}
+\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^6} \bar{z}^{\tau}z \right\}
-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} F_0\\
&\triangleq M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} F_0.
\end{align*}
It follows that
\begin{align*}
\left. \frac{d}{dt} E_t \right|_{t=0}=
\left. \frac{d}{dt}
\left\{J G\left(\frac{Id}{|\tilde{z}|^2}-\frac{\bar{\tilde{z}}^{\tau} \tilde{z}}{|\tilde{z}|^4}\right) \bar{G}^{\tau} \bar{J}^{\tau} \right\}
\right|_{t=0}=J\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{J}^{\tau}.
\end{align*}
It is easy to check that
\begin{align*}
zM\bar{z}^{\tau}=0, \quad zF_0\bar{z}^{\tau}=0, \quad
z\left( M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} F_0\right)\bar{z}^{\tau}=0.
\end{align*}
Without loss of generality, we can assume $B+\bar{B}^{\tau}$ is a diagonal matrix
$\diag(\lambda_0, \lambda_1, \cdots, \lambda_N)$.
Let $v=(v_0, v_1, \cdots, v_N)$ be a vector in $\ensuremath{\mathbb{C}}^{N+1}$ satisfying
\begin{align*}
z\bar{v}^{\tau}=\bar{v}_0z_0 +\bar{v}_1z_1 + \cdots + \bar{v}_N z_N=0.
\end{align*}
Then it is clear that
\begin{align*}
vM\bar{v}^{\tau}=\frac{v(B+\bar{B}^{\tau})\bar{v}^{\tau}}{|z|^2}, \quad
vF_0\bar{v}^{\tau}=\frac{|v|^2}{|z|^2}.
\end{align*}
Therefore, we have
\begin{align*}
&\quad v\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{v}^{\tau}\\
&=\frac{1}{|z|^4}\left\{ (\lambda_0 |v_0|^2+\cdots+\lambda_N |v_N|^2)(|z_0|^2+\cdots+|z_N|^2)
-(\lambda_0|z_0|^2+\cdots \lambda_N|z_N|^2)(|v_0|^2+\cdots+|v_N|^2)\right\}\\
&=\frac{1}{|z|^4} \left\{ \left[(\lambda_0-\lambda_0)|z_0|^2 +(\lambda_0-\lambda_1)|z_1|^2+\cdots
+(\lambda_0-\lambda_N)|z_N|^2 \right]|v_0|^2 \right.\\
&\quad + \left[(\lambda_1-\lambda_0)|z_0|^2 +(\lambda_1-\lambda_1)|z_1|^2+\cdots
+(\lambda_1-\lambda_N)|z_N|^2 \right]|v_1|^2 \\
&\quad +\cdots\\
&\quad \left.+\left[(\lambda_N-\lambda_0)|z_0|^2 +(\lambda_N-\lambda_1)|z_1|^2+\cdots
+(\lambda_N-\lambda_N)|z_N|^2 \right]|v_N|^2 \right\}\\
&\leq (\lambda_N-\lambda_0) \frac{|v|^2}{|z|^2}.
\end{align*}
Similarly, we have
\begin{align*}
v\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{v}^{\tau} \geq (\lambda_0-\lambda_N)\frac{|v|^2}{|z|^2}.
\end{align*}
Note that $z F_0 \bar{v}^{\tau}=0$. Therefore, we can apply the orthogonal decomposition with respect to $F_0$ to obtain that
for every vector $f=(f_0,f_1,\cdots, f_N) \in \ensuremath{\mathbb{C}}^{N+1}$, we have
\begin{align*}
\left| f\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{f}^{\tau} \right|
\leq (\lambda_N-\lambda_0) \frac{|f|^2}{|z|^2}=(\lambda_N-\lambda_0)fF_0\bar{f}^{\tau}.
\end{align*}
Let $u \in T_z^{(1,0)}\ensuremath{\mathbb{CP}}^N$. Then we have
\begin{align*}
\langle u,u \rangle_{\omega_{FS}}&=(uJ) F_0 \overline{(uJ)}^{\tau},\\
\left| \langle u, u \rangle_{\left.\frac{d}{dt}G^*(\omega_{FS}) \right|_{t=0}} \right|
&=\left|(uJ) \left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\overline{uJ}^{\tau} \right|\\
&\leq (\lambda_N-\lambda_0) (uJ) F_0 \overline{(uJ)}^{\tau}\\
&\leq (\lambda_N-\lambda_0)\langle u,u \rangle_{\omega_{FS}}.
\end{align*}
By the arbitrary choice of $u$, then (\ref{eqn:HB02_1}) follows directly from the above inequality.
\end{proof}
\begin{lemma}
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Let $\tilde{\omega}_t$ be the pull back of the Fubini-Study metric by orthonormal basis of $L$ with respect to $\omega_t$ and $h_t$.
Then we have the evolution inequality of $\tilde{\omega}_t$:
\begin{align}
-2A\tilde{\omega}_t \leq \frac{d}{dt} \tilde{\omega_t} \leq 2A \tilde{\omega}_t.
\label{eqn:HB03_1}
\end{align}
\label{lma:HB05_1}
\end{lemma}
\begin{proof}
Without loss of generality, it suffices to show (\ref{eqn:HB03_1}) at time $t=0$. \\
Suppose $\left\{ s_i \right\}_{i=0}^{N}$ is an orthonormal basis at time $0$, $\left\{ \tilde{s}_i \right\}_{i=0}^{N}$
is an orthonormal basis at time $t$. They are related by $\tilde{s}_i=s_jG_{ji}$. Fix $e_L$ a local representation of the line bundle $L$ around a point $x$ so that locally we have $s_j=z_j e_L$ and $\tilde{s}_j=\tilde{z}_j e_L=z_j G_{ji}e_L$. Then we have
\begin{align*}
& \tilde{\omega}_0=\sqrt{-1} \partial \bar{\partial} \log \left(|z_0|^2+|z_1|^2+\cdots +|z_N|^2 \right), \\
& \tilde{\omega}_t=\sqrt{-1} \partial \bar{\partial} \log \left( |\tilde{z}_0^2|+|\tilde{z}_1|^2+\cdots +|\tilde{z}_N|^2 \right).
\end{align*}
Let $\iota$ be the Kodaira embedding map induced by $\left\{ s_i \right\}_{i=0}^{N}$ at time $0$. Then it is clear that
\begin{align*}
\tilde{\omega}_0=\iota^* \omega_{FS}, \quad \tilde{\omega}_t=\iota^* (G^*\omega_{FS}).
\end{align*}
Therefore, we have
\begin{align*}
\left. \frac{d}{dt} \tilde{\omega}_t \right|_{t=0}=\iota^*\left( \left. \frac{d}{dt}G^*(\omega_{FS}) \right|_{t=0} \right).
\end{align*}
So (\ref{eqn:HB03_1}) is reduced to the estimate
\begin{align}
-2A \omega_{FS} \leq \left. \frac{d}{dt}G(t)^*(\omega_{FS}) \right|_{t=0} \leq 2A \omega_{FS}.
\label{eqn:HB05_5}
\end{align}
However, note that
\begin{align*}
\delta_{ik}=G_{ij}\bar{G}_{kl} \int_{M} \langle s_j, s_l\rangle_{h_t} \frac{\omega_t^n}{n!}.
\end{align*}
Taking derivative on both sides at time $0$ and denote $\dot{G}$ by $B$, we obtain
\begin{align*}
0=B_{ik}+\bar{B}_{ki} + \int_M (-\dot{\varphi}+n\lambda-R)\langle s_i,s_k\rangle_{h_0} \frac{\omega_0^n}{n!}.
\end{align*}
Therefore, for every $v\in \ensuremath{\mathbb{C}}^{N+1}$, the following inequality holds.
\begin{align}
\left| v_i (B_{ij} + \bar{B}_{ji}) \bar{v}_j \right| =\left|-v_i\bar{v}_j \int_M (-\dot{\varphi}+n\lambda-R)\langle s_i,s_j\rangle_{h_0} \frac{\omega_0^n}{n!} \right|\leq A|v|^2.
\label{eqn:HB05_1}
\end{align}
In particular, each eigenvalue of the Hermitian matrix $B+\bar{B}^{\tau}$ has absolute value bounded by $A$. Then (\ref{eqn:HB05_5}) follows from
Lemma~\ref{lma:HB02_1}.
\end{proof}
In view of Lemma~\ref{lma:HB05_1}, the following property is obvious now.
\begin{proposition}[\textbf{Bergman metric equivalence along time}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$. Then we have
\begin{align}
e^{-2A|t|} \tilde{\omega}_0 \leq \tilde{\omega}_t \leq e^{2A|t|} \tilde{\omega}_0.
\label{eqn:HB03_2}
\end{align}
\label{prn:HB05_2}
\end{proposition}
In general, we cannot hope a powerful estimate like (\ref{eqn:HB03_2}) holds for metrics $\omega_t$, since such an estimate will imply the Ricci curvature is uniformly bounded by $A$. However, if we only focus on points regular enough, then we do have a similar weaker estimate.
\begin{proposition}[\textbf{Flow metric equivalence along time}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in \mathcal{F}_r(M,0)$. Then we have
\begin{align}
\frac{1}{C} \omega_0(x) \leq \omega_t(x) \leq C\omega_0(x) \label{eqn:HB05_2}
\end{align}
for every $t \in [-1,1]$. Here $C$ is a constant depending only on $n,A$ and $r$.
\label{prn:HB05_1}
\end{proposition}
\begin{proof}
Without loss of generality, we assume $\mathbf{b}$ is uniformly bounded.
By short time two-sided pseudolocality, Theorem~\ref{thm:SL27_2} and rescaling, it suffices to show (\ref{eqn:HB05_2}) for $t=-1$ and $t=1$.
At time $0$, it is clear that $\omega_0(x)$ and $\tilde{\omega}_0(x)$ are uniformly equivalent. The volume form
$\omega_0^n$ is uniformly equivalent to $\omega_t^n$. By the stability of $\tilde{\omega}$, inequality (\ref{eqn:HB03_2}), it suffices to prove
the following two inequalities hold at point $x$.
\begin{align}
& \Lambda_{\omega_1} \tilde{\omega}_0 \leq C, \label{eqn:HB05_3}\\
& \Lambda_{\omega_0} \tilde{\omega}_{-1} \leq C. \label{eqn:HB05_4}
\end{align}
We shall prove the above two inequalities separately.
Let $w_0$ be defined as that before Lemma~\ref{lma:SK23_1}.
Let $w$ be the solution of $\square w=0$, initiating from $w_0$. By the heat kernel estimate and the uniform upper
bound of diameter of $B_{g(0)}(x,r)$ under metric $g(t)$(c.f. Lemma~\ref{lma:HA14_1}), we see that $w(x,1)$ is uniformly bounded away from $0$.
Then Lemma~\ref{lma:J02_1} applies and we obtain that
\begin{align*}
\Lambda_{\omega_1(x)} \tilde{\omega}_0(x) =F(x,1) \leq \frac{C}{w(x,1)} <C.
\end{align*}
So we finish the proof of (\ref{eqn:HB05_3}). The proof of (\ref{eqn:HB05_4}) is similar. Modular time shifting, the only difference is that we do not know
whether $x$ is very regular at time $t=-1$, so the construction of initial value of a heat equation may be a problem.
However, due to Proposition~\ref{prn:SC02_2}, we can always find a point $y_0 \in \mathcal{F}_{c_b \hslash}(M,-1) \cap B_{g(-1)}(x,\hslash)$. Consider the heat equation $w'$, starting from a cutoff function supported around $y_0$ at time $t=-1$.
In light of uniform diameter bound of $B_{g(-1)}(y_0,\hslash)$ under the metric $g(0)$, $w'(x,0)$ is uniformly bounded away from $0$. So we can follow the proof of
Lemma~\ref{lma:J02_1} to obtain that
\begin{align*}
\Lambda_{\omega_0(x)} \tilde{\omega}_{-1}(x) <\frac{C}{w'(x,0)}<C.
\end{align*}
Therefore, (\ref{eqn:HB05_4}) is proved.
\end{proof}
Note that due to the two-sided pseudolocality, Theorem~\ref{thm:SL27_2}, we now can use blowup argument, taking for granted that every convergence in regular part takes place in
smooth topology. Therefore, we can use the blowup argument in the proof of Proposition~\ref{prn:HE11_2},
based on the Liouville type theorem, Lemma~\ref{lma:HE11_1}.
Then the following corollary follows directly from Proposition~\ref{prn:HB05_1}.
\begin{corollary}[\textbf{Long-time regularity improvement in two time directions}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $r>0$, then
\begin{align*}
\mathcal{F}_r(M,0) \subset \bigcap_{-1 \leq t \leq 1} \mathcal{F}_{\delta}(M,t)
\end{align*}
for some $\delta=\delta(n,A,r)$.
\label{cly:HB05_1}
\end{corollary}
Now we are ready to prove Theorem~\ref{thmin:HC06_1}, the long-time, two-sided pseudolocality theorem.
\begin{proof}[Proof of Theorem~\ref{thmin:HC06_1}]
It follows from the combination of Corollary~\ref{cly:HB05_1} and Proposition~\ref{prn:HA08_3}.
\end{proof}
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Then for each time $t \in [-1,1]$ we have
\begin{align}
(M_i,x_i,g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M}(t),\bar{x}(t),\bar{g}(t)).
\label{eqn:HE15_1}
\end{align}
Let us see how are the two time slice limits $\bar{M}(0)$ and $\bar{M}(1)$ related.
Clearly, by Theorem~\ref{thmin:HC06_1}, the regular parts of $\bar{M}(0)$ and $\bar{M}(1)$ can be identified.
The relations among the singular parts at different time slices are more delicate.
For simplicity of notations, we denote $(\bar{M}(0), \bar{x}(0), \bar{g}(0))$ by $(\bar{M},\bar{x},\bar{g})$,
denote $(\bar{M}(1), \bar{x}(1), \bar{g}(1))$ by $(\bar{M}', \bar{x}', \bar{g}')$.
Let us also assume $\Vol(M_i)$ is uniformly bounded. Then it is clear that both $\bar{M}$
and $\bar{M}'$ are compact by the uniform non-collapsing caused by Sobolev constant bound.
In light of the uniform partial-$C^0$-estimate along the flow, without loss of generality,
we can assume that the Bergman function $\mathbf{b}$ is uniformly bounded below. By the fundamental estimates
in~\cite{DS}, we obtain that the map
\begin{align*}
Id_{0}: \quad (\bar{M}, \bar{x}, \bar{g}) \to (\bar{M}, \bar{x}, \tilde{\bar{g}})
\end{align*}
is a homeomorphism. Recall that $(\bar{M}, \bar{x}, \tilde{\bar{g}})$ is the limit of $(M_i, x_i, \tilde{g}_i(0))$, where
$\tilde{g}_i$ is the pull-back of Fubini-Study metric. Similarly, we have another homeomorphism map at time
$t=1$.
\begin{align*}
Id_{1}: \quad (\bar{M}', \bar{x}', \bar{g}') \to (\bar{M}', \bar{x}', \tilde{\bar{g}}').
\end{align*}
By Proposition~\ref{prn:HB05_2}, the pulled back Fubini-Study metrics $\tilde{g}_i(t)$ are uniformly equivalent
for $t \in [-1,1]$. It follows that there is a Lipschitz map $Id_{01}$ between two time slices, for the pulled back
Fubini-Study metrics:
\begin{align*}
Id_{01}: \quad (\bar{M}, \bar{x}, \tilde{\bar{g}}) \to (\bar{M}', \bar{x}', \tilde{\bar{g}}').
\end{align*}
Combining the previous steps and letting $\Psi=Id_{1}^{-1} \circ Id_{01} \circ Id_{0}$, we obtain that the map
\begin{align*}
\Psi: (\bar{M}, \bar{x}, \bar{g}) \to (\bar{M}', \bar{x}', \bar{g}')
\end{align*}
is a homeomorphism.
By analyzing each component identity map,
it is clear that $\Psi|_{\mathcal{R}(\bar{M})}$, where $\mathcal{R}(\bar{M})$ is the regular part of $\bar{M}$,
maps $\mathcal{R}(\bar{M})$ to $\mathcal{R}(\bar{M}')$, as a biholomorphic map.
Similarly, $\Psi|_{\mathcal{S}(\bar{M})}$ is a homeomorphism to $\mathcal{S}(\bar{M}')$.
Therefore, the variety structure of the $\bar{M}(t)$ does not depend on time.
We remark that the compactness of $\bar{M}$ is not essentially used here.
If $\bar{M}$ is noncompact, the above argument go through formally
if we replace the target embedding space $\ensuremath{\mathbb{CP}}^{N}$ by $\ensuremath{\mathbb{CP}}^{\infty}$.
This formal argument can be made rigorous by applying delicate localization technique.
However, in our applications, $\bar{M}$ is always compact except it is a bubble, i.e., a blowup limit.
In this situation, we have the extra condition $|R|+|\lambda| \to 0$,
then $\Psi$ can be easily chosen as identity map, due to Proposition~\ref{prn:HE15_1}.
From the above discussion, it is clear that the topology structure and variety structure of $\bar{M}(t)$ does not depend on time.
So we just denote $\bar{M}(t)$ by $\bar{M}$. Then we can denote the convergence (\ref{eqn:HE15_1}) by
\begin{align*}
(M_i,x_i,g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M},\bar{x}(t),\bar{g}(t))
\end{align*}
for each $t$.
Hence, the limit family of metric spaces can be regarded as a family of evolving metrics on the limit
variety. Therefore, the above convergence at each time $t$ can be glued together to obtain a global convergence
\begin{align*}
\left\{ (M_i,x_i,g_i(t)), -T_i < t < T_i\right\} \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left\{ (\bar{M}, \bar{x}, \bar{g}(t)), -\bar{T} <t <\bar{T}\right\}
\end{align*}
where $\displaystyle \bar{T}=\lim_{i \to \infty} t_i$. Clearly, $\bar{g}(t)$ satisfies the K\"ahler Ricci flow equation on the regular part of $\bar{M}$.
Recall that we typically denote the K\"ahler Ricci flow $\left\{ (M_i,x_i,g_i(t)), -T_i < t < T_i\right\}$ by $\mathcal{M}_i$. Then we obtain the convergence of
K\"ahler Ricci flows (with base points):
\begin{align}
(\mathcal{M}_i, x_i) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{\mathcal{M}}, \bar{x}).
\label{eqn:HE15_2}
\end{align}
If we further know the underlying space $\bar{M}$ is compact, then the notation can be even simplified as
\begin{align*}
\mathcal{M}_i \stackrel{\hat{C}^{\infty}}{\longrightarrow} \bar{\mathcal{M}}.
\end{align*}
\begin{remark}
The limit flow $\bar{\mathcal{M}}$ can be regarded as an intrinsic K\"ahler Ricci flow on the normal variety $\bar{M}$.
Actually, it is already clear that $\bar{\mathcal{M}}$ is at least a
weak super solution of Ricci flow, in the sense of R.J. McCann and P.M. Topping(\cite{MccTop}).
From the point of view of K\"ahler geometry, when restricted to the potential level,
the flow $\bar{\mathcal{M}}$ coincides with the weak K\"ahler Ricci flow solution defined by
Song and Tian(\cite{SongTian}), if $\bar{M}$ is compact.
\label{rmk:HE15_1}
\end{remark}
If we also consider the convergence of the line bundle structure, we can obviously generalize the convergence in (\ref{eqn:HE15_2})
as
\begin{align*}
& \left(\mathcal{LM}_i, x_i \right) \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left(\overline{\mathcal{LM}}, \bar{x}\right),
\quad \textrm{if $\bar{M}$ is non-compact}. \\
& \mathcal{LM}_i \stackrel{\hat{C}^{\infty}}{\longrightarrow} \overline{\mathcal{LM}}, \quad \textrm{if $\bar{M}$ is compact}.
\end{align*}
With these notations, we can formulate our compactness theorem as follows.
\begin{theorem}[\textbf{Polarized flow limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Then we have
\begin{align*}
\left(\mathcal{LM}_i, x_i \right) \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left(\overline{\mathcal{LM}}, \bar{x}\right),
\end{align*}
where $\overline{\mathcal{LM}}$ is a polarized K\"ahler Ricci flow solution on an analytic normal variety $\bar{M}$.
Moreover, if $\bar{M}$ is compact, then it is a projective normal variety.
\label{thm:HB07_1}
\end{theorem}
Notice that we have already proved Theorem~\ref{thmin:HC06_2} now.
\begin{proof}[Proof of Theorem~\ref{thmin:HC06_2}]
The limit polarized flow on variety follows from the combination of Theorem~\ref{thm:HB07_1} and Theorem~\ref{thm:HE19_1}.
The Minkowski dimension estimate of the singular set follows from Corollary~\ref{cly:HE25_2}.
\end{proof}
The properties of the limit spaces can be improved if extra conditions are available.
\begin{proposition}[\textbf{KE limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies
\begin{align}
\int_{-T_i}^{T_i}\int_{M_i} |R-n\lambda| dv dt \to 0.
\label{eqn:HA03_1}
\end{align}
Then $\overline{\mathcal{LM}}$ is a static, polarized K\"ahler Ricci flow solution. In other words, $\bar{g}(t) \equiv \bar{g}(0)$
and consequently are K\"ahler Einstein metric.
\label{prn:HB07_1}
\end{proposition}
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$ and $\lambda>0$. Then it is clear that $c_1(M)>0$, or $M$ is Fano.
Note that for every Fano manifold, we have a uniform bound $c_1^n(M) \leq C(n)$(c.f.~\cite{Deba}). This implies that
\begin{align*}
\frac{1}{A} \leq \Vol(M)=c_1^n(L)=\lambda^{-n} c_1^n(M) \leq C \lambda^{-n}.
\end{align*}
So $\lambda$ is bounded away from above. If we assume $\lambda$ is bounded away from zero, then
$\Vol(M)=c_1^{n}(L)$ is uniformly bounded. Consequently, $\diam(M)$ is uniformly bounded by non-collapsing, due to the Sobolev constant bound. Therefore, if we have a sequence of $\mathcal{LM}_i \in \mathscr{K}(n,A)$
with $\lambda_i>\lambda_0>0$, we can always assume
\begin{align*}
\lambda_i \to \bar{\lambda}>0, \quad \mathcal{LM} \stackrel{\hat{C}^{\infty}}{\longrightarrow} \overline{\mathcal{LM}}
\end{align*}
without considering the base points.
\begin{proposition}[\textbf{KRS limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies
\begin{align}
\lambda_i>\lambda_0>0, \quad
\mu\left(M_i, g(T_i), \frac{\lambda_i}{2} \right)-\mu\left(M_i, g(-T_i), \frac{\lambda_i}{2} \right) \to 0,
\label{eqn:HA03_2}
\end{align}
where $\mu$ is Perelman's $W$-functional. Suppose $\overline{\mathcal{LM}}$
is the limit of $\mathcal{LM}$.
Then $\overline{\mathcal{M}}$ is a gradient shrinking K\"ahler Ricci soliton. In other words, there is a smooth real valued function $\hat{f}$
defined on $\mathcal{R}(\bar{M}) \times (-\bar{T},\bar{T})$ such that
\begin{align}
\hat{f}_{jk}=\hat{f}_{\bar{j}\bar{k}}=0, \quad R_{j\bar{k}} + \hat{f}_{j\bar{k}}-\hat{g}_{j\bar{k}}=0.
\label{eqn:HB07_1}
\end{align}
\label{prn:HB07_2}
\end{proposition}
\begin{proof}
Without loss of generality, we may assume $\lambda_i=1$.
Let $\mathcal{LM} \in \mathscr{K}(n,A)$. At time $t=1$, let $u$ be the minimizer of Perelman's $\mu$-functional.
Then solve the backward heat equation $\square^* u= (-\partial_{t}-\Delta + R-n\lambda) u=0$.
Let $f$ be the function such that $(2\pi)^{-n}e^{-f}=u$. Then we have
\begin{align*}
&\quad \int_{-1}^{1}\int_M (2\pi)^{-n} \left\{\left| R_{j\bar{k}}+f_{j\bar{k}}-g_{j\bar{k}}\right|^2 + |f_{jk}|^2+|f_{\bar{j}\bar{k}}|^2\right\} e^{-f} dv\\
&\leq \mu\left(M,g(1),\frac{1}{2} \right)-\mu\left(M,g(-1),\frac{1}{2}\right)\\
&\leq \mu\left(M,g(T),\frac{1}{2} \right)-\mu\left(M,g(-T),\frac{1}{2}\right) \to 0.
\end{align*}
At time $t=1$, $f$ has good regularity estimate for it is a solution of an elliptic equation.
For $t \in (-1,1)$, we have estimate of $f$ from heat kernel estimate.
It is not hard to see that, on the space-time domain $\mathcal{R} \times (-1,1)$, $f$ converges to a limit function $\hat{f}$ satisfying (\ref{eqn:HB07_1}).
Clearly, the time interval of $(-1,1)$ can be replaced by $(-a,a)$ for every $a \in (1,\bar{T})$. For each $a$, we have a limit function
$\hat{f}^{(a)}$, which satisfies equation (\ref{eqn:HB07_1}) and therefore has enough a priori estimates.
Then let $a \to \bar{T}$ and take diagonal sequence limit, we obtain a limit function $\hat{f}^{(\bar{T})}$ which satisfies (\ref{eqn:HB07_1})
on $\mathcal{R} \times (-\bar{T}, \bar{T})$. Without loss of generality, we still denote $\hat{f}^{(\bar{T})}$ by $\hat{f}$.
Then $\hat{f}$ satisfies (\ref{eqn:HB07_1}) on $\mathcal{R} \times (-\bar{T}, \bar{T})$.
\end{proof}
\begin{remark}
It is an interesting problem to see whether $(\bar{M}, \bar{g}(0))$ is a conifold in Theorem~\ref{thm:HB07_1}.
This question has affirmative answer when we know $(\bar{M}, \bar{g}(0))$ has Einstein regular part, following the proof of
Theorem~\ref{thm:HE08_1} and Proposition~\ref{prn:SC30_1}.
In particular, the limit spaces in Proposition~\ref{prn:HB07_1} and Proposition~\ref{prn:HE15_1} are K\"ahler Einstein conifolds.
\label{rmk:HE20_1}
\end{remark}
\section{Applications}
In this section, we will focus on the applications of our structure theory to the study
of anti-canonical K\"ahler Ricci flows.
\subsection{Convergence of anti-canonical K\"ahler Ricci flows at time infinity}
Based on the structure theory, Theorem~\ref{thmin:SC24_3} can be easily proved.
\begin{proof}[Proof of Theorem~\ref{thmin:SC24_3}]
In view of the fundamental estimate of Perelman (c.f.~\cite{SeT}), in order (\ref{eqn:SK20_1}) to hold, we only need a Sobolev constant bound,
which was proved by Q. Zhang (c.f.\cite{Zhq1}) and R. Ye (c.f. \cite{Ye}).
Therefore, the truncated flow sequences locate in $\mathscr{K}(n,A)$ for a uniform $A$.
It follows from Theorem~\ref{thmin:HC06_2} that the limit K\"ahler Ricci flow exists on a compact projective normal variety.
The limit normal variety is $Q$-Fano since it has a limit anti-canonical polarization.
According to Proposition~\ref{prn:HB07_2},
the boundedness and monotonicity of Perelman's $\mu$-functional force the limit flow to be a K\"ahler Ricci soliton.
The volume estimate of $r$-neighborhood of $\mathcal{S}$ follows from Corollary~\ref{cly:HA11_1}
and estimate (\ref{eqn:SB13_1}).
\end{proof}
We continue to discuss applications beyond Theorem~\ref{thmin:SC24_3}.
The following property is well known to experts, we write it down here for the convenience of the readers.
\begin{proposition}[\textbf{Connectivity of limit moduli}]
Suppose $\mathcal{M}=\left\{ (M^n, g(t)), 0 \leq t <\infty \right\}$ is an anti-canonical K\"ahler Ricci flows on Fano manifold
$(M, J)$. Let $\mathscr{M}$ be the collection of all the possible limit space along this flow. Then $\mathscr{M}$ is connected.
\label{prn:SL06_1}
\end{proposition}
\begin{proof}
If the statement was wrong, we have two limit spaces $\bar{M}_a$ and $\bar{M}_b$, locating in different
connected components of $\mathscr{M}$. Let $\mathscr{M}_a$ be the connected component containing
$\bar{M}_a$. Since $\mathscr{M}_a$ is a connected component, it is open and closed.
So its closure $\overline{\mathscr{M}_a}$ is the
same as $\mathscr{M}_a$. Clearly, $\mathscr{M}_a$ is compact under the Gromov-Hausdorff topology.
Define
\begin{align}
&d(X,\mathscr{M}_a) \triangleq \inf_{Y \in \mathscr{M}_a} d_{GH}(X,Y), \label{eqn:SL06_7}\\
&\eta_a \triangleq \inf_{X \in \mathscr{M} \backslash \mathscr{M}_a} d(X, \mathscr{M}_a). \label{eqn:SL06_8}
\end{align}
Clearly, $\eta_a>0$ by the compactness of $\mathscr{M}_a$ and the fact that $\mathscr{M}_a$ is a connected component.
Without loss of generality, we can assume $(M, g(t_i))$ converges to $\bar{M}_a$,
$(M, g(s_i))$ converges to $\bar{M}_b$, for $t_i \to \infty$ and $s_i >t_i$.
For simplicity of notation, we denote $(M,g(t_i))$ by $M_{t_i}$, $(M,g(s_i))$ by $M_{s_i}$.
For large $i$, we have
\begin{align}
d_{GH}(M_{t_i}, \bar{M}_a)<\frac{\eta_a}{100}, \quad
d_{GH}(M_{s_i}, \bar{M}_b) < \frac{\eta_a}{100}. \label{eqn:SL06_6}
\end{align}
In particular, the above inequalities imply that
\begin{align*}
d(M_{t_i}, \mathscr{M}_a)<\frac{\eta_a}{100}, \quad d(M_{s_i}, \mathscr{M}_a)>\frac{99}{100}\eta_a.
\end{align*}
By continuity of the flow, we can find $\theta_i \in (t_i, s_i)$ such that
$d(M_{\theta_i}, \mathscr{M}_a)=\frac{1}{2}\eta_a$, whose limit form is
\begin{align}
d(\bar{M}_c, \mathscr{M}_a)=\frac{1}{2}\eta_a,
\label{eqn:SL06_10}
\end{align}
where $\bar{M}_c$ is the limit of $M_{\theta_i}$. However, (\ref{eqn:SL06_10}) contradicts with
(\ref{eqn:SL06_8}) and the fact $\eta_a>0$.
\end{proof}
Proposition~\ref{prn:SL06_1} can be generalized as follows.
\begin{proposition}[\textbf{KRS limit moduli}]
Suppose $\mathcal{M}_s=\left\{ (M_s^n, g_s(t)), 0 \leq t <\infty, s \in X \right\}$
is a smooth family of anti-canonical K\"ahler Ricci flows on Fano manifolds $(M_s, J_s)$, where
$X$ is a connected parameter space.
We call $(\bar{M}, \bar{g})$ as a limit space if $(\bar{M},\bar{g})$ is the Gromov-Hausdorff limit of $(M, g_{s_i}(t_i))$ for some
$t_i \to \infty$ and $s_i \to \bar{s} \in X$.
Suppose $\displaystyle f(s)=\lim_{t \to \infty} \mu \left(g_s(t), \frac{1}{2} \right)$ is an upper semi-continuous function on $X$.
Then we have the following properties.
\begin{itemize}
\item Every limit space is a K\"ahler Ricci soliton.
\item Let $\widetilde{\mathscr{M}}$ be the collection of all the limit spaces.
Then $\widetilde{\mathscr{M}}$ is connected under the Gromov-Hausdorff topology.
\end{itemize}
\label{prn:SL07_2}
\end{proposition}
\begin{proof}
We shall only show that every limit space is a K\"ahler Ricci soliton.
The connectedness of $\widetilde{\mathscr{M}}$ can be proved almost the same as Proposition~\ref{prn:SL06_1}.
So we leave the details to the readers.
Suppose $s_i \to \bar{s}$. Fix $\epsilon$, we can choose $T_{\epsilon}$ such that
\begin{align*}
\mu\left(g_{\bar{s}}(T_{\epsilon}), \frac{1}{2} \right)>f_{\bar{s}}-\epsilon.
\end{align*}
By the smooth convergence of $g_{s_i}(T_{\epsilon})$ and the upper semi-continuity of $f$, we have
\begin{align*}
\mu\left(g_{s_i}(T_{\epsilon}), \frac{1}{2} \right) >f_{s_i}-\epsilon
\end{align*}
for large $i$. Recall that $t_i \to \infty$. Therefore, it follows from the monotonicity of Perelman's functional that
\begin{align*}
\mu\left(g_{s_i}(T_{\epsilon}), \frac{1}{2} \right)<\mu\left(g_{s_i}(t_i-1), \frac{1}{2} \right)<\lim_{t \to \infty}\mu\left(g_{s_i}(t), \frac{1}{2} \right)=f_{s_i}.
\end{align*}
Hence, we have
\begin{align*}
0\leq \mu\left(g_{s_i}(t_i+1), \frac{1}{2} \right)-\mu\left(g_{s_i}(t_i-1), \frac{1}{2} \right) <\epsilon
\end{align*}
for large $i$. By the arbitrary choice of $\epsilon$, we obtain
\begin{align*}
\lim_{i \to \infty} \left\{ \mu\left(g_{s_i}(t_i+1), \frac{1}{2} \right)-\mu\left(g_{s_i}(t_i-1), \frac{1}{2} \right) \right\}=0.
\end{align*}
Therefore, $(M, g_{s_i}(t_i))$ converges to a K\"ahler Ricci soliton, in light of Proposition~\ref{prn:HB07_2}.
\end{proof}
The gap between singularity and regularity in Theorem~\ref{thmin:SC24_1} has a global version as follows.
\begin{proposition}[\textbf{Gap around smooth KE}]
Suppose $(\tilde{M},\tilde{g},\tilde{J})$ is a compact, smooth K\"ahler Einstein manifold which belongs to $\mathscr{K}(n,A)$
when regarded as a trivial polarized K\"ahler Ricci flow solution.
Then there exists an $\epsilon=\epsilon(n,A,\tilde{g})$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$ and $ d_{GH}((\tilde{M},\tilde{g}), (M,g(0))) < \epsilon$, then we have
\begin{align*}
\mathbf{vcr}(M, g(0)) >\frac{1}{2}\mathbf{vcr}(\tilde{M},\tilde{g}).
\end{align*}
\label{prn:SL07_3}
\end{proposition}
\begin{proof}
It follows from the continuity of canonical volume radius under the Cheeger-Gromov convergence.
\end{proof}
Proposition~\ref{prn:SL07_3} means that there is no singular limit space around any given smooth K\"ahler Einstein manifold.
Clearly, the single smooth K\"ahler Einstein manifold in this Proposition
can be replaced by a family of smooth K\"ahler Einstein manifolds with bounded geometry.
The gap between smooth and singular K\"ahler Einstein metrics can be conveniently used to carry out topology argument.
\begin{theorem}[\textbf{Convergence of KRF family}]
Suppose $\mathcal{M}_s=\left\{ (M_s^n, g_s(t), J_s), 0 \leq t <\infty, s \in X \right\}$
is a smooth family of anti-canonical K\"ahler Ricci flows on Fano manifolds $(M_s, J_s)$, where
$X$ is a connected parameter space. Moreover, we assume that
\begin{itemize}
\item The Mabuchi's K-energy is bounded from below along each flow.
\item Smooth K\"ahler Einstein metrics in all adjacent complex structures(c.f. Defintion 1.4 of~\cite{CS1}) have uniformly bounded Riemannian curvature.
\end{itemize}
Let $\Omega$ be the collection of $s$ such that the flow
$g_s$ has bounded Riemannian curvature. Then $\Omega=\emptyset$ or $\Omega=X$.
\label{thm:HC03_2}
\end{theorem}
\begin{proof}
It suffices to show that $\Omega$ is both open and closed in $X$.
The openness follows from the stability of K\"ahler Ricci flow around a given smooth K\"ahler Einstein metric, due to Sun and Wang (c.f.~\cite{SW}).
Suppose $s \in \Omega$, then the flow $g_s$ converges to some K\"ahler Einstein manifold $(M',g',J')$, which is the unique
K\"ahler Einstein metric in its small smooth neighborhood. By continuous dependence of flow on the initial data, and the stability of K\"ahler Ricci flow in a
very small neighborhood of $(M',g',J')$,
it is clear that $s$ has a neighborhood consisting of points in $\Omega$. Therefore, $\Omega$ is an open subset of $X$.
The closedness follows from Proposition~\ref{prn:SL07_2}.
Suppose $s_i \in \Omega$ and $s_i \to \bar{s} \in X$.
Due to the fact that the Mabuchi's K-energy is bounded from below along each K\"ahler Ricci flow we are concerning now, the
limit Perelman functional is always the same(c.f.\cite{CS1}). Therefore, we can apply Proposition~\ref{prn:SL07_2} to show
that every limit space is a possibly singular K\"ahler Einstein. However, along every $g_{s_i}$, we obtain a smooth
limit K\"ahler Einstein manifold
$(M',g',J')$, which has uniformly bounded curvature, as a K\"ahler Einstein manifold in an adjacent complex structure.
Note that the diameter of $M'$ is uniformly bounded by Myers theorem.
The volume of $M'$ is a topological constant. Therefore, the geometry of $(M',g')$ are uniformly bounded.
By a generalized version of Proposition~\ref{prn:SL07_3}, $(M',g',J')$ is uniformly bounded away from singular K\"ahler Einstein metrics.
Due to Proposition~\ref{prn:SL07_2}, the connectedness of $\mathscr{M}$ forces that the flow
$g_{\bar{s}}$ must converge to a smooth $(M',g',J')$. In particular, $g_{\bar{s}}$ has bounded curvature.
Threrefore, $\bar{s} \in \Omega$ and $\Omega$ is closed.
\end{proof}
The two assumptions in Theorem~\ref{thm:HC03_2} seem to be artificial. However, if $J_s$ is a trivial family or a test configuration family, by the unique
degeneration theorem of Chen-Sun(c.f.~\cite{CS1}), all the smooth K\"ahler Einstein metrics form an isolated family,
then the second condition is satisfied automatically. On the other hand, by the existence of K\"ahler Einstein metrics in the weak sense,
one can also obtain the lower bound of Mabuchi's $K$-energy(c.f.~\cite{BM},~\cite{DT},~\cite{C08}).
Consequently, Theorem~\ref{thm:HC03_2} can be applied to these special cases and obtain the following corollaries.
\begin{corollary}[\textbf{Convergence to given KE}, c.f. Tian-Zhu~\cite{TZ1}, Collins-Sz\'{e}kelyhidi~\cite{CoSz}]
Suppose $(M,J)$ is a Fano manifold with a K\"ahler Einstein metric $g_{KE}$. Then every anti-canonical K\"ahler Ricci flow on $(M,J)$ converges to
$(M,g_{KE},J)$.
\label{cly:HC04_0}
\end{corollary}
\begin{proof}
Let $\omega_{KE}$ be the K\"ahler Einstein metric form. Then every metric form $\omega$ can be written as $\omega_{KE}+\sqrt{-1} \partial \bar{\partial}\varphi$
for some smooth function $\varphi$. Define
\begin{align*}
\omega_s= \omega_{KE} + s \sqrt{-1} \partial \bar{\partial} \varphi, \quad s \in [0,1].
\end{align*}
It follows from Theorem~\ref{thm:HC03_2} that the K\"ahler Ricci flow from every $\omega_s$ has bounded curvature, and consequently
converges to $\omega_{KE}$, by the uniqueness theorem of Chen-Sun(c.f.~\cite{CS1}). In particular, the flow
start from $\omega$ converges to $\omega_{KE}$.
\end{proof}
\begin{corollary}[\textbf{Convergence of a test configuration}]
Suppose $\mathbf{M}$ is a smooth test configuration, i.e., a family of Fano manifolds $(M_s,J_s)$
parameterized by $s$ in unit disk $D \subset \ensuremath{\mathbb{C}}^1$ with a natural $C^*$-action.
Suppose each fiber is smooth and the central fiber $(M_0,g_0,J_0)$
admits K\"ahler Einstein metric $(M_0,g_{KE},J_0)$. Then each K\"ahler Ricci flow starting from $(M_s, g_s,J_s)$ for arbitrary $s \in D$
converges to $(M_0, g_{KE},J_0)$.
\label{cly:HC04_1}
\end{corollary}
\begin{proof}
Theorem~\ref{thm:HC03_2} can be applied for $X=D$.
The central K\"ahler Ricci flow converges by Corollary~\ref{cly:HC04_0}. Therefore, the K\"ahler Ricci flow on each fiber has bounded curvature and converge to
some smooth K\"ahler Einstein metric, which can only be $(M_0,g_{KE},J_0)$, due to the uniqueness theorem of Chen-Sun again.
\end{proof}
\begin{remark}
Corollary~\ref{cly:HC04_0} was announced by G.Perelman. The first written proof was given by Tian-Zhu in~\cite{TZ1} whenever there is no non-trivial holomorphic vector field.
The general case was proved by Collins-Sz\'{e}kelyhidi in~\cite{CoSz}.
The strategy of Corollary~\ref{cly:HC04_0} was inspired by that in~\cite{TZ2}.
Corollary~\ref{cly:HC04_0}-Corollary~\ref{cly:HC04_1} have the corresponding K\"ahler Ricci soliton versions. These generalizations will be discussed in a separate paper.
\label{rmk:HC09_1}
\end{remark}
\subsection{Degeneration of anti-canonical K\"ahler Ricci flows}
In this subsection, we shall prove Theorem~\ref{thmin:HA01_1} and related corollaries.
The following Theorem is due to Jiang(c.f.~\cite{Jiang}).
\begin{theorem}[\textbf{Jiang's estimate}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying \begin{align}
\norm{Ric^{-}}{C^0(M)} + |\log \Vol(M)| + C_S(M, g(0)) \leq F
\label{eqn:HA01_1}
\end{align}
at time $t=0$. Then we have
\begin{align}
|R|+ |\nabla \dot{\varphi}|^2 \leq \frac{C}{t^{n+1}}
\label{eqn:HA01_2}
\end{align}
for some constant $C=C(n,F)$.
\label{thm:HA01_1}
\end{theorem}
Note that (\ref{eqn:HA01_2}) implies a uniform bound of diameter at each time $t>0$, by the uniform bound of
Perelman's functional. Then one can easily deduce a uniform bound (depending on $t$) of $\norm{\dot{\varphi}}{C^1(M)}$.
Combing this with the Sobolev constant estimate along the flow(c.f.~\cite{Zhq1},~\cite{Ye}), we see that
\begin{align}
\norm{R}{C^0(M)} + \norm{\dot{\varphi}}{C^1(M)} + C_S(M, g(t)) \leq C(n,F,t)
\label{eqn:HA01_3}
\end{align}
for each $t>0$. Therefore, away from the initial time, we can always apply our structure theory.
\begin{theorem}[\textbf{Weak convergence with initial time}]
Suppose $\mathcal{M}_i=\{(M_i^n,g_i(t), J_i), 0 \leq t <\infty\}$ is a sequence of anti-canonical K\"ahler Ricci flow
solutions, whose initial time slices satisfy estimate (\ref{eqn:HA01_1}) uniformly. Then we have
\begin{align}
(\mathcal{M}_i, g_i) \stackrel{G.H.}{\longrightarrow} (\bar{\mathcal{M}}, \bar{g}),
\end{align}
where the limit is a weak K\"ahler Ricci flow solution on a $Q$-Fano normal variety $\bar{M}$, for time $t>0$.
Moreover, the convergence can be improved to
be in the $\hat{C}^{\infty}$-Cheeger-Gromov topology for each $t>0$, i.e.,
\begin{align}
(M_i, g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M}(t), \bar{g}(t))
\end{align}
for each $t>0$.
\label{thm:HA01_2}
\end{theorem}
Clearly, if $(M_i,g_i)$ is a sequence of almost K\"ahler Einstein metrics(c.f.~\cite{TW2}) in the anti-canonical classes, then
$(\bar{M}(0), \bar{g}(0))$ and $(\bar{M}(1), \bar{g}(1))$ are isometric to each other, due to Proposition~\ref{prn:HB07_1} and the estimate in~\cite{TW2}.
In this particular case, it is easy to see that partial-$C^0$-estimate holds uniformly at time $t=0$ for each $i$, at least intuitively.
Actually, by the work Jiang~\cite{Jiang}, it is now clear that partial-$C^0$-estimate at time $t=0$ only requires a uniform Ricci lower bound.
Note that the evolution equation of the anti-canonical K\"ahler Ricci flow is
\begin{align}
\dot{\varphi}=\log \frac{\omega_{\varphi}^n}{\omega^n} + \varphi - u_{\omega},
\label{eqn:HA01_5}
\end{align}
where $u_{\omega}$ is the Ricci potential satisfying the normalization condition
$\int_M e^{-u_{\omega}} \frac{\omega^n}{n!}=(2\pi)^n$.
By maximum principle and Green function argument, we have the following property(c.f.~\cite{Jiang}).
\begin{proposition}[\textbf{Potential equivalence}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying
(\ref{eqn:HA01_1}). At time $t=0$, let $\varphi=0$ and $u_{\omega}$ satisfy the normalization condition. Then we have
\begin{align}
C(1-e^{t}) \leq \varphi \leq C e^t
\label{eqn:HA01_4}
\end{align}
for a constant $C=C(n,F)$.
\label{prn:HA01_3}
\end{proposition}
Let $\mathbf{b}(\cdot, t)$ be the Bergman function at time $t$. By definition, at point $x \in M$ and time $t=0$,
we can find a holomorphic section $S \in H^0(M, K_M^{-1})$ such that
\begin{align*}
\int_{M} \norm{S}{h(0)}^2 \frac{\omega^n}{n!}=1, \quad\mathbf{b}(x,0)=\log \norm{S}{h(0)}^2(x).
\end{align*}
Note that $\norm{S}{h(1)}^2=\norm{S}{h(0)}^2e^{-\varphi(1)}$.
By (\ref{eqn:HA01_4}), it is clear that $\norm{S}{h(1)}^2$ and $\norm{S}{h(0)}^2$ are uniformly equivalent.
On the other hand, $\Delta \norm{S}{}^2 \geq -n \norm{S}{}^2$. At time $t=0$, applying Moser iteration implies that $\norm{S}{h(0)}^2 \leq C$.
Hence we obtain $\norm{S}{h(1)}^2 \leq C$.
At time $t=1$, let $\tilde{S}$ be the normalization of $S$, i.e., $\tilde{S}=\lambda S$ such that
$\int_{M} \norm{\tilde{S}}{h(1)}^2 \frac{\omega_1^n}{n!}=1$. Then we have
\begin{align*}
\lambda^{-2}= \int_{M} \norm{S}{h(1)}^2 \frac{\omega_1^n}{n!} \leq C.
\end{align*}
It follows that
\begin{align*}
\mathbf{b}(x,1) &\geq \log \norm{\tilde{S}}{h(1)}^2(x)= \log \norm{S}{h(1)}^2(x)+ \log \lambda^2\\
&=\log \norm{S}{h(0)}^2(x) -\varphi(1) + \log \lambda^2\\
&=\mathbf{b}(x,0)-\varphi(1) + \log \lambda^2\\
&\geq \mathbf{b}(x,0)-C.
\end{align*}
By reversing time, we can obtain a similar inequality with reverse direction.
Same analysis applies to $\mathbf{b}^{(k)}$ for each positive integer $k$.
So we have the following property.
\begin{proposition}[\textbf{Bergman function equivalence}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying
(\ref{eqn:HA01_1}). For each positive integer $k$, there exists $C=C(n,F,k)$ such that
\begin{align}
\mathbf{b}^{(k)}(x,0) -C \leq \mathbf{b}^{(k)}(x,1) \leq \mathbf{b}^{(k)}(x,0)+C
\label{eqn:HE27_1}
\end{align}
for all $x \in M$.
\label{prn:HA01_4}
\end{proposition}
In view of Theorem~\ref{thm:HA01_2}, partial-$C^0$-estimate holds at time $t=1$,
which induces the partial-$C^0$-estimate at time $t=0$, by Proposition~\ref{prn:HA01_4}.
Therefore, the following theorem is clear now.
\begin{theorem}[\textbf{Partial-$C^0$-estimate at initial time}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying
(\ref{eqn:HA01_1}). Then
\begin{align*}
\inf_{x \in M} \mathbf{b}^{(k_0)}(x,0) \geq -c_0
\end{align*}
for some positive integer $k_0=k_0(n,F)$ and positive number $c_0=c_0(n,F)$.
\label{thm:HA01_3}
\end{theorem}
By the Sobolev constant estimates for manifolds with uniform positive Ricci curvature, it is clear that
Theorem~\ref{thmin:HA01_1} follows from Theorem~\ref{thm:HA01_3} directly.
It is also clear that Corollary~\ref{clyin:HA01_2} follows from Theorem~\ref{thm:HA01_3}.
The proof of Corollary~\ref{clyin:HA03_1} is known in literature(c.f.~\cite{Sze}), provided the partial-$C^0$-estimate along
the K\"ahler Ricci flow. We shall be sketchy here.
In fact, due to the work of S. Paul(\cite{Paul1209},\cite{Paul1308}) and the argument in section 6 of Tian and Zhang(\cite{TZZ2}), one obtains that the $I$-functional is bounded along the flow.
Then the K\"ahler Ricci flow converges to a K\"ahler-Einstein metric, on the same Fano manifold.
It is an interesting problem to study the $K$-stability through the K\"ahler Ricci flow.
Based on Theorem~\ref{thmin:SC24_3}, the weak compactness of polarized K\"ahler Ricci flow, we are able to give an alternative K\"ahler Ricci flow proof of the stability theorem(Yau's conjecture) of Chen-Donaldson-Sun.
Interested readers are referred to~\cite{CSW} for the details.
\section{Introduction}
This paper is the continuation of the study in (\cite{CW3}) and (\cite{CW4}). In~\cite{CW3}, we developed
a weak compactness theory for non-collapsed Ricci flows with bounded scalar curvature and bounded half-dimensional curvature integral. This weak compactness theory is applied in~\cite{CW4} to prove the Hamilton-Tian conjecture of
complex dimension 2 and its geometric consequences.
However, the assumption of half dimensional curvature integral is restrictive.
It is not available for high dimensional anti-canonical K\"ahler Ricci flow, i.e., K\"ahler Ricci flow on a Fano manifold $(M,J)$, in the class $2\pi c_1(M,J)$.
In this paper, by taking advantage of the extra structures from K\"ahler geometry, we drop this curvature integral condition.
The present paper draw inspirations from two different sources.
One source is the structure theory of K\"ahler Einstein manifolds which was developed over last 20 years by many people,
notably, Anderson, Cheeger, Colding, Tian and more recently, Naber, Donaldson and Sun. The recent progress of the structure theory of
K\"ahler Einstein manifolds supplies many additional tools for our approach.
The other source is the seminal work of Perelman on the Ricci flow(c.f.~\cite{Pe1},~\cite{SeT}).
Actually, it was pointed out by Perelman already that his idea in \cite{Pe1} can be applied to study K\"ahler Ricci flow.
He said that
``\textit{present work has also some applications to the
Hamilton-Tian conjecture concerning K\"ahler-Ricci flow on K\"ahler manifold with
positive first Chern class: these will be discussed in a separate paper}".
\noindent We cannot help to wonder how far he will push the subject of Ricci flow if he continued to publicize
his works on arxiv.
Although ``\textit{this separate paper}" never appears, his fundamental estimates of K\"ahler Ricci flow on Fano manifolds is the base of our present
research. Besides Perelman's estimates,
we also note that the following technical results in the Ricci flow are important to the formation of this paper
over a long period of time: the Sobolev constant estimate by Q.S. Zhang(\cite{Zhq1}) and R. Ye (\cite{Ye}),
and the volume ratio upper bound estimate by Q.S. Zhang(\cite{Zhq3}) and Chen-Wang(\cite{CW5}).
Our key observation is that there is a ``canonical neighborhood" theorem for anti-canonical K\"ahler Ricci flows.
The idea of ``canonical neighborhood" originates from Theorem 12.1 of Perelman's paper~\cite{Pe1}.
For every 3-dimensional Ricci flow, Perelman showed that the space-time neighborhood of a high curvature point can be approximated by
a $\kappa$-solution, which is a model Ricci flow solution.
To be precise, a $\kappa$-solution is a 3-dimensional, $\kappa$-noncollapsed, ancient Ricci flow solution with bounded, nonnegative curvature operator.
By definition, it is not clear at all that the moduli of $\kappa$-solutions has compactness under (pointed-) smooth topology (modulo diffeomorphisms).
Perelman genuinely proved the compactness by delicate use of Hamilton-Ivey pinch and the geometry of nonnegatively curved 3-manifolds.
In light of the compactness of the moduli of $\kappa$-solutions, by a maximum principle type argument,
Perelman developed the ``canonical neighborhood" theorem,
which is of essential importance to his celebrated
solution of the Poincar\`{e} conjecture(c.f.~\cite{KL},~\cite{MT},~\cite{CZ}).
The idea of ``canonical neighborhood" is universal and can be applied in many different geometric settings.
In particular, there is a ``canonical neighborhood" theorem
for the anti-canonical K\"ahler Ricci flows, where estimates of many quantities, including scalar curvature, Ricci potential and Sobolev constant,
are available. Clearly, a ``canonical neighborhood" should be a neighborhood in space-time, behaving like a model space-time,
which is more or less the blowup limit of the given flow.
Therefore, it is natural to expect that the model space-time is the scalar flat Ricci flow solutions,
which must be Ricci flat, due to the
equation $ \frac{\partial}{\partial t} R= \Delta R + 2|Ric|^2$, satisfied by the scalar curvature $R$.
For this reason, the model space and model space-time can be identified, since the evolution on time direction is trivial.
It is also natural to expect that the model space has some K\"ahler structure. In other words, the model space should
be K\"ahler Ricci flat space, or Calabi-Yau space.
Now the first essential difficulty appears. A good model space should have a compact moduli.
For example, in the case of 3-dimensional Ricci flow, the moduli space of $\kappa$-solutions, which are the model space-times, has compactness in the smooth topology.
However, the moduli space of all the non-collapsed smooth Calabi-Yau space-times
is clearly not compact under the smooth topology.
A blowdown sequence of Eguchi-Hanson metrics is an easy example.
For the sake of compactness,
we need to replace the smooth topology by a weaker topology,
the pointed-Cheeger-Gromov topology.
At the same time,
we also need to enlarge the class of model spaces from complete Calabi-Yau manifolds to the Calabi-Yau spaces with mild singularities(c.f.~Definition~\ref{dfn:SC27_1}),
which we denote by $\widetilde{\mathscr{KS}}(n,\kappa)$.
Similar to the compactness theorem of Perelman's $\kappa$-solutions,
we have the compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{theoremin}[\textbf{Compactness of model moduli}]
$\widetilde{\mathscr{KS}}(n,\kappa)$ is compact under the pointed Cheeger-Gromov topology.
Moreover, each space $X \in \widetilde{\mathscr{KS}}(n,\kappa)$ is a Calabi-Yau conifold.
\label{thmin:HE21_1}
\end{theoremin}
The notion of conifold is well known to string theorist as
some special Calabi-Yau 3-folds with singularities(c.f.~\cite{Green}).
In this paper, by abusing notation,
we use it to denote a space whose singular part admits cone type tangent spaces.
The precise definition is given in Definition~\ref{dfn:HD20_1}.
Note that Calabi-Yau conifold is a generalization of Calabi-Yau orbifold.
The strategy to prove the compactness of
$\widetilde{\mathscr{KS}}(n,\kappa)$ follows the same route of the weak compactness theory
of K\"ahler Einstein manifolds, developed by Cheeger, Gromoll, Anderson, Colding, Tian, Naber, etc.
However, the analysis foundation on the singular spaces need to be carefully checked, which is done in section 2.
Theorem~\ref{thmin:HE21_1} is motivated by section 11 of Perelman's seminal paper~\cite{Pe1},
where Perelman proved the compactness of moduli space of $\kappa$-solutions and showed that
$\kappa$-solutions have many properties which are not obvious from definition.
By trivial extension, each
$X \in \widetilde{\mathscr{KS}}(n,\kappa)$ can be understood as a space-time $X \times (-\infty, \infty)$
satisfying Ricci flow equation.
Intuitively, the rescaled space-time structure in a given anti-canonical K\"ahler Ricci flow
should behave similar to that of $X \times (-\infty, \infty)$ for some $X \in \widetilde{\mathscr{KS}}(n,\kappa)$,
when the rescaling factor is large enough.
In order to make sense that two space-times are close to each other,
we need the Cheeger-Gromov topology for space-times, a slight generalization of the Cheeger-Gromov
topology for metric spaces.
When restricted on each time slice, this topology is the same as the usual Cheeger-Gromov topology.
Between every two different time slices, there is a natural homeomorphism map connecting them.
Therefore, the above intuition can be realized if we can show a sequence of Ricci flow
space-times blownup from a given K\"ahler Ricci flow converges to a limit space-time $X \times (-\infty, \infty)$,
in the pointed Cheeger-Gromov topology for space-times.
However, it is not easy to obtain the homeomorphism maps between different time slices in the limit.
Although it is quite obvious to guess that the homeomorphism maps among different time slices
are the limit of identity maps,
there exists serious technical difficulty to show the existence and regularity of the limit maps.
The difficulty boils down to a fundamental improvement of Perelman's pseudolocality theorem(Theorem 10.1 of~\cite{Pe1}).
Recall that Perelman's pseudolocality theorem says that Ricci flow cannot ``quickly" turn an almost Euclidean region into a very curved one.
It is a short-time, one-sided estimate in nature. We need to improve it to a long-time, two-sided estimate.
Not surprisingly, the rigidity of K\"ahler geometry plays an essential role for such an improvement.
The two-sided, long-time pseudolocality is an estimate in the time direction.
Modulo this time direction estimate and the weak compactness in the space direction, we can take limit for
a sequence of Ricci flows blownup from a given flow.
Then the canonical neighborhood theorem can be set up if we can show that the limit space-time locates in
$\widetilde{\mathscr{KS}}(n,\kappa)$,
following the same route as that in the proof of Theorem 12.1 of~\cite{Pe1}. \\
From the above discussion, it is clear that the strategy to prove the canonical neighborhood theorem is simple.
However, the technical difficulty hidden behind this simple strategy is not that simple.
We observe that the anti-canonical
K\"ahler Ricci flow has many additional structures, all of them should be used to carry out the proof of
the canonical neighborhood theorem.
In particular, over every anti-canonical K\"ahler Ricci flow, there is a natural anti-canonical polarization, which should play an important role, as done in~\cite{CW4}.
Although we are aiming at the anti-canonical case,
in this paper, however, we shall consider flows with more general polarizations.
We call $\mathcal{LM}=\left\{ (M^n, g(t), J, L, h(t)), t \ \in (-T,T) \subset \ensuremath{\mathbb{R}} \right\}$ a polarized K\"ahler Ricci flow if
\begin{itemize}
\item $\mathcal{M}=\left\{ (M^n, g(t), J), t \ \in (-T,T) \right\}$ is a K\"ahler Ricci flow solution.
\item $L$ is a Hermitian line bundle over $M$, $h(t)$ is a family of smooth metrics on $L$ whose curvature is $\omega(t)$, the metric form compatible with
$g(t)$ and the complex structure $J$.
\end{itemize}
Clearly, the first Chern class of $L$ is $[\omega(t)]$, which does not depend on time. So a polarized K\"ahler Ricci flow stays in a fixed integer K\"ahler class.
The evolution equation of $g(t)$ can be written as
\begin{align}
\frac{\partial}{\partial t} g_{i\bar{j}}=-R_{i\bar{j}}+ \lambda g_{i\bar{j}}, \label{eqn:K07_1}
\end{align}
where $\lambda=\frac{c_1(M)}{c_1(L)}$.
Since the flow stays in the fixed class, we can let $\omega_t=\omega_0 + \sqrt{-1} \partial \bar{\partial} \varphi$. Then $\dot{\varphi}$
is the Ricci potential, i.e.,
\begin{align*}
\sqrt{-1} \partial \bar{\partial} \dot{\varphi} = -Ric + \lambda g.
\end{align*}
Note the choice of $\varphi$ is unique up to adding a constant. So we can always modify the choice of $\varphi$ such that $\displaystyle \sup_M \dot{\varphi}=0$.
For simplicity, we denote $\mathscr{K}(n,A)$ as the collection of all the polarized K\"ahler Ricci flows
$\mathcal{LM}$ satisfying the following estimate
\begin{align}
\begin{cases}
&T \geq 2, \\
&C_S(M)+\frac{1}{\Vol(M)}+|\dot{\varphi}|_{C^0(M)} + |R-n\lambda|_{C^0(M)} \leq A,
\quad \textrm{for every time} \quad t \in (-T,T).
\end{cases}
\label{eqn:SK20_1}
\end{align}
Here $C_S$ means the Sobolev constant, $A$ is a uniform constant. In this paper, we study the structure of
polarized K\"ahler Ricci flows locating in the space $\mathscr{K}(n,A)$.
The motivation behind (\ref{eqn:SK20_1}) arises from the fundamental estimate of diameter, scalar curvature,
$C^1$-norm of Ricci potential, and Sobolev constant
along the anti-canonical K\"ahler Ricci flows(c.f.~\cite{SeT},~\cite{Zhq1},~\cite{Ye}).
Every polarized K\"ahler Ricci flow solution in $\mathscr{K}(n,A)$ has at least three structures: the metric space structure, the flow structure, the line bundle structure.
Same structures can be discussed on the model space-time in $\widetilde{\mathscr{KS}}(n,\kappa)$.
All the structures of a flow in $\mathscr{K}(n,A)$ can be modeled by the corresponding structures
in $\widetilde{\mathscr{KS}}(n,\kappa)$, which is the same meaning as the ``canonical neighborhood theorem".
We shall compare these structures term by term. \\
Under the (pointed-)Cheeger-Gromov topology at time $0$, let us compare the metric structure of a flow in
$\mathscr{K}(n,A)$ with a Calabi-Yau conifold in $\widetilde{\mathscr{KS}}(n,\kappa)$.
We shall show that $\mathscr{K}(n,A)$ and $\widetilde{\mathscr{KS}}(n,\kappa)$ behaves almost the same in this
perspective.
For simplicity of notation, we use $\stackrel{G.H.}{\longrightarrow}$ to denote the convergence in Gromov-Hausdorff topology. We use $\stackrel{\hat{C}^k}{\longrightarrow}$ to denote
the $C^k$-Cheeger-Gromov topology, i.e., the convergence is in the Gromov-Hausdorff topology,
and can be improved to be in $C^k$-topology (modulo diffeomorphisms) away from singularities.
We call a point being regular if it has a neighborhood with smooth manifold structure and call a point being singular if it is not regular(c.f. Proposition~\ref{prn:HA08_1} and Remark~\ref{rmk:HA07_1}).
\begin{theoremin}[\textbf{Metric space estimates}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$. Then we have
\begin{align*}
(M_i, x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M}, \bar{x}, \bar{g}).
\end{align*}
The limit space $\bar{M}$ has a classical regular-singular decomposition $\mathcal{R} \cup \mathcal{S}$ with the following properties.
\begin{itemize}
\item $\left(\mathcal{R}, \bar{g}\right)$ is a smooth, open Riemannian manifold. Moreover,
$\mathcal{R}$ admits a limit K\"ahler structure $\bar{J}$ such that $\left(\mathcal{R}, \bar{g}, \bar{J} \right)$ is an open K\"ahler manifold.
\item $\mathcal{S}$ is a closed set and $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$, where $\mathcal{M}$ means Minkowski dimension(c.f.~Definition~\ref{dfn:HE08_1}).
\item Every tangent space of $\bar{M}$ is an irreducible metric cone.
\item Let $\mathrm{v}$ be the volume density, i.e.,
\begin{align}
\mathrm{v}(y)=\limsup_{r \to 0} \omega_{2n}^{-1}r^{-2n}|B(y,r)|
\end{align}
for every point $y \in X$.
Then a point is regular if and only if $\mathrm{v}(y)=1$, a point is singular if and only if
$\mathrm{v}(y) \leq 1-2\delta_0$, where $\delta_0$ is a dimensional
constant determined by Anderson's gap theorem.
\end{itemize}
\label{thmin:SC24_1}
\end{theoremin}
The line bundle structure of $\mathscr{K}(n,A)$ is more or less dominated by the Bergman function.
Actually, for every positive integer $k$, we define the Bergman function $\mathbf{b}^{(k)}$ as follows
\begin{align}
\mathbf{b}^{(k)}(x,t)=\log \sum_{i=0}^{N_k} \norm{S_i^{(k)}}{h(t)}^2(x,t),
\label{eqn:HA03_3}
\end{align}
where $N_k=\dim_{\ensuremath{\mathbb{C}}} H^0(M, L^k)-1$, $\left\{S_i^{(k)} \right\}_{i=0}^{N_k}$ are orthonormal basis of $H^0(M, L^k)$ under the natural metrics $\omega(t)$ and $h(t)$.
Theorem~\ref{thmin:SC24_1} means that the metric structure of the center time slice of a K\"ahler Ricci flow in $\mathscr{K}(n,A)$ can be modeled by
non-collapsed Calabi-Yau manifolds with mild singularities. In particular, each tangent space of a point in the limit space is a metric cone.
The trivial line bundle structure on metric cone then implies an estimate of line bundle structure of the original manifold, due to
delicate use of H\"omander's $\bar{\partial}$-estimate, as done by Donaldson and Sun(c.f.~\cite{DS}).
\begin{theoremin}[\textbf{Line bundle estimates}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, then
\begin{align*}
\inf_{x \in M} \mathbf{b}^{(k_0)}(x,0) \geq -c_0
\end{align*}
for some positive number $c_0=c_0(n,A)$, and positive integer $k_0=k_0(n,A)$.
\label{thmin:HC08_1}
\end{theoremin}
In other words, Theorem~\ref{thmin:HC08_1} states that there is a uniform partial-$C^0$-estimate at time $t=0$. This estimate then implies variety structure of
limit space, as discussed in~\cite{Tian12} and~\cite{DS}.
Theorem~\ref{thmin:HC08_1} can be understood that the line bundle structure of $\mathscr{K}(n,A)$ is modeled after
that of $\widetilde{\mathscr{KS}}(n,\kappa)$.
Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1} deal only with one time slice.
In order to make sense of limit K\"ahler Ricci flow, we have to compare the limit spaces of different time slices.
For example, we choose $x_i \in M_i$, then we have
\begin{align*}
(M_i, x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M}, \bar{x}, \bar{g}), \quad (M_i, x_i, g_i(-1)) \longright{\hat{C}^{\infty}} (\bar{M}', \bar{x}', \bar{g}').
\end{align*}
How are $\bar{M}$ and $\bar{M}'$ related? If $\bar{x}$ is a regular point of $\bar{M}$, can we say $\bar{x}'$ is a regular point of $\bar{M}'$?
Note that Perelman's pseudolocality theorem cannot answer this question, due to its short-time, one-sided property. In order to relate different time slices,
we need to improve Perelman's pseudolocality theorem to the following long-time, two-sided estimate,
which is the technical core of the current paper.
\begin{theoremin}[\textbf{Time direction estimates}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Suppose $x_0 \in M$, $\Omega=B_{g(0)}(x_0,r)$, $\Omega'=B_{g(0)}(x_0, \frac{r}{2})$.
At time $t=0$, suppose the isoperimetric constant estimate $\mathbf{I}(\Omega) \geq (1-\delta_0) \mathbf{I}(\ensuremath{\mathbb{C}}^n)$ holds for $\delta_0=\delta_0(n)$,
the same constant in Theorem~\ref{thmin:SC24_1}. Then we have
\begin{align*}
|\nabla^k Rm|(x,t) \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0}, \quad x \in \Omega',
\quad t \in [-1,1],
\end{align*}
where $C_k$ is a constant depending on $n,A,r$ and $k$.
\label{thmin:HC06_1}
\end{theoremin}
Theorem~\ref{thmin:HC06_1} holds trivially on each space in $\widetilde{\mathscr{KS}}(n,\kappa)$, when regarded as
a static Ricci flow solution. Therefore, it can be understood as the time direction structure, or the flow structure of
$\mathcal{LM} \in \mathscr{K}(n,A)$ is similar to that of $\widetilde{\mathscr{KS}}(n,\kappa)$.
Theorem~\ref{thmin:HC06_1} removes the major stumbling block for defining a limit K\"ahler Ricci flow,
since it guarantees that the regular-singular decomposition of the limit space is independent of time.
Therefore, there is a natural induced K\"ahler Ricci flow structure on the regular part of the limit space. We denote its completion
by a limit K\"ahler Ricci flow solution, in a weak sense. Clearly, the limit K\"ahler Ricci flow naturally inherits a limit line
bundle structure, or a limit polarization, on the regular part.
Moreover, the limit underlying space does have a variety structure due to Theorem~\ref{thmin:HC08_1}.
With these structures in hand, we are ready to discuss the convergence theorem of polarized K\"ahler Ricci flows,
which is the main structure theorem of this paper.
\begin{theoremin}[\textbf{Weak compactness of polarized flows}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Then we have
\begin{align*}
\left(\mathcal{LM}, x_i \right) \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left(\overline{\mathcal{LM}}, \bar{x}\right),
\end{align*}
where $\overline{\mathcal{LM}}$ is a polarized K\"ahler Ricci flow solution on an analytic normal variety $\bar{M}$,
whose singular set $\mathcal{S}$ has Minkowski codimension at least $4$, with respect to each $\bar{g}(t)$.
Moreover, if $\bar{M}$ is compact, then it is a projective normal variety with at most log-terminal singularities.
\label{thmin:HC06_2}
\end{theoremin}
As it is developed for, our structure theory has applications in the study of anti-canonical K\"ahler Ricci flows.
Due to the fundamental estimate of Perelman and the monotonicity of his $\mu$-functional along each anti-canonical
K\"ahler Ricci flow, we can apply Theorem~\ref{thmin:HC06_2} directly and obtain the following theorem.
\begin{theoremin}[\textbf{Hamilton-Tian conjecture}]
Suppose $\left\{ (M^n, g(t)), 0 \leq t <\infty \right\}$ is an anti-canonical K\"ahler Ricci flow solution
on a Fano manifold $(M,J)$. For every $s>1$, define
\begin{align*}
&g_s(t) \triangleq g(t+s), \\
&\mathcal{M}_{s} \triangleq \{(M^n, g_s(t)), -s \leq t \leq s\}.
\end{align*}
Then for every sequence $s_i \to \infty$, by taking subsequence if necessary, we have
\begin{align}
\left(\mathcal{M}_{s_i}, g_{s_i}\right) \longright{\hat{C}^{\infty}} \left( \bar{\mathcal{M}}, \bar{g} \right), \label{eqn:SC13_111}
\end{align}
where the limit space-time $\bar{\mathcal{M}}$ is a K\"ahler Ricci soliton flow solution on a $Q$-Fano normal variety $(\bar{M},\bar{J})$.
Moreover, with respect to each $\bar{g}(t)$, there is a uniform $C$ independent of time such that
the $r$-neighborhood of the singular set $\mathcal{S}$ has measure not greater than $Cr^4$.
\label{thmin:SC24_3}
\end{theoremin}
Theorem~\ref{thmin:SC24_3} confirms the famous Hamilton-Tian conjecture, with more information
than that was conjectured(c.f.~Conjecture 9.1. of~\cite{Tian97} for the precise statement).
The two dimensional case was confirmed by the authors in~\cite{CW3}.
We note that in a recent paper \cite{TZZ2},
another approach to attack Hamilton-Tian conjecture in complex dimension $3$,
based on $L^4$-bound of Ricci curvature, was presented by Z.L. Zhang and G. Tian.
Their work in turn depends on the comparison geometry with integral Ricci bounded developed
by G.F. Wei and P. Peterson(\cite{PW1}).
For other important progress in K\"ahler Ricci flow, we refer interested readers to the following papers(far away from being complete): \cite{Sesum}, \cite{Zhq1}, \cite{Ye}, \cite{TZ2}, \cite{Zhq2}, \cite{SongWeinkove}, \cite{TZ2}, \cite{SongTian}, \cite{PhongSturm}, \cite{Tosa}, \cite{SzeKrf}, as well as references listed therein.\\
As corollaries of Theorem~\ref{thmin:SC24_3}, we can affirmatively answer some problems raised in~\cite{CW4}.
\begin{corollaryin}
Every anti-canonical K\"ahler Ricci flow is tamed, i.e., partial-$C^0$-estimate holds along the flow.
\label{clyin:SC24_4}
\end{corollaryin}
\begin{corollaryin}
Suppose $\{(M^n, g(t)), 0 \leq t < \infty \}$ is an anti-canonical K\"ahler Ricci flow on a Fano manifold $M$.
Then the flow converges to a K\"ahler Einstein metric
if one of the following conditions hold for every large positive integer $\nu$.
\begin{itemize}
\item $\alpha_{\nu,1}>\frac{n}{n+1}$.
\item $\alpha_{\nu,2}>\frac{n}{n+1}$ and $\alpha_{\nu,1} > \frac{1}{2- \frac{n-1}{(n+1) \alpha_{\nu,2}}}$.
\end{itemize}
\label{clyin:SC24_5}
\end{corollaryin}
Corollary~\ref{clyin:SC24_5} give rise to a method for searching Fano K\"ahler Einstein metrics in high dimension,
which generalize the 2-dimensional case due to Tian(c.f.~\cite{Tian90}).
Our structure theory can be applied to study a family of K\"ahler Ricci flows with some uniform initial conditions.
In this perspective, we have the following theorem.
\begin{theoremin}[\textbf{Partial-$C^0$-conjecture of Tian}]
For every positive constants $R_0,V_0$, there exists a positive integer $k_0$
and a positive constant $c_0$ with the following properties.
Suppose $(M,\omega, J)$ is a K\"ahler manifold satisfying
$Ric \geq R_0$ and $\Vol(M) \geq V_0$, $[\omega]=2\pi c_1(M,J)$.
Then we have
\begin{align*}
\inf_{x \in M} \mathbf{b}^{(k_0)}(x) > -c_0.
\end{align*}
\label{thmin:HA01_1}
\end{theoremin}
Theorem~\ref{thmin:HA01_1} confirms the partial-$C^0$-conjecture of Tian(c.f.~\cite{Tian90Kyo},\cite{Tian12}).
The low dimension case ($n \leq 3$) was proved by Jiang(\cite{Jiang}), depending on the partial-$C^0$-estimate along the flow,
developed by Chen-Wang(\cite{CW3},\cite{CW4}) in complex dimension 2
and Tian-Zhang(\cite{TZZ2}) in complex dimension 3.
In fact, a more general version of Theorem~\ref{thmin:HA01_1} is proved(c.f. Theorem~\ref{thm:HA01_3}).
As a corollary of Theorem~\ref{thmin:HA01_1}, we have
\begin{corollaryin}(c.f.~\cite{Sze})
The partial-$C^0$-estimate holds along the classical continuity path.
\label{clyin:HA01_2}
\end{corollaryin}
Following Corollary~\ref{clyin:SC24_4}, we obtain the following result, which was originally pointed out by
G. Sz\"{e}kelyhidi(c.f.~\cite{Sze}) along the classical continuity path.
\begin{corollaryin}
Suppose $(M,J)$ is a Fano manifold with $Aut(M,J)$ discrete. If it is stable in the sense of S.Paul(c.f.~\cite{Paul12}), then it admits a K\"ahler Einstein metric.
\label{clyin:HA03_1}
\end{corollaryin}
An important application of our structure theory is devoted to the study of the relationships among different stabilities.
By the work of Chen, Donaldson and Sun(c.f.~\cite{CDS3} and references therein), a long standing stability conjecture, going back to Yau(c.f. Problem 65 of~\cite{Yau93}) and critically contributed by Tian(c.f~\cite{Tian97}) and Donaldson(c.f.~\cite{Do02}), was confirmed.
We now know a Fano manifold is K-stable if and only if it admits K\"ahler Einstein metrics. A posteriori, we see that
the K-stability is equivalent to Paul's stability if the underlying manifold has discrete automorphism group.
It is an interesting problem to prove this equivalence a priori,
which will be discussed in a separate paper(c.f.~\cite{CW6}). \\
Let us quickly go over the relationships among the theorems.
Theorem~\ref{thmin:HE21_1} is the structure theorem of the model space $\widetilde{\mathscr{KS}}(n,\kappa)$.
Theorem~\ref{thmin:SC24_1}, Theorem~\ref{thmin:HC08_1} and Theorem~\ref{thmin:HC06_1}
combined together give the canonical neighborhood structure of the polarized K\"ahler Ricci flow
in $\mathscr{K}(n,A)$, in a strong sense.
The main structure theorem in this paper is Theorem~\ref{thmin:HC06_2}, the weak compactness theorem of polarized K\"ahler Ricci flows.
It is clear that Theorem~\ref{thmin:SC24_3} and Theorem~\ref{thmin:HA01_1} are
direct applications of Theorem~\ref{thmin:HC06_2}. The proof of Theorem~\ref{thmin:HC06_2} is based on the combination of
Theorem~\ref{thmin:SC24_1}, Theorem~\ref{thmin:HC08_1} and Theorem~\ref{thmin:HC06_1}. These three theorems deal with different structures
of $\mathscr{K}(n,A)$, including the Ricci flow structure, metric space structure, line bundle structure and variety structure.
The importance of these structures decreases in order, for the purpose of developing compactness.
However, all these structures are intertwined together.
Paradoxically, the proof of the compactness of these structures does not follow the same order, due to the lack of precise estimate of Bergman functions.
Instead of proving them in order, we cook up a concept called ``polarized canonical radius", which guarantees the convergence of all these structures
under this radius. The only thing we need to do then is to show that this radius cannot be too small. Otherwise, we can apply a maximum principle argument
to obtain a contradiction, which essentially arise from the monotonicity of Perelman's reduced volume and localized $W$-functional.\\
This paper is organized as follows. In section 2, we discuss the model space $\widetilde{\mathscr{KS}}(n,\kappa)$, which consists of non-collapsed Calabi-Yau spaces with mild singularities. By checking analysis foundation and repeating the weak compactness theory of K\"ahler Einstein manifolds, we prove the compactness of
$\widetilde{\mathscr{KS}}(n,\kappa)$ and show that every space in it is a conifold.
In other words, we prove Theorem~\ref{thmin:HE21_1} at the end of section 2.
We also develop some a priori estimates, which will be essentially used in the following sections.
In section 3, we define the ``canonical radius" and discuss the convergence of metric structures when canonical radius is uniformly bounded from below.
In section 4, we first set up a forward, long-time pseudolocality theorem based on the existence of
partial-$C^0$-estimate. Motivated by this pseudolocality theorem, we then refine the ``canonical radius" to ``polarized canonical radius" and discuss the convergence of
flow structure and line bundle structure under the assumption that polarized canonical radius is uniformly bounded from below. Finally, at the end of section 4, we use a maximum principle argument to
show that there is an a priori bound of the polarized canonical radius.
In section 5, we prove Theorem~\ref{thmin:SC24_1}-\ref{thmin:HC06_2}, together with some other
more detailed properties of the space $\mathscr{K}(n,A)$.
Up to this section, everything is developed for general polarized K\"ahler Ricci flow.
At last, in section 6, we focus on the anti-canonical K\"ahler Ricci flows.
Applying the general structure theory, we prove Theorem~\ref{thmin:SC24_3} and Theorem~\ref{thmin:HA01_1}. \\
\noindent {\bf Acknowledgment}
Both authors are very grateful to professor Simon Donaldson for his constant support.
The second author would like to thank Song Sun and Shaosai Huang for helpful discussions and suggestions.
Thanks also go to Weiyong He, Haozhao Li,Yuanqi Wang, Guoqiang Wu, Chengjian Yao, Hao Yin, Kai Zheng
for their valuable comments.
\section{Model space---Calabi-Yau Space with mild singularities}
The Model space of a polarized K\"ahler Ricci flow in $\mathscr{K}(n,A)$ consists of the space-time blowup limits from
flows in $\mathscr{K}(n,A)$. In this section, we shall discuss the properties of the model space, from the perspective of
metric space structure and the intrinsic Ricci flow structure.
\subsection{Singular Calabi-Yau space $\widetilde{\mathscr{KS}}(n,\kappa)$}
Let $\mathscr{KS}(n)$ be the collection of all the complete $n$-dimensional Calabi-Yau (K\"ahler Ricci flat) manifolds.
By Bishop-Gromov comparison, it is clear that the asymptotic volume ratio is well defined for every manifolds in the moduli space $\mathscr{KS}(n)$.
The gap theorem of Anderson (c.f. Gap Lemma 3.1 of~\cite{An90}) implies that the asymptotic volume ratio is strictly less than $1-2\delta_0$ whenever the underlying manifold is not $\ensuremath{\mathbb{C}}^n$, where
$\delta_0$ is a dimensional constant. We fix this constant and call it as Anderson constant in this paper.
Let $\mathscr{KS}(n,\kappa)$ be a subspace of $\mathscr{KS}(n)$, with every element has asymptotic volume ratio at least $\kappa$.
Clearly, $\mathscr{KS}(n,\kappa)$ is not compact under the pointed-Gromov-Hausdorff topology.
It can be compactified as a space $\overline{\mathscr{KS}}(n,\kappa)$.
However, this may not be the largest space that one can develop weak-compactness theory. So we extend the space $\overline{\mathscr{KS}}(n,\kappa)$ further to a possibly bigger compact space $\widetilde{\mathscr{KS}}(n,\kappa)$, which is defined as follows.
\begin{definition}
Let $\widetilde{\mathscr{KS}}(n,\kappa)$ be the collection of length spaces $(X,g)$ with the following properties.
\begin{enumerate}
\item $X$ has a disjoint regular-singular decomposition $X=\mathcal{R} \cup \mathcal{S}$, where $\mathcal{R}$ is the regular part, $\mathcal{S}$ is the singular part.
A point is called regular if it has a neighborhood which is isometric to a totally geodesic convex domain of some smooth Riemannian manifold. A point is called singular
if it is not regular.
\item The regular part $\mathcal{R}$ is a nonempty, open Ricci-flat manifold of real dimension $m=2n$.
Moreover, there exists a complex structure $J$ on $\mathcal{R}$ such that $(\mathcal{R}, g, J)$ is a K\"ahler manifold.
\item $\mathcal{R}$ is weakly convex, i.e., for every point $x \in \mathcal{R}$, there exists a measure zero set
$\mathcal{C}_x$ such that every point in $X \backslash \mathcal{C}_x$ can be connected to $x$
by a unique shortest geodesic. For convenience, we call $\mathcal{C}_x$ as the cut locus of $x$.
\item $\dim_{\mathcal{M}} \mathcal{S} < 2n-3$, where $\mathcal{M}$ means Minkowski dimension.
\item Let $\mathrm{v}$ be the volume density function,i.e.,
\begin{align}
\mathrm{v}(x) \triangleq \lim_{r \to 0} \frac{|B(x,r)|}{\omega_{2n} r^{2n}} \label{eqn:SC16_1}
\end{align}
for every $x \in X$. Then $\mathrm{v}\equiv 1$ on $\mathcal{R}$ and $\mathrm{v} \leq 1-2\delta_0$ on $\mathcal{S}$.
In other words, the function $\mathrm{v}$ is a criterion function for singularity.
Here $\delta_0$ is the Anderson constant.
\item The asymptotic volume ratio $\mathrm{avr}(X) \geq \kappa$. In other words, we have
\begin{align*}
\lim_{r \to \infty} \frac{|B(x,r)|}{\omega_{2n}r^{2n}} \geq \kappa
\end{align*}
for every $x \in X$.
\end{enumerate}
Let $\widetilde{\mathscr{KS}}(n)$ be the collection of metric spaces $(X,g)$ with all the above properties except the last one. Since Euclidean space is a special element,
we define
\begin{align*}
\widetilde{\mathscr{KS}}^{*}(n) \triangleq \widetilde{\mathscr{KS}}(n) \backslash \{(\ensuremath{\mathbb{C}}^n, g_{\ensuremath{\mathbb{E}}})\},
\quad \widetilde{\mathscr{KS}}^{*}(n,\kappa) \triangleq \widetilde{\mathscr{KS}}(n,\kappa) \backslash \{(\ensuremath{\mathbb{C}}^n, g_{\ensuremath{\mathbb{E}}})\}.
\end{align*}
\label{dfn:SC27_1}
\end{definition}
We use $\dim_{\mathcal{H}}$ to denote Hausdorff dimension, $\dim_{\mathcal{M}}$ to denote Minkowski dimension, or the
box-counting dimension. Since Minkowski dimension is not as often used as Hausdorff dimension, let us recall the definition of it quickly(c.f.~\cite{Falco}).
\begin{definition}
Suppose $E$ is a bounded subset of $X$.
$E_r$ is the $r$-neighborhood of $E$ in $X$. Then
the upper Minkowski dimension of $E$ is defined as the limit:
$\displaystyle \dim_{\mathcal{H}} X-\liminf_{r \to 0^+} \frac{\log |E_r|}{\log r}$.
We say $\dim_{\mathcal{M}} E \leq \dim_{\mathcal{H}} X-k$
if the upper Minkowski dimension of $E$ is not greater than $2n-k$.
Namely, we have
\begin{align*}
\liminf_{r \to 0^+} \frac{\log |E_r|}{\log r} \geq k.
\end{align*}
If $E$ is not a bounded set, we say $\dim_{\mathcal{M}} E \leq \dim_{\mathcal{H}} X-k$ if
$\dim_{\mathcal{M}} E \cap B \leq \dim_{\mathcal{H}} X-k$ for each unit geodesic ball $B \subset X$
satisfying $B\cap E \neq \emptyset$.
\label{dfn:HE08_1}
\end{definition}
In general, it is known that Hausdorff dimension is not greater than Minkowski dimension. Hence, we always have
$\dim_{\mathcal{H}}\mathcal{S} \leq \dim_{\mathcal{M}} \mathcal{S}$.
In our discussion, $X$ clearly has Hausdorff dimension $2n$. Therefore, $\dim_{\mathcal{M}} \mathcal{S}<2n-3$
implies that for each nonempty intersection $B(x_0,1) \cap \mathcal{S}$, its $r$-neighborhood has measure
$o(r^{3})$ for sufficiently small $r$. By virtue of the high codimension of $\mathcal{S}$ and the Ricci-flatness of $\mathcal{R}$,
in many aspects, each metric space $X \in \widetilde{\mathscr{KS}}(n,\kappa)$ can be treated as an intrinsic Ricci-flat space. We shall see that the geometry of $X$ is almost the same as that of Calabi-Yau manifold.
\begin{proposition}[\textbf{Bishop-Gromov volume comparison}]
Suppose $x_0 \in X$, then the volume ratio $\omega_{2n}^{-1}r^{-2n}|B(x_0,r)|$ is a monotonically non-increasing function of $r$,
with values in the interval $[\kappa, 1]$.
\label{prn:HD19_1}
\end{proposition}
\begin{proof}
By approximation, it is clear that we only need to prove the case when $x_0 \in \mathcal{R}$.
Away from the cut locus which is measure zero,
everything can be done in the polar coordinate of $x_0$, the same as that in the Riemannian manifold case.
Then the classical proof of Bishop-Gromov comparison works here without any change.
\end{proof}
\begin{corollary}[\textbf{Volume doubling}]
$X$ is a volume doubling metric space. More precisely, for every $x_0 \in X$ and $r>0$, we have
\begin{align*}
\frac{|B(x_0,2r)|}{|B(x_0,r)|} \leq \kappa^{-1}.
\end{align*}
\label{cly:HD20_1}
\end{corollary}
\begin{proposition}[\textbf{Segment inequality}]
For every nonnegative function $f \in L_{loc}^1(X)$, define
\begin{align*}
\mathcal{F}_f(x_1,x_2) \triangleq \inf_{\gamma} \int_0^{l} f(\gamma(s))ds,
\end{align*}
where the infimum is taken over all minimal geodesics $\gamma$, from $x_1$ to $x_2$ and $s$ denotes the arc length.
Suppose $p \in X$, $r>0$, $A_1$, $A_2$ are two subsets of $B(p,r)$.
Then we have
\begin{align}
\int_{A_1 \times A_2} \mathcal{F}_{f}(x_1,x_2) \leq 4^n r (|A_1|+|A_2|) \int_{B(p, 3r)} f.
\label{eqn:HD17_3}
\end{align}
\label{prn:HD17_1}
\end{proposition}
\begin{proof}
Fix a smooth point $x_1$, then away from cut locus, every point can be connected to $x_1$ by a unique geodesic.
Since $X \times X$ is equipped with the product measure, it is clear that away from a measure zero set, every point
$(x_1, x_2) \in X \times X$ has the property that $x_1$ and $x_2$ are smooth and can be joined by
a unique smooth shortest geodesic.
Then the proof of (\ref{eqn:HD17_3}) is reduced to the same status as the Riemannian manifold case.
Then the proof is well known in the literature. \end{proof}
Due to the work of Cheeger and Colding (c.f. Remark 2.82 of~\cite{CCWarp}),
the segment inequality implies the $(1,2)$-Poincar\`{e} inequality in general.
In our particular case, the Poincar\`{e} constant can be understood more precisely.
\begin{proposition}[\textbf{Bound of Poincar\`{e} constant}]
Suppose $f \in L_{loc}^1(X)$, $h$ is an upper gradient of $f$ in the sense of Cheeger(c.f.~Definition~\ref{dfn:HD18_1}).
Then for every geodesic ball $B(p,r) \subset X$ and real number $q \geq 1$, we have
\begin{align}
\fint_{B(p,r)} |f-\underline{f}| \leq 2 \cdot 6^{2n} \cdot r \left(\fint_{B(p,3r)} h^q \right)^{\frac{1}{q}},
\label{eqn:HD18_1}
\end{align}
where $\fint$ means the average, $\underline{f}$ is the average of $f$ on $B(p,r)$.
In particular, there is a uniform $(1,2)$-Poincar\`e constant on $X$.
\label{prn:HC29_2}
\end{proposition}
\begin{proof}
By H\"older inequality, it suffices to prove (\ref{eqn:HD18_1}) for $q=1$.
By adding a constant if necessary, we can always assume that $\underline{f} \equiv 0$ on $B(p,r)$. Then we have
\begin{align*}
|f(x_2)| |B(p,r)|=\left| \int_{B(p,r)} -f(x_2) d\mu_{x_1} \right|=\left| \int_{B(p,r)} \{f(x_1)-f(x_2)\} d\mu_{x_1} \right|.
\end{align*}
Integrating the above equation over $B(p,r)$ implies that
\begin{align*}
|B(p,r)| \int_{B(p,r)} |f(x_2)| d\mu_{x_2}&=\int_{B(p,r)} \left| \int_{B(p,r)} \{f(x_1)-f(x_2)\} d\mu_{x_1} \right| d\mu_{x_2}
\leq \int_{B(p,r) \times B(p, r)}|f(x_1)-f(x_2)|.
\end{align*}
Recall that $h$ is upper gradient function, hence $|f(x_1)-f(x_2)| \leq \mathcal{F}_{h}(x_1,x_2)$ by definition.
Applying segment inequality, i.e., (\ref{eqn:HD17_3}), to $h$, we obtain
\begin{align*}
\int_{B(p,r) \times B(p, r)}|f(x_1)-f(x_2)|
\leq \int_{B(p,r) \times B(p, r)} \mathcal{F}_{h}(x_1,x_2)
\leq 2^{2n+1} r |B(p,r)| \int_{B(p,3r)} h.
\end{align*}
Combining the previous inequalities gives us
\begin{align*}
\int_{B(p,r)} |f| \leq 2^{2n+1} r \int_{B(p,3r)} h.
\end{align*}
Consequently, by volume comparison, we have
\begin{align*}
\fint_{B(p,r)} |f| &\leq 2^{2n+1} r \cdot \frac{|B(p,3r)|}{|B(p,r)|} \fint_{B(p,3r)} h
\leq 2 \cdot 6^{2n} \cdot r \fint_{B(p,3r)} h,
\end{align*}
which is the same as (\ref{eqn:HD18_1}) as we choose $\underline{f} \equiv 0$.
\end{proof}
\begin{proposition}[\textbf{Bound of Sobolev constant}]
There is a uniform isoperimetric constant on $X$.
Consequently, a uniform $L^2$-Sobolev inequality hold on $X$.
\label{prn:HC29_3}
\end{proposition}
\begin{proof}
Due to the uniform non-collapsing condition and the weak convexity and Ricci-flatness of $\mathcal{R}$, the argument of Croke (c.f.~\cite{Croke}) applies.
So there is a uniform isoperimetric constant on $X$.
The $L^2$-Sobolev constant follows from the isoperimetric constant(c.f.~\cite{SchYau}).
\end{proof}
Note that for each $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, we lose smooth structure around $\mathcal{S}$.
In orbifold case, one can recover the smooth structure at a local ``covering" space. For our $X$, it is not known whether one has such a property. However, the good news is that the smooth structure does not play an essential role in many aspects. In the next subsection, we shall see that the analysis on $X$ is almost the same as that on manifold.
\subsection{Sobolev space, Dirichlet form and heat semigroup}
On a metric measure space, one can define
Sobolev space $H_{1,2}(X)$ following Cheeger(\cite{Cheeger99}), or $N^{1,2}(X)$
following Shanmugalingam(\cite{ShanNew}).
However, these two definitions coincide whenever volume doubling property and uniform $(1,2)$-Poincar\`{e} inequality holds,
in light of Theorem 4.10 of~\cite{ShanNew}, or the discussion on page 440 of~\cite{Cheeger99}.
In particular, for the space $(X,g,d\mu)$ which we are interested in, we have $N^{1,2}(X)=H_{1,2}(X)$ as Banach spaces.
For simplicity, we shall only use the notation $N^{1,2}(X)$ and follow the route of Cheeger.
\begin{definition}
Suppose $\Omega \subset X$. Let $f: \Omega \to [0,\infty]$ be an extended function.
An extended real function $h: \Omega \to [0,\infty]$ is called an upper gradient of $f$ on $\Omega$
if for every two points
$z_1,z_2 \in \Omega$ and all continuous rectifiable curves $c: [0,l] \to \Omega$, parameterized by arc length $s$, with $z_1,z_2$ end points, we have
\begin{align*}
|f(z_1)-f(z_2)| \leq \int_0^l h(c(s))ds.
\end{align*}
\label{dfn:HD18_1}
\end{definition}
\begin{definition}
The Sobolev space $N^{1,2}(X)$ is the subspace of $L^2(X)$ consisting of functions $f$ for which the norm
\begin{align}
\norm{f}{N^{1,2}}^2=\norm{f}{L^2}^2 + \inf_{f_i} \liminf_{i \to \infty}\norm{h_i}{L^2}^2 < \infty,
\label{eqn:HD13_1}
\end{align}
where the limit infimum is taken over all upper gradients $h_i$ of the functions $f_i$, which satisfies $\norm{f_i -f}{L^2(X)} \to 0$.
\end{definition}
Note that the above $N^{1,2}$-norm is equivalent to Cheeger's definition(c.f. equation (2.1) of ~\cite{Cheeger99}).
With this norm, we know $N^{1,2}(X)$
is complete(c.f. Theorem 2.7 of~\cite{Cheeger99}).
Clearly, it follows directly from the definition that zero function $f \in L^2(X)$ is the zero function in $N^{1,2}(X)$.
It is not surprising that $N^{1,2}(X)$ is the classical Sobolev space whenever $X$ is a smooth manifold.
This can be easily proved following the same argument of Theorem 4.5 of~\cite{ShanNew}, where the same conclusion was proved whenever
$X$ is a domain of Euclidean space. In particular, as Banach spaces, we have
\begin{align}
N^{1,2}(\mathcal{R}) \cong W^{1,2}(\mathcal{R}),
\end{align}
where $W^{1,2}(\mathcal{R})$ is the classical Sobolev space on the smooth manifold $\mathcal{R}$.
\begin{proposition}[\textbf{Smooth approximation}]
Suppose $\Omega$ is an open set of $X$, $f \in N^{1,2}(\Omega)$.
Then there is a sequence of $f_i \in C^{\infty}(\Omega \backslash \mathcal{S}) \cap N^{1,2}(\Omega)$,
$\supp f_i \subset \Omega \backslash \mathcal{S}$ such that
\begin{align}
\lim_{i \to \infty} \norm{f_i-f}{N^{1,2}(\Omega)}=0.
\label{eqn:HD15_1}
\end{align}
Moreover, if $f$ is also nonnegative, we can choose the approximation $f_i$ nonnegative.
If $\Omega$ is bounded, then $\supp f_i$ is a compact subset of $\overline{\Omega} \backslash \mathcal{S}$.
\label{prn:HD04_1}
\end{proposition}
\begin{proof}
It suffices to show the proof for the case when both $\diam(\Omega)$ and $\norm{f}{L^{\infty}}$ are bounded.
For otherwise, we can apply the bounded result for the truncated function $\min\{k, \max\{-k, f\}\}$ on $\Omega \cap B(x_0,k)$ for each $k$ and then use a
standard diagonal sequence argument, to reduce to the bounded case.
Since $\mathcal{S}$ has measure zero, $\Omega \backslash \mathcal{S}$ is a smooth manifold, we have
\begin{align*}
\norm{f_i-f}{N^{1,2}(\Omega)}=\norm{f_i-f}{N^{1,2}(\Omega \backslash \mathcal{S})}
=\norm{f_i-f}{W^{1,2}(\Omega \backslash \mathcal{S})}.
\end{align*}
Therefore, (\ref{eqn:HD15_1}) is equivalent to
\begin{align}
\lim_{i \to \infty} \norm{f_i-f}{W^{1,2}(\Omega \backslash \mathcal{S})}=0.
\label{eqn:HD15_2}
\end{align}
However, this sequence of $f_i$ can be constructed following a standard method,
as indicated by the proof of Theorem 2 of section 5.3.2 of Evans' book~\cite{Evans}.
For the convenience of the readers, we include a detailed construction of $f_i$ here.
For each positive integer $i$, define
\begin{align*}
\Omega_i \triangleq \{y \in \Omega| d(y,\mathcal{S})>2^{-i}\}, \quad
V_i \triangleq \Omega_{i+3} \backslash \overline{\Omega}_{i+1}, \quad
W_i \triangleq \Omega_{i+4} \backslash \overline{\Omega}_{i}.
\end{align*}
Also, choose open sets $V_0$ and $W_0$ such that
\begin{align*}
\overline{\Omega}_4 \cap \Omega \supset V_0 \supset \overline{\Omega}_2 \cap \Omega, \quad W_0 \supset \overline{\Omega}_6 \cap \Omega \supset V_0.
\end{align*}
Then we have
\begin{align*}
\Omega \backslash \mathcal{S}= \bigcup_{i=0}^{\infty} V_i=\bigcup_{i=0}^{\infty} W_i, \quad \overline{V}_i \cap \Omega_i \subset W_i, \quad \forall \; i \geq 0.
\end{align*}
Clearly, by composing with $d(\cdot, \mathcal{S})$, we can choose Lipschitz cutoff functions $\zeta_i$ such that $\zeta_i=1$ on $V_i$
and $\supp \zeta_i \subset W_i$, $|\nabla \zeta_i|<2^{i+5}$.
Set
\begin{align*}
\eta_i \triangleq \frac{\zeta_i}{\sum_j \zeta_j}.
\end{align*}
Clearly, $\eta_i$ is a kind of partition of unity subordinate to the covering $\bigcup_{i} W_i$. In other words, we have
\begin{align*}
\begin{cases}
0\leq \eta_i \leq 1, & \eta_i \in C_c^{1}(W_i), \quad \forall \; i \geq 1, \\
\sum_{i} \eta_i =1, & \textrm{on} \; \Omega \backslash \mathcal{S}.
\end{cases}
\end{align*}
Note that $\eta_0$ is special. It is only in $C^{1}(W_0)$ in general. However, it vanishes around $\partial W_0 \cap \Omega$.
For each $i \geq 0$, note that
$V_i \cap V_j =\emptyset$ if $|i-j| \geq 2$, $W_i \cap W_j= \emptyset$ if $|i-j| \geq 4$. Therefore, we have
\begin{align*}
0\leq \eta_i<1, \quad |\nabla \eta_i|<2^{i+10}.
\end{align*}
For each $i \geq 1$, we see that $\eta_i f \in W_0^{1,2}(W_i)$. Note that $W_i \subset \mathcal{R}$.
Applying convolution with smooth mollifiers(c.f. Theorem 1 of section 5.3.1 of~\cite{Evans}), we can choose a smooth function
$h_i \in C_c^{\infty}(W_i)$ such that
\begin{align*}
\norm{h_i-\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}^2=\norm{h_i-\eta_i f}{W^{1,2}(W_i)}^2 < 9^{-i-1}\epsilon^2.
\end{align*}
For $i=0$, we can choose $h_0 \in C^{\infty}(W_0)$ which vanishes around $\partial W_0 \cap \Omega$ such that the above inequality hold.
For each large $k$, we define
\begin{align*}
H_k \triangleq \sum_{i=0}^k h_i.
\end{align*}
It is easy to see that $H_k \in C^{\infty}(\cup_{i=0}^k W_i)
\subset C^{\infty}(\Omega \backslash \mathcal{S})$.
Moreover, we have estimate
\begin{align}
\norm{H_k-f}{W^{1,2}(\Omega \backslash \mathcal{S})}
&=\norm{\sum_{i=0}^k h_i -\sum_{i=1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}
=\norm{\sum_{i=0}^k (h_i-\eta_i f) -\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})} \notag\\
&\leq \sum_{i=0}^k \norm{h_i-\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}
+ \norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}.
\label{eqn:HD03_2}
\end{align}
However, the first term on the right hand side of the above inequality can be bounded as follows.
\begin{align}
\sum_{i=0}^k \norm{h_i-\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}
<\sum_{i=0}^{k} 3^{-i-1} \epsilon<\frac{1}{2}\epsilon.
\label{eqn:HD03_3}
\end{align}
On the other hand, note that $\sum_{i=k+1}^{\infty} \eta_i =1$ on $\bigcup_{i=k+5}^{\infty}W_i$, and it is supported on
$\bigcup_{i=k+1}^{\infty}W_i$. Thus, we have
\begin{align}
\norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega \backslash \mathcal{S})}^2
&\leq \norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\bigcup_{i=k+5}^{\infty}W_i)}^2
+\norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2 \notag\\
&=\norm{f}{W^{1,2}(\bigcup_{i=k+5}^{\infty}W_i)}^2
+\norm{\sum_{i=k+1}^{k+8}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2.
\label{eqn:HD03_1}
\end{align}
For simplicity of notation, define $\chi_k \triangleq \sum_{i=k+1}^{k+8} \eta_i$. Clearly, $0\leq \chi_k \leq 1$.
We have
\begin{align*}
\norm{\sum_{i=k+1}^{k+8}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
&=\norm{\chi_k f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
=\int_{\bigcup_{i=k+1}^{k+4}W_i} \chi_k^2 f^2 + \left| \left \langle \chi_k \nabla f + f \nabla \chi_k \right \rangle \right|^2\\
&\leq \int_{\bigcup_{i=k+1}^{k+4}W_i} f^2 + 2\chi_k^2 |\nabla f|^2 + 2f^2|\nabla \chi_k|^2\\
&\leq \left( 2 \int_{\bigcup_{i=k+1}^{k+4}W_i} f^2+|\nabla f|^2 \right)
+2\norm{f}{L^{\infty}(\Omega)}^2 \int_{\bigcup_{i=k+1}^{k+4}W_i} |\nabla \chi_k|^2.
\end{align*}
It is easy to see that $|\nabla \chi_k|<2^{k+20}$ by estimate of $\eta_k$.
By virtue of Minkowski codimension assumption, we obtain
\begin{align*}
\left|\bigcup_{i=k+1}^{k+4}W_i \right|< \left|\bigcup_{i=k+1}^{\infty}W_i \right|< C2^{-3k}
<C \left(2^{-k+5} \right)^3,
\end{align*}
which in turn implies that
\begin{align*}
\norm{\sum_{i=k+1}^{k+8}\eta_i f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
\leq 2\norm{f}{W^{1,2}(\bigcup_{i=k+1}^{k+4}W_i)}^2
+ C\norm{f}{L^{\infty}(\Omega)}^2 2^{-k}.
\end{align*}
Plug the above inequality into (\ref{eqn:HD03_1}), we obtain
\begin{align*}
\norm{\sum_{i=k+1}^{\infty}\eta_i f}{W^{1,2}(\Omega)}^2
\leq 2 \norm{f}{W^{1,2}(\bigcup_{i=k+1}^{\infty}W_i)}^2 + C\norm{f}{L^{\infty}(\Omega)}^2 2^{-k}.
\end{align*}
Together with (\ref{eqn:HD03_2}) and (\ref{eqn:HD03_3}), the above inequality implies that
\begin{align*}
\norm{H_k-f}{W^{1,2}(\Omega)} <\frac{1}{2}\epsilon
+2 \norm{f}{W^{1,2}(\bigcup_{i=k+1}^{\infty}W_i)}^2 + C\norm{f}{L^{\infty}(\Omega)}^2 2^{-k}.
\end{align*}
Recall that $f \in W^{1,2}(\Omega \backslash \mathcal{S})$, $|\bigcup_{i=k+1}^{\infty}W_i| \to 0$ as $k \to \infty$. So we can choose $k$ large
enough such that
\begin{align*}
\norm{H_k-f}{W^{1,2}(\Omega \backslash \mathcal{S})} <\epsilon.
\end{align*}
Let $\epsilon=\frac{1}{i}$, we denote the corresponding $H_k$ in the above inequality by $f_i$.
Clearly, $f_i$ is supported on $\Omega \backslash \mathcal{S}$ and is smooth.
Moreover, (\ref{eqn:HD15_2}), consequently (\ref{eqn:HD15_1}), follows from the above inequality.
It follows from the construction that $f_i \geq 0$ whenever $ f \geq 0$. Also, from the construction,
if $\Omega$ is bounded, $\supp f_i$ is a compact subset of $\overline{\Omega} \backslash \mathcal{S}$.
\end{proof}
\begin{corollary}[\textbf{Smooth functions with compact supports}]
$C_c^{\infty}(\mathcal{R}) \cap N^{1,2}(X)$ is dense in $N^{1,2}(X)$.
\label{cly:HE10_1}
\end{corollary}
\begin{proof}
Fix $f \in N^{1,2}(X)$, without loss of generality, we may assume that $f \in C^{\infty}(\mathcal{R})$ and $f$ vanishes around $\mathcal{S}$,
by Proposition~\ref{prn:HD04_1}.
Fix $x_0 \in \mathcal{R}$ and let $r(x)=d(x,x_0)$.
For each large $k$, let $\phi_k=\phi(r(x)-k)$, where $\phi$ is a smooth cutoff function on real axis such that $\phi \equiv 1$ on $(-\infty, 0)$ and
$\phi \equiv 0$ on $(1,\infty)$. Moreover, $|\phi'| \leq 2$.
Note that $\supp f \cap \overline{B(x_0, k+1)}$ is a compact subset of $B(x_0,k+2) \backslash \mathcal{S}$. By convolution with mollifier if necessary,
we can assume $\phi_k$ is smooth and on $\supp f \cap \supp \phi_k$, $\supp \phi_k \subset B(x_0,k+2)$, $\phi_k \equiv 1$ on $B(x_0,k-1)$.
Moreover, $|\nabla \phi_k|<4$ and $0 \leq \phi_k<2$.
Therefore, $\phi_k f \in C_c^{\infty}(\mathcal{R})$. It is easy to calculate
\begin{align*}
\norm{f-\phi_k f}{N^{1,2}(X)}^2&=\int_X (1-\phi_k)^2f^2 d\mu + \int_X \left| \nabla \{(1-\phi_k) f\} \right|^2 d\mu\\
&\leq \int_{X \backslash B(x_0,k-1)} (1-\phi_k)^2 f^2 du + 2\int_{X \backslash B(x_0, k-1)} \left\{(1-\phi_k)^2 |\nabla f|^2 + f^2|\nabla \phi_k|^2 \right\} d\mu\\
&\leq \int_{X \backslash B(x_0,k-1)} f^2 du + 2\int_{X \backslash B(x_0, k-1)} \left\{|\nabla f|^2 +16 f^2 \right\} d\mu\\
&\leq 33 \int_{X \backslash B(x_0, k-1)} \left\{|\nabla f|^2 + f^2 \right\} d\mu.
\end{align*}
Clearly, the right hand side of the above inequality goes to $0$ as $k \to \infty$, since $f \in N^{1,2}(X)$. Therefore, every $f \in N^{1,2}(X)$ can
be approximated by smooth functions with compact supports.
\end{proof}
In light of Proposition~\ref{prn:HD04_1}, we can define $N_0^{1,2}(\Omega)$ as the completion of all the functions in
$C_c^{\infty}(\Omega \backslash \mathcal{S}) \cap N^{1,2}(X)$, under the $N^{1,2}(\Omega)$-norm.
Note that a function $f$ in $N_0^{1,2}(\Omega)$ may not have compact support, with respect to $\Omega$.
However, $f|_{\partial \Omega}=0$, in the sense of traces.
\begin{proposition}[\textbf{Global continuous approximation}]
For each $f \in C_c(X)$, i.e., a continuous function with compact support, there exists a sequence of $f_i \in C_c(X) \cap N^{1,2}(X)$ such that
$\displaystyle \lim_{i \to \infty} \norm{f_i-f}{C(X)} \to 0$.
\label{prn:HD24_1}
\end{proposition}
\begin{proof}
For each $\epsilon>0$, $x \in X$, define $\phi_{\epsilon,x}$ to be the character equation of the geodesic ball
$B(x,\epsilon)$.
In other words, $\phi_{\epsilon, x} \equiv 1$ on $B(x,\epsilon)$ and $0$ on $X \backslash B(x,\epsilon)$. Define
$\psi_{\epsilon,x}$ to be $\frac{\phi_{\epsilon,x}}{|B(x,\epsilon)|}$. Clearly, we have
\begin{align}
\int_X \psi_{\epsilon,x}(y)d\mu_y=1.
\label{eqn:HD26_1}
\end{align}
Similar to Euclidean case, we define approximation functions as convolution of $f$ and $\psi_{\epsilon,\cdot}$ as follows:
\begin{align*}
f_{\epsilon}(x) \triangleq (\psi_{\epsilon}* f)(x)= \int_X f(y) \psi_{\epsilon, x}(y) d\mu_y.
\end{align*}
Fix $\epsilon>0$. Suppose $x_1,x_2$ are two points in $X$ with distance $\rho \in (0,\epsilon)$. Then we calculate
\begin{align*}
|f_{\epsilon}(x_1)-f_{\epsilon}(x_2)| &\leq \int_X |f|(y) \left| \psi_{\epsilon,x_1}(y)-\psi_{\epsilon,x_2}(y)\right| d\mu_y\\
&\leq \norm{f}{C(X)} \int_X \left| \frac{\phi_{\epsilon,x_1}}{|B(x_1,\epsilon)|} - \frac{\phi_{\epsilon,x_2}}{|B(x_2,\epsilon)|}\right| d\mu_y\\
&=\frac{\norm{f}{C(X)} }{|B(x_1,\epsilon)||B(x_2,\epsilon)|} \int_X \left\vert\phi_{\epsilon,x_1} |B(x_2,\epsilon)|
-\phi_{\epsilon,x_2} |B(x_1,\epsilon)| \right\vert d\mu_y\\
&\leq C(n,\kappa) \norm{f}{C(X)} \epsilon^{-4n}
\int_X \left\vert\phi_{\epsilon,x_1} |B(x_2,\epsilon)|
-\phi_{\epsilon,x_2} |B(x_1,\epsilon)| \right\vert d\mu_y.
\end{align*}
Notice that
\begin{align*}
&\quad \int_X \left\vert\phi_{\epsilon,x_1} |B(x_2,\epsilon)|
-\phi_{\epsilon,x_2} |B(x_1,\epsilon)| \right\vert d\mu_y\\
&=\int_X \left \vert \phi_{\epsilon,x_1}\left\{ |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\}
+|B(x_1,\epsilon)| \cdot (\phi_{\epsilon,x_1}-\phi_{\epsilon,x_2}) \right\vert d\mu_y\\
&\leq \int_X\phi_{\epsilon,x_1} \left \vert |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\vert d\mu_y
+|B(x_1,\epsilon)| \int_X |\phi_{\epsilon,x_1} -\phi_{\epsilon,x_2}| d\mu_y \\
&=|B(x_1,\epsilon)| \left\{ \left \vert |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\vert
+ \int_X |\phi_{\epsilon,x_1} -\phi_{\epsilon,x_2}| \right\}\\
&=|B(x_1,\epsilon)| \left\{ \left \vert |B(x_2,\epsilon)|-|B(x_1,\epsilon)| \right\vert
+|B(x_1,\epsilon) \backslash B(x_2,\epsilon)|+|B(x_2,\epsilon) \backslash B(x_1,\epsilon)| \right\}\\
&\leq 2 |B(x_1,\epsilon)| \left\{
|B(x_1,\epsilon) \backslash B(x_2,\epsilon)|+|B(x_2,\epsilon) \backslash B(x_1,\epsilon)| \right\}.
\end{align*}
By Bishop-Gromov volume comparison and non-collapsing condition, we have
\begin{align*}
&|B(x_2,\epsilon) \backslash B(x_1,\epsilon)|
\leq \left| B(x_1,\epsilon+\rho) \backslash B(x_1,\epsilon-\rho)\right| \leq C(n,\kappa) \epsilon^{2n-1} \rho, \\
&|B(x_1,\epsilon) \backslash B(x_2,\epsilon)|
\leq \left| B(x_2,\epsilon+\rho) \backslash B(x_2,\epsilon-\rho)\right| \leq C(n,\kappa) \epsilon^{2n-1} \rho.
\end{align*}
Thus, for each $\rho \in (0,\epsilon)$, we have estimate
\begin{align*}
|f_{\epsilon}(x_1)-f_{\epsilon}(x_2)|
\leq \frac{C(n,\kappa) \norm{f}{C(X)}}{\epsilon} \rho,
\end{align*}
which means that the Lipschitz constant of $f_{\epsilon}$ is uniformly bounded, for each fixed $\epsilon$.
In particular, $f_{\epsilon}$ locates in $C_c(X) \cap N^{1,2}(X)$.
It follows from (\ref{eqn:HD26_1}) that
\begin{align*}
\left| f_{\epsilon}(x) -f(x) \right|&=\left| \int_X \{f(y)-f(x)\} \psi_{\epsilon, x}(y) d\mu_y \right|
\leq \int_{B(x,\epsilon)} |f(y)-f(x)| \psi_{\epsilon,x}(y) d\mu_y
\leq \sup_{y \in B(x,\epsilon)} |f(y)-f(x)|.
\end{align*}
Note that $f$ is uniformly continuous since $f$ is continuous and $\supp(f)$ is contained in a compact subset of $X$.
Hence the right hand side of the above inequality converges to zero uniformly as $\epsilon \to 0$.
Therefore, $\psi_{2^{-i}} *f$ is a sequence of functions in $C_c(X) \cap N^{1,2}(X)$ and converges to $f$ in $C(X)$-norm.
\end{proof}
For each open set $\Omega \subset X$,
there is a restriction map $\pi: N^{1,2}(\Omega) \to N^{1,2}(\Omega \backslash \mathcal{S})$ in the obvious way.
Note that $\Omega \backslash \mathcal{S}=\Omega \cap \mathcal{R}$ is a smooth manifold,
hence $N^{1,2}(\Omega \backslash \mathcal{S})=W^{1,2}(\Omega \backslash \mathcal{S})$.
In general, the map $\pi$ is not surjective. However, in our special setting, $\mathcal{S}$ has high codimension,
we have much more information.
\begin{proposition}(\textbf{Identity is isometry})
Suppose $\Omega$ is an open set of $X$, then
the restriction map $\pi: N^{1,2}(\Omega) \to N^{1,2}(\Omega \backslash \mathcal{S})=W^{1,2}(\Omega \backslash \mathcal{S})$ is an isomorphic isometry.
\label{prn:HD13_1}
\end{proposition}
\begin{proof}
If we proved $\pi$ is an isomorphism, it is clear that $\pi$ is an isometry since $\mathcal{S}$ has measure zero.
Thus, we only need to focus on the proof of isomorphism. For simplicity, we assume $\Omega=X$.
Then $\Omega \backslash \mathcal{S}=X \backslash \mathcal{S}=\mathcal{R}$.
\textit{Injectivity}: Suppose $\pi(f)=0$. Then $\norm{f}{L^2(X)}=0$ since $\mathcal{S}$ has measure zero.
Due to the fact $f \in N^{1,2}(X)$, $\norm{f}{L^2(X)}=0$ implies that $\norm{f}{N^{1,2}(X)}=0$.
Therefore, $f$ is the zero element in $N^{1,2}(X)$.
\textit{Surjectivity}: For every $0 \neq \tilde{f} \in W^{1,2}(\mathcal{R})$, from the proof of Proposition~\ref{prn:HD04_1},
there is a sequence of smooth functions $\tilde{f}_i$ supported on $\mathcal{R}$ such that
$\displaystyle \norm{\tilde{f}_i -\tilde{f}}{W^{1,2}(\mathcal{R})} \to 0$.
In particular, $\tilde{f}_i$ is a Cauchy sequence in $W^{1,2}(\mathcal{R})$.
Since $\displaystyle \norm{\tilde{f}_i -\tilde{f}}{N^{1,2}(X)}=\norm{\tilde{f}_i -\tilde{f}}{W^{1,2}(\mathcal{R})}$,
it is clear that $\tilde{f}_i$ is a Cauchy sequence in $N^{1,2}(X)$.
Therefore, there is a function $f \in N^{1,2}(X)$, as the limit of $\tilde{f}_i$, by completeness of $N^{1,2}(X)$.
After we obtain $f$, it is clear that $\norm{f_i-f}{N^{1,2}(X)} \to 0$, which forces that
\begin{align*}
\norm{f_i-\pi(f)}{W^{1,2}(\mathcal{R})} \to 0.
\end{align*}
Therefore, $\pi(f)=\tilde{f}$.
\end{proof}
In light of Proposition~\ref{prn:HD13_1}, we can regard $N^{1,2}(X)$ as the same Banach space as $W^{1,2}(\mathcal{R})$.
However, $W^{1,2}(\mathcal{R})$ is a Hilbert space. This induces a natural inner product structure on $W^{1,2}(\mathcal{R})$ as follows:
\begin{align*}
\langle\langle f_1, f_2\rangle\rangle
= \int_{\mathcal{R}} \left\{ \pi(f_1) \pi(f_2) +\langle \nabla \pi(f_1), \nabla \pi(f_2)\rangle \right\}d\mu, \quad \forall \; f_1, f_2 \in N^{1,2}(X).
\end{align*}
For simplicity of notation, we shall not differentiate $f$ and $\pi(f)$. Under this convention, we have
\begin{align*}
\langle\langle f_1, f_2\rangle\rangle
= \int_{\mathcal{R}} \left\{ f_1f_2 +\langle \nabla f_1, \nabla f_2 \rangle \right\}d\mu, \quad \forall \; f_1, f_2 \in N^{1,2}(X).
\end{align*}
Therefore, $N^{1,2}(X)$ is isomorphic to $W^{1,2}(\mathcal{R})$ as a Hilbert space. For every $f_1,f_2 \in N^{1,2}(X)$, we define
a nonnegative, symmetric, bilinear form $\mathscr{E}$ as follows
\begin{align}
\mathscr{E}(f_1,f_2) \triangleq \int_{\mathcal{R}} \langle \nabla f_1, \nabla f_2 \rangle d\mu.
\label{eqn:HD26_2}
\end{align}
We want to show that $\mathscr{E}$ is a Dirichlet form.
Actually, it is clear that $\norm{f}{N^{1,2}(X)}^2=\norm{f}{L^2(X)}^2+\mathscr{E}(f,f)$.
Since $N^{1,2}(X)$ is complete, we know that $\mathscr{E}$ is closed by definition.
On the other hand, since $W^{1,2}(\mathcal{R})$ is dense
in $L^2(\mathcal{R})=L^2(X)$, $\mathcal{S}$ has measure zero,
it follows directly that $N^{1,2}(X)$ is dense in $L^2(X)$.
Furthermore, it is clear that
\begin{align}
\mathscr{E}(\min\{1,\max\{0,f\}\}, \min\{1,\max\{0,f\}\}) \leq \mathscr{E}(f,f), \quad \forall \; f \in N^{1,2}(X).
\label{eqn:HE26_1}
\end{align}
Therefore, $\mathscr{E}$ is a closed, nonnegative, symmetric, bilinear form on $N^{1,2}(X)$,
which is a dense subspace of $L^2(X)$, with unit contraction property (\ref{eqn:HE26_1}).
It follows from a standard definition (c.f.~\cite{FuOsTa} for definition of Dirichlet form) that $\mathscr{E}$ is a Dirichlet form.
Not surprisingly, this Dirichlet form $\mathscr{E}$ is much better than general Dirichlet form since
the underlying space $X$ has rich geometry. In fact, suppose $u \in N_0^{1,2}(\Omega)$ for some open set $\Omega \subset X$,
it is clear that $u \equiv 0$ on $\Omega$ if and only if $ \mathscr{E}(u,u)=0$.
This means that $\mathscr{E}$ is irreducible by direct definition.
Also, for every constant $c$, we have $\mathscr{E}(u,c)=0$, which is equivalent to say that $\mathscr{E}$ is strongly local.
Furthermore, it follows from Corollary~\ref{cly:HE10_1} that $N^{1,2}(X) \cap C_c(X)$ is dense in $N^{1,2}(X)$ with $N^{1,2}$-norm.
On the other hand, Proposition~\ref{prn:HD24_1} implies that $N^{1,2}(X) \cap C_c(X)$ is dense in $C_c(X)$ with uniform supreme norm.
Consequently, $N^{1,2}(X) \cap C_c(X)$ is a core of $\mathscr{E}$ and $\mathscr{E}$ is a regular Dirichlet form, following from the definition verbatim.
Putting all the above information together, we obtain the following property.
\begin{proposition}[\textbf{Existence of excellent Dirichlet form}]
On the Hilbert space $L^2(X)$, there exists a Dirichlet form $\mathscr{E}$ defined on a dense subspace
$N^{1,2}(X) \subset L^2(X)$, by formula (\ref{eqn:HD26_2}). Furthermore, the Dirichlet form $\mathscr{E}$
is irreducible, strongly local and regular.
\label{prn:HD26_1}
\end{proposition}
With respect to the Dirichlet form $\mathscr{E}$, there is a unique generator (c.f. Chapter 1 of~\cite{FuOsTa}) of $\mathscr{E}$, which we denote by $\mathcal{L}$.
In other words, $\mathcal{L}$ is a self-adjoint and non-positive definite operator in $L^2(X)$ with domain $Dom(\mathcal{L})$
which is dense in $N^{1,2}(X)$ such that
\begin{align}
\mathscr{E}(f,h)=-\int_X h \cdot \mathcal{L}f d\mu, \quad \forall \; f \in Dom(\mathcal{L}), \; h\in N^{1,2}(X).
\label{eqn:HD27_1}
\end{align}
Note that $C_c^{\infty}(\mathcal{R})$ is a dense subset of $Dom(\mathcal{L})$. Suppose $f \in C_c^{\infty}(\mathcal{R})$, $h \in N^{1,2}(X)=N^{1,2}(\mathcal{R})$,
it is clear that
\begin{align*}
\mathscr{E}(f,h)=\int_{\mathcal{R}} \langle \nabla f, \nabla h\rangle d\mu= -\int_X h \cdot \Delta f d\mu.
\end{align*}
Therefore, $\mathcal{L}$ is nothing but the extension of the classical Laplacian operator, with domain as the largest dense subset of $N^{1,2}(X)$ such that the integration by parts, i.e., equation (\ref{eqn:HD27_1}), holds.
For this reason, we shall just denote $\mathcal{L}$ by $\Delta$ in the future.
Based on the generator operator $\Delta$, there is an associated heat semigroup $\left(P_t\right)_{t \geq 0}= \left( e^{t\Delta}\right)_{t \geq 0}$, which acts on
$L^2(X)$ with the following properties(c.f. Chapter 1 of~\cite{FuOsTa}).
\begin{itemize}
\item Semi-group: $P_0=Id$; $P_t \circ P_s=P_{t+s}$, for every $t, s \geq 0$.
\item Generator: $\displaystyle \lim_{t \to 0^+} \norm{\frac{1}{t}(P_t f-f) -\Delta f}{L^2(X)}=0$, for every $f \in L^2(X) \cap Dom(\Delta)$.
\item $L^2$-contractive: $\norm{P_t f}{L^2(X)}^2 \leq \norm{f}{L^2(X)}^2$, for every $f \in L^2(X)$, $t>0$.
\item Strong continuous: $\displaystyle \lim_{t \to 0^{+}}\norm{P_t f-f}{L^2(X)}=0$, for every $f \in L^2(X)$.
\item Markovian: $\norm{P_t f}{L^{\infty}(X)} \leq \norm{f}{L^{\infty}(X)}$, for every $f \in L^2(X)\cap L^{\infty}(X)$, $t>0$.
\item Heat solution: $\Delta P_t f= \frac{\partial}{\partial t} P_t f$, for every $f \in L^2(X)$ and $t>0$.
\end{itemize}
The above properties are well known in semigroup theory on Banach spaces(c.f. Section 7.4 of \cite{Evans}). Actually, for every $f \in L^2(X)$,
one can also show that $P_t f$ is the unique square-integrable solution with initial value $f$(c.f. Proposition 1.2 of~\cite{Stu95} and references therein).
We call $(P_t)_{t \geq 0}$ as the heat semigroup as usual.
Associated with this heat semigroup, there exists a nonnegative kernel function, or fundamental solution, $p(t,x,y)$,
such that
\begin{align*}
P_t(f)(y)=\int_X f(x) p(t,x,y)d\mu_x, \quad \forall \; f \in L^2(X), \; t>0.
\end{align*}
Moreover, $p$ satisfies the symmetry $p(t,x,y)=p(t,y,x)$. Interested readers are referred to Proposition 2.3 and the discussion in Section 2.4(C) of~\cite{Stu95}
for more detailed information.
As usual, we call $p(t,x,y)$ as the heat kernel.
\begin{definition}
Suppose $u \in N_{loc}^{1,2}(\Omega)$. Define
\begin{align}
\int_{\Omega} \varphi \Delta u \triangleq -\mathscr{E}(u,\varphi)
\label{eqn:HD29_2}
\end{align}
for every $\varphi \in N_c^{1,2}(\Omega)$, i.e., $\varphi \in N^{1,2}(\Omega)$ and has compact support set in $\Omega$.
Similarly, (\ref{eqn:HD29_2}) can be applied if $u \in N^{1,2}(\Omega)$ and $\varphi \in N_0^{1,2}(\Omega)$.
\label{dfn:HD29_1}
\end{definition}
Suppose $u \in N_{loc}^{1,2}(\Omega) \cap C^{2}(\Omega \backslash \mathcal{S})$, then $\Delta u|_{\Omega \backslash \mathcal{S}}$
is a continuous function.
By taking value $\infty$ on $\mathcal{S}$, we can regard $\Delta u$ as an extended function on $\Omega$. Suppose $\Delta u \in L_{loc}^2(\Omega)$, then
for every smooth test function $\varphi_i \in N_c^{1,2}(\Omega)$, we have
\begin{align*}
\int_{\Omega \backslash \mathcal{S}} \varphi_i \Delta u=-\int_{\Omega \backslash \mathcal{S}} \langle \nabla u, \nabla \varphi_i \rangle.
\end{align*}
Let $\varphi$ be the limit of $\varphi_i$ in $N^{1,2}(\Omega)$. Taking limit of the above equation shows that
\begin{align*}
\int_{\Omega} \varphi \Delta u=-\int_{\Omega} \langle \nabla u, \nabla \varphi \rangle.
\end{align*}
Therefore, whenever $u \in N_{loc}^{1,2}(\Omega) \cap C^{2}(\Omega \backslash \mathcal{S})$ and the classical $\Delta u$ is in $L_{loc}^2(\Omega)$,
we see that the LHS and RHS of (\ref{eqn:HD29_2}) holds in the classical sense. Similar argument applies
if $u \in N^{1,2}(\Omega) \cap C^2(\Omega \backslash \mathcal{S})$, $\Delta u \in L^2(\Omega)$, $\varphi \in N_0^{1,2}(\Omega)$.
Therefore, Definition~\ref{dfn:HD29_1} is justified.
Now we assume $u \in N_c^{1,2}(\Omega)$.
Then in the weak sense, for every $\varphi \in N_c^{1,2}(\Omega)$, we can define
$\int_{\Omega} \varphi \Delta u$. It is not hard to see that $\int_{\Omega} \varphi \Delta u$ makes sense even if $\varphi$ is in $N^{1,2}(\Omega)$ only.
In fact, let $\chi$ be a cutoff function with value $1$ on $\Omega'$ and vanishes
around $\partial \Omega$, where $\Omega'$ contains the support of $u$.
By Definition~\ref{dfn:HD29_1}, we have
\begin{align*}
\int_{\Omega} (\chi \varphi) \Delta u= -\mathscr{E}(u, \chi \varphi)=-\int_{\Omega} \langle \nabla u, \nabla (\chi \varphi)\rangle
=-\int_{\Omega'} \langle \nabla u, \nabla \varphi \rangle=-\int_{\Omega} \langle \nabla u, \nabla \varphi \rangle
=-\mathscr{E}(u,\varphi).
\end{align*}
The above calculation does not depend on the particular choice of $\chi$. Consequently, we can define
$\int_{\Omega} \varphi \Delta u$ as $-\mathscr{E}(u,\varphi)$.
Summarizing the above discussion, we have the following property.
\begin{proposition}[\textbf{Integration by parts}]
Suppose $\Omega$ is a domain in $X$, $f_1 \in N_c^{1,2}(\Omega)$, $f_2 \in N^{1,2}(\Omega)$.
Then we have
\begin{align}
\int_{\Omega} f_2 \Delta f_1 d\mu = -\int_{\Omega} \langle \nabla f_1, \nabla f_2 \rangle d\mu=\int_{\Omega} f_1 \Delta f_2 d\mu.
\label{eqn:HD13_2}
\end{align}
Furthermore, if $f_2 \in N^{1,2}(\Omega) \cap C^2(\Omega \backslash \mathcal{S})$ and $\Delta f_2 \in L^2(\Omega \backslash \mathcal{S})$
as a classical function, then we can understand the integral
$\int_{\Omega} f_1 \Delta f_2 d\mu=\int_{\Omega \backslash \mathcal{S}} f_1 \Delta f_2 d\mu$ in the classical sense.
Similarly, if $f_1 \in N_c^{1,2}(\Omega) \cap C^2(\Omega \backslash \mathcal{S})$,
$\Delta f_1 \in L^2(\Omega \backslash \mathcal{S})$ as a classical function, then
$\int_{\Omega} f_2 \Delta f_1 d\mu=\int_{\Omega \backslash \mathcal{S}} f_2 \Delta f_1 d\mu$ can be understood in
the classical sense. If both $f_1$ and $f_2$ locate in $N_0^{1,2}(\Omega)$, then (\ref{eqn:HD13_2}) also holds.
\label{prn:HD13_2}
\end{proposition}
\begin{definition}
Suppose $u \in N_{loc}^{1,2}(\Omega)$, $f \in L_{loc}^2(\Omega)$, we say $\Delta u \geq f$ in the weak sense whenever
\begin{align}
\int_{\Omega}(-\Delta u +f)\varphi= \mathscr{E}(u,\varphi) + \int_{\Omega} f\varphi \leq 0
\label{eqn:HD29_1}
\end{align}
for every nonnegative test function $\varphi \in N_c^{1,2}(\Omega)$. We call $u$ subharmonic if $\Delta u \geq 0$ in the weak sense.
We call $u$ superharmonic if $-u$ is subharmonic. We call $u$ harmonic if $u$ is both subharmonic and superharmonic.
\label{dfn:HD29_2}
\end{definition}
Due to Proposition~\ref{prn:HD04_1}, for a $u \in N_{loc}^{1,2}(\Omega)$, in order to check (\ref{eqn:HD29_1}) for all $\varphi \in N_c^{1,2}(\Omega)$,
it suffices to check all smooth nonnegative test functions with supports in $\Omega \backslash \mathcal{S}$.
It is important to notice that the restriction of $\Delta$ on $\mathcal{R}$ is the classical Laplacian on Riemannian manifold.
In fact, if a function $u$ is harmonic in the above sense, then $u |_{\mathcal{R}}$ is harmonic function in the distribution sense.
By standard improving regularity theory of elliptic equations, we know our $u$ is smooth and $\Delta u=0$ in the classical sense.
Similarly, one can follow the standard route to define heat solution (sub solution, super solution) for the heat operator
$\square= \left( \frac{\partial}{\partial t} -\Delta \right)$ in the weak sense. We leave these details to interested readers.
It is quite clear that a weak heat solution is a smooth function when restricted on $\mathcal{R} \times (0,T]$, by
standard improving regularity theory of heat equations(c.f. Chapter 7 of~\cite{Evans}).
\subsection{Harmonic functions and heat flow solutions on model space}
Suppose $K$ is a compact subset of $\Omega \backslash \mathcal{S}$, it is clear that $K$ is also a compact subset of $\Omega$.
However, the reverse is not true. If $K$ is a compact subset of $\Omega$, then $K\backslash \mathcal{S}$ may not be a compact
subset of $\Omega \backslash \mathcal{S}$.
For this reason, we see that $N_{loc}^{1,2}(\Omega) \subset N_{loc}^{1,2}(\Omega \backslash \mathcal{S})$ and not equal if $\mathcal{S} \neq \emptyset$, even if
$\mathcal{S}$ has very high codimension.
However, if we restrict our attention only on bounded subharmonic functions, then the above difference will vanish.
\begin{proposition}[\textbf{Extension of bounded subharmonic functions}]
Suppose $\Omega$ is a bounded open domain in $X$, $u$ is a bounded subharmonic function on $\Omega \backslash \mathcal{S}$.
Then $u \in N_{loc}^{1,2}(\Omega)$ and it is subharmonic on $\Omega$.
\label{prn:HD16_2}
\end{proposition}
\begin{proof}
It suffices to prove that $u \in N_{loc}^{1,2}(\Omega)$.
Note that by definition, we only have $u \in N_{loc}^{1,2}(\Omega \backslash \mathcal{S})$, which is a superset of $N_{loc}^{1,2}(\Omega)$.
In fact, for each small $r>0$, one can construct a Lipschitz cutoff function
\begin{align}
\chi(x)=\phi \left(\frac{d(x,\mathcal{S})}{r} \right),
\label{eqn:HD16_1}
\end{align}
where $\phi$ is a cutoff function on $[0,\infty)$ which is equivalent to $1$ on $[0,1]$, $0$ on $[2, \infty)$, and $|\phi'| \leq 2$.
By the assumption of Minkowski codimension of $\mathcal{S}$, we have
\begin{align}
| \Omega \cap \supp \chi| \leq Cr^3, \quad \int_{\Omega} |\nabla \chi|^2 \leq C r^{-2}\int_{\Omega \cap \{\nabla \chi \neq 0\}} \chi^2 \leq Cr.
\label{eqn:HD16_2}
\end{align}
Fix a relatively compact subset $\Omega' \subset \Omega$, we can find a cutoff function $\eta$ which is identically $1$ on $\Omega'$ and vanishes around
$\partial \Omega$. Moreover, $|\nabla \eta| \leq C$, which depends on $\Omega'$ and $\Omega$.
By adding a constant if necessary, we can assume $u \geq 0$.
Note that $u$ is subharmonic on $\Omega \backslash \mathcal{S}$, $u\eta^2 (1-\chi)^2$ can be chosen as a test function. It follows from definition that
\begin{align*}
0&\leq \int_{\Omega \backslash \mathcal{S}} (\Delta u) u\eta^2(1-\chi)^2
=-\int_{\Omega \backslash \mathcal{S}} \langle \nabla u, \nabla (u\eta^2(1-\chi)^2) \rangle \\
& =-\int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 (1-\chi)^2
+ \int_{\Omega \backslash \mathcal{S}} u \langle \nabla u, -2(1-\chi)^2\eta \nabla \eta+ 2\eta^2(1-\chi) \nabla \chi \rangle.
\end{align*}
Note that $u,\eta,\nabla \eta$ are bounded. Then H\"{o}lder inequality applies.
\begin{align*}
\frac{1}{2} \int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 (1-\chi)^2
&\leq C+ C\int_{\Omega \backslash \mathcal{S}}\eta^2 (1-\chi) |\nabla u||\nabla \chi|\\
&\leq C+ \frac{1}{4}\int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 (1-\chi)^2 + C\int_{\Omega \backslash \mathcal{S}} \eta^2 |\nabla \chi|^2.
\end{align*}
Recall the definition of $\chi$ in (\ref{eqn:HD16_1}) and estimate (\ref{eqn:HD16_2}).
Let $r \to 0$, the above inequality yields that
\begin{align*}
\int_{\Omega \backslash \mathcal{S}} |\nabla u|^2\eta^2 \leq C,
\end{align*}
which forces that $\displaystyle \int_{\Omega' \backslash \mathcal{S}} |\nabla u|^2 \leq C$.
Hence $u \in W^{1,2}(\Omega' \backslash \mathcal{S})$ since $u$ is bounded. This is the same to say $u \in N^{1,2}(\Omega')$.
By the arbitrary choice of $\Omega'$, we have proved that $u \in N_{loc}^{1,2}(\Omega)$.
\end{proof}
We now move on to the discussion of heat kernels.
\begin{proposition}[\textbf{Heat Kernel estimates}]
The exists a unique heat kernel $p(t,x,y)$ of $X$, with respect to the Dirichlet form $\mathscr{E}=\langle \nabla \cdot, \nabla \cdot \rangle$.
Moreover, $p(t,x,y)$ satisfies the following properties.
\begin{itemize}
\item Stochastically completeness. In other words, we have
\begin{align*}
\int_X p(t,x,y) d\mu_x=1
\end{align*}
for every $x \in X$.
\item The Gaussian estimate holds. In other words, there exists a constant $C$ depending only on $n,\kappa$ such that
\begin{align}
\frac{1}{C}t^{-n} e^{-\frac{d^2(x,y)}{3t}} \leq p(t,x,y) \leq Ct^{-n} e^{-\frac{d^2(x,y)}{5t}}
\label{eqn:HC31_7}
\end{align}
for every $x,y \in X$ and $t>0$.
\item For each positive integer $j$, there is a constant $C=C(n,\kappa,j)$ such that
\begin{align}
\left| \left( \frac{\partial}{\partial t} \right)^j p(t,x,y)\right| \leq Ct^{-n-j} e^{-\frac{d^2(x,y)}{5t}}
\label{eqn:HD08_1}
\end{align}
for every $x,y \in X$ and $t>0$.
\label{prn:HC29_7}
\end{itemize}
\label{prn:HD07_1}
\end{proposition}
\begin{proof}
Since $X$ satisfies the doubling property and has a uniform $(1,2)$-Poincar\`{e} constant, by Corollary~\ref{cly:HD20_1}
and Proposition~\ref{prn:HC29_2},
the existence of the heat Kernel follows from the work of Sturm (c.f.Proposition 2.3 of~\cite{Stu95}).
The uniqueness of the heat kernel follows from the uniqueness of the heat semigroup.
The stochastic completeness is guaranteed by the doubling property, see Theorem 4 and the following remarks of~\cite{Stu94}.
The Gaussian estimate follows from Corollary 4.2 and Corollary 4.10 in Sturm's paper~\cite{Stu96}, where $C$
depends on the volume doubling condition and the $(1,2)$-Poincar\`{e} constant.
The heat kernel derivative estimate, inequality (\ref{eqn:HD08_1}), follows from Corollary 2.7. of ~\cite{Stu95}, whose proof follows the same line
as Theorem 6.3 of~\cite{Saloff}.
\end{proof}
In general, the estimates of heat kernel only hold almost everywhere with respect to the measure $d\mu$.
However, in the current situation, when restricted on $\mathcal{R} \times (0,\infty)$, $p$ is clearly a smooth function.
Therefore, (\ref{eqn:HC31_7}) and (\ref{eqn:HD08_1}) actually hold true everywhere away from $\mathcal{S}$.
Note that $\Delta p=\frac{\partial}{\partial t} p$ clearly locates in $L^2(X) \cap C^{\infty}(\mathcal{R})$.
Hence integration by parts (Proposition~\ref{prn:HD13_2}) applies. Then by standard radial cutoff function construction and direct calculation,
Proposition~\ref{prn:HD07_1}
yields the following estimates immediately.
\begin{corollary}[\textbf{Off-diagonal integral estimates of heat kernel}]
For every $r>0, t>0$, and $x_0 \in X$, we have
\begin{align}
&\int_{X \backslash B(x_0,2r)} p^2 d\mu_x<Ct^{-n}e^{-\frac{r^2}{5t}}, \label{eqn:HD08_7}\\
&\int_{X \backslash B(x_0,2r)} |\nabla p(t,x,x_0)|^2 d\mu_x
< C \left( \frac{1}{t} + \frac{1}{r^2}\right) t^{-n} e^{-\frac{r^2}{5t}}, \label{eqn:HD08_4}
\end{align}
for some $C=C(n,\kappa)$. Consequently, we have
\begin{align}
\int_0^{t} \int_{X \backslash B(x_0,2r)} \left( p^2+|\nabla p|^2 \right) d\mu_x ds
< C \int_0^t \left(1+\frac{1}{s} + \frac{1}{r^2}\right) s^{-n} e^{-\frac{r^2}{5s}} ds.
\label{eqn:HD08_3}
\end{align}
\end{corollary}
By virtue of Proposition~\ref{prn:HD04_1}, smooth functions are dense in $N^{1,2}(X)$. Then it is easy to see that $(X,g,d\mu)$, together with the heat process,
has non-negative Ricci curvature in the sense of Bakry-Emery, i.e., $X \in CD(0,\infty)$ by the notation of Bakry-Emery(c.f.~\cite{BaEm},~\cite{Bak}).
The following Proposition is nothing but part of Proposition 2.1 of~\cite{Bak}.
\begin{proposition}[\textbf{Weighted Sobolev inequality}]
For every function $f \in N^{1,2}(X)$, every $t>0$ and every $y \in X$, we have
\begin{align}
\int_X f^2(x) p(t,x,y) d\mu_x - \left(\int_X f(x) p(t,x,y)d\mu_x \right)^2 \leq 2t \int_X |\nabla f|^2(x) p(t,x,y) d\mu_x.
\label{eqn:HD05_3}
\end{align}
In other words, for every $t>0$, with respect to the probability measure $p(t,x,y)d\mu_x$, $L^2$-Sobolev inequality holds
with the uniform Sobolev constant $\frac{1}{2t}$.
\label{prn:HD07_2}
\end{proposition}
On a Riemannian manifold with proper geometry bound, the heat kernel can be regarded as a solution starting from a $\delta$-function.
This property also holds for every $X \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{proposition}[\textbf{$\delta$-function property of heat kernel}]
Suppose $w$ is a function on $[0,t] \times X$, differentiable along the time direction,
$w(s,\cdot) \in N_c^{1,2}(X)$ for each $s \in [0,t]$.
Moreover, we assume $\displaystyle \limsup_{s \to 0^{+}} \norm{w(s,\cdot)}{L^1(X)}<\infty$, $w$ is continuous at $(0,x_0)$.
Then we have
\begin{align}
&-w(0,x_0) +\int_{X} w(t,x)p(t,x,x_0)d\mu_x=\int_0^{t} \int_X \left\{ \left(\frac{\partial}{\partial s}+\Delta \right) w(s,x) \right\} p(s,x,x_0) d\mu_x ds.
\label{eqn:HD09_1}
\end{align}
Consequently, equation (\ref{eqn:HD09_1}) holds for functions $w(s,x)+a(s)$ where $a$ is a differentiable function of time.
\label{prn:HD09_1}
\end{proposition}
\begin{proof}
Clearly, (\ref{eqn:HD09_1}) holds if $w(s,\cdot) \equiv a(s)$. Therefore, it suffices to show (\ref{eqn:HD09_1}) when $w(s,\cdot) \in N_c^{1,2}(X)$ for each $s$.
For simplicity of notation, we denote $p(t,x,x_0)$ by $p$ and assume $d\mu_x$ as the default measure. It follows from integration by parts that
\begin{align*}
\frac{d}{dt} \int_{X} wp
=\int_X \left\{ \left(\frac{\partial}{\partial t}+\Delta \right) w \right\} p + \int_X w \left\{\left(\frac{\partial}{\partial t}-\Delta \right) p\right\}
=\int_X \left\{ \left(\frac{\partial}{\partial t}+\Delta \right) w \right\} p.
\end{align*}
For each $\epsilon>0$, we can find $\delta$ small enough such that $|w(s,x)-w(0,x_0)|<\epsilon$ whenever $0<s<\delta^2$ and $d(x,x_0)<\delta$.
Then the heat kernel estimate implies that
\begin{align*}
&\quad \left|-w(0,x_0)+\lim_{t \to 0^+} \int_X w(t,x)p(t,x,x_0)d\mu_x \right|\\
&= \left|\lim_{t \to 0^+} \int_X \left\{w(t,x)-w(0,x_0)\right\}p(t,x,x_0)d\mu_x \right|\\
&\leq \lim_{t \to 0^{+}} \left\{ |w(0,x_0)|\int_{X \backslash B(x_0,\delta)} p(t,x,x_0)d\mu_x
+ \int_{X \backslash B(x_0,\delta)} |w(t,x)| p(t,x,x_0)d\mu_x +\epsilon \right\}\\
&\leq \epsilon.
\end{align*}
By arbitrary choice of $\epsilon$, we have $\displaystyle \lim_{t \to 0^+} \int_X w(t,x)p(t,x,x_0)d\mu_x =w(0,x_0)$. Plugging this relationship into the integration of previous equation,
we obtain (\ref{eqn:HD09_1}).
\end{proof}
Based on the excellent properties of heat kernels, from Proposition~\ref{prn:HC29_7} to Proposition~\ref{prn:HD09_1}, we are ready to generalize the celebrated Cheng-Yau estimate (c.f.~\cite{ChYau}) to our setting.
We basically follow the paper~\cite{KoRaSh}.
However, due to the essential importance of this estimate and the excellent geometry of our underlying space, we write down a simplified proof here.
\begin{proposition} [\textbf{Cheng-Yau type gradient estimate}]
Suppose $\Omega=B(x_0,4r)$ for some $r>0$ and $x_0 \in \mathcal{R}$.
Suppose $u \in L^{\infty}(\Omega) \cap N^{1,2}(\Omega)$ and $u$ satisfies the equation
\begin{align}
\Delta u=h
\label{eqn:HE03_3}
\end{align}
for some $h \in L^2(\Omega) \cap C^{\frac{1}{2}}(\Omega \backslash \mathcal{S})$.
Then we have
\begin{align}
|\nabla u|(x_0) \leq \frac{C}{r} \left( \norm{u}{L^{\infty}(\Omega)} + \norm{h}{L^2(\Omega)}\right)
\label{eqn:HD01_1}
\end{align}
for a constant $C=C(n,\kappa)$.
\label{prn:HC29_6}
\end{proposition}
\begin{proof}
Without loss of generality, we assume $r=1$ and $\norm{u}{L^{\infty}(\Omega)}=1$.
Let $\chi$ be a Lipschitz cutoff function such that $\chi \equiv 1$ on $B(x_0,1)$ and vanishes outside $B(x_0, 1.5)$ such that $|\nabla \chi| \leq 5$.
Define
\begin{align*}
&a(t) \triangleq P_t(u\chi)(x_0), \\
& w(t,x) \triangleq u(x)\chi(x)-a(t), \\
&J(t) \triangleq \frac{1}{t} \int_0^{t} \int_X |\nabla w(s,x)|^2 p(s,x,x_0) d\mu_x ds, \quad \forall \; t>0,\\
&J(0) \triangleq \lim_{t \to 0^{+}} J(t)=|\nabla w(0, x_0)|^2=|\nabla u(x_0)|^2.
\end{align*}
From the definition of $w(t,x)$, it is clear that $\displaystyle \int_X w(t,x) p(t,x,x_0)d\mu_x=0$. Therefore, inequality (\ref{eqn:HD05_3}) applies to obtain
\begin{align}
\int_X w^2(t,x) p(t,x,x_0)d\mu_x \leq 2t\int_X |\nabla w|^2p(t,x,x_0) d\mu_x.
\label{eqn:HD09_3}
\end{align}
Note that
\begin{align}
&|a(t)|= \left|\int_X u\chi p(t,x,x_0)d\mu_x\right| \leq \int_X |u\chi| p(t,x,x_0)d\mu_x
\leq \int_X p(t,x,x_0)d\mu_x \leq 1, \label{eqn:HF20_1}\\
&|w(t,x)|= |u(x)\chi(x) -a(t)| \leq |u(x)\chi(x)|+|a(t)| \leq 2, \quad \forall \; x \in X. \label{eqn:HD06_5}
\end{align}
By virtue of the definition of $J$, it is clear that
\begin{align}
|\nabla u(x_0)|^2=J(0)=-\int_0^{1} J'(t) dt +J(1).
\label{eqn:HD06_6}
\end{align}
However, in light of (\ref{eqn:HD09_3}), we have
\begin{align}
J'(t)&=-\frac{1}{t} J(t)+\frac{1}{t} \int_X |\nabla w|^2p(t,x,x_0) d\mu_x
\geq-\frac{1}{t} J(t)+\frac{1}{2t^2} \int_X w^2(t,x) p(t,x,x_0)d\mu_x \notag\\
&=\frac{1}{t^2} \left(-tJ(t) +\frac{1}{2}\int_X w^2(t,x) p(t,x,x_0)d\mu_x \right)
=\frac{F(t)}{t^2}, \label{eqn:HD09_8}
\end{align}
where we used the definition
\begin{align}
F(t) \triangleq -tJ(t) +\frac{1}{2}\int_X w^2(t,x) p(t,x,x_0)d\mu_x. \label{eqn:HD09_6}
\end{align}
Therefore, according to (\ref{eqn:HD06_5}) and (\ref{eqn:HD09_8}), we see that (\ref{eqn:HD06_6}) can be rewritten as
\begin{align}
|\nabla u(x_0)|^2 \leq -\int_0^1 \frac{F(s)}{s^2} ds -F(1) +\frac{1}{2} \int_X w^2p
\leq 2-\int_0^1 \frac{F(s)}{s^2} ds -F(1).
\label{eqn:HD09_7}
\end{align}
Consequently, in order to estimate $|\nabla u(x_0)|$, it suffices to estimate $F(t)$.
We now focus on the estimate of $F(t)$. First, we observe that
\begin{align*}
w^2(x,t)=u\chi (u\chi -2a(t)) +a^2(t).
\end{align*}
The first part of the right hand side of the above equation is a function in $N_c^{1,2}(X)$ for each $t$. It is Lipschitz continuous around $(0,x_0)$,
since $x_0$ is a smooth point and the standard improving regularity theory of elliptic functions applies here.
The second part is a differentiable function of time.
Therefore, (\ref{eqn:HD09_1}) applies for $w^2$ and we obtain
\begin{align}
\int_0^t \int_X \left\{ \left( \frac{\partial}{\partial s} + \Delta \right) w^2(s,x)\right\}p(s,x,x_0)d\mu_x ds
=\int_X w^2(t,x)p(t,x,x_0) d\mu_x,
\label{eqn:HD06_7}
\end{align}
since $w(0,x_0)=0$. On the other hand, the fact that $u$ is a weak solution of (\ref{eqn:HE03_3}) implies that
\begin{align*}
|\nabla w|^2=\frac{1}{2}\Delta w^2-w\Delta w
=\frac{1}{2}\Delta w^2-w(u\Delta \chi +\chi h+2 \langle \nabla u, \nabla \chi \rangle)
\end{align*}
in the weak sense.
Plugging the above equation and (\ref{eqn:HD06_7}) into the definition of $J(t)$, we have
\begin{align*}
tJ(t)
&=\int_0^t\int_X \left\{ \frac{1}{2}\left( \frac{\partial}{\partial s}+\Delta\right) w^2
-w(u\Delta \chi+\chi h +2 \langle \nabla u, \nabla \chi \rangle) \right\} p(s,x,x_0)d\mu_x ds \notag\\
&=\int_0^t\int_X \left\{ \frac{1}{2}w^2
-w(u\Delta \chi +\chi h+2 \langle \nabla u, \nabla \chi \rangle) \right\} p(s,x,x_0)d\mu_x ds.
\end{align*}
From the definition of $F(t)$, the above equation is the same as
\begin{align*}
&\quad F(t)-\int_0^t\int_X w\chi h p(s,x,x_0) d\mu_x ds \\
&=\int_0^{t}\int_X w(u\Delta \chi+2 \langle \nabla u, \nabla \chi \rangle) p(s,x,x_0) d\mu_x ds\\
&=\int_0^t \int_X -up\langle\nabla w, \nabla \chi \rangle -uw \langle \nabla p, \nabla \chi\rangle + pw \langle \nabla u, \nabla \chi\rangle d\mu_x ds\\
&=\int_0^t \int_X (w-u)p \langle \nabla u, \nabla \chi\rangle d\mu_x ds
+\int_0^t\int_X -uw \langle \nabla p, \nabla \chi \rangle d\mu_x ds\\
&=\int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left\langle (w-u) p\nabla u-uw \nabla p, \nabla \chi \right\rangle d\mu_x ds.
\end{align*}
Similar to (\ref{eqn:HF20_1}) and (\ref{eqn:HD06_5}), we have $|w-u|\leq 2$ and $|uw| \leq 2$.
By the choice of $\chi$ and the H\"{o}lder inequality, it is clear that
\begin{align}
&\quad \left|F(t)-\int_0^t\int_X w\chi h p d\mu_x ds\right|\\
&\leq 10 \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} (p|\nabla u| + |\nabla p|) d\mu_x ds \notag\\
&\leq C\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla u| ^2+1\right) \right)^{\frac{1}{2}}
\cdot \left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla p| ^2+p^2\right) \right)^{\frac{1}{2}} \notag\\
&\leq C\left(1+ \norm{h}{L^2(\Omega)} \right) \sqrt{t}\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla p| ^2+p^2\right) \right)^{\frac{1}{2}}.
\label{eqn:HD09_9}
\end{align}
Note that in the last step of the above inequality, we used the following Caccioppoli-type inequality:
\begin{align*}
\int_{B(x_0,2)} |\nabla u|^2 d\mu_x \leq C \int_{B(x_0,4)} (u^2+h^2) d\mu_x \leq C(n,\kappa) \left(1+\norm{h}{L^2(\Omega)} \right)^2,
\end{align*}
which can be proved by multiplying equation (\ref{eqn:HE03_3}) on both sides by $\tilde{\chi}^2 u$ and doing integration by parts, for some cutoff function $\tilde{\chi}$.
On the other hand, note that
\begin{align*}
\left| \int_0^t\int_X w\chi h p d\mu_x ds\right| &=\left| \int_0^t\int_{B(x_0,2) \backslash B(x_0,1)} w\chi h p d\mu_x ds\right|\\
&\leq 2 \int_0^t\int_{B(x_0,2) \backslash B(x_0,1)} |h| p d\mu_x ds\\
&\leq 2\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} h^2 d\mu_x dt \right)^{\frac{1}{2}}
\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} p^2d\mu_x dt \right)^{\frac{1}{2}}\\
&\leq 2\sqrt{t} \norm{h}{L^2(\Omega)} \left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla p| ^2+p^2\right) \right)^{\frac{1}{2}}.
\end{align*}
Plugging the above inequality into (\ref{eqn:HD09_9}), we obtain
\begin{align*}
|F(t)| \leq C\left(1+ \norm{h}{L^2(\Omega)} \right) \sqrt{t}\left( \int_0^t \int_{B(x_0,2) \backslash B(x_0,1)} \left(|\nabla p| ^2+p^2\right) \right)^{\frac{1}{2}}.
\end{align*}
Combining the above inequality, (\ref{eqn:HD08_3}) and the fact $r=\frac{1}{2}$, $0<t<1$, we obtain the desired estimate of $F(t)$:
\begin{align}
\frac{|F(t)|}{1+ \norm{h}{L^2(\Omega)}}<C\sqrt{t} \left( \int_0^t \frac{1}{s} e^{-\frac{1}{20s}}ds\right)^{\frac{1}{2}}< C\sqrt{t} \left( \int_0^t \frac{1}{s^2}e^{-\frac{1}{20s}} ds\right)^{\frac{1}{2}}
< Ct^{\frac{1}{2}}e^{-\frac{1}{40t}},
\label{eqn:HD09_4}
\end{align}
which implies
\begin{align}
\frac{1}{1+ \norm{h}{L^2(\Omega)}} \int_0^{1}\frac{|F(s)|}{s^2} ds< C \int_0^1 s^{-\frac{3}{2}} e^{-\frac{1}{40s}}ds<C\int_0^1\frac{1}{s^2} e^{-\frac{1}{40s}}ds<C.
\label{eqn:HD09_5}
\end{align}
Therefore, we obtain the estimate of $|\nabla u(x_0)|$ by plugging (\ref{eqn:HD09_4}) and (\ref{eqn:HD09_5}) into (\ref{eqn:HD09_7}).
\end{proof}
An interesting application of Cheng-Yau type inequality is the following Liouville theorem.
\begin{corollary}[\textbf{Liouville theorem}]
Suppose $u$ is a bounded harmonic function on $\mathcal{R}$, then $u \equiv C$.
\label{cly:HE08_1}
\end{corollary}
\begin{proof}
By virtue of Proposition~\ref{prn:HD16_2}, the extension property of subharmonic functions, we know that $u \in N_{loc}^{1,2}(X)$.
In light of Proposition~\ref{prn:HD04_1}, the dense property of smooth functions in $N_0^{1,2}(\Omega)$, it is clear that $\Delta u=0$ on $X$ in the weak sense.
Fix $x_0 \in \mathcal{R}$ and a large $r>0$, by Cheng-Yau estimate in Proposition~\ref{prn:HC29_6},
we have
\begin{align*}
|\nabla u|(x_0) < \frac{C}{r}
\end{align*}
for some uniform constant $C$. Let $r \to \infty$, we see that $|\nabla u|(x_0)=0$. It follows that $\nabla u \equiv 0$ on $\mathcal{R}$
by the arbitrary choice of $x_0 \in \mathcal{R}$. Consequently, $u \equiv C$ on $\mathcal{R}$.
\end{proof}
\begin{proposition}[\textbf{Estimates for Dirichlet problem solution}]
Suppose $\Omega$ is a bounded open set of $X$, $f $ is a continuous function in $N^{1,2}(\Omega)$. Then we have the following properties.
\begin{itemize}
\item There is a unique solution $u \in N^{1,2}(\Omega)$ solving the Dirichlet problem
\begin{align}
\Delta u=0, \quad \textrm{in} \; \Omega; \quad
(u-f)|_{\partial \Omega}=0
\label{eqn:HE03_2}
\end{align}
in the weak sense of traces. In other words, $\Delta u=0$ in the weak sense and $u-f \in N_0^{1,2}(\Omega)$.
\item Weak maximum principle holds for $u$, i.e.,
\begin{align}
\sup_{x \in \Omega} u(x)=\sup_{x \in \partial \Omega} u(x), \quad \inf_{x \in \Omega} u(x)=\inf_{x \in \partial \Omega} u(x).
\end{align}
\item Strong maximum principle holds for $u$, i.e., if there is an interior point $x_0 \in \Omega$ such that
$\displaystyle u(x_0)= \sup_{x \in \Omega} u(x)$ or $\displaystyle u(x_0)=\inf_{x \in \Omega} u(x)$, then $u$ is a constant.
\end{itemize}
\label{prn:HD11_1}
\end{proposition}
\begin{proof}
The existence and uniqueness of the Dirichlet problem follows from Theorem 7.12 and Theorem 7.14 of Cheeger's work~\cite{Cheeger99},
where a much more general case was considered. The weak maximum principle follows from the uniqueness.
Also, the weak maximum principle was proved by Shanmugalligam in~\cite{ShanHar}.
The strong maximum principle follows from Corollary 6.4 of~\cite{KiSh}. However, we shall give a simpler proof here since our underlying space
has much better geometry.
Without loss of generality, we assume $u \geq 0$ on $\partial \Omega$ and $u$ is not a constant.
It suffices to show that $u>0$ in $\Omega$. Clearly, by classical harmonic function theory on Riemannian manifold and continuity of
$u$(c.f.~Proposition~\ref{prn:HD16_1}), it is clear that
$u>0$ on $\Omega \cap \mathcal{R}$. Therefore, we only need to show that $u>0$ on $\Omega \cap \mathcal{S}$.
We argue by contradiction. If this statement were wrong, we can find a point $y_0 \in \Omega \cap \mathcal{S}$ such that
$u(y_0)=0$. Choose $r$ small enough such that $B(y_0,2r) \subset \Omega$.
For each small $\epsilon$, choose $\tau$ small enough such that
\begin{align*}
\left| \Omega_{\tau} \cap B(y_0,2r) \right| < \epsilon |B(y_0,2r)|,
\end{align*}
where $\Omega_{\tau}=\{x \in \Omega| u(x) \leq \tau\}$. Note that $\tau$ can be chosen since $\Omega_0 \cap B(y_0,2r)$ is a subset of
$\mathcal{S}$ which has zero measure, and $u$ is continuous.
Now consider the function $\tau-u$, which is obviously harmonic. Let $(\tau-u)^{+}$ be $\max\{\tau-u, 0\}$. Then
$(\tau-u)^{+}$ is a bounded, continuous, subharmonic function in $N^{1,2}(B(y_0,2r))$. So Moser iteration applies to obtain
\begin{align*}
\sup_{B(y_0,r)} |(\tau-u)^{+}|^2 \leq C \int_{B(y_0,2r)} |(\tau-u)^{+}|^2=C \int_{\Omega_{\tau} \cap B(y_0,2r)} |(\tau-u)^{+}|^2 \leq C\epsilon r^{2n} \tau^2
\end{align*}
for some $C=C(n,\kappa)$. Choose $\epsilon$ small enough such that $C\epsilon^2 r^{2n}<\frac{1}{4}$. Then we have
\begin{align*}
\sup_{B(y_0,r)} |(\tau-u)^{+}| <\frac{\tau}{2},
\end{align*}
which implies that $u>\frac{\tau}{2}$ on $B(y_0,r)$. In particular, $u(y_0) \geq \frac{\tau}{2}>0$, which contradicts the assumption $u(y_0)=0$.
\end{proof}
Clearly, the essential stuff in the proof of the strong maximum principle of Proposition~\ref{prn:HD11_1} is a delicate use of elliptic Moser iteration.
In equation (\ref{eqn:HE03_2}), if we replace the operator $\Delta$ by $\square$, the heat operator, then one can easily obtain a strong maximum principle for heat equation solutions, based on a parabolic Moser iteration. The details are left to the interested readers.
On the other hand, if we replace the right hand side of equation (\ref{eqn:HE03_2}) by a function $h$, we can also obtain uniqueness and existence of solutions.
\begin{proposition}[\textbf{Existence and uniqueness of Poisson equation solution}]
Suppose $\Omega$ is a bounded open set of $X$, $f \in N^{1,2}(\Omega)$, $h \in L^2(\Omega)$.
Then there exists a unique $u \in N^{1,2}(\Omega)$ such that
\begin{align}
\Delta u=h, \quad \textrm{in} \; \Omega; \quad
(u-f)|_{\partial \Omega}=0.
\label{eqn:HE03_1}
\end{align}
\label{prn:HE03_1}
\end{proposition}
\begin{proof}
First, let us consider the Poisson equation
\begin{align*}
\Delta v=h, \; \textrm{in} \; \Omega; \quad v \in N_0^{1,2}(\Omega).
\end{align*}
By standard functional analysis, the existence of the above equation is guaranteed by Riesz representation theorem.
The uniqueness follows from the irreducibility of $\mathscr{E}$.
Second, it is obvious that there is a bijective map between the solution $u$ of (\ref{eqn:HE03_1})
and harmonic solution $w$ of (\ref{eqn:HE03_2}) by $u=w+v$. Therefore, the existence and uniqueness of
(\ref{eqn:HE03_1}) follows from Proposition~\ref{prn:HD11_1}.
\end{proof}
Combining the strong maximum principle for harmonic functions in Proposition~\ref{prn:HD11_1} with the heat kernel estimates, we obtain the
strong maximum principle for subharmonic functions.
\begin{proposition} [\textbf{Strong maximum principle for subharmonic functions}]
Suppose $\Omega$ is a bounded domain of $X$, $u$ is a continuous subharmonic function in $N^{1,2}(\Omega)$. Then we have
\begin{align}
\sup_{x \in \Omega} u(x) = \sup_{x \in \partial \Omega} u(x).
\label{eqn:HC29_2}
\end{align}
In other words, the weak maximum principle holds for subharmonic functions. Moreover, if $\displaystyle \sup_{x \in \Omega}u(x)$ is achieved at
some point $x_0 \in \Omega$, then $u$ is a constant. Namely, the strong maximum principle holds for subharmonic functions.
\label{prn:HC29_5}
\end{proposition}
\begin{proof}
The weak maximum principle is well known in literature.
For example, see Lemma 4 of~\cite{Stu94} and the reference therein.
We provide a simple proof here for the convenience of the readers.
Let $v$ be the unique harmonic function such that $(u-v)|_{\partial \Omega}=0$. Let $w=u-v$.
Then we have
\begin{align*}
\Delta w \geq 0, \quad w \in N_0^{1,2}(\Omega).
\end{align*}
We shall show that $w \leq 0$ in $\Omega$ by a contradiction argument.
Suppose $\displaystyle \sup_{x \in \Omega} w>0$, then we can assume this value is $2\epsilon_0$ and is achieved at an interior point $x_0$.
Define
\begin{align*}
\tilde{w} \triangleq
\begin{cases}
\max \{w, \epsilon_0\}-\epsilon_0, & \textrm{on} \quad \Omega,\\
0, &\textrm{on} \quad X \backslash \Omega.
\end{cases}
\end{align*}
Clearly, $\tilde{w}$ is subharmonic and support of $\tilde{w}$ is a compact subset of $\Omega$. By trivially extending $\tilde{w}$ in the time direction, we can apply (\ref{eqn:HD09_1}) to obtain
\begin{align*}
-\tilde{w}(x_0) +\int_{\Omega} \tilde{w}(x)p(t,x,x_0)d\mu_x=\int_0^{t} \int_{\Omega} p(s,x,x_0) \Delta \tilde{w}(x) d\mu_x ds.
\end{align*}
Note that integration by parts works here.
Let $t \to \infty$, the heat kernel upper bound (\ref{eqn:HC31_7}) implies that
\begin{align}
0< \epsilon_0=\tilde{w}(x_0)=-\int_0^{\infty} \int_{\Omega} p(s,x,x_0) \Delta \tilde{w}(x) d\mu_x ds \leq 0,
\label{eqn:HD16_5}
\end{align}
which is impossible.
The strong maximum principle can be proved as that in Proposition~\ref{prn:HD11_1}. Actually, $-w$ is a nonnegative superharmonic function on $\Omega$.
If $w$ is not a constant, then we can regard $-w$ as $u$ in the second part of the proof of Proposition~\ref{prn:HD11_1}.
Then everything goes through since only Moser iteration for subharmonic function is used there.
\end{proof}
\begin{proposition}[\textbf{Removing singularity of harmonic functions}]
Suppose $\Omega$ is an open domain in $X$, $u$ is a bounded harmonic function on $\Omega \backslash \mathcal{S}$.
Then $u$ can be regarded as a harmonic function on $\Omega$. Moreover, on each compact subset of $\Omega$, $u$ is uniformly Lipschitz continuous.
In particular, $u$ can be extended continuously over the singular set $\Omega \cap \mathcal{S}$.
\label{prn:HD16_1}
\end{proposition}
\begin{proof}
In light of Proposition~\ref{prn:HD16_2}, we see that $u \in N_{loc}^{1,2}(\Omega)$. Since $\Delta u=0$ on $\Omega \backslash \mathcal{S}$, we see that
\begin{align*}
\int_{\Omega} \langle \nabla u, \nabla \varphi \rangle=0
\end{align*}
for every smooth test function supported on $\Omega \backslash \mathcal{S}$. However, such functions are dense in $N_0^{1,2}(\Omega)$, so the above
equation actually holds form every $\varphi \in N_c^{1,2}(\Omega)$. Therefore, $u$ is harmonic in $\Omega$ by definition.
The Lipschitz continuity follows from Proposition~\ref{prn:HC29_6} and the density and weak convexity of $\mathcal{R}$.
\end{proof}
Note that the weak convexity of $\mathcal{R}$ is important for that $u$ can be extended over singularities. For otherwise, the limit of $u(x_i)$ for $x_i \to x_0$ may depends
on the choice of sequence $\{x_i\}$, where $x_0 \in \mathcal{S}, x_i \in \mathcal{R}$. If $\mathcal{R}$ is not convex, there is an easy counter example
of Proposition~\ref{prn:HD16_1}. Let $X$ be the union of two cones $C(S^3/\Gamma)$, for some finite group $\Gamma \in ISO(S^3)$, by identifying two vertices.
In this case, $\mathcal{S}$ is the isolated vertex $O$. Let $u$ be $1$ on one branch and $0$ on the other, then it is clear that $u$ is a harmonic function
on $X \backslash \mathcal{S}$. However, $u$ can not take a value at $O$ so that $u$ is continuous. Of course, convexity of $\mathcal{R}$ is only a sufficient
condition to guarantee the continuity extension. It can be replaced by other weaker conditions.
Moreover, based on Proposition~\ref{prn:HC29_6}, one can obtain uniform gradient estimate of $|\nabla p(t,\cdot, x_0)|$, which depends only
on $t,n$ and $\kappa$. Hence the heat kernel $p(t,\cdot, x_0)$ is a continuous function on $X \times (0,\infty)$.
Therefore, by approximation, the estimate in Proposition~\ref{prn:HD07_1} holds on every point on
$X$, even if this point is singular.
\subsection{Approximation functions of distance}
\begin{proposition}[\textbf{Almost super-harmonicity of distance function}] Suppose $x_0 \in X$, $r(x)=d(x,x_0)$. Then we have
\begin{align}
\Delta r^2 \leq 4n
\label{eqn:HC29_1}
\end{align}
in the weak sense. In other words, for every nonnegative $\chi \in N_c^{1,2}(X)$, we have
\begin{align}
-\int_X \left\langle \nabla r^2, \nabla \chi \right\rangle \leq \int_X 4n \chi.
\label{eqn:HD16_3}
\end{align}
\label{prn:HC29_1}
\end{proposition}
\begin{proof}
Let us first assume $x_0 \in \mathcal{R}$.
Clearly, away from the generalized cut locus, we have
$\Delta r^2 \leq 4n$ in the classical sense. Therefore, $\Delta r^2 \leq 4n$ on $\mathcal{R}$ in the distribution sense,
same as the smooth Riemannian manifold case. Since smooth cutoff functions supported on $\mathcal{R}$
are dense in $N_0^{1,2}(X)$, we see that for every $\chi \in N_c^{1,2}(X)$, inequality (\ref{eqn:HD16_3}) holds true.
Now suppose $x_0 \in \mathcal{S}$, we can choose regular points $x_i \to x_0$. Let $r_i=d(x_i, \cdot)$, then for each nonnegative function $\chi \in N_c^{1,2}(X)$, we have
\begin{align*}
-\int_X \left\langle \nabla r_i^2, \nabla \chi \right\rangle \leq \int_X 4n \chi.
\end{align*}
Let $\Omega$ be a bounded open set containing the support of $\chi$.
Then $r_i^2$ weakly converges to a unique limit in $N^{1,2}(\Omega)$,
$r_i^2$ strongly converges to $r^2$ in $L^2(\Omega)$. This means that
$r^2$ is the weak limit of $r_i^2$ in $N^{1,2}(\Omega)$. It follows that
\begin{align*}
-\int_X \left\langle \nabla r^2, \nabla \chi \right\rangle
=-\lim_{i \to \infty} \int_X \left\langle \nabla r_i^2, \nabla \chi \right\rangle
\leq \int_X 4n \chi.
\end{align*}
\end{proof}
In view of Proposition~\ref{prn:HC29_1}, we can obtain many rigidity theorems.
\begin{lemma}[\textbf{Cheeger-Gromoll type splitting}]
Suppose $X$ contains a straight line $\gamma$. Then there is a length space $N$ such that
$X$ is isometric to $N \times \ensuremath{\mathbb{R}}$ as metric space product.
\label{lma:HE04_1}
\end{lemma}
\begin{proof}
The proof is almost the same as the classical one. However, we shall take this as an opportunity to check the analysis tools developed in previous subsections.
Actually, fix $x_0 \in \gamma$, we can divide $\gamma$ into two rays $\gamma^{+}$ and $\gamma^{-}$.
Accordingly, there are Buseman functions $b^{+}$ and $b^{-}$. Proposition~\ref{prn:HC29_1} implies that $\Delta r \leq \frac{2n-1}{r}$ in the weak sense,
which in turn forces both $b^{+}$ and $b^{-}$ to be subharmonic functions. By triangle inequality, we know $b^{+}+b^{-} \geq 0$ globally and
achieve $0$ on $x_0$. It follows from strong maximum principle, by Proposition~\ref{prn:HC29_5}, that $b^{+}+b^{-} \equiv 0$.
Consequently, $b^{+}$ is harmonic. Then Weitzenb\"{o}ck formula implies that in the weak sense, we have
\begin{align*}
0=\frac{1}{2}\Delta |\nabla b^{+}|^2= |Hess_{b^{+}}|^2.
\end{align*}
Since $b^{+}$ is harmonic, it is harmonic on $\mathcal{R}=X \backslash \mathcal{S}$. By standard improving regularity theory of harmonic functions on
smooth manifold, we see that $b^{+}$ is a smooth function on $\mathcal{R}$ satisfying
\begin{align*}
|\nabla b^{+}|^2 \equiv 1, \quad |Hess_{b^{+}}|^2 \equiv 0.
\end{align*}
Up to this step, everything is the same as the classical case.
However, since the regular part $\mathcal{R}$ is not complete, the following argument is slightly different.
On $\mathcal{R}$, since $\mathcal{L}_{\nabla b^{+}} g=2Hess_{b^{+}}=0$, we see that $\nabla b^{+}$ is a Killing field.
The flow generated by $\nabla b^{+}$ preserves metrics, and in particular the volume element. By weak convexity of $\mathcal{R}$ and the essential gap
of volume density between regular and singular points, we see that the flow generated by $\nabla b^{+}$ preserves regularity.
Let $N$ be the level set $b^{+}=0$, $N'=N \cap \mathcal{R}$. Then is clear that
\begin{align*}
\mathcal{R} \cong N' \times \ensuremath{\mathbb{R}}.
\end{align*}
Take metric completion on both sides, we obtain $X \cong N \times \ensuremath{\mathbb{R}}$ as metric space product.
\end{proof}
\begin{lemma}[\textbf{Metric cone rigidity}]
Suppose $x_0 \in X$, $\Omega=B(x_0,1)$. Then the following conditions are equivalent.
\begin{enumerate}
\item Volume ratio same on scales $0.5$ and $1$, i.e., we have
\begin{align}
2^{2n}|B(x_0,0.5)|=|B(x_0,1)|. \label{eqn:HE10_1}
\end{align}
\item $\Omega$ is a volume cone, i.e., for every $0<r_1<r_2<1$, we have
\begin{align}
r_1^{-2n}|B(x_0,r_1)|=r_2^{-2n}|B(x_0,r_2)|. \label{eqn:HE04_1}
\end{align}
\item $\frac{r^2}{2}$ is the unique weak solution of the Poisson equation
\begin{align}
\Delta u=2n, \; \textrm{in} \; \Omega; \quad \left.\left(u-\frac{r^2}{2} \right)\right|_{\partial \Omega}=0.
\label{eqn:HE03_4}
\end{align}
\item $\frac{r^2}{2}$ induces local metric cone structure on $\Omega$. In other words, on $\Omega \backslash \mathcal{S}$, we have
\begin{align*}
Hess_{\frac{r^2}{2}} -g \equiv 0.
\end{align*}
\end{enumerate}
\label{lma:HE04_2}
\end{lemma}
\begin{proof}
\textit{1 $\Rightarrow$ 2:}
First, let us assume $x_0$ is a smooth point. Then the precise Bishop-Gromov volume estimate holds in the polar coordinate at $x_0$.
Let $A(r)$ be the normalized ``area" ratio of geodesic sphere $\partial B(x_0, r)$, i.e.,
\begin{align*}
A(r) \triangleq \frac{|\partial B(x_0,r)|}{r^{2n-1}},
\end{align*}
then
$A(r)$ is a non-increasing function due to the Bishop volume comparison.
Note that
\begin{align*}
\frac{d}{dr} \left( \frac{|B(x_0,r)|}{r^{2n}}\right)=\frac{|\partial B(x_0,r)|}{r^{2n}} -\frac{2n}{r} \frac{|B(x_0,r)|}{r^{2n}}
=\frac{2n}{r} \left\{\frac{A(r)}{2n} -\frac{|B(x_0,r)|}{r^{2n}} \right\} \geq 0.
\end{align*}
So (\ref{eqn:HE10_1}) implies that
\begin{align*}
\frac{A(r)}{2n}-\frac{|B(x_0,r)|}{r^{2n}} \equiv 0, \quad \forall \; r \in (0.5, 1).
\end{align*}
In particular, we have
\begin{align*}
|B(x_0,1)|=\frac{A(1)}{2n}=\int_0^{1} A(1) r^{2n-1}dr.
\end{align*}
On the other hand, calculation of volume in polar coordinates yields that
\begin{align*}
|B(x_0,1)|=\int_0^{1} A(r) r^{2n-1} dr.
\end{align*}
So we have
\begin{align*}
\int_0^{1} (A(r)-A(1)) r^{2n-1} dr=0.
\end{align*}
Note that $A(r)$ is a non-increasing function. So the above equality means that
\begin{align*}
A(1) \equiv A(r) \equiv \lim_{r \to 0} A(r)= 2n \omega_{2n}.
\end{align*}
It follows that $|B(x_0,r)| \equiv \omega_{2n}r^{2n}$ for every $0<r<1$. In particular, $B(x_0,1)$ is a volume cone.
Actually, in this case, we can proceed to show that $B(x_0,1)$ is isometric to the unit ball in Euclidean space $\ensuremath{\mathbb{C}}^n$.
It is clear that to obtain volume cone, it suffices to have the Bishop-Gromov ``area" ratio monotonicity.
Suppose $x_0$ is singular, if we regard $A(r)r^{2n-1}$ as the derivative of volume of $B(x_0,r)$,
we still have the Bishop-Gromov ``area" ratio monotonicity, by approximating of $x_0$ with smooth points.
We leave the details to the interested readers.
\textit{2 $\Rightarrow$ 3:} Suppose $u$ is the unique solution of the Poisson equation (\ref{eqn:HE03_4}), we need to show that
$u\equiv \frac{r^2}{2}$. By uniqueness of weak solutions, it suffices to show that $\frac{r^2}{2}-u$ is harmonic on $\Omega$, i.e., for every
$\varphi \in N_c^{1,2}(\Omega)$, we have $\int_{\Omega} \varphi \Delta \left( \frac{r^2}{2}-u\right)=0$.
By rescaling, we can also assume $0 \leq \varphi \leq 1$.
Fix such a $\varphi$, we can choose $\epsilon$ small such that the support of $\varphi$ is contained in $B(x_0, 1-\epsilon)$. Define
\begin{align*}
\eta(x)=
\begin{cases}
1, & \textrm{if} \quad d(x,x_0)<1-\epsilon, \\
\frac{1-d(x,x_0)}{\epsilon}, & \textrm{if} \quad 1-\epsilon \leq d(x,x_1) \leq 1.
\end{cases}
\end{align*}
Note that $\eta \in N_0^{1,2}(\Omega)$, $\frac{r^2}{2}-u$ is superharmonic on $\Omega$. It follows from integration by parts that
\begin{align*}
\int_X \eta \Delta\left( \frac{r^2}{2} -u\right) =-2n\int_{\Omega} \eta +\frac{1}{\epsilon} \int_{B(x_0,1) \backslash B(x_0,1-\epsilon)} r \geq -O(\epsilon),
\end{align*}
where we used volume cone condition in the last step.
Thus, we have
\begin{align*}
0 \geq \int_{\Omega} \varphi \Delta \left( \frac{r^2}{2}-u\right)
=\int_{\Omega} \eta \Delta \left( \frac{r^2}{2}-u\right)+\int_{\Omega} (\varphi-\eta) \Delta \left( \frac{r^2}{2}-u\right)
\geq \int_{\Omega} \eta \Delta \left( \frac{r^2}{2}-u\right) \geq -O(\epsilon).
\end{align*}
Let $\epsilon \to 0$, we obtain $\int_{\Omega} \varphi \Delta \left( \frac{r^2}{2}-u\right)=0$. Consequently, $\frac{r^2}{2}-u$ is harmonic by the arbitrary choice
of $\varphi$.
\textit{3 $\Rightarrow$ 4:}
Since $\frac{r^2}{2}$ solves the Poisson equation with right hand side a constant, by standard bootstrapping argument for elliptic equation,
we see that $\frac{r^2}{2}$ is a smooth function on $\Omega \backslash \mathcal{S}$.
Clearly, we have $\left|\nabla \frac{r^2}{2}\right|^2=\frac{r^2}{2}$. Taking Laplacian on both sides, Weitzenb\"{o}ck formula yields that
$\displaystyle \left|Hess_{\frac{r^2}{2}} \right|^2=2n$, which in turn implies that
\begin{align*}
\left|Hess_{\frac{r^2}{2}} -g\right|^2= \left|Hess_{\frac{r^2}{2}} \right|^2-2\Delta \frac{r^2}{2}+2n
= \left|Hess_{\frac{r^2}{2}} \right|^2-2n=0.
\end{align*}
Therefore, on $\Omega \backslash \mathcal{S}$, we have $Hess_{\frac{r^2}{2}} -g\equiv 0$ in the classical sense.
Consequently, $\nabla \frac{r^2}{2}$ is a conformal Killing field. Then similar to the argument at the end of the proof of Lemma~\ref{lma:HE04_1},
we see that the flow generated by $\nabla \frac{r^2}{2}$ preserves regularity. Hence it is clear that $\Omega \backslash \mathcal{S}$ has a local
metric cone structure, whose completion implies that $\overline{\Omega}$ is a unit ball in a metric cone.
\textit{4 $\Rightarrow$ 1:}
For each $0<r<1$, note that $|B(x_0,r)|=|B(x_0,r) \backslash \mathcal{S}|$. Note the flow generated by $\nabla \frac{r^2}{2}$ preserves regularity.
More precisely, we have
\begin{align*}
\mathcal{L}_{\nabla \frac{r^2}{2}} g=g, \quad \mathcal{L}_{\nabla \frac{r^2}{2}} d\mu= 2n d\mu.
\end{align*}
Then (\ref{eqn:HE04_1}) follows from the integration of the above equation along flow lines.
\end{proof}
\begin{lemma}[\textbf{K\"ahler cone splitting}]
Suppose $X \in \widetilde{\mathscr{KS}}^*(n,\kappa)$ is a metric cone with vertex $x_0$. Then we can find a metric cone $C(Z)$ with vertex $z^*$ such that
\begin{align*}
X= \ensuremath{\mathbb{C}}^{n-k} \times C(Z), \quad x_0=(0, z^*), \quad 2 \leq k \leq n.
\end{align*}
Moreover, there is no straight line in $C(Z)$ passing through $z^*$.
\label{lma:HD30_1}
\end{lemma}
\begin{proof}
It suffices to show that if $X$ splits off a real straight line $\ensuremath{\mathbb{R}}$, then it splits off a complex line $\ensuremath{\mathbb{C}}$.
In fact, if there is a straight line passing through $x_0$, we can find a function $h$ which is the Buseman function determined by the line.
Therefore, $\nabla h$ is a parallel vector field with $|\nabla h| \equiv 1$. The K\"ahler condition implies that
$J \nabla h$ is another parallel vector field satisfying $|J \nabla h| \equiv 1$ on $\mathcal{R}=X \backslash \mathcal{S}$.
On the regular set, define function
\begin{align*}
u=\left\langle J\nabla h, \nabla \frac{r^2}{2} \right\rangle,
\end{align*}
where $r$ is the distance to the vertex $x_0$.
Metric cone condition implies that $Hess_{\frac{r^2}{2}}=g$. Since $J\nabla h$ is parallel, we see that
\begin{align*}
\nabla u= Hess_{\frac{r^2}{2}} (J\nabla h, \cdot)=J \nabla h.
\end{align*}
Recall that $Hess_h \equiv 0$. Taking gradient of the above equation implies that $Hess_u \equiv 0$. This forces that $\nabla u=J \nabla h$
is also a splitting direction. Note that although $u$ is only defined on $\mathcal{R}$, which is not complete.
However, we can bypass this difficulty as done in the proof of Lemma~\ref{lma:HE04_1}, since $\nabla u$ is a Killing field preserving regularity.
Therefore, we obtain a splitting factor $\ensuremath{\mathbb{C}}$. Since $J\nabla u=-\nabla h$, the space spanned by $\nabla u$ and
$J \nabla u$ is closed under the $J$-action. This induces the $J$-action closedness of the split linear space, which then must be $\ensuremath{\mathbb{C}}^{n-k}$
for some integer $k$. Because $X$ is not $\ensuremath{\mathbb{C}}^n$, we know the singular set is not empty, whose dimension restriction forces that $k \geq 2$.
\end{proof}
For each rigidity property in Lemma~\ref{lma:HE04_1}-Lemma~\ref{lma:HD30_1}, there should exist an ``almost" version.
For example, Lemma~\ref{lma:HE04_2} basically says that a volume cone implies a metric cone.
Hence the ``almost" version is that for a unit geodesic ball $B(x_0,1)$ whose volume ratio function $r^{-2n}|B(x_0,r)|$ is very close to a constant
function on $[0,1]$, then after proper rescaling, each ball $B(x_0, r)$ is very close to $B(x_0,1)$ in the Gromov-Hausdorff topology.
The basic idea is expressed clearly in~\cite{CCWarp}.
We only interpret what they did. Actually, if volume ratio is almost a constant, then it is expected that $|Hess_{\frac{r^2}{2}}-g|$ has a small
$L^2$-norm. However, since the regularity of distance function $r$ is bad, one should replace $\frac{r^2}{2}$ by an approximation function, which is
very close to $\frac{r^2}{2}$ in $N^{1,2}$-norm on one hand, and has excellent regularity on the other hand.
Such approximation function is nothing but the solution of the Poisson equation (\ref{eqn:HE03_4}).
For the purpose of developing ``almost" rigidity properties, one need some technical preparation, which will be listed as Lemmas.
Note that the space $\widetilde{\mathscr{KS}}(n,\kappa)$ has scaling invariance.
Therefore, we can always let the scale we are interested in to be $1$, to simplify the notations.
In view of Proposition~\ref{prn:HC29_1}, we can define many auxiliary radial functions, as in the classical case for Riemannian manifold(c.f.~\cite{CCWarp}).
For each $0<r<R<\infty$, define
\begin{align}
&\underline{U}(r) \triangleq \frac{r^2}{4n},
\quad
\underline{G}(r) \triangleq \frac{r^{2-2n}}{2n(2n-2)\omega_{2n}}, \\
&\underline{U}_R \triangleq \frac{r^2-R^2}{4n}, \quad \underline{G}_R \triangleq \frac{r^{2-2n}-R^{2-2n}}{2n(2n-2)\omega_{2n}}, \\
&\underline{L}_R \triangleq \frac{r^{2-2n}R^{2n}-R^{2}}{2n(2n-2)} + \frac{r^2-R^2}{4n}.
\label{eqn:HE08_1}
\end{align}
Then by Proposition~\ref{prn:HC29_1} and direct calculation, we have the following lemma.
\begin{lemma}[\textbf{Existence of good radial comparison functions}]
Suppose $x_0 \in X$. Let $r(x)=d(x,x_0)$ and
define $\underline{U}_1(x)=\underline{U}_1(r(x))$, $\underline{G}_1=\underline{G}_1(r(x))$ and $\underline{L}_1(x)=\underline{L}(r(x))$
as done in (\ref{eqn:HE08_1}). Then we have
\begin{align*}
&\Delta \underline{U}_1 \leq 1, \textrm{on} \; X; \quad \underline{U}_1|_{\partial B(x_0,1)}=0. \\
&\Delta \underline{G}_1 \geq 0, \textrm{on} \; B(x_0,1) \backslash \{x_0\}; \quad \underline{G}_1 |_{\partial B(x_0,1)}=0. \\
&\Delta \underline{L}_1 \geq 0, \textrm{on} \; B(x_0,1) \backslash \{x_0\}; \quad \underline{L}_1 |_{\partial B(x_0,1)}=0.
\end{align*}
\label{lma:HE08_1}
\end{lemma}
Lemma~\ref{lma:HE08_1} is used to improve the maximum principle. Same as that done by Abresch-Gromoll (c.f. Proposition 2.3 of~\cite{AbGr}), we
obtain the following estimate of excess function.
\begin{lemma}[\textbf{Abresch-Gromoll type estimate}]
Suppose $x_0 \in X$,
$\gamma$ is a line segment centered at $x_0$ with length $2$, end points $p_{+}$ and $p_{-}$.
Let $e(x)$ be the excess function $d(x,p_{+})+d(x,p_{-})-2$.
Then we have
\begin{align}
\sup_{x \in B(x_0,\epsilon)} e(x) \leq C \epsilon^{\frac{2n}{2n-1}} \label{eqn:HE08_2}
\end{align}
for each $\epsilon \in (0, 1)$ and some universal constant $C=C(n)$.
\label{lma:HE05_1}
\end{lemma}
Lemma~\ref{lma:HE08_1} can also be applied to construct good cutoff functions.
\begin{lemma}[\textbf{Cutoff functions on annulus}]
Suppose $x_0 \in X$, $0<\rho<1<\infty$.
Then there exists a function $\phi: X \to [0,1]$ such that
\begin{align*}
&\phi \in C^{\infty}(B(x_0,1) \backslash \mathcal{S}), \quad \supp \phi \Subset B(x_0, 1), \quad \phi \equiv 1 \; \textrm{on} \; B(x_0,\rho), \\
&|\nabla \phi| \leq c(n, \rho), \qquad \quad |\Delta \phi| \leq c(n, \rho), \quad \textrm{on} \; B(x_0,1) \backslash \mathcal{S}.
\end{align*}
Furthermore, for each pair $\rho_1, \rho_2$ satisfying $0<\rho_1<\frac{1}{2}<2<\rho_2<\infty$, there exists a function $\phi: X \times [0,1]$ such that
\begin{align*}
&\phi \in C^{\infty}((B(x_0,\rho_2) \backslash \overline{B(x_0, \rho_1 )}) \cap \mathcal{R}),
\quad \supp \phi \Subset B(x_0, \rho_2) \backslash \overline{B(x_0,\rho_1)}, \\
&\phi \equiv 1 \; \textrm{on} \; B\left(x_0,\frac{\rho_2}{2} \right) \backslash \overline{B(x_0, 2\rho_1)}, \\
&|\nabla \phi| \leq c(n, \rho_1,\rho_2), \qquad \quad |\Delta \phi| \leq c(n, \rho_1,\rho_2),
\quad \textrm{on} \; (B(x_0,\rho_2) \backslash \overline{B(x_0, \rho_1 )}) \cap \mathcal{R}.
\end{align*}
\label{lma:HE04_3}
\end{lemma}
The proof of Lemma~\ref{lma:HE04_3} is based on the maximum principle, solvability of Poinsson equation and the fact that $\Delta \underline{L}_1 \geq 1$
and $\Delta \underline{U}_{R'} \geq 1$ for each $R'>0$.
With these properties, one can compare $\underline{L}_1$ with the Poinsson equation solution $f$ which has same boundary value as $\underline{L}_1$.
Then construct cutoff function based on the value of $f$. Since the proof follows that of~\cite{CCWarp} verbatim, we omit the details here.
\begin{lemma}[\textbf{Harmonic approximation of local Buseman function}]
There exists a constant $c=c(n)$ with the following properties.
Suppose $x_0 \in X$,
$\gamma$ is a line segment centered at $x_0$ with length $2$, end points $p_{+}$ and $p_{-}$, $\epsilon$ is an arbitrary small positive number,
say $0<\epsilon<0.1$.
In the ball $B(x_0,4\epsilon)$, define local Buseman functions
\begin{align*}
b_{+}(x)=d(x, p_{+})-d(x_0, p_{+}), \quad b_{-}(x)=d(x,p_{-})-d(x_0, p_{-}).
\end{align*}
Let $u_{\pm}$ be the harmonic functions in $B(x_0,4\epsilon)$ such that $\left.\left( u_{\pm}-b_{\pm} \right) \right|_{\partial B(x_0,4\epsilon)}=0$.
Let $u$ be one of $u_{\pm}$ and $b$ be the corresponding $b_{\pm}$ respectively. Then we have
\begin{itemize}
\item $|u-b| \leq c\epsilon^{1+\alpha}$.
\item $\fint_{B(x,\epsilon)} |\nabla (u-b)|^2 \leq c\epsilon^{\alpha}$.
\item $\fint_{B(x,\epsilon)} |Hess_{u}|^2 < c\epsilon^{-2+\alpha}$.
\end{itemize}
Here $\alpha=\alpha(n)$ is a universal constant, which can be chosen as $\frac{1}{2n-1}$.
\label{lma:HE05_2}
\end{lemma}
\begin{proof}
For simplicity, we assume $u=u_{+}$ and $b=b_{+}$.
The pointwise estimate of $|u-b|$ follows from maximum principle and the the excess estimate (\ref{eqn:HE08_2}), same as traditional case.
We proceed to show the integral estimate of $|\nabla (u-b)|$.
Note that $b(x)=r(x)-d(x_0,p_{+})$, where $r(x)=d(x, p_{+})$. It follows from rescaling that
\begin{align*}
\fint_{B(x_0,4\epsilon)} |\Delta b| = \fint_{B(x_0,4\epsilon)} |\Delta r| < C \epsilon^{-1}.
\end{align*}
Actually, by the fact $\Delta r \leq \frac{2n-1}{r}$, the estimate of $\int |\Delta r|$ is reduced to the estimate of
$\int \Delta r$. However, $\int \Delta r$ can be bounded by integration by parts,
modulo some technical discussion around the generalized cut locus and singular set $\mathcal{S}$. Due to the high
codimension of $\mathcal{S}$, the integral of $\Delta r$ around of $\mathcal{S}$ can be ignored. Then we return
to the smooth manifold case, which is discussed clearly in~\cite{Cheeger01}.
Clearly, $u-b \in N_0^{1,2}(B(x_0,4\epsilon))$. Hence integration by parts, Proposition~\ref{prn:HD13_2},
applies and we have
\begin{align*}
\fint_{B(x_0,4\epsilon)} |\nabla (u-b)|^2 &= \fint_{B(x_0,4\epsilon)} (u-b) \Delta (b-u)= \fint_{B(x_0,4\epsilon)} (u-b) \Delta b
\leq C\epsilon^{1+\alpha} \fint_{B(x_0,4\epsilon)} |\Delta b| < C\epsilon^{\alpha}.
\end{align*}
Note that $u$ is harmonic in $B(x_0,4\epsilon)$. Weitzenb\"{o}ck formula implies that
\begin{align*}
\frac{1}{2} \Delta \left(|\nabla u|^2-1\right)=\frac{1}{2} \Delta |\nabla u|^2=|Hess_{u}|^2 \geq 0
\end{align*}
in the classical sense on $B(x_0, 4\epsilon) \backslash \mathcal{S}$. By extension property of subharmonic function, Proposition~\ref{prn:HD16_2}, we see that
$|\nabla u|^2 \in N_{loc}^2(B(x_0, 4\epsilon))$. Let $\phi$ be a cutoff function vanishes on $\partial B(x_0, 4\epsilon)$ and equivalent to $1$ on $B(x_0,\epsilon)$,
with $\epsilon |\nabla \phi|$
and $\epsilon^2 |\Delta \phi|$ bounded as in Lemma~\ref{lma:HE04_3}. Clearly, $\phi \in N_c^{1,2}(B(x_0, 4\epsilon))$.
Therefore, it follows from integration by parts, Proposition~\ref{prn:HD13_2}, that
\begin{align*}
2\fint_{B(x_0, 4\epsilon)} \phi |Hess_u|^2=\fint_{B(x_0, 4\epsilon)} \phi \Delta \left(|\nabla u|^2-1\right)=\fint_{B(x_0, 4\epsilon)} \left(|\nabla u|^2-1 \right) \Delta \phi.
\end{align*}
Consequently, we obtain
\begin{align*}
\fint_{B(x_0,\epsilon)} |Hess_{u}|^2 \leq \fint_{B(x_0, 4\epsilon)} \phi |Hess_u|^2 \leq C \epsilon^{-2} \fint_{B(x_0, 4\epsilon)} \left||\nabla u|^2-1\right| \leq C\epsilon^{\alpha-2}.
\end{align*}
\end{proof}
Note that Lemma~\ref{lma:HE05_2} implies almost splitting property already. Therefore, it is generalization of Lemma~\ref{lma:HE04_1}, the splitting property.
Not surprisingly, one can use Lemma~\ref{lma:HE05_2} to prove Lemma~\ref{lma:HE04_1}, at least formally.
Actually, if there is a line with length $2L$ centered at $x_0$, then in the unit ball $B(x_0,1)$, it follows from Lemma~\ref{lma:HE05_2} that
\begin{align*}
|u-b|<cL^{-\alpha}, \quad \fint_{B(x_0,1)} |\nabla(u-b)|<cL^{-\alpha}, \quad \fint_{B(x_0,1)} |Hess_u|^2<cL^{-\alpha}.
\end{align*}
Let $L \to \infty$, we see that $Hess_{u} \equiv 0$ on $\mathcal{R}$.
From the proof of Lemma~\ref{lma:HE05_2}, it is clear that the key to obtain smallness of $|Hess_u|^2$ is the integration by parts, which is checked in our case.
For smooth Riemannian manifold, the approximation in Lemma~\ref{lma:HE05_2} was improved by Colding and Naber in~\cite{ColdNa}.
The essential difference is that they chose parabolic approximation functions, instead of harmonic approximations.
Suppose $\gamma$ is a line segment with length $2$, centered around $x_0$, with end points $p_{+}$ and $p_{-}$.
Then one can construct cutoff functions $\psi$ such that it vanishes outside
$B(x_0, 8)$ and inside $B(p_{+}, 0.1)$ and $B(p_{-}, 0.1)$, and equals $1$ on $B(x_0,4) \backslash (B(p_{+}, 0.2) \cup B(p_{-}, 0.2))$. Moreover, we have pointwise
bound of $|\Delta \psi|$ and $|\nabla \psi|$. Then for $b_{\pm}$, we can run heat flow starting from $\psi b_{\pm}$ to obtain solution $h_{t,\pm}$.
Then the function $h_{t,\pm}$ is a better approximation function of $b_{\pm}$ on the scale around $\sqrt{t}$.
The extra technical tools needed for Colding-Naber's argument beyond the harmonic approximation consists of an a priori bound of heat kernel, and the construction of
cutoff function with the properties as mentioned above. However, in light of Proposition~\ref{prn:HD07_1} and Lemma~\ref{lma:HE04_3}, both tools are available in our setting.
Therefore, we can parallelly move the parabolic approximation estimate, Theorem 2.19 of~\cite{ColdNa}, to our case.
\begin{lemma}[\textbf{Parabolic approximation of local Buseman function}]
There exist two constants $c=c(n), \bar{\epsilon}=\bar{\epsilon}(n)$ with the following properties.
Suppose $x_0 \in X$, $\gamma$ is a line segment whose center point locates in $B(x_0,0.2)$, with end points $p_{+}$ and $p_{-}$, with length $2$.
Let $h_{t}$ be the heat approximation of $b$ which is one of $b_{\pm}$.
Suppose the excess value $d(x_0,p_{+})+d(x_0,p_{-})-2<\epsilon^2$ for some $\epsilon \in (0,\bar{\epsilon})$.
Then there exists $\lambda \in [0.5, 2]$ such that
\begin{itemize}
\item $|h_{\lambda \epsilon^2} - b| \leq c \epsilon^2$.
\item $\fint_{B(x,\epsilon)} ||\nabla h_{\lambda \epsilon^2}|^2-1| \leq c\epsilon$.
\item $\int_{0.1}^{1.9} \fint_{B(x,\epsilon)} ||\nabla h_{\lambda \epsilon^2}|^2-1| \leq c\epsilon^2$.
\end{itemize}
Most importantly, we have
\begin{align*}
\int_{0.1}^{1.9} \fint_{B(\gamma(s), \epsilon)} |Hess_{h_{\lambda \epsilon^2}}|^2 \leq c.
\end{align*}
\label{lma:HE04_4}
\end{lemma}
Note that we did not formulate the parabolic approximation in the most precise way. For example, $\gamma$ need not to be a geodesic, an $\epsilon$-geodesic
suffices. Interested readers are referred to \cite{ColdNa} for the most general version.
According to the discussion form Lemma~\ref{lma:HE04_3} to Lemma~\ref{lma:HE04_4}, it is quite clear that the integral estimate of approximation functions
can be obtained in the same way as the Riemannian manifold case, provided the following properties.
\begin{itemize}
\item Almost super-harmonicity of distance functions, Proposition~\ref{prn:HC29_1}.
\item Bishop-Gromov volume comparison, Proposition~\ref{prn:HD19_1}.
\item Strong maximum principle for subharmonic functions, Proposition~\ref{prn:HC29_5}.
\item Integration by parts, Proposition~\ref{prn:HD13_2}.
\item Existence of excellent cutoff function, Lemma~\ref{lma:HE04_3}.
\end{itemize}
Since all of these properties are checked in our situation, we can follow the route of Cheeger-Colding to obtain the following properties, almost line by line.
\begin{lemma}[\textbf{Approximation slices}]
Suppose $x_0 \in X$, $\gamma_1, \gamma_2, \cdots, \gamma_k$ are $k$ line segments with length $2L>>2$ such that the center point of
$\gamma_k$ locates in $B(x_0,1)$ for each $k$. Suppose the end points of $\gamma_k$ are $p_{i,+}$ and $p_{i,-}$.
Let $b_{i,\pm}$ be the corresponding local Buseman functions with respect to $\gamma_i$. Let $u_i$ be the harmonic function
on $B(x_i,4)$ with the same value as $b_{i,+}$ on $\partial B(x_0,4)$. Then we have
\begin{align*}
\int_{B(x_0,1)} \left\{ \sum_{1 \leq i \leq k} |\nabla u_i -1|^2 + \sum_{1 \leq i<j\leq k}|\langle \nabla u_i, \nabla u_j\rangle| + \sum_{1 \leq i \leq k} |Hess_{u_i}|^2\right\}
\leq CL^{-\alpha},
\end{align*}
where $\alpha=\alpha(n)>0$.
\label{lma:HE07_1}
\end{lemma}
Let $\vec{u}=(u_1, u_2, \cdots, u_k)$, we can regard $\vec{u}$ as an almost submersion from $B(x_0,1)$ to its image on
$\ensuremath{\mathbb{R}}^k$. Consequently, slice argument can be set up as that in~\cite{CCT}. The slice argument together with the Chern-Simons theory can improve the behavior of the singular set $\mathcal{S}$. A more fundamental application of the slice argument is to set up the following volume convergence property, as done in~\cite{ColdVol}.
\begin{proposition}[\textbf{Volume continuity}]
For every $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$ and $\epsilon>0$, there is a constant $\xi=\xi(X,\epsilon)$ such that
\begin{align*}
\left| \log \frac{|B(y_0,1)|}{|B(x_0,1)|} \right|<\epsilon
\end{align*}
for any $(Y,y_0,h) \in \widetilde{\mathscr{KS}}(n,\kappa)$ satisfying $d_{GH}((X,x_0,g), (Y,y_0,h))<\xi$.
\label{prn:HD22_1}
\end{proposition}
Apply the same argument as in~\cite{CCWarp}, we obtain the rigidity of almost volume cones.
\begin{proposition}[\textbf{Almost volume cone implies almost metric cone}]
For each $\epsilon>0$, there exists $\xi=\xi(n,\epsilon)$ with the following properties.
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$ satisfies $\displaystyle \frac{|B(x_0,2)|}{ |B(x_0,1)|} \geq (1-\epsilon)2^{2n}$,
then there exists a metric cone over a length space $Z$, with vertex $z^*$ such that
\begin{align*}
\diam(Z) < \pi + \xi, \quad
d_{GH} \left( B(x_0,1), B(z^*, 1)\right) <\xi.
\end{align*}
\label{prn:HD22_2}
\end{proposition}
Similar to Lemma 9.14 of~\cite{CCT}, we obtain the almost K\"ahler cone splitting, based on Proposition~\ref{prn:HD22_2}.
\begin{proposition}[\textbf{Almost K\"ahler cone splitting}]
For each $\epsilon>0$, there exists $\xi=\xi(n,\epsilon)$ with the following properties.
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x_0 \in X$, $b$ is a Lipschitz function on $B(x_0,2)$ satisfying
\begin{align*}
\sup_{B(x_0,2) \backslash \mathcal{S}} |\nabla b| \leq 2, \quad \fint_{B(x_0,2) \backslash \mathcal{S}} |Hess_b|^2 \leq \epsilon^2.
\end{align*}
Suppose also $\displaystyle \frac{|B(x_0,2)|}{|B(x_0,1)|} \geq (1-\epsilon) 2^{2n}$, i.e., $B(x_0,1)$ is an almost volume cone.
Then there exists a Lipschitz function $\tilde{b}$ on $B(x_0,1)$ such that
\begin{align*}
\sup_{B(x_0,1) \backslash \mathcal{S}} \left|\nabla \tilde{b}\right| \leq 3,
\quad \fint_{B(x_0,1) \backslash \mathcal{S}} \left|\nabla \tilde{b}-J \nabla b\right|^2 \leq \xi.
\end{align*}
\label{prn:HE08_1}
\end{proposition}
\subsection{Volume radius}
Anderson's gap theorem implies that one can improve regularity of the very interior part of a geodesic ball whenever the volume ratio of the geodesic ball is very close
to the Euclidean one. This suggests us to define the volume radius as follows.
\begin{definition}
Let $\delta_0$ be the Anderson constant.
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x_0 \in X$.Then we define
\begin{align*}
&\Omega_{x_0} \triangleq \left\{ r| r>0, r^{-2n}|B(x_0,r)|\geq (1-\delta_0) \omega_{2n} \right\}.\\
&\mathbf{vr}(x_0) \triangleq
\begin{cases}
\sup \Omega_{x_0}, &\textrm{if} \; \Omega_{x_0} \neq \emptyset, \\
0, &\textrm{if} \; \Omega_{x_0}=\emptyset.
\end{cases}
\end{align*}
We call $\mathbf{vr}(x_0)$ the volume radius of the point $x_0$.
\label{dfn:SC17_1}
\end{definition}
According to this definition, a point is regular if and only if its volume radius is positive.
On the other hand, if the space is not $\ensuremath{\mathbb{C}}^n$, then every point has a finite volume radius by a generalized Anderson's gap theorem.
\begin{proposition}[\textbf{Euclidean space by $\mathbf{vr}$}]
Suppose $X \in \widetilde{\mathscr{KS}}(n)$ and $\mathbf{vr}(x_0)=\infty$ for some $x_0 \in X$, then $X$ is isometric to the Euclidean space $\ensuremath{\mathbb{C}}^{n}$.
\label{prn:SC17_1}
\end{proposition}
\begin{proof}
Fix an arbitrary point $x \in X$, then volume comparison implies that
\begin{align*}
\mathrm{v}(x) \geq \lim_{r \to \infty} \omega_{2n}^{-1} r^{-2n} |B(x,r)|=\mathrm{avr}(X) \geq 1-\delta_0.
\end{align*}
Therefore, $x$ is a regular point. Since $x$ is arbitrarily chosen, we see that $X \in \mathscr{KS}(n)$. Then the statement follows from Anderson's gap theorem.
\end{proof}
A local version of Proposition~\ref{prn:SC17_1} is the following local Harnack inequality of $\mathbf{vr}$.
\begin{proposition}[\textbf{Local Harnack inequality of volume radius}]
There is a constant $\tilde{K}=\tilde{K}(n)$ with the following properties.
Suppose $x \in X \in \widetilde{\mathscr{KS}}^{*}(n)$, $r=\mathbf{vr}(x)>0$, then we have
\begin{align}
\tilde{K}^{-1}r \leq \mathbf{vr} \leq \tilde{K}r
\label{eqn:SC26_8}
\end{align}
in the ball $B(x, \tilde{K}^{-1}r)$. Moreover, for every $\rho \in (0, \tilde{K}^{-1}r)$, $y \in B(x, \tilde{K}^{-1}r)$, we have
\begin{align}
&\omega_{2n}^{-1}\rho^{-2n} |B(y,\rho)| \geq 1-\frac{\delta_0}{100}, \label{eqn:SC26_9}\\
& |Rm|(y) \leq \tilde{K}^{2} r^{-2}, \label{eqn:SC27_3}\\
& inj(y) \geq \tilde{K}^{-1} r. \label{eqn:SC27_4}
\end{align}
\label{prn:SC26_7}
\end{proposition}
\begin{proof}
It follows from Bishop volume comparison, Anderson's gap theorem and a compactness argument.
\end{proof}
On a Ricci-flat geodesic ball, it is well known that $|Rm|$ bound implies
bound of $|\nabla^k Rm|$ for each positive integer $k$ in a smaller geodesic ball.
So (\ref{eqn:SC27_3}) immediately yields the following corollary.
\begin{corollary}[\textbf{Improving regularity property of volume radius}]
There is a small positive constant $c_a=c_a(n)$ with the following properties.
Suppose $x \in X \in \widetilde{\mathscr{KS}}^{*}(n)$, $\mathbf{vr}(x) \geq r>0$, then we have
\begin{align}
r^{2+k}|\nabla^k Rm|(y) \leq c_a^{-2}, \quad \forall \; y \in B(x,c_a r), \quad 0 \leq k \leq 5.
\label{eqn:SL23_8}
\end{align}
\label{cly:SL23_1}
\end{corollary}
In Ricci bounded geometry, harmonic radius (c.f.~\cite{An90}) plays an important role. A point $x$ is defined to have harmonic radius at least $r$ if on the smooth geodesic ball $B(x,r)$, there exists a harmonic diffeomorphism
$\Psi=(u_1, u_2, \cdots, u_{2n}): B(x,r) \to \Omega \subset \ensuremath{\mathbb{R}}^{2n}$ such that
\begin{align*}
\frac{1}{2}\delta_{ij} \leq g_{ij}=g(\nabla u_i, \nabla u_j) \leq 2 \delta_{ij}, \quad
r^{\frac{3}{2}}\norm{g_{ij}}{C^{1,\frac{1}{2}}} \leq 2.
\end{align*}
Then harmonic radius is defined as the supreme of all the possible $r$'s mentioned above. For convenience, we use
$\mathbf{hr}$ to denote harmonic radius. This definition can be easily moved to our case when the underlying space
is in $\widetilde{\mathscr{KS}}^*(n)$. We define $\mathbf{hr}$ to be $0$ on the singular part of the underlying space.
It is clear from the definitions and Proposition~\ref{prn:SC26_7} that volume radius and harmonic radius can bound each other, i.e., they are equivalent. The following Proposition is obvious.
\begin{proposition}[\textbf{Equivalence of volume and harmonic radius}]
Suppose $x \in X \in \widetilde{\mathscr{KS}}^{*}(n)$, then we have
\begin{align*}
\frac{1}{C} \mathbf{hr}(x) \leq \mathbf{vr}(x) \leq C \mathbf{hr}(x)
\end{align*}
for some uniform constant $C=C(n)$.
\label{prn:HA09_1}
\end{proposition}
Note that the regularity requirement of the underlying space to define volume radius is much weaker than that to define harmonic radius a priori. Therefore, Proposition~\ref{prn:HA09_1} already implies a regularity improvement.
We shall set up the compactness theory based on volume radius,
since volume radius may be applicable to more general metric measure spaces.
Let $X \in \widetilde{\mathscr{KS}}^*(n)$ and decompose it as $X=\mathcal{R} \cup \mathcal{S}$. Then $\mathbf{vr}$ is a positive finite function on $\mathcal{R}$ and equals $0$ on $\mathcal{S}$.
\begin{proposition}[\textbf{Rigidity of volume ratio}]
Suppose $X \in \widetilde{\mathscr{KS}}(n)$. If for two concentric geodesic balls $B(x_0,r_1) \subset B(x_0,r_2)$
centered at a regular point $x_0$, we have
\begin{align}
\omega_{2n}^{-1}r_1^{-2n}|B(x_0,r_1)|= \omega_{2n}^{-1}r_2^{-2n}|B(x_0,r_2)|,
\label{eqn:SL23_1}
\end{align}
then the ball $B(x_0,r_2)$ is isometric to a geodesic ball or radius $r_2$ in $\ensuremath{\mathbb{C}}^n$.
Furthermore, if $X \in \mathscr{KS}(n)$, then we can further conclude that $X$ is Euclidean.
\label{prn:SC17_3}
\end{proposition}
\begin{proof}
From the proof of Lemma~\ref{lma:HE04_2}, it is clear that $B(x_0,r_2)$ is a volume cone with constant volume
ratio $\omega_{2n}$.
Observe the change of volume element along each smooth geodesic emanating from $x_0$, in the polar coordinate.
By the volume density gap between regular and singular points,
the optimal volume ratio of $B(x_0,1)$ forces that it does not contain any singular point.
Then the situation is the same as the smooth Riemannian case. Clearly, a smooth Ricci-flat geodesic ball with volume
ratio $\omega_{2n}$ is isometric to a Euclidean ball of the same radius.
If $X\in \mathscr{KS}(n)$, by analyticity of metric tensor, it is clear that $X$ is flat and hence $\ensuremath{\mathbb{C}}^n$ due to its non-collapsing property at infinity.
\end{proof}
\begin{proposition}[\textbf{Continuity of volume radius}]
$\mathbf{vr}$ is a continuous function on $X$ whenever $X \in \widetilde{\mathscr{KS}}(n)$.
\label{prn:HE07_2}
\end{proposition}
\begin{proof}
Since $\mathbf{vr} \equiv \infty$ on $\ensuremath{\mathbb{C}}^n$, which is obvious continuous. So we can assume $X \in \widetilde{\mathscr{KS}}^{*}(n)$ without loss of generality.
By Proposition~\ref{prn:SC17_1}, we know $\mathbf{vr}$ is a finite function on $X$.
So we assume $\mathbf{vr}$ is a function with value in $[0,\infty)$. It is also easy to see that $\mathbf{vr}$ is continuous at singular points.
We know that a point $x_0$ is singular if and only if $\mathbf{vr}(x_0)=0$. Clearly, for every sequence $x_i \to x_0$, we must have
$\displaystyle \lim_{i \to \infty} \mathbf{vr}(x_i)=0$.
Otherwise, we have a sequence $x_i$ converging to $x_0$ and
$\displaystyle \lim_{i \to \infty} \mathbf{vr}(x_i) \geq \xi>0$. However, we note that $x_0 \in B(x_i, \tilde{K}^{-1} \xi)$ for
large $i$. Therefore, $x_0$ is forced to be regular by the improving regularity property of volume radius. Contradiction.
Therefore, discontinuity point must admit positive $\mathbf{vr}$ if it does exist. Suppose $x_0$ is a discontinuous point of $\mathbf{vr}$.
Then we can find a sequence of points $x_i \in X$ such that
\begin{align*}
&x_0=\lim_{i \to \infty} x_i, \\
&0<\mathbf{vr}(x_0)=r_0<\infty, \\
&\lim_{i \to \infty} \mathbf{vr}(x_i) \neq r_0.
\end{align*}
Clearly, $\log \mathbf{vr}(x_i)$ are uniformly bounded by Proposition~\ref{prn:SC26_7}.
So we can assume $\mathbf{vr}(x_i)$ converge to a positive number $\bar{r}$. By volume continuity, we clearly have
\begin{align*}
\omega_{2n}^{-1}r_0^{-2n}|B(x_0,r_0)|=1-\delta_0=\lim_{i \to \infty} \omega_{2n}^{-1}\mathbf{vr}(x_i)^{-2n}|B(x_i,\mathbf{vr}(x_i))|=
\omega_{2n}^{-1}\bar{r}^{-2n}|B(x_0,\bar{r})|.
\end{align*}
Since $\bar{r} \neq r_0$, we obtain from Proposition~\ref{prn:SC17_3} that $B(x_0,r_0)$ is a ball in a metric cone centered at the vertex. Note that $x_0$ is a regular
point since $\mathbf{vr}(x_0)>0$. Therefore, $B(x_0,r_0)$ is the standard ball in $\ensuremath{\mathbb{C}}^{n}$ with radius $r_0$.
Consequently, the normalized volume ratio of $B(x_0,r_0)$ is $1$, which contradicts the fact that $\mathbf{vr}(x_0)=r_0$ and the definition of volume radius.
\end{proof}
The volume radius has better property. It satisfies Harnack inequality in the interior of a shortest geodesic.
The H\"older continuity estimate of Colding-Naber (c.f.~\cite{ColdNa}) can be interpreted by volume radius as follows.
\begin{proposition}[\textbf{Global Harnack inequality of volume radius}]
For every small constant $c$, there is a constant $\epsilon=\epsilon(n,\kappa,c)$ with the following properties.
Suppose $(X,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x,y \in X$, $\gamma$ is shortest, unit speed geodesic connecting $x$ and $y$, with smooth interior parts.
Suppose $\gamma(0)=x, \gamma(L)=y$, $L \leq r$.
If $\mathbf{vr}(y)>c r$, then we have
\begin{align}
\mathbf{vr}(\gamma(t))> \epsilon r, \quad \forall \; t \in [cL, L].
\label{eqn:SC26_4}
\end{align}
In particular, if $\min\{\mathbf{vr}(x), \mathbf{vr}(y)\}>cr$, then we have
\begin{align}
\mathbf{vr}(\gamma(t))>\epsilon r, \quad \forall \; t \in [0,L].
\label{eqn:SC26_44}
\end{align}
\label{prn:SC26_2}
\end{proposition}
\begin{proof}
Clearly, (\ref{eqn:SC26_44}) follows from (\ref{eqn:SC26_4}). Therefore, it suffices to prove (\ref{eqn:SC26_4}) only.
Up to a normalization, we can assume $r=L=1$.
So $\gamma$ is the shortest geodesic connecting $x,y$ such that $\gamma(0)=x, \gamma(1)=y$.
By assumption, we have $\mathbf{vr}(y)>c$.
By local Harnack inequality of volume radius, Proposition~\ref{prn:SC26_7},
there exists $\bar{\epsilon}=\bar{\epsilon}(n, c)$ such that $\mathbf{vr}>\bar{\epsilon}$ for each $\gamma(t)$
with $t \in [1-\bar{\epsilon}, 1]$. Clearly, in the middle part of $\gamma$, i.e., for every $t \in [\bar{\epsilon}, 1-\bar{\epsilon}]$, we have
$|\Delta r|< \frac{C}{\bar{\epsilon}}$ for a universal $C=C(n)$, where $r$ is the distance to $\gamma(0)$.
Because of the segment inequality (Proposition~\ref{prn:HD17_1}) and the parabolic approximation (Lemma~\ref{lma:HE04_4}),
we can follow the proof of Proposition 3.6 and Theorem 1.1 of~\cite{ColdNa} verbatim.
Similar to the statement in the proof of Theorem 1.1 on page 1213 of~\cite{ColdNa}, we can find constants $\bar{s}=\bar{s}(n,c,\bar{\epsilon}), \bar{r}=\bar{r}(n,c,\bar{\epsilon})$
such that for every $t_1, t_2 \in [\bar{\epsilon},1-\bar{\epsilon}]$ satisfying $|t_1-t_2|<\bar{s}$ and every $r \in (0,\bar{r})$, we have
\begin{align*}
1-\frac{\delta_0}{100} \leq \frac{|B(\gamma(t_1), r)|}{|B(\gamma(t_2),r)|} \leq 1+\frac{\delta_0}{100}.
\end{align*}
Then it is easy to see that if the volume radius is uniformly bounded below at $t_1$, it must be uniformly bounded below at $t_2$.
Actually, suppose the volume radius at $\gamma(t_1)$ is greater than $r_1$ for some $r_1 \in (0,\bar{r})$,
by inequality (\ref{eqn:SC26_9}) in Proposition~\ref{prn:SC26_7}, we have
$\omega_{2n}^{-1}r^{-2n}|B(\gamma(t_1),r)| \geq 1-\frac{\delta_0}{100}$ for every $r \in [0, \frac{r_1}{\tilde{K}}]$.
Put this information into the above inequality implies that
\begin{align*}
\omega_{2n}^{-1}r^{-2n}|B(\gamma(t_2),r)| \geq \frac{1-\frac{\delta_0}{100}}{1+\frac{\delta_0}{100}}>1-\delta_0,
\quad \forall \; r \in \left[0, \frac{r_1}{\tilde{K}} \right].
\end{align*}
Therefore, the volume radius of $\gamma(t_2)$ is at least $\frac{r_1}{\tilde{K}}$. From this induction, it is clear that
\begin{align*}
\mathbf{vr}(\gamma(t)) \geq
\tilde{K}^{-\frac{1-\bar{\epsilon}-t}{\bar{s}}} \mathbf{vr}(\gamma(1-\bar{\epsilon}))
>\bar{\epsilon} \tilde{K}^{-\frac{1-\bar{\epsilon}-t}{\bar{s}}}.
\end{align*}
Let $\epsilon$ be the number on the right hand side of the above inequality when $t=\bar{\epsilon}$.
Then $\epsilon=\epsilon(\bar{\epsilon}, \tilde{K}, \bar{s})=\epsilon(n,\kappa,c)$
and we finish the proof of (\ref{eqn:SC26_4}).
\end{proof}
In general, if $X$ is only a metric space, we even do not know whether $\mathbf{vr}$ is semi-continuous. The continuity of $\mathbf{vr}$ on $X$ whenever
$X \in \widetilde{\mathscr{KS}}(n,\kappa)$ makes $\mathbf{vr}$ a convenient tool to study the geometry of $X$. By Proposition~\ref{prn:SC26_7},
one can improve regularity on a scale proportional to $\mathbf{vr}$. So it is convenient to decompose the space $X$ according to the function $\mathbf{vr}$.
\begin{definition}
Suppose $(X,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$. Define
\begin{align}
&\mathcal{F}_{r}(X) \triangleq \left\{ x \in X | \mathbf{vr}(x) \geq r \right\}, \label{eqn:SB25_1}\\
&\mathcal{D}_{r}(X) \triangleq \left( \mathcal{F}_{r}(X) \right)^{c}=\left\{ x \in X | \mathbf{vr}(x) < r \right\}. \label{eqn:SB25_2}
\end{align}
We call $\mathcal{F}_{r}(X)$ the $r$-regular part of $X$, $\mathcal{D}_{r}(X)$ the $r$-singular part of $X$.
\label{dfn:SB25_1}
\end{definition}
From Definition~\ref{dfn:SB25_1}, it is clear that
\begin{align}
& \mathcal{R}(X)= \bigcup_{r>0} \mathcal{F}_{r}(X), \label{eqn:SB25_4}\\
& \mathcal{S}(X)= \bigcap_{r>0} \mathcal{D}_{r}(X). \label{eqn:SB25_5}
\end{align}
We observe that the volume radius of each point is related to its distance to singular set by the following property.
\begin{proposition}[\textbf{$\mathbf{vr}$ bounded from above by distance to $\mathcal{S}$}]
Suppose $(X,x,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then
\begin{align}
& \left\{ x | d(x, \mathcal{S}) \geq r \right\} \supset \mathcal{F}_{\tilde{K}r}, \label{eqn:SB25_14} \\
& \left\{x | d(x,\mathcal{S}) < r \right\} \subset \mathcal{D}_{\tilde{K}r}. \label{eqn:SB25_13}
\end{align}
\end{proposition}
\begin{proof}
Choose an arbitrary point $x \in \mathcal{F}_{\tilde{K}r}$, then $ {\mathbf{vr}}(x) \geq \tilde{K} r.$ It follows from Proposition~\ref{prn:SB25_2} that
$\mathbf{vr}(y) \geq r>0$ for every point $y \in B(x, r)$. Therefore, every point in $B(x,r)$ is regular. So $d(x,\mathcal{S})>r$. This proves
(\ref{eqn:SB25_14}) by the arbitrary choice of $x \in \mathcal{F}_{\tilde{K}r}$. Taking complement of (\ref{eqn:SB25_14}), we obtain (\ref{eqn:SB25_13}).
\end{proof}
\subsection{Compactness of $\widetilde{\mathscr{KS}}(n, \kappa)$}
As a model space, $\widetilde{\mathscr{KS}}(n,\kappa)$ should have compactness.
However, we need first to obtain a weak compactness, then we improve regularity further to obtain the
genuine compactness.
It is not hard to see the weak compactness theory of Anderson-Cheeger-Colding-Tian-Naber can be generalized to apply on $\widetilde{\mathscr{KS}}(n,\kappa)$
without fundamental difficulties, almost verbatim.
Actually, the key of Anderson-Cheeger-Colding-Tian-Naber theory is that one can approximate the distance function by harmonic function, or heat flow solution,
which have much better regularity for developing integral estimates. These estimates are justified by the technical preparation in previous subsections.
\begin{proposition}[\textbf{Weak compactness}]
Suppose $(X_i,x_i,g_i) \in \widetilde{\mathscr{KS}}(n,\kappa)$, by taking subsequences if necessary,
we have
\begin{align*}
(X_i, x_i, g_i) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{X},\bar{x},\bar{g})
\end{align*}
for some length space $\bar{X}$ which satisfies all the properties of spaces in
$\widetilde{\mathscr{KS}}(n,\kappa)$ except the 3rd and 4th property, i.e., the weak convexity of $\mathcal{R}$
and the Minkowski dimension estimate $\mathcal{S}$.
However, the Hausdorff dimension of $\mathcal{S}$ is not greater than $2n-4$.
\label{prn:HA05_1}
\end{proposition}
\begin{proof}
Note that each space in $\widetilde{\mathscr{KS}}(n,\kappa)$ satisfies volume doubling property. Therefore, if
there exists a sequence $(X_i,x_i,g_i) \in \widetilde{\mathscr{KS}}(n,\kappa)$, by standard ball packing argument, it is clear that
\begin{align*}
(X_i, x_i, g_i) \stackrel{G.H.}{\longrightarrow} (\bar{X},\bar{x},\bar{g})
\end{align*}
for some length space $\bar{X}$. Then let us list the properties satisfied by $\bar{X}$.
By Proposition~\ref{prn:HD22_1}, $\bar{X}$ inherits a natural measure from the limit process, which is a measure compatible with
the limit metric structure, as that in~\cite{CC1}. Then the volume convergence follows, almost tautologically.
It follows directly from this property and the volume comparison that $\bar{X}$ satisfies
Property 6 in Definition~\ref{dfn:SC27_1}.
In the limit space $\bar{X}$, we can define regular points as the collection
of points where every tangent space is $\ensuremath{\mathbb{R}}^{2n}$, singular points as those points which are not regular.
Let $\mathcal{R}(\bar{X})$ and $\mathcal{S}(\bar{X})$ be the regular and singular part of $\bar{X}$ respectively.
We automatically obtain the regular-singular decomposition $\bar{X}=\mathcal{R}(\bar{X}) \cup \mathcal{S}(\bar{X})$.
By a version of Anderson's gap theorem(c.f. Proposition~\ref{prn:SC17_1}) and volume convergence,
a blowup argument shows that each regular point
has a small neighborhood which has a smooth manifold structure. Clearly, this manifold is Ricci-flat with an attached
limit K\"ahler structure. So we proved Property 1 and Property 2, except the non-emptiness of $\mathcal{R}$.
Note that each tangent space of $\bar{X}$ is a volume cone, due to the volume convergence
and Bishop-Gromov volume comparison, which can be established as that in~\cite{CC1}.
The it follows from Proposition~\ref{prn:HD22_2} that every volume cone is actually a metric cone. Then an induction argument can be applied, like that in~\cite{CC1}, to obtain the stratification of singularities $\mathcal{S}=\mathcal{S}_1\cup \mathcal{S}_2 \cdots \cup \mathcal{S}_{2n}$, where $\mathcal{S}_k$
is the union of singular points whose tangent space can split-off at least $(2n-k)$-straight lines.
In particular, generic points of $\bar{X}$ have tangent spaces $\ensuremath{\mathbb{R}}^{2n}$. In other words, generic points
are regular, so $\mathcal{R} \neq \emptyset$ and we finish the proof of Property 2.
The K\"ahler condition
guarantees that each tangent cone exactly splits off $\ensuremath{\mathbb{C}}^k$, by Proposition~\ref{prn:HE08_1}, as done in~\cite{CCT}.
So the stratification of singular set can be improved as
$\mathcal{S}=\mathcal{S}_2 \cup \mathcal{S}_4 \cup \cdots \mathcal{S}_{2n}$.
By Lemma~\ref{lma:HE07_1}, we can apply slice argument as that in \cite{CCT} and \cite{Cheeger03}.
Consequently, Chern-Simons theory implies that codimension 2 singularity cannot appear, due to the fact that generic slice is a smooth surface with boundary, and the Ricci curvature's restriction on such a surface is zero.
Therefore, $\mathcal{S}=\mathcal{S}_4 \cup \cdots \mathcal{S}_{2n}$,
which means $\dim_{\mathcal{H}} \mathcal{S} \leq 2n-4$.
Let $\bar{y} \in \mathcal{S}(\bar{X})$. Suppose $y_i \in X_i$ satisfies $y_i \to \bar{y}$. Then either there is a uniform $\xi$ such that
every point in each $B(y_i,\xi)$ are regular(but without uniform curvature bound as $i$ increase), or we can choose $y_i$ such that every $y_i$ is singular. In the first case,
we can use a blowup argument and Anderson's gap theorem to show that the volume density of $\bar{y}$ is strictly less
that $1-2\delta_0$. In the second case, we can use volume comparison and convergence to show
$\mathrm{v}(\bar{y}) \leq 1-2\delta_0$. So we proved Property 5.
We have checked all the properties of $\bar{X}$ as claimed. We now need to improve the convergence topology from Gromov-Hausdorff topology.
However, this improvement follows from volume convergence and the improving regularity property of volume radius, Corollary~\ref{cly:SL23_1}.
\end{proof}
From the above argument, it is clear that no new idea is needed beyond the traditional theory,
when technical lemmas and propositions in the previous sections are available. Actually, weak compactness can be established under even weaker conditions, which is checked in details by Huang in~\cite{Huang}.
Based on the weak compactness, we immediately obtain an $\epsilon$-regularity property, as that in~\cite{CCT}.
\begin{proposition}[\textbf{$\epsilon$-regularity}]
There exists an $\epsilon=\epsilon(n,\kappa)$ with the following properties.
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, $x_0 \in X$.
Suppose
\begin{align*}
d_{GH}\left( (B(x_0,1), B((z_0^*,0), 1))\right)<\epsilon
\end{align*}
where $(z_0^*,0) \in C(Z_0) \times \ensuremath{\mathbb{R}}^{2n-3}$ for some metric cone $C(Z_0)$ with vertex $z_0^*$. Then we have
\begin{align*}
\mathbf{vr}(x_0)>\frac{1}{2}.
\end{align*}
\label{prn:HD20_1}
\end{proposition}
\begin{proof}
Otherwise, there is a sequence of $\epsilon_i \to 0$ and $x_i \in X_i$ violating the statement.
By weak compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$, we can assume $x_i \to x$ and $z_i^* \to z^*$ with the following identity holds.
\begin{align*}
d_{GH}\left( (B(x,1), B((z^*,0), 1))\right)=0.
\end{align*}
In particular, the tangent cone at $x$ is exactly the cone $C(Z) \times \ensuremath{\mathbb{R}}^{2n-3}$, which must be $\ensuremath{\mathbb{C}}^n$ by the complex rigidity.
Therefore, $B(x,1)$ is the unit ball in $\ensuremath{\mathbb{C}}^n$. Thus, the volume convergence implies that for large $i$,
$2^{2n}\left|B \left(x_i, \frac{1}{2} \right) \right|$ can be very close to $1$. In particular, $\mathbf{vr}(x_i)>\frac{1}{2}$
by the definition of volume radius. However, this contradicts our assumption.
\end{proof}
Then we are able to move the integral estimate of~\cite{CN} to $X \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{proposition}[\textbf{Density estimate of regular points}]
For every $0<p<2$, there is a constant $E=E(n,\kappa,p)$ with the following properties.
Suppose $(X,x,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then we have
\begin{align}
r^{2p-2n} \int_{B(x,r)} \mathbf{vr}(y)^{-2p} dy \leq E(n,\kappa,p). \label{eqn:SB25_15}
\end{align}
\label{prn:SB25_2}
\end{proposition}
\begin{proof}
In light of Proposition~\ref{prn:SC26_7}, it is clear that volume radius and harmonic radius are uniformly equivalent. Therefore, Proposition~\ref{prn:SB25_2} is nothing but a singular version of the Cheeger-Naber estimate(c.f. the second inequality of part 2 of Corollary 1.26 in~\cite{CN}).
As pointed out by Cheeger and Naber, their estimate holds for Gromov-Hausdorff limit for Ricci-flat manifolds.
Actually, going through their proof, it is clear that the smooth structure of the underlying space is not used.
Intuitively, if Bishop-Gromov volume comparison holds, then most geodesic balls are almost volume cones, hence almost metric cones.
However, if a cone is very close to a cone which splits off at least $(2n-3)$-lines,
then it must be Euclidean space by the $\epsilon$-regularity property. This intuition was quantified in~\cite{CN}, by the method they called quantitative calculus,
which does not depends on smooth structure by its nature.
We note that the quantitative calculus argument of~\cite{CN} works when we have the following properties.
\begin{itemize}
\item Bishop-Gromov volume comparison, by Proposition~\ref{prn:HD19_1}.
\item Weak compactness of $ \widetilde{\mathscr{KS}}(n,\kappa)$, by Proposition~\ref{prn:HA05_1}.
\item Volume convergence, by Proposition~\ref{prn:HD22_1}.
\item Almost volume cone implies almost metric cone, by Proposition~\ref{prn:HD22_2}.
\item $\epsilon$-regularity, by Proposition~\ref{prn:HD20_1}.
\end{itemize}
Since all these properties hold on $ \widetilde{\mathscr{KS}}(n,\kappa)$, the proof follows that of~\cite{CN} verbatim.
\end{proof}
An immediate consequence of Proposition~\ref{prn:SB25_2} is the following volume estimate of neighborhood of
singular set.
\begin{corollary}[\textbf{Volume estimate of singular neighborhood}]
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $0<\rho<<1$.
Then for each $0<p<2$, we have
\begin{align*}
\left| \{x| d(x,\mathcal{S})<\rho, x \in B(x_0, 1)\} \right|< C\rho^{2p},
\end{align*}
for some $C=C(n,\kappa,p)$.
\label{cly:HD19_2}
\end{corollary}
\begin{proof}
It follows from Definition~\ref{dfn:SB25_1} that
\begin{align*}
(2r)^{-2p}\left|B(x_0,1) \cap \mathcal{F}_{r} \backslash \mathcal{F}_{2r}\right|
= \int_{B(x_0,1) \cap \mathcal{F}_{r} \backslash \mathcal{F}_{2r}} (2r)^{-2p}
< \int_{B(x_0,1) \cap \mathcal{F}_{r} \backslash \mathcal{F}_{2r}} \mathbf{vr}^{-2p}<E(n,\kappa,p),
\end{align*}
which implies that
\begin{align*}
\left|B(x_0,1) \cap \mathcal{D}_{2r} \backslash \mathcal{D}_{r}\right| < 2^{2p} E r^{2p}
\Rightarrow |B(x_0,1) \cap \mathcal{D}_{2r}|<\frac{E}{1-4^{-p}} (2r)^{2p}.
\end{align*}
By virtue of (\ref{eqn:SB25_13}), we have
\begin{align*}
\left|B(x_0,1) \cap \{x |d(x,\mathcal{S})<\rho\} \right| \leq |B(x_0,1) \cap \mathcal{D}_{\tilde{K}\rho}|< \frac{E}{1-4^{-p}} \tilde{K}^{2p} \rho^{2p}<C\rho^{2p}.
\end{align*}
\end{proof}
Now we are ready to prove the compactness theorem.
\begin{theorem}[\textbf{Compactness}]
$\widetilde{\mathscr{KS}}(n,\kappa)$ is compact under the pointed Cheeger-Gromov topology.
\label{thm:HD19_1}
\end{theorem}
\begin{proof}
Suppose $(X_i, x_i,g_i) \in \widetilde{\mathscr{KS}}(n,\kappa)$, we already know, by Proposition~\ref{prn:HA05_1},
that $(X_i,x_i,g_i)$ converges to a limit space $(\bar{X},\bar{x},\bar{g})$, which satisfies almost all the properties of
spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$, except the weak convexity of $\mathcal{R}$ and the
Minkowski dimension estimate of $\mathcal{S}$.
However, fix every two points $\bar{y},\bar{z} \in \mathcal{R} \subset \bar{X}$, we can find a sequence of points
$y_i, z_i \in X_i$ such that $y_i \to \bar{y}$ and $z_i \to \bar{z}$. It is clear that
$\mathbf{vr}(y_i) \to \mathbf{vr}(\bar{y})$ and $\mathbf{vr}(z_i) \to \mathbf{vr}(\bar{z})$.
It follows from the global Harnack inequality of volume radius, Proposition~\ref{prn:SC26_2}, that each shortest geodesic
$\gamma_i$ connecting $y_i$ and $z_i$ is uniformly regular. Consequently, the limit shortest geodesic $\bar{\gamma}$
connecting $\bar{y}$ and $\bar{z}$ is a smooth shortest geodesic. Therefore, we have actually proved that
$\mathcal{R}$ is convex, rather than weakly convex.
Furthermore, if we repeatedly use the first inequality in Proposition~\ref{prn:SC26_2} and smooth convergence determined by volume radius,
one can see that a shortest geodesic $\bar{\gamma}$ with smooth interior can be obtained, even if we drop the condition $\bar{y} \in \mathcal{R}$.
In other words, if $\bar{z} \in \mathcal{R}$, $\bar{y} \in \bar{X}$, then there is a shortest geodesic $\bar{\gamma}$ connecting them, with smooth interior.
This means that $\mathcal{R}$ is strongly convex.
By convexity of $\mathcal{R}$, it is clear that the limit space $\bar{X}$ has Bishop-Gromov volume comparison.
By virtue of volume convergence and the same argument in Proposition~\ref{prn:HE07_2},
we see that $\mathbf{vr}$ is a continuous function under the pointed Cheeger-Gromov topology. In other words, for every point $\bar{z} \in \bar{X}$, and points
$z_i \in X_i$ satisfying $z_i \to \bar{z}$, we have $\displaystyle \mathbf{vr}(\bar{z})=\lim_{i \to \infty} \mathbf{vr}(z_i)$.
For each $r>0$, by density estimate, Proposition~\ref{prn:SB25_2},
we see that inequality (\ref{eqn:SB25_15}) holds for every $B(x_i,r)$ uniformly.
Take limit, by the convergence of volume radius, we obtain (\ref{eqn:SB25_15}) holds on $(\bar{X},\bar{x},\bar{g})$, for
each $p\in (1.5, 2)$. Then it follows from Corollary~\ref{cly:HD19_2} and the definition of Minkowski dimension (c.f. Definition~\ref{dfn:HE08_1})
that $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$.
\end{proof}
\begin{theorem}[\textbf{Space regularity improvement}]
Suppose $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, then $\mathcal{R}$ is strongly convex, and
$\dim_{\mathcal{M}}\mathcal{S} \leq 2n-4$.
Suppose $x_0 \in \mathcal{S}$, $Y$ is a tangent space of $X$ at $x_0$.
Then $Y$ is a metric cone in $\widetilde{\mathscr{KS}}(n,\kappa)$ with the splitting
\begin{align*}
Y=\ensuremath{\mathbb{C}}^{n-k} \times C(Z)
\end{align*}
for some $k \geq 2$, where $C(Z)$ is a metric cone without lines.
\label{thm:HE08_1}
\end{theorem}
\begin{proof}
The strong convexity of $\mathcal{R}$ and $\dim_{\mathcal{M}}\mathcal{S} \leq 2n-4$ follows from the argument in the proof of Theorem~\ref{thm:HD19_1}.
Moreover, by Theorem~\ref{thm:HD19_1}, we know each tangent space, as a pointed Gromov-Hausdorff limit,
must locate in $\widetilde{\mathscr{KS}}(n,\kappa)$. Since $Y$ is a volume cone, due to volume convergence,
the splitting of $Y$ follows from Lemma~\ref{lma:HD30_1}.
\end{proof}
Because of Theorem~\ref{thm:HD19_1} to Theorem~\ref{thm:HE08_1}, it seems reasonable to make the following definition for the simplicity of notations.
\begin{definition}
A length space $(X^n,g)$ is called a conifold of complex dimension $n$ if the following properties are satisfied.
\begin{enumerate}
\item $X$ has a disjoint regular-singular decomposition $X=\mathcal{R} \cup \mathcal{S}$, where $\mathcal{R}$ is the regular part, $\mathcal{S}$ is the singular part.
A point is called regular if it has a neighborhood which is isometric to a totally geodesic convex domain of some smooth Riemannian manifold. A point is called singular
if it is not regular.
\item The regular part $\mathcal{R}$ is a nonempty, open manifold of real dimension $2n$.
Moreover, there exists a complex structure $J$ on $\mathcal{R}$ such that $(\mathcal{R}, g, J)$ is a K\"ahler manifold.
\item $\mathcal{R}$ is strongly convex, i.e., for every two points $x \in \mathcal{R}$ and $y \in X$, one can find a shortest geodesic $\gamma$ connecting $x$, $y$ whose every interior point is in $\mathcal{R}$. In particular, $\mathcal{R}$ is geodesic convex.
\item $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$, where $\mathcal{M}$ means Minkowski dimension.
\item Every tangent space of $x \in \mathcal{S}$ is a metric cone of Hausdorff dimension $2n$. Moreover, if $Y$ is a tangent cone of $x$, then the unit ball
$B(\hat{x},1)$ centered at vertex $\hat{x}$ must satisfy
\begin{align*}
|B(\hat{x},1)|_{d\mu} \leq (1-\delta_0) \omega_{2n}
\end{align*}
for some uniform positive number $\delta_0=\delta_0(n)$.
Here $d\mu$ is the $2n$-dimensional Hausdorff measure, $\omega_{2n}$ is the volume of unit ball in $\ensuremath{\mathbb{C}}^n$.
\end{enumerate}
\label{dfn:HD20_1}
\end{definition}
Roughly speaking, a conifold is a space which is almost a manifold away from a small singular set, where every tangent space is a metric cone.
Note that we abuse notation here since the conifold has different meaning
in the literature of string theory(c.f.~\cite{Green}).
It is easy to see that every K\"ahler orbifold with singularity codimension not less than $4$ is a conifold in our sense.
With this terminology, we see that $\widetilde{\mathscr{KS}}(n,\kappa)$ is nothing but the collection of Calabi-Yau conifold
with Euclidean volume growth, i.e.,
\begin{align*}
\lim_{r \to \infty} \frac{|B(x,r)|_{d\mu}}{\omega_{2n}r^{2n}} \geq \kappa, \quad \forall \; x \in X.
\end{align*}
Then Theorem~\ref{thm:HD19_1} can be interpreted as that the moduli space of non-collapsed Calabi-Yau conifold is compact, under the pointed Cheeger-Gromov topology. Theorem~\ref{thm:HE08_1} can be understood as that a ``weakly" Calabi-Yau conifold is really a conifold, due to an intrinsic improving regularity property
originates from the intrinsic Ricci flatness of the underlying space.
The property of the moduli space $\widetilde{\mathscr{KS}}(n,\kappa)$ is quite clear now.
\begin{proof}[Proof of Theorem~\ref{thmin:HE21_1}]
It follows directly from Theorem~\ref{thm:HD19_1}, Theorem~\ref{thm:HE08_1} and Definition~\ref{dfn:HD20_1}.
\end{proof}
Actually, along the route to prove Theorem~\ref{thm:HE08_1}, we shall be able to improve the regularity of the
spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$ even further. For example, we believe the following statement is true.
\begin{conjecture}
At every point $x_0$ of a Calabi-Yau conifold $X \in \widetilde{\mathscr{KS}}(n,\kappa)$, the tangent space is unique.
\label{cje:HE09_1}
\end{conjecture}
The above problem is only interesting when $n>2$ and away from generic singular point.
Note that if $X$ is a limit space of a sequence of Ricci flat manifolds, then the uniqueness of tangent cone is a well known problem, due to the work of Cheeger, Colding and Tian.
Clearly, similar questions can be asked for general K\"ahler Einstein conifold.
It is not hard to see that a compact K\"ahler Einstein conifold is a projective variety.
Due to its independent interest, we shall discuss this issue in another separate paper.
\subsection{Space-time structure of $\widetilde{\mathscr{KS}}(n)$}
Every space $X \in \widetilde{\mathscr{KS}}(n)$ can be regarded as a trivial Ricci flow solution.
Therefore, Perelman's celebrated work \cite{Pe1} can find its role in the study of $X$.
Let us briefly recall some fundamental functionals defined for the Ricci flow by Perelman.
Suppose $\{(X^m, g(t)), -T \leq t \leq 0\}$ is a Ricci flow solution on a smooth complete Riemannian manifold
$X$ of real dimension $m$. Suppose $x,y \in X$. Suppose $\boldsymbol{\gamma}$ is a space-time curve
parameterized by $\tau=-t$ such that
\begin{align*}
\boldsymbol{\gamma}(0)=(x,0), \quad \boldsymbol{\gamma}(\bar{\tau})=(y,-\bar{\tau}).
\end{align*}
Let $\gamma$ be the space-projection curve of $\boldsymbol{\gamma}$.
In other words, we have
\begin{align*}
\boldsymbol{\gamma}(\tau)=(\gamma(\tau), -\tau).
\end{align*}
By the way, for the simplicity of notations, we always use bold symbol of a Greek character to denote a space-time curve. The corresponding space projection will be denoted by the normal Greek character.
Following Perelman, the Lagrangian of the space-time curve $\boldsymbol{\gamma}$ is defined as
\begin{align*}
\mathcal{L}(\boldsymbol{\gamma})=\int_0^{\bar{\tau}} \sqrt{\tau} (R+|\dot{\gamma}|^2)_{g(-\tau)} d\tau.
\end{align*}
Among all such $\boldsymbol{\gamma}$'s that connected $(x,0)$, $(y,-\bar{\tau})$ and parameterized by $\tau$,
there is at least one smooth curve $\boldsymbol{\alpha}$ which minimizes the Lagrangian.
This curve is called a shortest reduced geodesic. The reduced distance between $(x,0)$ and $(y,-\bar{\tau})$ is
defined as
\begin{align*}
l((x,0),(y,-\bar{\tau}))=\frac{\mathcal{L}(\boldsymbol{\alpha})}{2\sqrt{\bar{\tau}}}.
\end{align*}
Let $V=\dot{\alpha}$. Then $V$ satisfies the equation
\begin{align*}
\nabla_V V +\frac{V}{2\tau} + 2Ric(V, \cdot) + \frac{\nabla R}{2}=0,
\end{align*}
which is called the reduced geodesic equation. It is easy to check that $\dot{\alpha}=V=\nabla l$.
The reduced volume is defined as
\begin{align*}
\mathcal{V}((x,0), \bar{\tau})=\int_{X} (4\pi \bar{\tau})^{-\frac{m}{2}} e^{-l} dv.
\end{align*}
It is proved by Perelman that $(4\pi \tau)^{-\frac{m}{2}} e^{-l}dv$, the reduced volume element, is
monotony non-increasing along each reduced geodesic emanating from $(x,0)$.
Suppose the Ricci flow solution mentioned above is static, i.e., $Ric \equiv 0$. Then it is easy to check that
\begin{align}
\begin{cases}
&\mathcal{L}(\boldsymbol{\alpha})=\frac{d^2(x,y)}{2\sqrt{\bar{\tau}}}, \\
&l((x,0),(y,-\bar{\tau}))=\frac{d^2(x,y)}{4\bar{\tau}}, \\
&\nabla_V V +\frac{V}{2\tau}=0, \\
&|\dot{\alpha}|^2=|V|^2=|\nabla l|^2=\tau l, \\
&\mathcal{V}((x,0), \bar{\tau})=\int_{X} (4\pi \bar{\tau})^{-\frac{m}{2}} e^{-\frac{d^2}{4\bar{\tau}}} dv.
\end{cases}
\label{eqn:SL25_6}
\end{align}
Now we assume $X \in \widetilde{\mathscr{KS}}(n)$.
By a trivial extension in an extra time direction, we obtain a static, eternal singular K\"ahler Ricci flow solution.
Since distance structure is already known, we can define reduced distance, reduced volume, etc, following the equation
(\ref{eqn:SL25_6}). Clearly, this definition coincides with the original one when $X$ is smooth.
The following theorem is important to bridge the Cheeger-Colding's structure theory to the Ricci flow theory.
\begin{theorem}[\textbf{Volume ratio and reduced volume}]
Suppose $X \in \widetilde{\mathscr{KS}}(n)$, $x \in X$. Let $X \times (-\infty, 0]$ have the obvious
static space-time structure. Then we have
\begin{align}
&\mathrm{avr}(X)=\lim_{\tau \to \infty} \mathcal{V}((x,0), \tau). \label{eqn:SL25_4}\\
&\mathrm{v}(x)=\lim_{\tau \to 0} \mathcal{V}((x,0), \tau). \label{eqn:SL25_7}
\end{align}
\label{thm:SL25_1}
\end{theorem}
\begin{proof}
The proof relies on the volume cone structure at local tangent space, or tangent space at infinity. So
the proof of (\ref{eqn:SL25_4}) and (\ref{eqn:SL25_7}) are almost the same. For simplicity, we will only
prove (\ref{eqn:SL25_4}) and leave the proof for (\ref{eqn:SL25_7}) to the readers.
Clearly, the real dimension of $X$ is $m=2n$.
For each $\epsilon$ small, we have
\begin{align*}
m \omega_m \mathrm{avr}(X)+\epsilon
> H^{-m+1}|\partial B(x,H)| >m \omega_m \mathrm{avr}(X)-\epsilon,
\end{align*}
whenever $H$ is large enough. Note that
\begin{align*}
\mathcal{V}((x,0), H^2)&=(4\pi)^{-\frac{m}{2}} H^{-m} \int_0^{\infty} |\partial B(x,r)| e^{-\frac{r^2}{4H^2}} dr, \\
1&=(4\pi)^{-\frac{m}{2}} H^{-m} \int_0^{\infty} m \omega_m r^{m-1} e^{-\frac{r^2}{4H^2}} dr.
\end{align*}
So we have
\begin{align*}
&\quad \mathcal{V}((x,0), H^2)-\mathrm{avr}(X)\\
&=(4\pi)^{-\frac{m}{2}} H^{-m} \int_0^{\infty} \left\{|\partial B(x,r)| -m\omega_m\mathrm{avr}(X) r^{m-1}\right\}
e^{-\frac{r^2}{4H^2}} dr.
\end{align*}
We can further decompose the last integral as follows.
\begin{align*}
&\quad \left| \int_0^{\epsilon H} \left\{|\partial B(x,r)| -m\omega_m\mathrm{avr}(X)r^{m-1} \right\}
e^{-\frac{r^2}{4H^2}} dr \right|\\
&\leq m\omega_m \int_0^{\epsilon H} r^{m-1} e^{-\frac{r^2}{4H^2}} dr
=m\omega_m H^m \int_0^{\epsilon} s^{m-1}e^{-\frac{s^2}{4}} ds, \\
&\quad \left| \int_{\epsilon H}^{\infty} \left\{|\partial B(x,r)| -m\omega_m\mathrm{avr}(X)r^{m-1} \right\}
e^{-\frac{r^2}{4H^2}} dr \right|\\
&\leq \epsilon \int_{\epsilon H}^{\infty} r^{m-1}e^{-\frac{r^2}{4H^2}} dr
<\epsilon H^m \int_{0}^{\infty} e^{-\frac{s^2}{4}} ds=\epsilon H^m \pi^{\frac{1}{2}}.
\end{align*}
Therefore, we have
\begin{align*}
\left|\mathcal{V}((x,0), H^2)-\mathrm{avr}(X)\right|
<(4\pi)^{-\frac{m}{2}} \left\{m\omega_m \int_0^{\epsilon} s^{m-1}e^{-\frac{s^2}{4}} ds + \epsilon \pi^{\frac{1}{2}}\right\}.
\end{align*}
Since the above inequality holds for every $H$ large enough, we see that
\begin{align*}
\left|\lim_{\tau \to \infty} \mathcal{V}((x,0),\tau)-\mathrm{avr}(X) \right|
<(4\pi)^{-\frac{m}{2}} \left\{m\omega_m \int_0^{\epsilon} s^{m-1}e^{-\frac{s^2}{4}} ds + \epsilon \pi^{\frac{1}{2}}\right\}.
\end{align*}
Let $\epsilon \to 0$, we obtain (\ref{eqn:SL25_4}).
\end{proof}
Theorem~\ref{thm:SL25_1} says that when we study the asymptotic behavior of $X$, the volume ratio and reduced volume
play the same role. Note that volume ratio is monotone along radius direction on a manifold with nonnegative Ricci curvature, which property plays an essential role in Cheeger-Colding's theory.
Since reduced volume is monotone along Ricci flow, Theorem~\ref{thm:SL25_1} suggests that Cheeger-Colding's theory can be transplanted to the Ricci flow case.
\section{Canonical radius}
In section 2, we established the compactness of the model space $\widetilde{\mathscr{KS}}(n,\kappa)$,
following the route of Anderson-Cheeger-Colding-Tian-Naber.
It is clear that the volume ratio's monotonicity is essential to this route. However, most K\"ahler manifolds do not have this monotonicity.
For example, if we take out a time slice from a K\"ahler Ricci flow solution, there is no obvious reason at all that volume ratio monotonicity holds on it.
Therefore, in order to set up weak compactness for general K\"ahler manifolds, we have to give up the volume ratio monotonicity and search for a new route.
This will be done in this section.
\subsection{Motivation and definition}
Let us continue the discussion in Section 2.6.
As a consequence of the weak compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$, we have
density estimate of volume radius, Proposition~\ref{prn:SB25_2}. For simplicity of notation, we fix $p_0=1.8$.
Actually, any $p \in (1.5, 2)$ works as well. Define
\begin{align}
\mathbf{E} \triangleq E(n,\kappa,p_0)+200 \omega_{2n} \kappa^{-1}. \label{eqn:SC17_11}
\end{align}
Here we adjust the number $E(n,\kappa,p_0)$ to a much larger number, to reserve spaces for later use.
Then Proposition~\ref{prn:SB25_2} implies
\begin{align}
r^{2p_0-2n} \int_{B(x,r)} \mathbf{vr}(y)^{-2p_0} dy < \mathbf{E}. \label{eqn:SL23_5}
\end{align}
The above inequality contains a lot of information. For example, it immediately implies that in every unit ball,
there exists a fixed sized sub-ball with uniform regularity.
\begin{proposition}[\textbf{Generic regular sub-ball}]
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then we have
\begin{align}
\mathcal{F}_{c_b r} \cap B(x_0,r) \neq \emptyset, \label{eqn:SC29_10}
\end{align}
where
\begin{align}
c_b \triangleq \left(\frac{\omega_{2n} \kappa}{4 \mathbf{E}} \right)^{\frac{1}{2p_0}}.
\label{eqn:SL23_6}
\end{align}
\label{prn:HE11_1}
\end{proposition}
\begin{proof}
Let $\mathbf{vr}$ achieve maximum value at $y_0$ in the ball closure $\overline{B(x_0,r)}$.
By inequality (\ref{eqn:SL23_5}), we have
\begin{align*}
\mathbf{vr}(y_0)^{-2p_0} \leq \fint_{B(x,r)} \mathbf{vr}(y)^{-2p_0} dy
\leq (\omega_{2n} \kappa)^{-1} r^{-2n} \int_{B(x,r)} \mathbf{vr}(y)^{-2p_0} dy
\leq (\omega_{2n} \kappa)^{-1} r^{-2p_0} \mathbf{E}.
\end{align*}
It follows that
\begin{align*}
\mathbf{vr}(y_0) \geq \left(\frac{\omega_{2n}\kappa}{\mathbf{E}} \right)^{\frac{1}{2p_0}}r
>c_br.
\end{align*}
By continuity of $\mathbf{vr}$, there must exist a point $z \in B(x_0,r)$ such that $\mathbf{vr}(z) >c_b r$.
In other words, we have $z \in \mathcal{F}_{cr} \cap B(x_0,r)$. So (\ref{eqn:SC29_10}) holds.
\end{proof}
Let $\mathbf{E}$ and $c_b$ be the constants defined in (\ref{eqn:SC17_11}) and (\ref{eqn:SL23_6}).
Then we can choose a small constant $\epsilon_b$ such that
\begin{align}
\epsilon_b \triangleq \epsilon\left(n,\kappa,\frac{c_b}{100} \right)
\label{eqn:SL23_7}
\end{align}
by the dependence in (\ref{eqn:SC26_44}) of Proposition~\ref{prn:SC26_2}.
Combining the estimates in $\widetilde{\mathscr{KS}}(n,\kappa)$, we obtain the following theorem.
\begin{theorem}[\textbf{A priori estimates in model spaces}]
Suppose $(X,x_0,g) \in \widetilde{\mathscr{KS}}(n,\kappa)$, $r$ is a positive number. Then the following estimates
hold.
\begin{enumerate}
\item Strong volume ratio estimate: $\kappa \leq \omega_{2n}^{-1}r^{-2n}|B(x_0,r)| \leq 1$.
\item Strong regularity estimate: $r^{2+k}|\nabla^k Rm|\leq c_a^{-2}$ in the ball $B(x_0, c_a r)$ for every $0 \leq k \leq 5$ whenever $\mathbf{vr}(x_0) \geq r$.
\item Strong density estimate: $\displaystyle r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}(y)^{-2p_0} dy \leq \mathbf{E}$.
\item Strong connectivity estimate: Every two points $y_1,y_2 \subset B(x_0,r) \cap \mathcal{F}_{\frac{1}{100}c_b r}(X)$
can be connected by a shortest geodesic $\gamma$ such that
$\gamma \subset \mathcal{F}_{\epsilon_b r} (X)$.
\end{enumerate}
\label{thm:SL21_1}
\end{theorem}
We shall show that a weak compactness of $\widetilde{\mathscr{KS}}(n,\kappa)$ can be established using the estimates in Theorem~\ref{thm:SL21_1},
without knowing the volume ratio monotonicity.
For this new route of weak compactness theory, we define a scale called canonical radius with respect to $\widetilde{\mathscr{KS}}(n,\kappa)$.
Under the canonical radius, rough estimates like that in Theorem~\ref{thm:SL21_1} are satisfied.
In this section, we focus on the study of smooth complete K\"ahler manifold.
Every such a manifold is denoted by $(M^n, g, J)$, where $n$ is the complex dimension.
The Hausdorff dimension, or real dimension of $M$ is $m=2n$.
We first need to make sense of the rough volume radius, without the volume ratio monotonicity.
\begin{definition}
Denote the set $\left\{ r \left| 0<r<\rho, \omega_{2n}^{-1}r^{-2n}|B(x_0,r)|\geq 1-\delta_0 \right. \right\}$ by
$I_{x_0}^{(\rho)}$ where $x_0 \in M$, $\rho$ is a positive number.
Clearly, $I_{x_0}^{(\rho)} \neq \emptyset$ since $M$ is smooth. Define
\begin{align*}
\mathbf{vr}^{(\rho)}(x_0) \triangleq \sup I_{x_0}^{(\rho)}.
\end{align*}
For each pair $0<r \leq \rho$, define
\begin{align*}
&\mathcal{F}_{r}^{(\rho)}(M) \triangleq \left\{ x \in M | \mathbf{vr}^{(\rho)}(x) \geq r \right\}, \\
&\mathcal{D}_{r}^{(\rho)}(M) \triangleq \left\{ x \in M | \mathbf{vr}^{(\rho)}(x) < r \right\}.
\end{align*}
\label{dfn:SC24_1}
\end{definition}
\begin{definition}
A subset $\Omega$ of $M$ is called $\epsilon$-regular-connected on the scale $\rho$ if every two points $x,y \in \Omega$ can be connected by a
rectifiable curve $\gamma \subset \mathcal{F}_{\epsilon}^{(\rho)}$ and $|\gamma| < 2d(x,y)$.
For notational simplicity, if the scale is clear in the context, we shall just say
$\Omega$ is $\epsilon$-regular-connected.
\label{dfn:SC15_3}
\end{definition}
Inspired by the estimates in Theorem~\ref{thm:SL21_1}, we can define the concept of canonical radius as follows.
\begin{definition}
We say that the canonical radius (with respect to model space $\widetilde{\mathscr{KS}}(n,\kappa)$) of a point $x_0 \in M$ is
not less than $r_0$ if for every $r < r_0$, we have the following properties.
\begin{enumerate}
\item Volume ratio estimate: $\kappa \leq \omega_{2n}^{-1}r^{-2n}|B(x_0,r)| \leq \kappa^{-1}$.
\item Regularity estimate: $r^{2+k}|\nabla^k Rm|\leq 4 c_a^{-2}$ in the ball $B(x_0, \frac{1}{2}c_a r)$ for every $0 \leq k \leq 5$ whenever
$\omega_{2n}^{-1}r^{-2n}|B(x_0,r)| \geq 1-\delta_0$.
\item Density estimate: $\displaystyle r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}^{(r)}(y)^{-2p_0} dy \leq 2\mathbf{E}$.
\item Connectivity estimate: $B(x_0,r) \cap \mathcal{F}_{\frac{1}{50}c_b r}^{(r)}(M)$ is $\frac{1}{2}\epsilon_b r$-regular-connected on the scale $r$.
\end{enumerate}
Then we define canonical radius of $x_0$ to be the supreme of all the $r_0$ with the properties mentioned above.
We denote the canonical radius by $\mathbf{cr}(x_0)$.
For subset $\Omega \subset M$, we define the canonical radius of $\Omega$ as the infimum of all $\mathbf{cr}(x)$ where $x \in \Omega$.
We denote this canonical radius by $\mathbf{cr}(\Omega)$.
\label{dfn:SC02_1}
\end{definition}
\begin{remark}
In Definition~\ref{dfn:SC02_1}, the first condition(volume ratio estimate) is used to guarantee the existence of Gromov-Hausdorff limit. The second condition (regularity estimate) is for the purpose of improving regularity. The third condition (density estimate), together with the second condition(regularity estimate),
implies that the regular part is almost dense(c.f.~Theorem~\ref{thm:HE11_1}). The fourth condition(connectivity estimate) is defined to assure that the regular part is connected (c.f.~Proposition~\ref{prn:SB27_1}).
\label{rmk:SB16_1}
\end{remark}
Because of the regularity estimate of Definition~\ref{dfn:SC02_1}, it is useful to define the concept of canonical volume radius as follows.
\begin{definition}
Suppose $\rho_0=\mathbf{cr}(x_0)$. Then we define
\begin{align}
\mathbf{cvr}(x_0) \triangleq \mathbf{vr}^{(\rho_0)}(x_0). \label{eqn:SC24_4}
\end{align}
We call $\mathbf{cvr}(x_0)$ the canonical volume radius of the point $x_0$.
\label{dfn:SB25_2}
\end{definition}
\begin{remark}
For every compact smooth manifold $M$, there is an $\eta>0$ such that
every geodesic ball with radius less than $\eta$ must have normalized volume radius at least $1-\delta_0$.
Then it is easy to see that $\displaystyle r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}^{(r)}(y)^{-2p_0} dy$ is a continuous
function with respect to $x_0$ and $r$. Therefore, if $\rho_0=\mathbf{cr}(x_0)$ is a finite positive number, we have
\begin{align}
\displaystyle \rho_0^{2p_0-2n} \int_{B(x_0, \rho_0)} \mathbf{vr}^{(\rho_0)}(y)^{-2p_0} dy \leq 2\mathbf{E}.
\label{eqn:SL25_1}
\end{align}
If $r \leq \mathbf{cr}(M)$, then $\mathbf{vr}^{(r)} \leq \mathbf{cvr}$ as functions. Therefore, we have
\begin{align}
r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{cvr}(y)^{-2p_0} dy \leq
r^{2p_0-2n} \int_{B(x_0, r)} \mathbf{vr}^{(r)}(y)^{-2p_0} dy \leq 2\mathbf{E}.
\label{eqn:SL25_2}
\end{align}
\label{rmk:SL22_1}
\end{remark}
Let $r_0$ be $\mathbf{cvr}(x_0)$. By Definition~\ref{dfn:SB25_2}, it is clear that $r_0 \leq \mathbf{cr}(x_0)$.
If $r_0=\mathbf{cvr}(x_0)<\mathbf{cr}(x_0)$, then we have
\begin{align}
&\omega_{2n}^{-1}r_0^{-2n}|B(x_0,r_0)| = 1-\delta_0, \label{eqn:SC16_2}\\
&\omega_{2n}^{-1}r^{-2n}|B(x_0,r)| < 1-\delta_0, \quad \forall \; r \in (r_0, \mathbf{cr}(x_0)). \label{eqn:SC16_3}
\end{align}
If $r_0=\mathbf{cvr}(x_0)=\mathbf{cr}(x_0)$, then we only have
\begin{align}
\omega_{2n}^{-1} r_0^{-2n}|B(x_0,r_0)| \geq 1-\delta_0. \label{eqn:SC17_1}
\end{align}
It is possible that equality (\ref{eqn:SC16_2}) does not hold on the scale $r_0$ in this case.
\begin{remark}
The three radii functions, $\mathbf{cr}$,$\mathbf{vr}$ and $\mathbf{cvr}$ are all positive functions on the interior part of $M$.
However, we do not know whether they are continuous in general.
\label{rmk:DA25_1}
\end{remark}
We shall use canonical radius as a tool to study the weak-compactness theory of K\"ahler manifolds.
\subsection{Rough estimates when canonical radius is bounded from below}
We assume $\mathbf{cr}(M) \geq 1$ in the following discussion of this subsection.
Under this condition, we collect important estimates for the development of weak-compactness.
For simplicity of notation, we denote
\begin{align}
\mathcal{F}_{r} \triangleq \mathcal{F}_{r}^{(\mathbf{cr}(M))}, \quad
\mathcal{D}_{r} \triangleq \mathcal{D}_{r}^{(\mathbf{cr}(M))}.
\label{eqn:SL27_1}
\end{align}
Note that this definition can be regarded as the generalization of the corresponding definition for metric spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$.
It coincides the original one since $\mathbf{cr}(M)=\infty$ whenever $M \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\begin{proposition}
For every $0<r\leq \rho_0 \leq 1$, $x_0 \in M$, we have
\begin{align}
&\left| B(x_0,\rho_0) \cap \mathcal{D}_r \right| < 4\mathbf{E} \rho_0^{2n-2p_0} r^{2p_0}, \label{eqn:SC02_2} \\
&\left| B(x_0,\rho_0) \cap \mathcal{F}_r \right| > \left(\kappa \omega_{2n} - 4\mathbf{E}r^{2p_0}\rho_0^{-2p_0} \right) \rho_0^{2n}.
\label{eqn:SC02_4}
\end{align}
In particular, there exists at least one point $z \in B(x_0,\rho_0)$ such that
\begin{align}
\mathbf{cvr}(z) > c_b \rho_0,
\label{eqn:SC02_5}
\end{align}
where $c_b=\left(\frac{\kappa \omega_{2n}}{4 \mathbf{E}} \right)^{\frac{1}{2p_0}}$.
\label{prn:SC02_2}
\end{proposition}
\begin{proof}
Recall that $\mathbf{vr}^{(\mathbf{cr}(M))} \geq \mathbf{vr}^{(\rho_0)}$. By density estimate(c.f. Definition~\ref{dfn:SC02_1}), we have
\begin{align*}
r^{-2p_0} \left|B(x_0,\rho_0) \cap \mathcal{F}_{\frac{r}{2}} \backslash \mathcal{F}_{r} \right|
\leq \int_{B(x_0,\rho_0) \cap \mathcal{F}_{\frac{r}{2}} \backslash \mathcal{F}_{r}}\left\{ \mathbf{vr}^{(\mathbf{cr}(M))}\right\}^{-2p_0}
\leq \int_{B(x_0,\rho_0)} \left\{ \mathbf{vr}^{(\rho_0)}\right\}^{-2p_0} \leq 2\mathbf{E} \rho_0^{2n-2p_0}.
\end{align*}
Note that $\mathcal{F}_{\frac{r}{2}} \backslash \mathcal{F}_{r}=\mathcal{D}_{r} \backslash \mathcal{D}_{\frac{r}{2}}$, so we have
\begin{align}
\left|B(x,\rho_0) \cap \mathcal{D}_{r} \backslash \mathcal{D}_{\frac{r}{2}} \right| \leq 2\mathbf{E} \rho_0^{2n-2p_0} r^{2p_0}.
\label{eqn:SC02_3}
\end{align}
Since $\mathcal{D}_r=\bigcup_{k=0}^{\infty} \mathcal{D}_{2^{-k}r} \backslash \mathcal{D}_{2^{-k-1}r}$, it follows from (\ref{eqn:SC02_3}) that
\begin{align*}
\left| B(x,\rho_0) \cap \mathcal{D}_r \right| \leq \frac{2\mathbf{E}}{1-4^{-p_0}} \rho_0^{2n-2p_0} r^{2p_0},
\end{align*}
which implies (\ref{eqn:SC02_2}) since $p_0=1.8$. Together with the $\kappa$-non-collapsing condition, (\ref{eqn:SC02_2}) yields (\ref{eqn:SC02_4}).
Let $r=c_b \rho_0$, then (\ref{eqn:SC02_2}) implies
\begin{align*}
\left| B(x_0,\rho_0) \cap \mathcal{F}_{c_b \rho_0} \right| >0.
\end{align*}
In particular, $B(x_0,\rho_0) \cap \mathcal{F}_{c_b \rho_0} \neq \emptyset$. In other words, we can find a point $z \in B(x_0, \rho_0)$ satisfying
$\mathbf{vr}^{(\mathbf{cr}(M))}>c_b \rho_0$ and consequently inequality (\ref{eqn:SC02_5}).
\end{proof}
\begin{corollary}
Suppose $x_0 \in M$, $H \geq 1 \geq r$, then we have
\begin{align}
\left| B(x_0,H) \cap \mathcal{D}_r \right| \leq \left(\frac{2^{2n+2}\left|B(x_0, 2H) \right|}{\kappa \omega_{2n}} \right) r^{2p_0} \mathbf{E}.
\label{eqn:SC06_1}
\end{align}
\label{cly:SC04_2}
\end{corollary}
\begin{proof}
Try to fill the ball $B(x_0,H)$ with balls $B(y_i,\frac{1}{2})$ such that $y_i \in B(x_0,H)$ until no more such balls can squeeze in. Clearly, we have
\begin{align*}
B(x_0,H) \subset \bigcup_{i=1}^{N} B(y_i,1), \quad \bigcup_{i=1}^{N} B \left(y_i, \frac{1}{2} \right) \subset B \left(x_0,H+\frac{1}{2} \right) \subset B(x_0,2H).
\end{align*}
On one hand, the balls $B\left(y_i,\frac{1}{2}\right)$ are disjoint to each other. So we have
\begin{align}
N\kappa \omega_{2n} \left( \frac{1}{2}\right)^{2n} \leq \sum_{i=1}^{N} \left|B(y_i,1)\right|
\leq \left|B(x_0,2H)\right|, \quad
\Rightarrow \quad N \leq \frac{2^{2n} |B(x_0,2H)|}{\kappa \omega_{2n}}.
\label{eqn:SC06_2}
\end{align}
On the other hand, $B(x_0,H)$ is covered by $\displaystyle \bigcup_{i=1}^{N} B(y_i,1)$. So we have
\begin{align*}
\left| B(x_0,H) \cap \mathcal{D}_r \right| \leq \sum_{i=1}^{N} \left| B(y_i,1) \cap \mathcal{D}_r \right| \leq 4N\mathbf{E} r^{2p_0} \leq \left(\frac{2^{2n+2}\left|B(x_0, 2H) \right|}{\kappa \omega_{2n}} \right) r^{2p_0} \mathbf{E},
\end{align*}
where we used (\ref{eqn:SC06_2}) and (\ref{eqn:SC02_2}).
\end{proof}
\begin{proposition}
For every $r \leq 1$, two points $x,y \in \mathcal{F}_{r}$ can be connected by a curve $\gamma \subset \mathcal{F}_{\frac{1}{2}\epsilon_b r}$ with length $|\gamma| <3d(x,y)$.
\label{prn:SB27_1}
\end{proposition}
\begin{proof}
By rescaling if necessary, we can assume $r=1$. Then $\mathbf{cr}(M) \geq 1$.
Suppose $x,y \in \mathcal{F}_{1}$. If $d(x,y) \leq 1$, then there is a curve connecting $x,y$ and it satisfies the requirements, by the $\frac{1}{2}\epsilon_b$-regular connectivity property of the canonical radius. So we assume $H=d(x,y) >1$ without loss of generality.
Let $\beta$ be a shortest geodesic connecting $x,y$ such that $\beta(0)=x$ and $\beta(H)=y$.
Let $N$ be an integer locating in $[2H, 2H+1]$. Define
\begin{align*}
s_i=\frac{Hi}{N}, \quad x_i=\beta(s_i).
\end{align*}
Clearly, $x_0=x, x_N=y$, which are both in $\mathcal{F}_1 \subset \mathcal{F}_{\frac{c_b}{50}}$.
For each $1 \leq i \leq N-1$, $x_i$ may not locate in $\mathcal{F}_{\frac{c_b}{50}}$. However,
in the ball $B(x_i, \frac{1}{20})$, there exists a point $x_i'$ such that
\begin{align*}
\mathbf{vr}(x_i') \geq \frac{1}{2} c_b \cdot \frac{1}{20}=\frac{c_b}{40}> \frac{c_b}{50}.
\end{align*}
Clearly, we have
\begin{align*}
d(x_i', x_{i+1}') \leq d(x_i', x_i) + d(x_i, x_{i+1}) + d(x_{i+1}, x_{i+1}') \leq \frac{H}{N} + \frac{1}{10} < \frac{3}{5} \leq 1.
\end{align*}
Since $\mathbf{cr}(M) \geq 1$, one can apply $\frac{1}{2}\epsilon_b$-regular connectivity property of the canonical radius
to find a curve $\beta_i$ connecting $x_i'$ and $x_{i+1}'$ such that
$\beta_i \subset \mathcal{F}_{\frac{1}{2}\epsilon_b}$. Moreover, we have
\begin{align*}
|\beta_i| \leq 2d(x_i', x_{i+1}') \leq 2\left(\frac{H}{N} + \frac{1}{10} \right).
\end{align*}
Concatenating all $\beta_i$'s, we obtain a curve $\gamma$ connecting $x=x_0,y=x_{N}$ and $\gamma \subset \mathcal{F}_{\frac{1}{2} \epsilon_b}$.
Furthermore, we have
\begin{align*}
|\gamma|=\sum_{i=0}^{N-1} |\beta_i| \leq 2N \left( \frac{H}{N} + \frac{1}{10} \right)= 2H +\frac{1}{15}N \leq 2H + \frac{2H+1}{5} \leq \frac{12}{5}H + \frac{1}{5}
< \frac{13}{5}H <3H.
\end{align*}
\end{proof}
\begin{corollary}
For every $x \in M$, $0<r \leq 1$, we can find a curve $\gamma$ connecting $\partial B(x, \frac{r}{2})$
and $\partial B(x,r)$ such that
\begin{align*}
\gamma \subset \mathcal{F}_{\frac{1}{2}\epsilon_br}, \quad |\gamma| \leq 2r.
\end{align*}
In particular, we have
\begin{align*}
\partial B(x, r) \cap \mathcal{F}_{\frac{\epsilon_b}{2}r} \neq \emptyset.
\end{align*}
\label{cly:SL24_1}
\end{corollary}
\begin{proof}
Let $\beta$ be a shortest geodesic connecting $x$ and some point $y \in \partial B(x, \frac{9}{8}r)$.
Let $z$ be the intersection point of $\beta$ and $\partial B(x, \frac{3}{8}r)$. Let $y', z'$ be regular points
around $y,z$, i.e., we require
\begin{align*}
y' \in B\left(y,\frac{r}{8} \right) \cap \mathcal{F}_{\frac{c_b}{8}r}, \quad
z' \in B\left(z,\frac{r}{8} \right) \cap \mathcal{F}_{\frac{c_b}{8}r}.
\end{align*}
Clearly, triangle inequality implies that
\begin{align*}
d(y',z') \leq \frac{6}{8}r + \frac{1}{8}r + \frac{1}{8}r=r \leq 1.
\end{align*}
Since $\mathbf{cr}(M) \geq 1$, by connectivity estimate, there is a curve $\alpha$ connecting $y'$ and $z'$ such that
\begin{align*}
|\alpha| \leq 2r, \quad \alpha \subset \mathcal{F}_{\frac{\epsilon_b}{2}r}.
\end{align*}
Note that $z' \in B(x,\frac{r}{2})$ and $y' \in B(x,r)^{c}$. The connectedness of $M$ guarantees that
$\alpha$ must have intersection with both $\partial B(x,\frac{r}{2})$ and $\partial B(x,r)$.
So we can truncate $\alpha$ to obtain a curve $\gamma$ which connects $\partial B(x,\frac{r}{2})$ and $\partial B(x,r)$. Clearly, we have
\begin{align*}
\gamma \subset \alpha \subset \mathcal{F}_{\frac{\epsilon_b}{2}r},
\quad |\gamma| \leq |\alpha| \leq 2r.
\end{align*}
\end{proof}
\begin{proposition}
Suppose $x \in M$, $0<r \leq 1$. Then for every point $y \in \mathcal{F}_{\frac{1}{2}\epsilon_br} \cap \partial B(x,r)$.
There is a curve $\gamma$ connecting $x$ and $y$ such that
\begin{itemize}
\item $|\gamma|<10r$.
\item For each nonnegative integer $i$,
$\gamma \cap B(x,2^{-i}r) \backslash B\left(x, 2^{-i-1}r \right)$ contains a component which connects
$\partial B(x,2^{-i}r)$ and $\partial B(x, 2^{-i-1}r)$ and is contained in
$\mathcal{F}_{2^{-i-3}\epsilon_b^2r}$.
\end{itemize}
\label{prn:SL24_1}
\end{proposition}
\begin{proof}
Choose $y_i$ be a point on $\partial B(x, 2^{-i}r) \cap \mathcal{F}_{2^{-i-1}\epsilon_b r}$.
By Proposition~\ref{prn:SB27_1}, for each $i \geq 0$, there is a curve $\gamma_i$ connecting
$y_i$ and $y_{i+1}$ such that
\begin{align*}
|\gamma_i| < 9 \cdot 2^{-i-1} r, \quad \gamma_i \subset \mathcal{F}_{2^{-i-3}\epsilon_b^2 r}.
\end{align*}
Concatenate all the $\gamma_i$'s to obtain $\gamma$. Then $\gamma$ satisfies all the properties.
\end{proof}
For the purpose of improving regularity, we need to study the behavior of $\mathbf{cvr}$.
Similar to $\mathbf{vr}$ on spaces in $\widetilde{\mathscr{KS}}(n,\kappa)$ (c.f. Proposition~\ref{prn:SC26_7}),
$\mathbf{cvr}$ satisfies a local Harnack inequality.
\begin{proposition}
There is a constant $K=K(n,\kappa)$ with the following properties.
Suppose $x \in X$, $r=\mathbf{cvr}(x)<\frac{1}{K}$, then for every point $y \in B(x, K^{-1}r)$, we have
\begin{align}
&K^{-1}r \leq \mathbf{cvr}(y) \leq Kr, \label{eqn:SC24_1}\\
&\omega_{2n}^{-1}\rho^{-2n} |B(y,\rho)| \geq 1-\frac{1}{100}\delta_0, \quad \forall \; \rho \in (0, K^{-1}r), \label{eqn:SC24_2}\\
& |Rm|(y) \leq K^{2} r^{-2}, \label{eqn:SC29_6}\\
& inj(y) \geq K^{-1} r. \label{eqn:SC29_7}
\end{align}
\label{prn:SC24_1}
\end{proposition}
\begin{corollary}
For every $r \in (0,1]$, $\mathcal{F}_{r}(M)$ is a closed set. Moreover,
$\mathbf{cvr}$ is an upper-semi-continuous function on $\mathcal{F}_{r}(M)$.
\label{cly:SC24_1}
\end{corollary}
\begin{proof}
Fix $r \in (0,1]$. Suppose $x_i \in \mathcal{F}_{r}(M)$ converges to a point $x \in M$.
Let $r_i=\mathbf{cvr}(x_i)$. We need to show $\mathbf{cvr}(x) \geq r$. Clearly, this follows directly
if $\displaystyle \lim_{i \to \infty} r_i=\infty$ by Proposition~\ref{prn:SC24_1}. Without loss of generality, we may assume
$r_i$ is uniformly bounded from above. Use Proposition~\ref{prn:SC24_1} again, we see that
$r_i$ is uniformly bounded away from zero. Let $r$ be a limit of $r_i$. Then we have
\begin{align}
|B(x,r)|=\lim_{i \to \infty} |B(x_i,r_i)| \geq \lim_{i \to \infty} (1-\delta_0) \omega_{2n} r_i^{2n}
= (1-\delta_0) \omega_{2n} r^{2n},
\end{align}
which implies $\displaystyle \mathbf{cvr}(x) \geq r= \lim_{i \to \infty} r_i$ by definition of canonical volume radius and the fact that
$\displaystyle r=\lim_{i \to \infty} r_i\leq 1 \leq \mathbf{cvr}$.
Consequently, we have $x \in \mathcal{F}_{r}(M)$. Therefore, $\mathcal{F}_{r}(M)$ is a closed set by the arbitrary choice of $\{x_i\}$.
From the above argument, we have already seen that
\begin{align}
\mathbf{cvr}(x) \geq \lim_{i \to \infty} \mathbf{cvr}(x_i),
\end{align}
which means that $\mathbf{cvr}$ is an upper-continuous function on $\mathcal{F}_{r}(M)$.
\end{proof}
Clearly, the conclusion in the above corollary is weaker than that in Proposition~\ref{prn:HE07_2}, since here we do not have a rigidity property like Proposition~\ref{prn:SC17_3}. However, even if
$\displaystyle \mathbf{cvr}(x)>\lim_{i \to \infty} \mathbf{cvr}(x_i)$, the local Harnack inequality of $\mathbf{cvr}$ guarantees that $\displaystyle \mathbf{cvr}(x) < K \lim_{i \to \infty} \mathbf{cvr}(x_i)$. So $\mathbf{cvr}$ is better than
general semi-continuous function. For example, in the decomposition $M=\mathcal{F}_r \cup \mathcal{D}_r$,
every point $y \in \partial \mathcal{F}_r$ satisfies $r \leq \mathbf{cvr}(y) \leq Kr$. In many situations, it is convenient to just regard $K=1$, i.e., $\mathbf{cvr}$ being continuous, without affecting the effectiveness of the argument.
\subsection{K\"ahler manifolds with canonical radius bounded from below}
Similar to the traditional theory, volume convergence is very important. However, in the current situation, the volume convergence can be proved in a much easier way.
\begin{proposition}[\textbf{Volume convergence}]
Suppose $(M_i,g_i,J_i)$ is a sequence of K\"ahler manifolds satisfying $\mathbf{cr}(M_i) \geq r_0$. Then
we have \begin{align*}
(M_i, x_i, g_i) \longright{G.H.} (\bar{M}, \bar{x}, \bar{g}).
\end{align*}
Moreover, the volume (2n-dimensional Gromov-Hausdorff measure) is continuous under this convergence, i.e., for every fixed $\rho_0>0$, we have
\begin{align*}
|B(\bar{x}, \rho_0)|= \lim_{i \to \infty} |B(x_i,\rho_0)|.
\end{align*}
\label{prn:SC12_1}
\end{proposition}
\begin{proof}
The existence of the Gromov-Hausdorff limit space follows from the volume doubling property and the standard ball-packing argument. Fix $r<<\rho_0$, then it follows from the definition of $\mathcal{F}_r$ that the convergence
on $B(x_i, \rho_0) \cap \mathcal{F}_r$ can be improved to $C^4$-topology. Then the volume converges trivially on this part. On the other hand, the volume of $B(x_i, \rho_0) \cap \mathcal{D}_r$ is bounded by $C r^{2p_0}$, which tends to
zero as $r \to 0$. So the volume convergence of geodesic balls $B(x_i,\rho_0)$ follows from the combination of
the two factors mentioned above.
\end{proof}
Now we are able to show the weak compactness theorem.
\begin{theorem}[\textbf{Rough weak compactness}]
Same condition as Proposition~\ref{prn:SC12_1}.
Denote $\mathcal{R} \subset \bar{M}$ as the set of regular points, i.e., the points with some small neighborhoods which have $C^4$-Riemannian manifolds structure.
Denote $\mathcal{S} \subset \bar{M}$ be the set of singular points, i.e., the points which are not regular.
Then we have the regular-singular decomposition $\bar{M}=\mathcal{R} \cup \mathcal{S}$ with the following properties.
\begin{itemize}
\item The regular part $\mathcal{R}$ is an open, path connected $C^4$-Riemannian manifold.
Furthermore, for every two points $x,y \in \mathcal{R}$, there exists a curve $\gamma$ connecting $x,y$ satisfying
\begin{align}
\gamma \subset \mathcal{R}, \quad |\gamma| \leq 3d(x,y). \label{eqn:HE11_1}
\end{align}
\item The singular part $\mathcal{S}$ satisfies the Minkowski dimension estimate
\begin{align}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-2p_0.
\label{eqn:SB13_4}
\end{align}
\end{itemize}
\label{thm:HE11_1}
\end{theorem}
\begin{proof}
Clearly, the pointed-Gromov-Hausdorff limit of $(M_i, x_i, g_i)$ exists by the volume ratio estimates.
Let $\left( \bar{M}, \bar{x}, \bar{g} \right)$ be
the limit space. For each $r \leq r_0$, define
\begin{align}
&\mathcal{R}_{r} \triangleq \left\{ \bar{y} \in \bar{M}
\left| \textrm{There exists} \; y_i \in M_i \; \textrm{such that} \; y_i \to \bar{y} \; \textrm{and} \; \liminf_{i \to \infty} \mathbf{cvr}(y_i) \geq r \right. \right\}. \label{eqn:HA11_3}\\
&\mathcal{S}_{r} \triangleq \left( \mathcal{R}_{r} \right)^{c}. \label{eqn:HA11_4}
\end{align}
It is easy to see that $\mathcal{R}_{r}$ is a closed subset of $\bar{M}$. Define
\begin{align*}
\mathcal{R}' \triangleq \bigcup_{0<r\leq r_0} \mathcal{R}_{r}.
\end{align*}
Recall that the regular set $\mathcal{R} \subset \bar{M}$ is defined as the collection of points which have small neighborhoods with manifolds structure. We will show that $\mathcal{R}'$ is nothing but $\mathcal{R}$, i.e., $\displaystyle \mathcal{R} = \bigcup_{0<r\leq r_0} \mathcal{R}_{r}$.
Actually, by regularity estimate property of canonical radius, for every fixed $r \in (0,r_0)$, every point
$\bar{y} \in \mathcal{R}_{r}$, we see that the convergence to $B(\bar{y}, \frac{1}{3}c_a r)$ can be improved to be in the $C^{4}$-topology. Clearly, $B(\bar{y}, \frac{1}{3}c_a r)$ has a manifold structure. So $\mathcal{R}_r \subset \mathcal{R}$. Let $r \to 0$, we have $\displaystyle \bigcup_{0<r\leq r_0} \mathcal{R}_{r} \subset \mathcal{R}$.
On the other hand, suppose $\bar{y} \in \mathcal{R}$. Then there is a ball $B(\bar{y},r)$ with a manifold structure.
By shrinking $r$ if necessary, we can assume that the volume ratio of this ball is very close to the Euclidean one.
Note that the volume ($2n$-dimensional Hausdorff measure) converges when $(M_i, x_i, g_i)$ converges to $(\bar{M}, \bar{x}, \bar{g})$.
Suppose $y_i \to \bar{y}, y_i \in M_i$. Then we have $\omega_{2n}^{-1}r^{-2n}|B(y_i,r)|>1-\delta_0$ for large $i$. By definition, this means that
$\mathbf{cvr}(y_i,0) \geq r$. It follows from the regularity estimates that
$\displaystyle \bar{y} \in \mathcal{R}_{r} \subset \mathcal{R}'$.
By the arbitrary choice of $\bar{y}$, we obtain $\mathcal{R} \subset \mathcal{R}'$.
So we finish the proof of
\begin{align}
\mathcal{R}=\mathcal{R}'=\bigcup_{0<r\leq r_0} \mathcal{R}_{r}.
\label{eqn:SC06_3}
\end{align}
Fix $r<r_0$. Let $\rho_0=r_0$ and take limit of (\ref{eqn:SC02_2}), we obtain
\begin{align}
|B(\bar{y}, r_0) \cap \mathcal{S}_{r}| \leq 4\mathbf{E}r_0^{2n-2p_0} r^{2p_0}, \label{eqn:B10_1}
\end{align}
for every $\bar{y} \in \bar{M}$. Suppose $y \in \mathcal{R}_r \subset \bar{M}$.
The regularity estimate property of canonical radius yields that every point in $B(y, \frac{1}{4}c_a r)$ is regular.
So $d(y,\mathcal{S})\geq \frac{1}{4}c_a r$.
It follows that
\begin{align*}
\mathcal{R}_{r} \subset \left\{x \in \bar{M} \left| d(x, \mathcal{S}) \geq \frac{1}{4}c_a r \right.\right\}
\Leftrightarrow \mathcal{S}_{r} \supset \left\{x \in \bar{M} \left| d(x, \mathcal{S}) <\frac{1}{4}c_a r \right. \right\}.
\end{align*}
Therefore, we have
\begin{align}
\{x \in \bar{M} | d(x, \mathcal{S}) <r\} \subset \mathcal{S}_{4c_a^{-1} r}, \label{eqn:SB13_1}
\end{align}
whenever $r$ is very small. Combining (\ref{eqn:B10_1}) and (\ref{eqn:SB13_1}) yields
\begin{align}
\left| B(\bar{y},r_0) \cap \left\{x \in \bar{M} | d(x, \mathcal{S}) <r \right\} \right| \leq 4^{2p_0+1}\mathbf{E}c_a^{-2p_0} r_0^{2n-2p_0} r^{2p_0}=C r^{2p_0}.
\label{eqn:SB13_2}
\end{align}
Since (\ref{eqn:SB13_2}) holds for every small $r$ and every $\bar{y} \in \bar{M}$, it yields (\ref{eqn:SB13_4}) directly.
It follows from definition that $\mathcal{R}$ is an open $C^4$-manifold. The path connectedness of $\mathcal{R}$ clearly follows from (\ref{eqn:HE11_1}).
Now we proceed to show (\ref{eqn:HE11_1}). Fix $x,y \in \mathcal{R}$, let $r=\sup\{\rho | x \in \mathcal{R}_{\rho}, y \in \mathcal{R}_{\rho}\}$.
Clearly, $r>0$.
So we can choose sequence $x_i, y_i \in \mathcal{F}_{\frac{r}{2}}(M_i)$ such that $x_i \to x$, $y_i \to y$.
Let $\gamma_i$ be a curve connecting $x_i, y_i$ constructed by the method described in Proposition~\ref{prn:SB27_1}.
Clearly, $\gamma_i \subset \mathcal{F}_{\frac{1}{4}\epsilon_b r}$ and $|\gamma_i|<3d(x_i,y_i)$. Note that the convergence of
$\mathcal{F}_{\frac{1}{4}\epsilon_b r}$ to its limit set is in $C^{4}$-topology. Clearly, the limit curve $\gamma$ satisfies (\ref{eqn:HE11_1}).
\end{proof}
\begin{remark}
The definition of regular points in Theorem~\ref{thm:HE11_1} is stronger than the classical one, i.e., a point is regular if and only if every tangent space at this point is
isometric to $\ensuremath{\mathbb{C}}^{n}$. Therefore, some regular points in the classical definition may be singular in our definition.
Of course, at last, our definition coincides with the classical one after we set up sufficient estimates(c.f. Remark~\ref{rmk:HA07_1}).
\label{rmk:HA01_1}
\end{remark}
The properties of the limit space $\bar{M}$ in Theorem~\ref{thm:HE11_1} are not good enough. For example, we do not know if every tangent space is a metric cone, we do not know if $\mathcal{R}$
is convex. In general, one should not expect these to hold. However, if $(M_i, g_i, J_i)$ is a blowup sequence from given K\"ahler Ricci flow solutions with proper geometric bounds, we shall show that $\bar{M}$ do have the mentioned good properties.
\section{Polarized canonical radius}
In this section, we shall improve the regularity of the limit pace $\bar{M}$ in Theorem~\ref{thm:HE11_1}, under the help of K\"ahler geometry and the Ricci flow.
The Ricci flow has reduced volume and local $W$-functional monotonicity, discovered by Perelman. These monotonicities will be used to show
that each tangent space is a metric cone, and the regular part $\mathcal{R}$ is weakly convex, under some natural geometric conditions. However, the weak-compactness we developed in last section only deals with the metric structure. On $\bar{M}$, we cannot see a Ricci flow structure. In order to make use of the intrinsic monotonicity of the Ricci flow, we need a weak compactness of Ricci flows, not just the weak compactness of time slices. However, along the Ricci flow, the metric at different time slices cannot be compared obviously if no estimate of Ricci curvature is known. This is one of the fundamental difficulty to develop the weak compactness theory of the Ricci flows.
We overcome this difficulty by taking advantage of the rigidity of K\"ahler geometry.
\subsection{A rough long-time pseudolocality theorem for polarized K\"ahler Ricci flow}
Suppose $\mathcal{LM}=\left\{ (M^n, g(t), J, L, h(t)), t \ \in I \subset \ensuremath{\mathbb{R}} \right\}$ a polarized K\"ahler Ricci flow.
Let $\mathbf{b}$ be the Bergman function with respect to $\omega(t)$ and $h(t)$, i.e.,
\begin{align*}
\mathbf{b}(x,t) =\log \sum_{k=0}^{N} \norm{S_k}{h(t)}^2, \quad \int_{M} \langle S_k, S_l\rangle \omega(t)^n = \delta_{kl},
\end{align*}
where $N=\dim (H^0(L))-1$, $\{S_k\}_{k=0}^{N}$ are holomorphic sections of $L$.
By pulling back the Fubini-Study metric through the natural holomorphic embedding, we have
\begin{align*}
\tilde{\omega}= \iota^*(\omega_{FS})= \omega+ \sqrt{-1} \partial \bar{\partial} \mathbf{b}.
\end{align*}
Let $\omega_0=\omega(0), \mathbf{b}_0=\mathbf{b}(0)$. Then
\begin{align*}
\omega(t)=\omega_0 + \sqrt{-1}\partial \bar{\partial} \varphi,
\quad
\tilde{\omega}(t)=\omega_0 + \sqrt{-1}\partial \bar{\partial} (\varphi + \mathbf{b}).
\end{align*}
Clearly, $\varphi(0)=0$.
In this section, we focus on polarized K\"ahler Ricci flow $\mathcal{LM}$ satisfying the following estimate
\begin{align}
\norm{\dot{\varphi}}{C^0(M)} + \norm{\mathbf{b}}{C^0(M)} + \norm{R}{C^0(M)} +|\lambda| +C_S(M) \leq B
\label{eqn:SK23_1}
\end{align}
for every time $t \in I$. Let $\mathbf{b}^{(k)}$ be the Bergman function of the line bundle $L^{k}$ with the naturally induced metric.
Then a standard argument implies that
\begin{align}
\norm{\dot{\varphi}}{C^0(M)} + \norm{\mathbf{b}^{(k)}}{C^0(M)} + \norm{R}{C^0(M)} +|\lambda| +C_S(M) \leq B^{(k)}
\label{eqn:SK23_2}
\end{align}
for a constant $B^{(k)}$ depending on $B$ and $k$. Define
\begin{align*}
&\tilde{\omega}^{(k)} \triangleq \frac{1}{k} \left( \iota^{(k)} \right)^* (\omega_{FS}), \\
&F^{(k)} \triangleq \Lambda_{\omega_t} \tilde{\omega}_{0}^{(k)}=n- \Delta \left(\varphi-\mathbf{b}_0^{(k)} \right).
\end{align*}
In this section, the existence time of the polarized K\"ahler Ricci flow is always infinity, i.e., $I=[0,\infty)$.
\begin{lemma}[\textbf{Integral bound of trace}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}).
Suppose $u$ is a positive, backward heat equation solution, i.e.,
\begin{align*}
\square^* u=(-\partial_{t}-\Delta +R-\lambda n) u=0,
\end{align*}
and $\int_{M} u dv \equiv 1$. Then for every $t_0>0$, we have
\begin{align}
\int_0^{t_0} \int_{M} F^{(k)}u dv dt \leq \left(n+2B^{(k)} \right)t_0 + 2B^{(k)}.
\label{eqn:SK07_2}
\end{align}
\label{lma:I29_1}
\end{lemma}
\begin{proof}
For simplicity of notation, we only give a proof of the case $k=1$ and denote $F=F^{(1)}$. Note that $\mathbf{b}=\mathbf{b}^{(1)}, B=B^{(1)}$.
The proof of general $k$ follows verbatim.
Direct calculation shows that
\begin{align*}
\int_0^{t_0} \int_{M} Fu dv &=nt_0- \int_{0}^{t_0} \int_{M} \{\Delta (\varphi-\mathbf{b}_0)\} u dv\\
&=nt_0-\int_0^{t_0}\int_{M} (\varphi-\mathbf{b}_0) (\Delta u) dv\\
&=nt_0+ \int_{0}^{t_0} \int_{M} (\varphi-\mathbf{b}_0) \left( \dot{u}-Ru +\lambda n u \right) dv\\
&=nt_0 + \int_{0}^{t_0} \left[ \frac{d}{dt}\left( \int_{M} (\varphi-\mathbf{b}_0) u dv \right) - \int_{M}\dot{\varphi} u dv\right]dt\\
&=nt_0+ \left. \int_{M} (\varphi-\mathbf{b}_0) u dv \right|_0^{t_0} - \int_0^{t_0} \int_{M} \dot{\varphi} u dv dt.
\end{align*}
Note $|\varphi| \leq Bt_0$ at time $t_0$, then (\ref{eqn:SK07_2}) follows from the above inequality and (\ref{eqn:SK23_2}).
\end{proof}
We shall proceed to improve the integral estimate (\ref{eqn:SK07_2}) of $F^{(k)}$ to pointwise estimate, under local geometry bounds.
Before we go into details, let us first fix some notations.
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow solution satisfying (\ref{eqn:SK23_1}), $x_0 \in M$.
In this subsection, we shall always assume
\begin{align}
\Omega \triangleq B_{g(0)}(x_0,r_0), \quad \Omega' \triangleq B_{g(0)}(x_0, (1-\delta)r_0),
\quad \Omega'' \triangleq B_{g(0)}(x_0, (1-2\delta)r_0).
\label{eqn:SL26_3}
\end{align}
See Figure~\ref{figure:fourballs} for intuition. Then we define
\begin{align}
w_0 \triangleq \phi \left(\frac{2(d-1+2\delta)}{\delta} \right),
\label{eqn:SL26_4}
\end{align}
where $d=d_{g(0)}(x_0, \cdot)$, $\phi$ is a cutoff function, which equals one on $(-\infty, 1]$, decreases to $0$ on $(1,2)$. Moreover, $(\phi')^2 \leq 10\phi$.
Note that such $\phi$ exists by considering the behavior of $e^{-\frac{1}{s}}$ around $s=0$.
Clearly, $w_0$ satisfies
\begin{align}
\begin{cases}
&|\nabla w_0|^2 \leq \frac{40}{\delta^2}w_0, \\
&w_0 \equiv 1, \quad \textrm{on} \quad \Omega'', \\
&w_0 \equiv 0, \quad \textrm{on} \quad (\Omega')^c.
\end{cases}
\label{eqn:SL26_5}
\end{align}
\begin{lemma}[\textbf{Pointwise bound of trace}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$, $\Omega'$ is defined by (\ref{eqn:SL26_3}).
Suppose $\frac{1}{2} \omega_0 \leq \tilde{\omega}_0^{(k)} \leq 2 \omega_0$ on $\Omega'$.
Let $w$ be a solution of heat equation $\square w=\left( \frac{\partial}{\partial t} -\Delta \right)w=0$, initiating from a cutoff function $w_0$
satisfying (\ref{eqn:SL26_5}). Then for every $t_0>0$ and $y_0 \in M$, we have
\begin{align}
F^{(k)}(y_0,t_0) w(y_0, t_0) \leq C \label{eqn:SK23_3}
\end{align}
where $C=C(B,k,\delta,t_0)$.
\label{lma:J02_1}
\end{lemma}
\begin{proof}
For simplicity of notation, we only give a proof for the case $k=1$ and denote $F=F^{(1)}$, $B=B^{(1)}$, $H=\frac{40}{\delta^2}$.
The proof of general $k$ follows verbatim.
Note that $0 \leq w_0 \leq 1$, since $w$ is the heat solution, it follows from maximum principle that $0 \leq w \leq 1$.
On the other hand, according to the choice of $w_0$, we have $|\nabla w|^2 -Hw \leq 0$ at the initial time. Direct calculation implies that
\begin{align*}
\square \left\{ e^{\lambda t}\left(|\nabla w|^2 -Hw\right) \right\}
=-e^{\lambda t}\left\{|\nabla \nabla w|^2 +|\nabla \bar{\nabla} w|^2 +Hw\right\} \leq 0.
\end{align*}
Therefore, $|\nabla w|^2 -Hw \leq 0$ is preserved along the flow by maximum principle.
In other words, we always have
\begin{align*}
w|\nabla \log w|^2 \leq H, \quad 0 \leq w \leq 1
\end{align*}
on the space-time $M \times [0,\infty)$.
In light of parabolic Schwartz lemma(c.f.~\cite{SongWeinnotes} and references therein), we obtain
\begin{align*}
\square \log F \leq B F-\lambda.
\end{align*}
Note that
\begin{align*}
\omega(t)=\omega_0+\sqrt{-1} \partial \bar{\partial} \varphi=\tilde{\omega}_0+\sqrt{-1} \partial \bar{\partial} (\varphi-\mathbf{b}_0)
=\tilde{\omega}_0+\sqrt{-1} \partial \bar{\partial}\tilde{\varphi},
\end{align*}
where we denote $\varphi-\mathbf{b}_0$ by $\tilde{\varphi}$ for simplicity of notation. It is obvious that $\dot{\tilde{\varphi}}=\dot{\varphi}$.
Direct calculation shows that
\begin{align*}
&\square \tilde{\varphi}=\dot{\varphi}-\Delta \tilde{\varphi}= F-n+\dot{\varphi}, \\
&\square (\log F - B\tilde{\varphi})\leq B(n-\dot{\varphi})-\lambda \leq B(n+\norm{\dot{\varphi}}{C^0(\mathcal{M})})+|\lambda|
\leq C.
\end{align*}
Let $u$ be the solution of $\square^*u=0$, starting from a $\delta$-function from $(y_0, t_0)$. Then we calculate
\begin{align*}
\frac{d}{dt} \int_{M} Fe^{-B\tilde{\varphi}} wu dv
&=\int_{M} \square(Fe^{-B\tilde{\varphi}}w) u dv - \int_{M} Fe^{-B\tilde{\varphi}}w \square^*u dv\\
&=\int_{M} \square(Fe^{-B\tilde{\varphi}}w) u dv\\
&=\int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}w)-|\nabla \log (Fe^{-B\tilde{\varphi}}w)|^2 \right\} udv\\
&\leq \int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}w) \right\} udv\\
&=\int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}) + \square \log w \right\} udv\\
&=\int_{M} Fe^{-B\tilde{\varphi}}w \left\{\square \log (Fe^{-B\tilde{\varphi}}) + |\nabla \log w|^2 \right\} udv\\
&\leq C \int_{M} Fe^{-B\tilde{\varphi}} w u dv + H \int_{M} Fe^{-B\tilde{\varphi}} u dv.
\end{align*}
It follows that
\begin{align*}
\frac{d}{dt} \left\{e^{-Ct} \int_{M} Fe^{-B\tilde{\varphi}} w u dv \right\} \leq H e^{-Ct} \int_{M} Fu dv.
\end{align*}
Integrate the above inequality and apply Lemma~\ref{lma:I29_1}, we have
\begin{align*}
e^{-Ct_0} F(y_0,t_0)w(y_0, t_0) e^{-B\tilde{\varphi}(y_0, t_0)} &\leq \left. \int_{M} Fe^{-B\tilde{\varphi}}wu dv \right|_{t=0} + H \int_0^{t_0} \int_{M} Fu dv dt\\
&\leq C \left. \int_{\Omega'} Fu dv \right|_{t=0}+C\\
&\leq C \left. \int_{\Omega'} u dv \right|_{t=0} + C \leq C.
\end{align*}
Therefore, (\ref{eqn:SK23_3}) follows directly from the above inequality.
\end{proof}
\begin{figure}
\begin{center}
\psfrag{t0}[c][c]{$t=0$}
\psfrag{t1}[c][c]{$t=t_0$}
\psfrag{x0}[c][c]{$x_0$}
\psfrag{A1}[c][c]{$(M,x_0,g(t_0))$}
\psfrag{A2}[c][c]{$(M,x_0,g(0))$}
\psfrag{B1}[c][c]{$\Omega=\textrm{blue}$}
\psfrag{B2}[c][c]{$\Omega'=\textrm{green}$}
\psfrag{B3}[c][c]{$\Omega''=\textrm{red}$}
\psfrag{B4}[c][c]{$B_{g(t_0)}(x_0,r)=\textrm{black}$}
\includegraphics[width=0.5 \columnwidth]{fourballs}
\caption{Different domains}
\label{figure:fourballs}
\end{center}
\end{figure}
\begin{lemma}[\textbf{Lower bound of heat solution}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$,
notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}).
Suppose $\Omega'' \subset B_{g(t)}(x_0, r)$ for some $t>0$ and $r>0$. Then in the geodesic ball $B_{g(t)}(x_0, r)$, we have
\begin{align*}
w(y,t)>c
\end{align*}
for some constant $c=c(n,B,k,\delta,r_0,r,t)$.
\label{lma:SK23_1}
\end{lemma}
\begin{proof}
By the construction of $w_0$ and maximum principle, it is clear that $0 <w \leq 1$ when $t>0$.
Let $P$ be the heat kernel function, then we can write
\begin{align}
w(x_0,t)=\int_M P(x_0,t; y,0) w_0(y) dv_y \geq \int_{\Omega''} P(x_0,t;y,0)w_0(y)dv_y.
\label{eqn:SK23_4}
\end{align}
In light of the Sobolev constant bound and scalar curvature bound, one has the on-diagonal bound
\begin{align*}
\frac{1}{C} t^{-n} \leq P(x,t;x,0) \leq C t^{-n},
\end{align*}
which combined with the gradient estimate of heat equation(c.f. Theorem 3.3 of~\cite{Zhq2}) implies that
\begin{align*}
P(x,t;y,0) \geq \frac{1}{C} t^{-n}
\end{align*}
where $C=C(B,d_{g(t)}(x,y))$. Plugging this estimate into (\ref{eqn:SK23_4}) implies that
\begin{align*}
w(x_0,t) \geq \frac{|\Omega''|}{Ct^n}.
\end{align*}
Note that $C_S$ bound forces $|\Omega''|$ is bounded from below. Since $0<w\leq 1$,
then (\ref{eqn:SK23_4}) follows from the above inequality and the gradient estimate of heat equation.
\end{proof}
The following two lemmas show that K\"ahler geometry is much more rigid than Riemannian geometry.
\begin{lemma}[\textbf{Fubini-Study approximation}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$,
notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}).
Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$. Then there exists an integer $k=k(B,r_0, \delta)$ such that
\begin{align}
\frac12 \omega_0 \leq \tilde{\omega}_0^{(k)} \leq 2\omega_0 \label{eqn:SL29_3}
\end{align}
on $\Omega'$.
\label{lma:SK25_1}
\end{lemma}
\begin{proof}
This follows essentially from the peak section method of Tian(c.f.~\cite{Tian90JDG},\cite{Lu}). We give a proof here for the convenience of the readers.
Fix arbitrary $x \in \Omega'$, $V \in T_x^{(1,0)}M$ with unit norm. In order to prove (\ref{eqn:SL29_3}), it suffices to show that
\begin{align}
\frac{1}{2} \leq \tilde{\omega}(V,JV) \leq 2. \label{eqn:SL29_1}
\end{align}
Around $x$, we can always choose a normal coordinate (K-coordinate, c.f.~\cite{Lu}) chart around $x$ such that
\begin{align*}
V=\frac{\partial}{\partial z_1}, \quad
g_{i\bar{j}}(x)=\delta_{i\bar{j}}, \quad
\frac{\partial^{p_1+p_2+\cdots +p_n}}{\partial z_1^{p_1} \partial z_2^{p_2} \cdots \partial z_n^{p_n}} g_{i\bar{j}}(x)=0
\end{align*}
for any nonnegative integers $p_1,p_2, \cdots, p_n$ with $p=p_1+p_2+p_3+\cdots p_n>0$. Moreover,
there exists a
local holomorphic frame $e_L$ of $L$ around $x$ such that the local representation $a$ of the Hermitian metric $h$ has the
properties
\begin{align*}
a(x)=1, \quad \frac{\partial^{p_1+p_2+\cdots +p_n}}{\partial z_1^{p_1}\partial z_2^{p_2} \cdots \partial z_n^{p_n}}a(x)=0
\end{align*}
for any nonnegative integers $p_1,p_2, \cdots, p_n$ with $p=p_1+p_2+p_3+\cdots p_n>0$.
Suppose $\{S_0^{k}, \cdots S_{N_k}^{k}\}$ is an orthonormal basis of $H^0(M, L^k)$,
where $N_k=\dim_{\ensuremath{\mathbb{C}}} H^0(M,L^k)-1$. Around $x$, we can write
\begin{align*}
S_0^{k}=f_0^{k} e_L, \cdots, S_{N_k}^{k}=f_{N_k}^{k} e_L.
\end{align*}
Rotate basis if necessary(c.f.~\cite{Tian90JDG}), we can assume
\begin{align*}
&f_i^{k}(x)=0, \quad \forall \; i \geq 1, \\
&\frac{\partial f_i^k}{\partial z_j}(x)=0, \quad \forall \; i \geq j+1.
\end{align*}
Recall that
\begin{align*}
\tilde{\omega}^{(k)}&=\omega_0 + \frac{1}{k} \sqrt{-1} \partial \bar{\partial} \log \sum_{j=0}^{N_k} \norm{S_j^{k}}{}^2
=\frac{1}{k} \sqrt{-1} \partial \bar{\partial} \log \sum_{j=0}^{N_k} |f_j^k|^2.
\end{align*}
So we have
\begin{align}
\tilde{\omega}^{(k)}(V,JV)
=\frac{1}{k} \frac{\partial^2 \log \sum_{j=0}^{N_k} |f_j^k|^2}{\partial z_1 \bar{\partial} z_1}
=\frac{|\frac{\partial f_1^k}{\partial z_1}|^2}{k|f_0^k|^2}.
\label{eqn:SL29_2}
\end{align}
Because of (\ref{eqn:SL29_1}) and (\ref{eqn:SL29_2}), the problem boils down to a precise estimate of
$\frac{\partial f_1^k}{\partial z_1}$ and $f_0^k$.
As pointed out by Tian in \cite{Tian90JDG}, the peak section method is local in nature.
The global information of the underlying manifold is only used in the step of H\"{o}rmander's estimate. However, in our case, we have
\begin{align*}
&\sqrt{-1} \partial \bar{\partial} \dot{\varphi}+Ric = \lambda g, \quad |\dot{\varphi}| + |\lambda| \leq B.
\end{align*}
Due to the uniformly bounded geometry (up to $C^2$-norm of $g$) inside $\Omega'$ and the uniform bound of
$\sqrt{-1} \partial \bar{\partial} \dot{\varphi}+Ric$ on the whole manifold $M$, Lemma 1.2 of~\cite{Tian90JDG}
follows directly and can be written as follows.
\textit{ For an $n$-tuple of integers $(p_1,p_2, \cdots, p_n) \in \ensuremath{\mathbb{Z}}_{+}^n$ and an integer $p'>p=p_1+p_2+\cdots+p_n$, there
exists an $k_0=k_0(n,B,r_0,\delta)$ such that for $k>k_0$, there is a unit norm holomorphic section $S \in H^0(M, L^k)$ satisfying}
\begin{align*}
\int_{M \backslash \{|z|^2 \leq \frac{(\log k)^2}{k}\}} \norm{S}{}^2 dv \leq \frac{1}{k^{2p'}}.
\end{align*}
Then the same argument as in~\cite{Tian90JDG} implies that(c.f. Lemma 3.2 of~\cite{Tian90JDG})
\begin{align*}
&\left|f_0^k(x)-\sqrt{\frac{(n+k)!}{k!}} \left\{1+\frac{1}{2(k+n+1)!} (R(x)-n^2-n) \right\} \right|<\frac{C}{k^2},\\
&\left|\frac{\partial f_1^k}{\partial z_1}(x)-\sqrt{\frac{(n+k+1)!}{k!}}\left\{ 1+\frac{1}{2(k+n+1)}(R(x)-n^2-3n-2)\right\} \right| <\frac{C}{k^2},
\end{align*}
for some $C=C(n,B,r_0,\delta)$. Here $R$ is the complex scalar curvature. Plug the above estimate into (\ref{eqn:SL29_2}),
we obtain (\ref{eqn:SL29_1}), whenever $k$ is larger than a big constant, which depends only on $n,B,r_0,\delta$.
\end{proof}
\begin{lemma}[\textbf{Liouville type theorem}]
Every complete K\"ahler Ricci flat metric $\tilde{g}$ on $\ensuremath{\mathbb{C}}^n$ must be a Euclidean metric if there is a constant $C$ such that
\begin{align}
\frac{1}{C} \delta_{i\bar{j}}(z) \leq \tilde{g}_{i\bar{j}}(z) \leq C \delta_{i\bar{j}}(z), \quad \forall \; z \in \ensuremath{\mathbb{C}}^n.
\label{eqn:SL26_1}
\end{align}
\label{lma:HE11_1}
\end{lemma}
\begin{proof}
The original proof of this lemma goes back to the famous paper of E. Calabi~\cite{Ca58} and Pogorelov~\cite{Po78} on real Monge Ampere equation. For complex
Monge Ampere equation, this is initially due to Riebesehl-Schulz~\cite{RS84} where higher derivatives are used heavily.
We say a few words here for the convenience of the readers,
using the Schauder estimate of Evans-Krylov.
Actually, it is not difficult to see that the problem boils down to the study of a
global pluri-subharmonic function $u$ in $\ensuremath{\mathbb{C}}^n$ such that
\begin{align}
\begin{cases}
&\det \left({\partial^2 u \over {\partial z_i\partial \bar z_j}}\right) = 1, \\
&C^{-1} (\delta_{i\bar j}) < \left({\partial^2 u \over {\partial z_i\partial \bar z_j}}\right) < C (\delta_{i\bar j}).
\end{cases}
\label{eqn:SL26_2}
\end{align}
In order to show the metric $\tilde{g}$ is Euclidean, it suffices to show that $u$ is a global quadratic polynomial.
Without loss of generality, we may assume that $u(0) = D u(0) = 0$.
For every positive integer $k$, we can define a function $u^{(k)}$ in the unit ball by
\begin{align*}
u^{(k)} (z) = {u(kz)\over {k^2}}.
\end{align*}
Clearly, $u^{(k)}$ satisfies (\ref{eqn:SL26_2}). Note that $\norm{u^{(k)}}{C^2}$ is uniformly bounded,
in the unit ball $B(0,1)$. By standard Evans-Krylov theorem, there exists a
uniform constant $C$ such that
\begin{align*}
[D^2 u^{(k)}]_{C^\alpha(B(0,\frac{1}{2}))} \leq C
\end{align*}
for every $k$. Putting back the scaling factor, the above inequality is equivalent to
\begin{align*}
[D^2 u]_{C^\alpha(B(0, \frac{k}{2}))} \leq C k^{-\alpha},\qquad \forall \quad k=1,2,\cdots.
\end{align*}
Let $k\rightarrow \infty$, we have $[D^2 u]_{C^\alpha(\ensuremath{\mathbb{C}}^n)} = 0$. Therefore, $D^2 u$ is a constant matrix,
$u$ is a quadratic polynomial. So we finish the proof.
\end{proof}
\begin{proposition}[\textbf{Ball containing relationship implies regularity improvement}]
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$,
notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}).
Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$.
Moreover, we assume
\begin{align*}
\Omega'' \subset B_{g(t)}(x_0,r) \subset \Omega'
\end{align*}
for every $0\leq t \leq t_0$.
Then the following estimates hold.
\begin{itemize}
\item In the geodesic ball $B_{g(t)}(x_0,r)$, we have
\begin{align}
\frac{1}{C} \omega_0 \leq \omega_t \leq C \omega_0
\label{eqn:SK23_6}
\end{align}
for some constant $C=C(n,B,k,\delta,r_0,r,t)$.
\item In the geodesic ball $B_{g(t)}(x_0,r-\xi)$, we have
\begin{align}
|Rm| \min\left\{ |t_0-t|, |d_t-r+\xi|^2\right\} \leq C \label{eqn:SK23_5}
\end{align}
for each small $\xi$ and some constant $C=C(n,B,k,\delta,r_0,r,t_0,\xi)$.
\end{itemize}
\label{prn:HE11_2}
\end{proposition}
\begin{proof}
Note that Perelman's strong version of pseudolocality theorem, i.e., Theorem 10.3 of~\cite{Pe1}, can be modified and applied here.
In fact, the almost Euclidean volume ratio condition in that theorem can be replaced by $\kappa$-noncollapsing condition.
Since one has injectivity radius estimate when curvature and volume ratio bounds are available,
thanks to the work of Cheeger, Gromov and Taylor, in~\cite{ChGrTa}.
By shrinking the ball to some fixed smaller size, one can get back the condition of almost Euclidean volume ratio.
Up to a covering argument, we can apply this strong version pseudolocality theorem to show that
$|Rm|$ is uniformly bounded on $\Omega' \times [0,s_0]$ for some positive $s_0=s(n,\kappa,\delta)$.
Then (\ref{eqn:SK23_6}) and (\ref{eqn:SK23_5}) follows trivially. For this reason, we can assume $t_0>s_0$.
We first prove estimate (\ref{eqn:SK23_6}). Due to Fubini-Study metrics' approximation, Lemma~\ref{lma:SK25_1}, it is clear that one can regard
$\omega_0$ and $\tilde{\omega}_0^{(k)}$ as the same metric on $\Omega'$. Therefore, it follows from the combination of Lemma~\ref{lma:J02_1}
and Lemma~\ref{lma:SK23_1} that $F^{(k)}$ is bounded from above, which implies $\Lambda_{\omega_t} \omega_0 \leq C$.
Recall that the volume element $\omega_0^n$ and $\omega_t^n$ are uniformly equivalent, due to the uniform bound of $|R|+|\lambda|$ and the evolution equation
\begin{align*}
\frac{\partial}{\partial t} \log \omega_t^n= n\lambda-R.
\end{align*}
Consequently, (\ref{eqn:SK23_6}) follows.
Then we proceed to prove inequality (\ref{eqn:SK23_5}).
If (\ref{eqn:SK23_5}) does not hold uniformly, let $(y,s)$ be the maximum point of the left hand side of (\ref{eqn:SK23_5}) which is large enough.
Clearly, $s$ is bounded below by the above argument.
Blowup at $(y,s)$ so that the new $|Rm|$ becomes $1$, we obtain a limit smooth un-normalized K\"ahler Ricci flow solution with $R \equiv 0$, which must be
K\"ahler Ricci flat. Also, it is not hard to see that this convergence happens in smooth topology by the choice of $(y,s)$ and the application of Perelman's pseudolocality
and Shi's curvature estimate. By the same scale blowup at $(y,0)$, we obtain nothing but $\ensuremath{\mathbb{C}}^n$. Recall we have (\ref{eqn:SK23_6}), so we obtain a nontrivial K\"ahler Ricci flat metric
$\tilde{g}_{i\bar{j}}$ on $\ensuremath{\mathbb{C}}^n$ such that (\ref{eqn:SL26_1}) holds for some $C$.
This contradicts Lemma~\ref{lma:HE11_1}.
\end{proof}
The rough estimate (\ref{eqn:SK23_6}) and (\ref{eqn:SK23_5}) can be improved when $|R|+|\lambda|$ is very small.
When curvature tensor is bounded in the space-time, one can estimate the Ricci curvature in terms of scalar curvature.
Let $|R|+|\lambda|$ tend to zero, we see that the Ricci curvature tends to zero at the space-time where $|Rm|$ is bounded. By adjusting $\xi$ if necessary, we obtain that in the limit, $B_{g(t)}(x_0,(1-\xi)r)$ is isometric to
$B_{g(0)}(x_0, (1-\xi)r)$ for every $0<t<t_0$. By adjusting $\xi$ and applying Perelman's pseudolocality theorem, we see the convergence at time $t=t_0$ is also smooth since curvature derivatives are all bounded in
the ball $B_{g(t_0)}(x_0,(1-\xi)r)$ at time $t_0$.
\begin{proposition}[\textbf{Volume element derivative small implies ball containing relationship}]
For every $r_0, T$ and small $\xi$, there exists an $\epsilon$ with the following property.
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$, notations fixed by (\ref{eqn:SL26_3}) and (\ref{eqn:SL26_4}). Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$.
If $\displaystyle \sup_{\mathcal{M}} (|R| + |\lambda|) <\epsilon$, then for every $t \in [0,T]$ we have
\begin{align}
&\Omega'' \subset B_{g(t)}\left(x_0, 1-\frac{3}{2}\delta \right) \subset \Omega', \label{eqn:SK25_2} \\
&\left(1- \xi \right)\omega(0) \leq \omega(t) \leq \left(1 + \xi \right) \omega(0), \quad \textrm{in} \; \Omega''.
\label{eqn:SK20_2}
\end{align}
\label{prn:HE11_3}
\end{proposition}
\begin{proof}
If the statement was wrong, we can find a tuple $(n,B,\delta,r_0,T)$ and $\epsilon_i \to 0$ such that the property does not hold for every $\epsilon_i \to 0$.
For each $\epsilon_i$, let $t_i \in [0,T)$ be the critical time such that the properties hold on $[0,t_i]$.
By the strong version of Perelman's pseudolocality theorem, it is clear that $t_i$ is uniformly bounded away from $0$.
At time $t=0$, let $w_{0,i}$ be the function defined like that in (\ref{eqn:SL26_4}), but around base point $x_i \in M_i$.
Let $w_i$ be heat solution starting from $w_{0,i}$.
Note that $t_i$ is uniformly bounded from below.
By virtue of Lemma~\ref{lma:SK23_1}, $w_i(\cdot, t_i)$ is uniformly bounded from below
in the ball $B_{g_i(t_i)}(x_i, 1-\frac{3}{2}\delta)$.
In view of Lemma~\ref{lma:SK25_1}, we can choose $k$ such that
\begin{align*}
\frac12 \omega_{0,i} \leq \tilde{\omega}_{0,i}^{(k)} \leq 2\omega_{0,i}
\end{align*}
on $\Omega_i'=B_{g_i(0)}(x_i, 1-\delta)$. So Proposition~\ref{prn:HE11_2} applies.
Applying Perelman's pseudolocality in strong version, together with the estimate of the type $|Ric| \leq \sqrt{|Rm||R|}$ in ~\cite{Wa2},
it is clear that the open ball $B_{g_i(0)}(x_0, (1-\frac{3}{2}\delta)r_0)$
and the open ball $B_{g(t_i)}(x_0, (1-\frac{3}{2}\delta)r_0)$ are almost isometric to each other.
In other words, let $t$ be the limit of $t_i$, $g$ be the limit flow, $x_0$ be the limit of $x_i$, we then have
that $g$ is a static flow on $B_{g(0)}(x_0,1-\frac{3}{2}\delta)$.
Then it follows from inequality
(\ref{eqn:SK23_6}) that
\begin{align*}
\Omega'' \Subset \overline{B_{g_i(t_i)}\left(x_0, 1-\frac{3}{2}\delta \right)} \Subset \Omega'.
\end{align*}
Moreover, the metric $g_i(t)$ is very close to $g_i(0)$ for every $0 \leq t \leq t_i$, could be as close as possible. Therefore, both (\ref{eqn:SK25_2}) and (\ref{eqn:SK20_2}) holds for every $t \in [0,t_i+\eta_i]$
for some small positive $\eta_i$.
This means $t_i$ is not the critical time, which contradicts our assumption.
\end{proof}
Combine Proposition~\ref{prn:HE11_2} and Proposition~\ref{prn:HE11_3}, by further applying the argument in Proposition~\ref{prn:HE11_2},
the following theorem is clear now.
\begin{theorem}[\textbf{Rough long-time pseudolocality theorem for polarized K\"ahler Ricci flow}]
For every group of numbers $\delta, \xi, r_0,T$, there exists an $\epsilon=\epsilon(n,B,\delta, \xi, r_0,T)$
with the following properties.
Suppose $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK23_1}), $x_0 \in M$.
Suppose $|Rm| \leq r_0^{-2}$ in $\Omega$ at time $t=0$, where $\Omega=B_{g(0)}(x_0,r_0)$.
If $\displaystyle \sup_{\mathcal{M}} (|R| + |\lambda|) <\epsilon$, then for every $t \in [0,T]$ we have
\begin{align}
& B_{g(t)}(x_0, (1-2\delta)r_0) \subset \Omega, \\
& |Rm|(\cdot, t) \leq 2r_0^{-2}, \quad \textrm{in} \; B_{g(t)}(x_0, (1-2\delta)r_0), \\
& \left(1-\xi \right) g(0) \leq g(t) \leq \left(1 +\xi \right) g(0), \quad \textrm{in}
\; B_{g(t)}(x_0, (1-2\delta)r_0).
\end{align}
\label{thm:K8_1}
\end{theorem}
\subsection{Motivation and definition}
In previous subsection, we see that the assumption (\ref{eqn:SK23_1}) helps a lot to relate different time slices of
the K\"ahler Ricci flow solution. However, why is this assumption reasonable? This question will be answered in this
subsection.
\begin{proposition}[\textbf{Weak continuity of Bergman function}]
There is a big integer constant $k_0=k_0(n,A)$ and small constant $\epsilon=\epsilon(n,A)$
with the following property.
Suppose $(M,g,J,L,h)$ is a polarized K\"ahler manifold, taken out from a polarized K\"ahler Ricci flow
in $\mathscr{K}(n,A)$ as a central time slice. In particular, we have
\begin{align}
Osc_{M} \dot{\varphi} + C_S(M) + |\lambda| \leq B,
\label{eqn:SK27_1}
\end{align}
where $B=B(n,A)$.
If $\mathbf{cr}(M) \geq 1$, then
\begin{align}
\sup_{1 \leq k \leq k_0} \mathbf{b}^{(k)}(x) > -k_0
\label{eqn:SK27_2}
\end{align}
whenever $d_{GH}((M,x,g), (\tilde{M}, \tilde{x}, \tilde{g}))<\epsilon$ for some space
$(\tilde{M}, \tilde{x}, \tilde{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{prn:SK27_1}
\end{proposition}
\begin{proof}
Note that the model moduli space $\widetilde{\mathscr{KS}}(n,\kappa)$ has compactness
under the pointed Gromov-Hausdorff topology.
Suppose the statement was wrong, then there is a sequence of polarized K\"ahler manifolds,
whose underlying K\"ahler manifolds converge to
$(\bar{M},\bar{x},\bar{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$,
satisfying the estimate (\ref{eqn:SK27_1}) and violating (\ref{eqn:SK27_2}) for $k_i \to \infty$.
To be explicit, we have
\begin{align}
&(M_i,x_i,g_i) \longright{G.H.} (\bar{M},\bar{x},\bar{g}), \label{eqn:K8_3} \\
& \sup_{1 \leq j \leq k_i} \mathbf{b}^{(j)}(x_i) \to -\infty, \quad k_i \to \infty. \label{eqn:K8_4}
\end{align}
The canonical radius assumption makes sure that the topology of the convergence can be improved to the
$\hat{C}^4$-Cheeger-Gromov topology. Since $\bar{x}$ is a point in the model space,
every tangent space at $\bar{x}$ is a possibly singular Calabi-Yau metric cone. Note that $\dot{\varphi}$ is bounded,
so we can apply the general H\"ormander's estimate(c.f.~\cite{CW4} for this particular case).
Similar to the argument of Sun and Donaldson(c.f.~\cite{DS}), one can find positive integer $q=q(\bar{x})$, and real numbers $r=r(\bar{x})$, $C=C(\bar{x})$
such that
\begin{align*}
\inf_{y \in B(x_i, r)} \mathbf{b}^{(q)}(y) \geq -C.
\end{align*}
In particular, we have
\begin{align*}
\mathbf{b}^{(q)}(x_i) \geq -C, \quad \Rightarrow \quad \sup_{j \leq k_i} \mathbf{b}^{(j)} (x_i) \geq -C,
\end{align*}
which contradicts (\ref{eqn:K8_4}), the assumption.
\end{proof}
Proposition~\ref{prn:SK27_1} means that the Bergman function has a weak continuity under the Cheeger-Gromov
convergence if the limit space is the model space. Inspired by this property, we can define the polarized canonical radius as follows.
\begin{definition}
Suppose $\left( M,g,J,L,h\right)$ is a polarized K\"ahler manifold satisfying (\ref{eqn:SK27_1}), $x \in M$. We say the polarized canonical radius of $x$ is not less than $1$ if
\begin{itemize}
\item $\mathbf{cr}(x) \geq 1$.
\item $\displaystyle \sup_{1 \leq j \leq 2k_0}\mathbf{b}^{(j)}(x) \geq -2k_0$.
\end{itemize}
For every $r=\frac{1}{j}, j \in \ensuremath{\mathbb{Z}}^{+}$, we say the polarized canonical radius of $x$ is not less than $r$ if the rescaled polarized manifold $\left( M, j^2 g, J, L^j, h^j\right)$
has polarized canonical radius at least $1$ at the point $x$.
Fix $x$, let $\mathbf{pcr}(x)$ be the supreme of all the $r$ with the above property and call it
as the polarized canonical radius of $x$.
\label{dfn:SK27_1}
\end{definition}
We can define the polarized canonical radius of a manifold as the infimum of the polarized canonical radii of all points in that manifold.
Similarly, we can define the polarized canonical radius of time slices of a flow.
Note that from the above definition, $\mathbf{pcr}$ is always the reciprocal of a positive integer.
Under this terminology, the continuity of Bergman function implies the following corollary.
\begin{corollary}[\textbf{Weak equivalence of $\mathbf{cr}$ and $\mathbf{pcr}$}]
There is a small constant $\epsilon=\epsilon(n,B,\kappa)$ with the following property.
Suppose $(M,g,J,L,h)$ is a polarized K\"ahler manifold satisfying (\ref{eqn:SK27_1}) and
$\mathbf{cr}(M) \geq 1$. Then
\begin{align}
\mathbf{pcr}(x) \geq 1
\label{eqn:SK27_4}
\end{align}
whenever $d_{GH}((M,x,g), (\tilde{M}, \tilde{x}, \tilde{g}))<\epsilon$ for some space
$(\tilde{M}, \tilde{x}, \tilde{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{cly:SK27_1}
\end{corollary}
\subsection{K\"ahler Ricci flow with lower bound of polarized canonical radius}
Suppose the polarized canonical radius is uniformly bounded from below, then the convergence theory is much better
than that in section 3. This is basically because of the rough long-time pseudolocality theorem, Theorem~\ref{thm:K8_1}.
\begin{proposition}[\textbf{Improving regularity in forward time direction}]
For every $r_0>0$, $r \in (0,r_0)$ and $T_0>0$, there is an $\epsilon=\epsilon(n,A,r_0,r,T_0)$ with the following properties.
If $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK20_1}) and
\begin{align}
\mathbf{pcr}(\mathcal{M}^{t}) \geq r_0, \quad \forall \; t \in [0,T_0],
\label{eqn:SL04_1}
\end{align}
then \begin{align}
\mathcal{F}_{r}(M, 0) \subset \bigcap_{0\leq t \leq T_0} \mathcal{F}_{\frac{r}{K}}(M, t) \label{eqn:SC19_3}
\end{align}
whenever $\displaystyle \sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
Here $K$ is the constant in Proposition~\ref{prn:SC24_1}.
\label{prn:SC09_3}
\end{proposition}
\begin{proof}
It follows directly from Theorem~\ref{thm:K8_1}, the long time pseudolocality theorem for polarized K\"ahler Ricci flow with partial $C^0$-estimate.
\end{proof}
\begin{proposition}[\textbf{Improving regularity in backward time direction}]
For every $r_0>0$, $r \in (0,r_0)$ and $T_0>0$, there is an $\epsilon=\epsilon(n,A,r_0,r,T_0)$ with the following properties.
If $\mathcal{LM}$ is a polarized K\"ahler Ricci flow satisfying (\ref{eqn:SK20_1}) and (\ref{eqn:SL04_1}), then \begin{align}
\bigcup_{0 \leq t \leq T_0} \mathcal{F}_{r}(M, t) \subset \mathcal{F}_{\frac{r}{K}}(M,0) \label{eqn:SK27_6}
\end{align}
whenever $\displaystyle \sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
\label{prn:SK27_3}
\end{proposition}
\begin{proof}
At time $0$, $\mathcal{F}_r(M,0) \subset \mathcal{F}_{\frac{r}{K}}(M,0)$ trivially. Suppose $t_0>0$ is the first time such that
(\ref{eqn:SK27_6}) start to fail. It suffices to show that $t_0 >T_0$ whenever $\epsilon$ is small enough. Otherwise, at time $t_0 \in (0,T_0]$,
we can find a point $x_0 \in (\partial \mathcal{F}_{\frac{r}{K}}(M,0)) \cap (\partial \mathcal{F}_{r}(M,t_0))$. In other words, we have
\begin{align*}
\mathbf{cvr}(x_0,0)=\frac{r}{K}, \quad \mathbf{cvr}(x_0,t_0)=r.
\end{align*}
In particular, we have
\begin{align}
\left|B_{g(0)}\left(x_0,\frac{r}{K} \right) \right|_0 = \left(1-\delta_0 \right) \omega_{2n} \left(\frac{r}{K} \right)^{2n}. \label{eqn:SL16_1}
\end{align}
Let $\xi$ be a small number which will be fixed later. Let $\Omega_{\xi}(x_0,t_0)$ be the subset of unit sphere of tangent space of $T_{x_0}(M, g(t_0))$ such that every geodesic (under metric $g(t_0)$) emanating from $x_0$ along the direction in $\Omega_{\xi}(x_0,t_0)$ does not hit points in $\mathcal{D}_{\xi}(M,0)$ before
distance $\frac{r}{K}$. By canonical radius assumption, $|Rm|_{g(t_0)}$ is uniformly bounded in $B_{g(t_0)}(x_0,\frac{r}{K})$(See Figure~\ref{figure:backpseudo} for intuition).
By long-time pseudolocality theorem (c.f. Proposition~\ref{prn:SK27_3}), $B_{g(t_0)}(x_0,\frac{r}{K^3})$ has empty intersection with $\mathcal{D}_{\xi}(M,0)$
when $\xi<<\frac{r}{K^3}$. Note that every geodesic (emanating from $x_0$) entering $\mathcal{D}_{\xi}(M,0)$ must hit $\partial \mathcal{D}_{\xi}(M,0)$ first,
where $\mathbf{cvr}(\cdot, 0)=\xi$. So every point in $\partial \mathcal{D}_{\xi}(M,0)$ will be uniformly regular at time $t_0$, in light of the long-time pseudolocality.
At time $t_0$, observing from $x_0$, the set which stays behind $\partial \mathcal{D}_{\xi}(M,0)$ must have small measure.
Since $B_{g(t_0)}(x_0,\frac{r}{K})$ has uniformly bounded curvature,
it is clear that $\Omega_{\xi}(x_0,t_0)$ is an almost full measure subset of $S^{2n-1}$. Actually, we have
\begin{align*}
|\Omega_{\xi}(x_0,t_0)| \geq 2n\omega_{2n} \cdot \left(1-C \xi^{2p_0} \right)
\end{align*}
whenever $\epsilon$ is sufficiently small. On the other hand, we see that every geodesic (under metric $g(t_0)$) emanating from $\Omega_{\xi}(x_0,t_0)$ is almost geodesic at time $t=0$ (under metric $g(0)$), when $\epsilon$ small enough. Therefore,
$|B_{g(0)}(x_0,\frac{r}{K})|_{0}$ is almost not less than $|B_{g(t_0)}(x_0,\frac{r}{K})|_{t_0}$. Note that the volume ratio of
$B_{g(t_0)}(x_0,\frac{r}{K})$ is at least $(1-\frac{\delta_0}{100})\omega_{2n}$.
Suppose we choose $\xi$ small (according to $\delta_0$) and $\epsilon$ very small (based on $\xi,\delta_0,A,T_0$), we obtain
\begin{align*}
\left|B_{g(0)}\left(x_0,\frac{r}{K} \right) \right|_0 \geq \left(1-\frac{\delta_0}{2} \right) \omega_{2n} \left(\frac{r}{K} \right)^{2n},
\end{align*}
which contradicts (\ref{eqn:SL16_1}).
\end{proof}
\begin{figure}
\begin{center}
\psfrag{0}[c][c]{$0$}
\psfrag{t}[c][c]{$t$}
\psfrag{x0}[c][c]{$x_0$}
\psfrag{M1}[c][c]{$(M,x_0,g(t_0))$}
\psfrag{M2}[c][c]{$(M,x_0,g(0))$}
\psfrag{B1}[c][c]{$yellow: B_{g(t_0)}(x_0,r)$}
\psfrag{B2}[c][c]{$blue: B_{g(t_0)}(x_0,\frac{r}{K})$}
\psfrag{B3}[c][c]{$red: B_{g(t_0)}(x_0,\frac{r}{K^3})$}
\psfrag{B4}[c][c]{$green: B_{g(0)}(x_0,r)$}
\psfrag{B5}[c][c]{$black: \mathcal{D}_{\xi}(M,0)$}
\includegraphics[width=0.5 \columnwidth]{backpseudo}
\caption{Find a geodesic ball with almost Euclidean volume ratio}
\label{figure:backpseudo}
\end{center}
\end{figure}
\begin{definition}
Let $\mathscr{K}(n,A)$ be the collection of polarized K\"ahler Ricci flows satisfying (\ref{eqn:SK20_1}).
For every $r \in (0,1]$, define
\begin{align*}
\mathscr{K}(n,A;r)
\triangleq \left\{\mathcal{LM} \left| \mathcal{LM} \in \mathscr{K}(n,A), \mathbf{pcr}(M \times [-1,1]) \geq r \right. \right\}.
\end{align*}
\label{dfn:SC23_1}
\end{definition}
Clearly, $\mathscr{K}(n,A) \supset \mathscr{K}(n,A;r_2) \supset \mathscr{K}(n,A;r_1)$ whenever $1\geq r_2 >r_1>0$.
Since every polarized K\"ahler Ricci flow $\mathcal{LM} \in \mathscr{K}(n,A)$ has a smooth compact underlying manifold, we see $\mathcal{LM} \in \mathscr{K}(n,A;r)$
for some very small $r$, which depends on $\mathcal{LM}$. Therefore, it is clear that
\begin{align*}
\bigcup_{0<r<1} \mathscr{K}(n,A;r) = \mathscr{K}(n,A).
\end{align*}
Fix $r>0$, we shall first make clear the structure of $\mathscr{K}(n,A;r)$ under the help of polarized canonical radius.
Then we show that the canonical radius can actually been bounded a priori. In other words, there exists a uniform small constant $\hslash$ (Planck scale)
such that
\begin{align*}
\mathscr{K}(n,A) = \mathscr{K}(n,A;\hslash),
\end{align*}
which will be proved in Theorem~\ref{thm:SC28_1}.
\begin{proposition}[\textbf{Limit space-time with static regular part}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies the following properties.
\begin{itemize}
\item $\mathbf{pcr}(M_i \times [-T_i,T_i]) \geq r_0$ for each $i$.
\item $\displaystyle \lim_{i \to \infty} \sup_{\mathcal{M}_i} (|R|+|\lambda|)=0$.
\end{itemize}
Suppose $x_i \in M_i$ and
$\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i,0)>0$, then
\begin{align*}
(M_i,x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g}).
\end{align*}
Moreover, we have
\begin{align*}
(M_i,x_i, g_i(t)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g})
\end{align*}
for every $t \in (-\bar{T}, \bar{T})$, where $\displaystyle \bar{T}=\lim_{i \to \infty} T_i>0$.
In particular, the limit space does not depend on time.
\label{prn:SL04_1}
\end{proposition}
\begin{proof}
It follows from the combination of Proposition~\ref{prn:SC09_3} and Proposition~\ref{prn:SK27_3} that the limit
space does not depend on time. From the definition of canonical radius, the convergence locate in
$\hat{C}^4$-topology for each time. However, this can be improved to $\hat{C}^{\infty}$-topology by using Perelman's
pseudolocality theorem and Shi's estimate.
\end{proof}
In Proposition~\ref{prn:SL04_1},
we show that the limit flow exists and is static in the regular part, whenever we have $|R|+|\lambda| \to 0$.
It is possible that the limit points in the singular part $\mathcal{S}$ are moving as time evolves. However, this possibility
will be ruled out finally(c.f. Proposition~\ref{prn:HE15_1}).
\subsubsection{Tangent structure of the limit space}
In this subsection, we shall show that the tangent space of each point in the limit space has a metric cone structure, provided polarized canonical radius is uniformly bounded below. Basically, the cone structure is induced from
the localized $W$-functional's monotonicity. Up to a parabolic rescaling, we can assume $\lambda=0$ without loss
of generality.
\begin{proposition}[\textbf{Local $W$-functional}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;r_0)$ satisfies
$\sup_{\mathcal{M}_i} (|R|+|\lambda|) \to 0$.
Let $u_i$ be the fundamental solution of the
backward heat equation $\left[-\frac{\partial}{\partial t} - \triangle + R \right] u_i =0$ based at the space-time point $(x_i, 0)$. Then $u_i$ converges to
a limit solution $\bar{u}$ on $\mathcal{R} \times (-1, 0]$, i.e.,
\begin{align*}
\left[-\frac{\partial}{\partial t} - \Delta + R \right] \bar{u}=0.
\end{align*}
Moreover, we have
\begin{align}
\iint_{\mathcal{R} \times (-1, 0]} 2|t| \left|Ric+\nabla \nabla \bar{f} +\frac{\bar{g}}{2t} \right|^2 \bar{u} dv_{\bar{g}} \leq C,
\label{eqn:SC16_4}
\end{align}
where $C=C(n,A)$, $\bar{u}=(4\pi |t|)^{-n} e^{-\bar{f}}$.
\label{prn:SC16_1}
\end{proposition}
\begin{proof}
Fix $r>0$. Choose a point $\bar{y} \in \mathcal{R}_{r}$ and a time $\bar{t}<0$. Without loss of generality, we assume that
there is a sequence of points $(y_i, t_i)$ converging to $(\bar{y}, \bar{t})$. Note that $d_{g_i(0)}(y_i, x_i)$ is uniformly bounded. So by the heat kernel estimate(c.f.~\cite{Zhq2},~\cite{CaHa}), it is clear that $u_i$ is uniformly bounded around $(y_i, t_i)$. Then simple regularity argument from heat equation shows that all derivatives of $u_i$ is
uniformly bounded around $(y_i, t_i)$. Therefore, there is a limit solution $\bar{u}$ around $(\bar{y}, \bar{t})$. By the arbitrary choice of $r,\bar{y}, \bar{t}$. It is clear
that there is a smooth heat solution $\bar{u}$ defined on $\mathcal{R} \times (-1, 0)$.
By Perelman's calculation, for each flow $g_i$, we have
\begin{align}
\int_{-1}^{0}\int_{M_i} 2|t| \left|Ric_{g_i}+\nabla \nabla f_i +\frac{g_i}{2t} \right|^2 u_i dv_{g_i}=-\mu(M_i, g_i(t_i), 1) \leq C,
\label{eqn:SK17_1}
\end{align}
since Sobolev constant is uniformly bounded. By passing to limit, (\ref{eqn:SC16_4}) follows.
\end{proof}
\begin{theorem}[\textbf{Tangent cone structure}]
Suppose $\mathcal{LM}_i$ is a sequence of polarized K\"ahler Ricci flow solutions in $\mathscr{K}(n,A;r_0)$,
$x_i \in M_i$. Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of
$(M_i,x_i,g_i(0))$, $\bar{y}$ be an arbitrary point of $\bar{M}$.
Then every tangent space of $\bar{y}$ is an irreducible metric cone.
\label{thm:SC09_1}
\end{theorem}
\begin{proof}
Suppose $\hat{Y}$ is a tangent space of $\bar{M}$ at the point $\bar{y}$, i.e., there are scales $r_k\to 0$ such that
\begin{align*}
(\hat{Y}, \hat{y}, \hat{g})=\lim_{k \to \infty} (\bar{M}, \bar{y}, \bar{g}_k)
\end{align*}
where $\bar{g}_k= r_k^{-2}\bar{g}$. By taking subsequence if necessary, we can assume $ (\hat{Y}, \hat{y}, \hat{g})$
as the limit space of $(M_{i_k}, y_{i_k}, \tilde{g}_{i_k})$ where $\tilde{g}_{i_k}=r_i^{-2}g_{i_k}(0)$.
Denote $\hat{\mathcal{R}}$ as the regular part of $\hat{Y}$. Then on the space-time
$\hat{\mathcal{R}} \times (-\infty, 0]$, there is
a smooth limit backward heat solution $\hat{u}$. For every compact subset $K \subset \mathcal{R}$ and positive number $H$, it follows from Cheeger-Gromov convergence and
the estimate (\ref{eqn:SK17_1}) that
\begin{align*}
\iint_{K \times [-H, 0]} 2|t| \left| Ric + \nabla \nabla \hat{f} + \frac{\hat{g}}{2t}\right|^2 \hat{u} dv =0,
\end{align*}
which implies
\begin{align*}
\iint_{\hat{\mathcal{R}} \times (-\infty, 0]} 2|t| \left| Ric + \nabla \nabla \hat{f} + \frac{\hat{g}}{2t}\right|^2 \hat{u} dv =0.
\end{align*}
Note that $\hat{\mathcal{R}}$ is Ricci flat. So there is a smooth function $\hat{f}$ defined on $\hat{\mathcal{R}} \times (-\infty, 0]$ such that
\begin{align}
\nabla \nabla \hat{f} + \frac{\hat{g}}{2t} \equiv 0.
\label{eqn:SL13_1}
\end{align}
In particular, $\nabla \nabla \hat{f}= \frac{\hat{g}}{2}$ at time $t=-1$. At this time slice, we can calculate the Lie derivative
\begin{align*}
\mathcal{L}_{\nabla \hat{f}} \hat{g}=2\nabla \nabla \hat{f}= \hat{g}.
\end{align*}
This equation means that $\nabla \hat{f}$ is a conformal Killing vector field, which induces local cone structure around each regular integral curve of $\nabla \hat{f}$.
We shall show that a global cone structure can be obtained due to the high co-dimension of the singular set $\mathcal{S}$.
Suppose $a$ is a positive number, we define $\Omega_a$ to be the collection of points where $\hat{f} \leq a$. Without loss of generality, we assume $\partial \Omega_1 \neq \emptyset$.
Clearly, $\hat{f} \equiv 1$ on $\partial \Omega_1$.
Let $z$ be a regular point of $\partial \Omega_1$, $\gamma$ be a flow line generated by $\nabla{\hat{f}}$ and pass through $z$.
If $\gamma$ does not hit singularity in $\Omega_1 \backslash \Omega_{\frac{1}{2}}$, then it has a neighborhood which is diffeomorphic to
$D \times (\frac{1}{2}, 1]$ for some open set $D \subset \ensuremath{\mathbb{R}}^{2n-1}$.
Moreover, the metric $\hat{g}$ in this neighborhood can be written as $ \hat{g}=d \hat{f} \otimes d \hat{f} + \hat{f} d^2 \theta$.
Note that $\frac{1}{2} \leq \hat{f} \leq 1$, along $\gamma$, the metric at different level surfaces are uniformly comparable. Collect all the flow curves passing through a
point in $\mathcal{S} \cap \Omega_1 \backslash \Omega_{\frac{1}{2}}$ and let $E$ be the union set of all such curves.
Since $\dim_{\mathcal{M}} \mathcal{S}<2n-3$, it is clear that $\dim_{\mathcal{H}}(E)<2n-2$. In particular, $E$ is a measure zero set.
Clearly, $\Omega_1 \backslash (\Omega_{\frac{1}{2}} \cup E)$ has a cone annulus structure. Taking metric completion, we see that
$\Omega_1 \backslash \Omega_{\frac{1}{2}}$ has a cone annulus structure. Replacing $[\frac{1}{2}, 1]$ by $[\frac{1}{2}\epsilon, \epsilon^{-1}]$ for every small, but finite
$\epsilon$, the same argument go through to show that $\Omega_{\epsilon^{-1}} \backslash \Omega_{\frac{1}{2}\epsilon}$ has a cone annulus structure.
Since $\epsilon$ is arbitrary, the closure of $\hat{Y} \backslash \Omega_0$ clearly has a metric cone structure, possibly with more than one ends.
However, the connectedness of $\hat{\mathcal{R}}$ excludes this possibility and guarantees
that the closure of $\hat{Y} \backslash \Omega_0$ is nothing but $\hat{Y}$.
Therefore, $\hat{Y}$ is an irreducible metric cone.
\end{proof}
\subsubsection{Improved estimates in $\mathscr{K}(n,A;r_0)$}
In this subsection, we shall improve the limit space structure by the fact that every tangent space is a metric cone.
For simplicity, we assume $r_0=1$.
\begin{proposition}
Suppose $\mathcal{LM}_i$ is a sequence of polarized K\"ahler Ricci flow solutions in $\mathscr{K}(n,A;1)$,
$x_i \in M_i$. Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i,x_i,g_i(0))$.
Let $\mathcal{S}$ be the singular part of $\bar{M}$. Then
\begin{align}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-2p_0, \quad
\dim_{\mathcal{H}}\mathcal{S} \leq 2n-4. \label{eqn:SL15_1}
\end{align}
\label{prn:SL15_1}
\end{proposition}
\begin{proof}
The Minkowski dimension estimate follows from Theorem~\ref{thm:HE11_1}.
By induction, it is clear that every tangent cone's singularity has an integer Hausdorff dimension. On the other hand, the Minkowski dimension of singularity is
at most $2n-2p_0$. This forces that every tangent cone's singularity has Hausdorff dimension $2n-4$ at most,
which in turn implies $\dim_{\mathcal{H}} \mathcal{S}\leq 2n-4$.
\end{proof}
After we set up the tangent cone structure, we can improve Proposition~\ref{prn:SB27_1}.
\begin{proposition}
Same condition as in Proposition~\ref{prn:SL15_1}.
For every two points $x,y \in \mathcal{R}$ and every small positive number $\epsilon>0$, there exists a rectifiable curve connecting $x,y$ such that
\begin{itemize}
\item $\gamma$ locates in $\mathcal{R}$.
\item $|\gamma| \leq (1+\epsilon)d(x,y)$.
\end{itemize}
\label{prn:SC17_4}
\end{proposition}
\begin{proof}
The proof is very similar to the proof o Proposition~\ref{prn:SB27_1}.
Fix $\epsilon>0$. Fix two points $x,y \in \mathcal{R}$. Let $\beta$ be a shortest geodesic connecting $x,y$. For every point $z \in \beta$, there is a radius $r_z$ such that
every two points $u,v$ in $B(z,r_z)$ can be approximated by two points in $u',v' \in \mathcal{F}_{\epsilon r_z} \cap B(z,r_z)$ which can be connected by a smooth
curve $\alpha$ such that
\begin{align*}
|\alpha| \leq (1+\epsilon) d(u,v).
\end{align*}
Since $\bigcup_{z \in \beta} B(z, \frac{1}{4}r_z)$ is a cover of $\beta$, we can find a finite cover. By deleting interior extra ball, we can make this cover the smallest one, i.e., missing any ball will cause the loss of the cover. Furthermore, we can assume $z_0=x, z_N=y$.
Suppose $\beta \subset \cup_{k=1}^N B\left(z_k, \frac{1}{4}r_{z_k} \right)$. Now replace $z_k$ by $z_k'$
such that $z_k' \in \mathcal{F}_{\epsilon r_{z_k}}$. Clearly, $z_k'$ and $z_{k+1}'$ can be connected by a smooth curve $\gamma_k$ satisfying
\begin{align*}
|\gamma_k| \leq (1+\epsilon) d(z_k, z_{k+1}).
\end{align*}
By concatenating all the curves $\gamma_k$, we obtain the curve $\gamma$ such that
\begin{align*}
|\gamma|=\sum_{k=0}^{N-1} |\gamma_k| \leq (1+\epsilon) \sum_{k=0}^{N} d(z_k, z_{k+1}) = (1+\epsilon) d(x,y).
\end{align*}
\end{proof}
\begin{lemma}
There is an $\epsilon=\epsilon(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A;1)$,
$x,y\in M$ and $r=d_0(x,y)<1$. Suppose $y \in \mathcal{F}_{\frac{\epsilon_b}{2} r}(M,0)$. Then we have
\begin{align}
l((x,0),(y,-r^2))< 100 \label{eqn:SL25_3}
\end{align}
whenever $\sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
\label{lma:SL13_1}
\end{lemma}
\begin{proof}
Let $y_0=y$.
According to the construction in Proposition~\ref{prn:SL24_1}, there exists a point
$y_1 \in \partial B_{g(0)}(x,\frac{r}{2}) \cap \mathcal{F}_{\frac{\epsilon_b r}{4}}(M,0)$ and a curve
$\gamma_1 \subset \mathcal{F}_{\frac{\epsilon_b^2}{8} r}(M,0)$ connecting $y_0,y_1$, with length less than
$\frac{9}{2}r$.
Suppose $|R|+|\lambda|$ is small enough, then
$\displaystyle \gamma_1 \subset \bigcap_{-r^2 \leq t \leq 0} \mathcal{F}_{\frac{\epsilon_b^2 r}{16}}(M,0)$.
So $\gamma_1$ can be lifted as a space-time curve connecting $(y_1, -\frac{r^2}{4})$
and $(y_0, -r^2)$. Reparameterize $\gamma_1$ by $\tau$, after a proper adjustment, we have
\begin{align*}
\int_{\frac{r^2}{4}}^{r^2} \sqrt{\tau} |\dot{\gamma}_1|_{g(-\tau)}^2 d\tau < 100 r.
\end{align*}
Following the same procedure, we can find $\gamma_2$ connecting $y_1$ to
$y_2 \in \partial B_{g(0)}(x,\frac{r}{4}) \cap \mathcal{F}_{\frac{\epsilon_b r}{8}}(M,0)$ with
$\displaystyle \gamma_2 \subset \bigcap_{-\frac{r^2}{4} \leq t \leq 0} \mathcal{F}_{\frac{\epsilon_b^2 r}{32}}(M,0)$.
By a proper reparameterization of $\tau$, we can regard $\gamma_2$ as a space-time curve connecting
$(y_1,-\frac{r^2}{4})$ and $(y_2, -\frac{r^2}{16})$, and it satisfies the estimate
\begin{align*}
\int_{\frac{r^2}{16}}^{\frac{r^2}{4}} \sqrt{\tau} |\dot{\gamma}_2|_{g(-\tau)}^2 d\tau < 100 \cdot \frac{r}{2}.
\end{align*}
Repeating this process, we can find curve $\gamma_k$ connecting $(y_k, -\frac{r^2}{4^k})$
and $(y_{k+1}, -\frac{r^2}{4^{k+1}})$. Concatenating all $\gamma_k$'s together, we obtain a space-time curve
$\gamma$ connecting $(x,0)$ and $(y, -r^2)$ such that
\begin{align*}
\int_0^{r^2} \sqrt{\tau} |\dot{\gamma}|_{g(-\tau)}^2 d\tau < 100 \sum_{k=0}^{\infty} \frac{r}{2^k}=200r.
\end{align*}
It follows that
\begin{align*}
l((x,0), (y,-r^2)) < \frac{200r}{2\sqrt{r^2}}=100.
\end{align*}
\end{proof}
\begin{lemma}
For every group of numbers $0<\xi<\eta<1<H$, there is a big constant $C=C(\eta,H)$
and a small constant $\epsilon=\epsilon(n,A,H,\eta,\xi)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A;1)$, $x\in \mathcal{F}_{\eta}(M,0)$.
Let $\Omega_{\xi}$ be the collection of points $y \in M$ such that every
shortest reduced geodesic $\boldsymbol{\beta}$ connecting $(x,0)$ and $(y,-1)$ satisfies
\begin{align}
\beta \cap \mathcal{D}_{\xi}(M,0) \neq \emptyset.
\label{eqn:SL24_4}
\end{align}
Then
\begin{align}
|B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap \Omega_{\xi}| < C \xi^{2p_0-1}
\label{eqn:SL24_5}
\end{align}
whenever $\sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
\label{lma:SL14_1}
\end{lemma}
\begin{proof}
From the argument in Lemma~\ref{lma:SL13_1}, it is not hard to obtain the following bound
\begin{align}
l((x,0),(z,-1)) < C, \quad \forall \; z \in B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0), \label{eqn:SL24_6}
\end{align}
where $C=C(\eta,H)$.
Suppose $z \in B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0)$, $\boldsymbol{\beta}$ is a shortest reduced geodesic connecting $(x,0)$ and $(z,-1)$. Let $\beta$ be the corresponding space curve.
Note that $x$ and $z$, the two end points of $\beta$, locate outside of $ \mathcal{D}_{\xi}(M,0)$.
Therefore if $z \in \Omega_{\xi}$, then (\ref{eqn:SL24_4}) is satisfied. By continuity, we have
\begin{align*}
\beta \cap \partial \mathcal{F}_{\xi}(M,0) =\beta \cap \partial \mathcal{D}_{\xi}(M,0) \neq \emptyset.
\end{align*}
Let $\bar{\tau}$ be the smallest $\tau$ such that $\boldsymbol{\beta}(\tau) \in \partial \mathcal{F}_{\xi}(M,0)$.
In light of (\ref{eqn:SL24_6}), the bound of reduced distance,
we have $\bar{\tau} \in \left[\frac{\eta^4}{C}, 1-\frac{\eta^2}{C} \right]$ for some uniform big constant $C$.
Moreover, we have the estimate
\begin{align}
d_{g(0)}(x,\beta{\bar{\tau}}) < C, \quad |\beta'(\bar{\tau})|_{g(-\bar{\tau})} < C
\label{eqn:SL24_3}
\end{align}
for some constant $C=C(H,\eta)$.
Note that the reduced volume element $(4\pi \tau)^{-n}e^{-l}$ is decreasing along $\boldsymbol{\beta}$.
Regard $\partial \mathcal{F}_{\xi}(M,0) \times [-1, 0]$ as
a hypersurface in the space-time.
Recall that $\Omega_{\xi}$ is the collection of points $y \in M$ such that every
shortest reduced geodesic $\boldsymbol{\beta}$ connecting $(x,0)$ and $(y,-1)$ satisfies (\ref{eqn:SL24_4}).
Therefore, by mapping $y$ to $\beta(\bar{\tau})$, we set up a projection map (maybe multi-valued on a measure zero set)
from $\Omega_{\xi}$ to $\partial \mathcal{F}_{\xi}(M,0) \times [-1, 0]$.
Restricted this projection on $\Omega_{\xi} \cap B_{g(0)}(y, \frac{c_a\eta}{4})$, its image must locate in
$\left\{\partial \mathcal{F}_{\xi}(M,0) \cap B(x,C) \right\} \times [-1, 0]$.
Then the bound (\ref{eqn:SL24_6}), the estimate (\ref{eqn:SL24_3}),
together with the monotonicity of reduced volume element along $\boldsymbol{\beta}$,
imply the following inequality.
\begin{align*}
\left|B_{g(0)}(x,H) \cap \mathcal{F}_{\eta}(M,0) \cap \Omega_{\xi} \right|&\leq C\left| \left\{\partial \mathcal{F}_{\xi}(M,0) \cap B_{g(0)}(x, C) \right\} \times [-1, 0] \right|\\
&\leq C \left| \partial \mathcal{F}_{\xi}(M,0) \cap B_{g(0)}(x,C) \right| \\
&\leq C\xi^{2p_0-1}.
\end{align*}
\end{proof}
The limit version of Lemma~\ref{lma:SL14_1} is the following property.
\begin{lemma}
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;1)$ satisfies
\begin{align}
\lim_{i \to \infty} \left( \frac{1}{T_i} + \frac{1}{\Vol(M_i)} + \sup_{\mathcal{M}_i} (|R|+|\lambda|) \right)=0. \label{eqn:SL06_1}
\end{align}
Suppose $x_i \in M_i$.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i, x_i, g_i(0))$, $\mathcal{R}$ be the regular part of $\bar{M}$ and $\bar{x} \in \mathcal{R}$.
Suppose $\bar{t}<0$ is a fixed number. Then every $(\bar{y}, \bar{t})$ can be connected to $(\bar{x}, 0)$ by a smooth reduced geodesic, whenever $\bar{y}$ is away from a measure zero set.
\label{lma:SK27_4}
\end{lemma}
\begin{proof}
Without loss of generality, let $\bar{t}=-1$.
Fix $\bar{y}$ a regular point in $\bar{M}$. Then we can fix $\eta>0$ such that both $\bar{x}$ and $\bar{y}$ are uniformly regular. For every fixed $\xi$ small, Lemma~\ref{lma:SL14_1}
implies that away from a set of measure $C\xi^{2p_0-1}$, every point around $(\bar{y}, -1)$ can be connected to $(\bar{x}, 0)$ by $\xi$-regular reduced geodesics.
Let $\xi \to 0$, we see that away from measure zero set, every point $(\bar{y},-1)$ can be connected to $(\bar{x},0)$ by a smooth reduced geodesic.
\end{proof}
By natural projection to the time slice $t=0$, we obtain the following property.
\begin{proposition}
Same condition as Lemma~\ref{lma:SK27_4}.
Then away from a measure zero set, every point in $\mathcal{R}$ can be connected to $\bar{x}$ with a smooth shortest geodesic. Consequently, $\mathcal{R}$ is weakly convex.
\label{prn:SC30_1}
\end{proposition}
\begin{proof}
It suffices to show that the space-projection of a smooth, shortest reduced geodesic is shortest.
If it is not shortest, then by Proposition~\ref{prn:SC17_4}, we have a shorter smooth curve connecting the two ends points, whose space-time
lifting will be curve with reduced length less than the given reduced distance, contradiction.
\end{proof}
The rough estimate in Lemma~\ref{lma:SL13_1} can be improved as the following proposition.
\begin{proposition}
Same condition as Proposition~\ref{prn:SC30_1}.
Suppose $(y_i,t_i) \in \mathcal{M}_i$ converges to $(\bar{y}, \bar{t})$, which is regular and $\bar{t}<0$.
Then we have
\begin{align}
\lim_{i \to \infty} l((x_i,0), (y_i,t_i))=\frac{d_0^2(\bar{x}, \bar{y})}{4|\bar{t}|}=l((\bar{x},0), (\bar{y},\bar{t}))
\label{eqn:SK27_9}
\end{align}
where $l$ is Perelman's reduced distance. Therefore, reduced distance is continuous function under Cheeger-Gromov topology whenever $\bar{y}$ is regular.
\label{prn:SL14_1}
\end{proposition}
\begin{proof}
Without loss of generality, we assume $t_i \equiv -1$, $d_0(x_i,y_i) \equiv 1$.
We first show
\begin{align}
\lim_{i \to \infty} l((x_i,0), (y_i,t_i)) \leq \frac{1}{4}.
\label{eqn:SL14_5}
\end{align}
If $x_i$ are uniformly regular, then there is a limit smooth geodesic connecting $\bar{x}$ and $\bar{y}$, which can
be lifted to a smooth reduced geodesic connection $(\bar{x},0)$ and $(\bar{y},-1)$ with reduced length $\frac{1}{4}$.
Then (\ref{eqn:SL14_5}) follows trivially. So we focus on the case when $\bar{x}$ is a singular point.
Choose a smooth point $\bar{z}$ very close to $\bar{x}$, say $\delta$-away from $\bar{x}$ under metric $\bar{g}(0)$.
From Lemma~\ref{lma:SL13_1}, the reduced length from $(x_i, 0)$ to $(z_i, -\delta^2)$ is uniformly less than $100$. So we have space-time curves $\boldsymbol{\alpha}_i$ connecting these two points such that
\begin{align*}
\int_0^{\delta^2} \sqrt{\tau} |\dot{\boldsymbol{\alpha}}_i|^2 d\tau< 200\delta.
\end{align*}
Note that $(\bar{z},-\delta^2)$ and $(\bar{y}, -1)$ can be connected by a space-time curve $\boldsymbol{\beta}$ such that
\begin{align*}
\int_{\delta^2}^{1} \sqrt{\tau} |\dot{\boldsymbol{\beta}}|^2 d\tau< \frac{1}{2} + 100\delta
\end{align*}
if $\delta$ is small enough.
So for large $i$, we have space-time curve $\boldsymbol{\beta}_i$ connecting $(z_i,-\delta^2)$ and $(y_i, -1)$ such that
\begin{align*}
\int_{\delta^2}^{1} \sqrt{\tau} |\dot{\boldsymbol{\beta}_i}|^2 d\tau < \frac{1}{2} + 200\delta.
\end{align*}
Concatenating $\boldsymbol{\alpha}_i$ and $\boldsymbol{\beta}_i$ to obtain $\boldsymbol{\gamma}_i$ such that
\begin{align*}
\int_{\delta^2}^{1} \sqrt{\tau} |\dot{\boldsymbol{\gamma}_i}|^2 d\tau < \frac{1}{2} + 400\delta,
\end{align*}
which implies $l((x_i,0),(y_i,-1)) < \frac{1}{4} + 200\delta$ for large $i$. Thus (\ref{eqn:SL14_5}) follows by letting $i \to \infty$ and $\delta \to 0$.
Then we show the equality holds. Otherwise, there exists a small $\epsilon$ such that
\begin{align*}
\lim_{i \to \infty} l((x_i,0), (y_i, -1)) < \frac{1}{4} -\epsilon.
\end{align*}
Note that $(y_i,-1)$ is uniformly regular. So we can find small $\delta$ such that
\begin{align*}
l((x_i,0), (z, -1-\delta^2)) < \frac{1}{4}-\frac{1}{2}\epsilon, \quad \forall \;
z \in B_{g(-1-\delta^2)}(y_i, \epsilon \delta)
\end{align*}
By Lemma~\ref{lma:SK27_4}, we obtain a point $(\bar{z}, -1-\delta^2)$, which can be connected to
$(\bar{x},0)$ by a smooth reduced geodesic, with reduced length smaller than $\frac{1}{4}-\frac{1}{2}\epsilon$.
Projecting this reduced geodesic to time zero slice, we obtain a curve connecting $\bar{x}$ and $\bar{z}$ with
\begin{align*}
d_0^2(\bar{x}, \bar{y}) < 4(1+\delta^2) \cdot (\frac{1}{4}-\frac{1}{2}\epsilon)=(1+\delta^2)(1- 2\epsilon)<1-\epsilon
\end{align*}
if we choose $\delta$ sufficiently small. This is impossible since $d_0(\bar{x},\bar{y})=1$. Therefore, we have
\begin{align*}
\lim_{i \to \infty} l((x_i,0), (y_i, -1)) = \frac{1}{4}.
\end{align*}
\end{proof}
Since singular set has measure zero, it is clear that
\begin{align}
\mathcal{V}((\bar{x},0), |\bar{t}|) \leq \lim_{i \to \infty} \mathcal{V}((x_i,0),|\bar{t}|).
\label{eqn:SK27_10}
\end{align}
We shall improve the above inequality as equality.\\
\begin{lemma}
For every positive $\eta$ and $H$, there exists an $\epsilon=\epsilon(n,A,\eta,H)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in \mathcal{F}_{\eta}(M,0)$. Then we have
\begin{align}
\left|\mathcal{V}((x,0),1)-(4\pi)^{-n}\int_{B_{g(0)}(x,H)} e^{-l}dv \right| \leq 2a(H), \label{eqn:SL18_1}
\end{align}
whenever $\sup_{\mathcal{M}} (|R|+|\lambda|)<\epsilon$.
Here $a$ is a positive function defined as
\begin{align}
a(H) \triangleq (4\pi)^{-n} \int_{\{|\vec{w}|>\frac{H}{100}\} \subset \ensuremath{\mathbb{R}}^{2n}} e^{-\frac{|\vec{w}|^2}{4}} dw. \label{eqn:SL18_2}
\end{align}
\label{lma:SL18_1}
\end{lemma}
\begin{proof}
The line bundle structure is not used in the following proof. So up to a parabolic rescaling if necessary, we can assume $\lambda=0$.
For every $y \in M$, there is at least one shortest reduced geodesic $\boldsymbol{\gamma}$ connecting $(x,0)$ and $(y,-1)$.
By standard ODE theory, the limit $\displaystyle \lim_{\tau \to 0} \sqrt{\tau} \gamma'(\tau)$ is unique as a vector in $T_{x} M$,
which is called the reduced tangent vector of $\boldsymbol{\gamma}$.
Away from a measure zero set, every $(y,-1)$ can be connected to $(x,0)$ by a
unique shortest reduced geodesic. For simplicity for our argument, we may assume this measure zero set is empty, since measure zero
set does not affect integral at all.
So there is a natural injective map from $M$ to $T_xM$, by mapping $y$ to the corresponding
reduced tangent vector $\vec{w}$.
We define
\begin{align*}
\Omega(H) \triangleq \{ y \in M | |\vec{w}|>H \}.
\end{align*}
It follows from the monotonicity of reduced element along reduced geodesic that
\begin{align*}
\int_{\Omega(H)} (4\pi)^{-n} e^{-l} dv \leq \int_{\{|\vec{w}| > H\} \subset \ensuremath{\mathbb{R}}^{2n}} (4\pi)^{-n} e^{-\frac{|\vec{w}|^2}{4}} dw.
\end{align*}
Choose $\xi<\eta$, with size to be determined.
Suppose $\boldsymbol{\gamma}$ is a reduced geodesic connecting $(x,0)$ to $(y,-1)$ for some $y \in M$.
It is clear that $\gamma(0)$ is in the interior part of $\mathcal{F}_{\xi}(M,0)$. Let $\tau$ to be the first time such that $\gamma(\tau)$ touches
the boundary of $\mathcal{F}_{\xi}(M,0)$. Then we see that $\boldsymbol{\gamma}([0,\tau])$ locates in a space-time domain
with uniformly bounded geometry, Ricci curvature very small. In particular, the reduced distance between $(x,0)$ and $\boldsymbol{\gamma}(\tau)$
is comparable to the length of $\vec{w}$, which is the reduced tangent vector of $\boldsymbol{\gamma}$ at $(x,0)$.
If $|\vec{w}|<H$, then we see that
\begin{align*}
\frac{H^2}{4} > \frac{|\vec{w}|^2}{4} \sim \frac{d_{g(0)}^2(x, \gamma(\tau))}{4\tau}> \frac{c_a^2 \eta^2}{100\tau},
\quad \Rightarrow \quad \tau > \frac{c_a^2 \eta^2}{25H^2}.
\end{align*}
Note that $\boldsymbol{\gamma}([0,\tau])$ is in a space-time region where Ricci curvature is almost flat, geometry is uniformly bounded.
So the lower bound of $\tau$ and the upper bounded of $|\vec{w}|$ imply an upper bound of $d_{g(0)}(x,\gamma(\tau))$.
Say $d_{g(0)}(x,\gamma(\tau))<H'$.
Around $\boldsymbol{\gamma}$, there is a natural projection (induced by reduced geodesic) from the space-time hypersurface
$\partial \mathcal{F}_{\xi}(M,0) \times [-1, -\frac{c_a^2 \eta^2}{25H^2}]$,
to the time slice $M \times \{-1\}$. At point $\boldsymbol{\gamma}(\tau)$, $\boldsymbol{\gamma}$ has space-time tangent vector $(\gamma', -1)$,
with $\tau|\gamma'(\tau)|^2$ is almost less than $\frac{H^2}{4}$. Together with the lower bound of $\tau$, we obtain an upper bound of $|\gamma'(\tau)|$.
Up to a constant depending on $H,\eta$, the volume element of $\partial \mathcal{F}_{\xi}(M,0) \times [-1,-\frac{c_a^2 \eta^2}{25H^2}]$
is comparable to the reduced volume element
$(4\pi \tau)^{-n} e^{-l}$ of $M$, around the point $\boldsymbol{\gamma}(\tau)$.
Note that the reduced volume element is monotone along each reduced geodesic. This implies that the projection map mentioned above
``almost" decreases weighted hypersurface volume element, if we equip $\{B(x,H') \cap \partial \mathcal{F}_{\xi}(M,0)\} \times [-1, -\frac{\eta^2}{4H^2}]$ with the natural
weighted volume element $e^{-l}|d \sigma \wedge dt|$. Let $\Omega_{\xi}'$ be the collection of all $y$'s such that $(y,-1)$ cannot be connected to $(x,0)$ by a shortest reduced geodesic $\boldsymbol{\gamma}$ which locates completely in $\mathcal{F}_{\xi}(M,0) \times [-1,0]$.
Then we have
\begin{align*}
\int_{\Omega_{\xi}'} e^{-l}(4\pi \tau)^{-n} dv &\leq C \int_{\frac{c_a^2 \eta^2}{25H^2}}^1 \int_{B(x,H') \cap \partial \mathcal{F}_{\xi}(M,0)} e^{-l} d\sigma d\tau\\
&\leq C \int_{B(x,H') \cap \partial \mathcal{F}_{\xi}(M,0)} d\sigma\\
&\leq C \xi^{2p_0-1}
\end{align*}
where $C=C(n,H,H',\eta)=C(n,H,\eta)$. By choosing $\xi$ small enough, we have
\begin{align}
\int_{\Omega_{\xi}'} e^{-l}(4\pi \tau)^{-n} dv \leq (4\pi)^n a(H). \label{eqn:SL18_3}
\end{align}
Note that
\begin{align*}
&\Omega_{100 H} \cap B_{g(0)}(x,H) \subset \Omega_{\xi}', \\
&M \backslash (\Omega_{\xi}' \cup B_{g(0)}(x,H)) \subset \Omega_{\frac{H}{100}}.
\end{align*}
Therefore, we have
\begin{align*}
&\quad (4\pi)^n \mathcal{V}((x,0),1)\\
&=\int_{M} e^{-l}dv\\
&=\int_{M \backslash (\Omega_{\xi}' \cup B_{g(0)}(x,H))} e^{-l}dv + \int_{\Omega_{\xi}'} e^{-l}dv + \int_{B_{g(0)}(x,H) \backslash \Omega_{\xi}'} e^{-l}dv\\
&\leq \int_{|\vec{w}|>\frac{H}{100}} e^{-\frac{|\vec{w}|^2}{4}} dw + \int_{\Omega_{\xi}'} e^{-l}dv + \int_{B_{g(0)}(x,H)} e^{-l}dv\\
&\leq \int_{|\vec{w}|>\frac{H}{100}} e^{-\frac{|\vec{w}|^2}{4}} dw + C \xi^{2p_0-1} + \int_{B_{g(0)}(x,H)} e^{-l}dv\\
&\leq 2 (4\pi)^n a(H) + \int_{B_{g(0)}(x,H)} e^{-l}dv.
\end{align*}
Then (\ref{eqn:SL18_1}) follows from the above inequality directly.
\end{proof}
\begin{lemma}
Suppose $\mathcal{M}=\{(M, g(t)), -\tau \leq t \leq 0\}$ is an unnormalized K\"ahler Ricci flow solution.
Suppose $x,y$ are two points in $M$, $d=d_{g(0)}(x,y)$. Then we have
\begin{align}
|\mathcal{V}((x,0),\tau) - \mathcal{V}((y,0),\tau)|<(4n+1) (e^{\frac{d}{2}}-1).
\label{eqn:SL15_2}
\end{align}
In particular, the reduced volume changes uniformly continuously with respect to the base point.
\label{lma:SL14_2}
\end{lemma}
\begin{proof}
Recall the definition of reduced volume
\begin{align*}
\mathcal{V}((x,0),\tau)= (4\pi \tau)^{-n} \int_M e^{-l} dv.
\end{align*}
Let $x$ move along a unit speed Riemannian geodesic $\alpha$, with respect to the metric $g(0)$.
Let $x=\alpha(0)$, $s$ be parameter of $\alpha$, $\vec{u}=\alpha'$.
For simplicity of notation, we denote $\mathcal{V}((\alpha(s),0),\tau)$ by $\mathcal{V}_s$.
It can be calculated directly the first variation of $l$ is $\langle\vec{u}, \vec{w} \rangle$ where $\vec{w}$ is the tangent vector of the reduced geodesic at time $t=0$.
Therefore, we have
\begin{align*}
\left|\frac{d}{ds} \mathcal{V}((\alpha(s),0),\tau) \right|&=\left| (4\pi \tau)^{-n} \int_M \langle \vec{u}, \vec{w} \rangle e^{-l}dv \right|\\
&\leq (4\pi \tau)^{-n} \int_M \frac{1+|\vec{w}|^2}{2} e^{-l}dv\\
&=\frac{1}{2}\mathcal{V} + \frac{1}{2} \int_{\ensuremath{\mathbb{R}}^{2n}} |\vec{w}|^2e^{-\frac{|\vec{w}|^2}{4}} J dw\\
&\leq \frac{1}{2}\mathcal{V} + \frac{(4\pi)^{-n}}{2} \int_{\ensuremath{\mathbb{R}}^{2n}} |\vec{w}|^2e^{-\frac{|\vec{w}|^2}{4}}dw,
\end{align*}
where $J$ is the Jacobian determinant of the reduced exponential map, which is always not greater than $1$, due to Perelman's argument in Section 7 of \cite{Pe1}.
Plugging the identity
\begin{align*}
(4\pi)^{-n} \int_{\ensuremath{\mathbb{R}}^{2n}} |\vec{w}|^2 e^{-\frac{|\vec{w}|^2}{4}} dw = 4n
\end{align*}
into the above inequality implies
\begin{align*}
\left|\frac{d}{ds} \mathcal{V} \right| \leq \frac{1}{2}\mathcal{V} +2n,
\end{align*}
which can be integrated as
\begin{align*}
(-\mathcal{V}_0+4n) (1-e^{-\frac{s}{2}})
\leq \mathcal{V}_s -\mathcal{V}_0
\leq (\mathcal{V}_0 + 4n) (e^{\frac{s}{2}}-1).
\end{align*}
Note that $0<\mathcal{V}_0\leq 1, s>0$. So we obtain
\begin{align*}
|\mathcal{V}_s -\mathcal{V}_0| \leq (4n+1) (e^{\frac{s}{2}}-1),
\end{align*}
which yields (\ref{eqn:SL15_2}) by letting $s=d$.
\end{proof}
The above argument clearly works for every Riemannian Ricci flow.
Note that the reduced volume is continuous for geodesic balls of each fixed scale under the Cheeger-Gromov convergence.
Combining this continuity together with the estimate in Lemma~\ref{lma:SL18_1} and Lemma~\ref{lma:SL14_2},
we can improve (\ref{eqn:SK27_10}) as an equality.
\begin{proposition}
Same condition as Proposition~\ref{prn:SC30_1}, $\bar{t}<0$ is a finite number.
Then we have
\begin{align}
\mathcal{V}((\bar{x},0), |\bar{t}|) = \lim_{i \to \infty} \mathcal{V}((x_i,0),|\bar{t}|).
\label{eqn:SL09_1}
\end{align}
Therefore, reduced volume is a continuous function under the Cheeger-Gromov convergence.
\label{prn:SL14_2}
\end{proposition}
Then we can study the gap property of the singularities.
\begin{proposition}
Same condition as in Theorem~\ref{thm:SC09_1}.
Suppose $\bar{y} \in \mathcal{S}(\bar{M})$, then we have
\begin{align}
\mathrm{v}(\bar{y})=\lim_{r \to 0} \omega_{2n}^{-1} r^{-2n}|B(\bar{y},r)| \leq 1-2\delta_0.
\label{eqn:SC30_4}
\end{align}
\label{prn:SC17_7}
\end{proposition}
\begin{proof}
Due to the tangent cone structure(c.f. Theorem~\ref{thm:SL25_1}), we have
\begin{align}
\mathrm{v}(\bar{y})=\lim_{r \to 0} \omega_{2n}^{-1} r^{-2n}|B(\bar{y},r)|=\lim_{r \to 0} \mathcal{V}((\bar{y},0), r^2).
\label{eqn:SL14_6}
\end{align}
Let $y_i \to \bar{y}$ under the metric $g_i(0)$. By rearranging points if necessary, we can assume $y_i$ has the local
minimum canonical volume radius $\rho_i$. By rescaling $\rho_i$ to $1$, we obtain new Ricci flows $\tilde{g}_i$.
Take limit of $(M_i,y_i,\tilde{g}_i(0))$, we have a complete, Ricci flat
eternal Ricci flow solution. By the continuity of canonical volume radius(c.f. Corollary~\ref{cly:SL27_1}), this limit space is not flat. So it has normalized
asymptotic volume ratio less than $1-2\delta_0$, according to Anderson's gap theorem. Then the infinity tangent cone
structure implies the asymptotic reduced volume is the same as the asymptotic reduced volume ratio. So it is
at most $1-2\delta_0$. Therefore, there exists a big constant $H$ such that
\begin{align*}
\mathcal{V}_{\tilde{g}_i}((y_i,0),H) < 1-2\delta_0.
\end{align*}
Note that $H\rho_i^2<r$ for each fixed $r$ and the corresponding large $i$. Recall the scaling invariant property of reduced volume, we can apply the reduced volume monotonicity to obtain
\begin{align*}
\mathcal{V}_{g_i}((y_i,0),r^2) \leq \mathcal{V}_{\tilde{g}_i}((y_i,0), \rho_i^{-2} r^2)
\leq \mathcal{V}_{\tilde{g}_i}((y_i,0),H) < 1-2\delta_0.
\end{align*}
The continuity of reduced volume (Proposition~\ref{prn:SL14_2}) then implies that
\begin{align*}
\mathcal{V}((\bar{y},0),r^2) \leq 1-2\delta_0
\end{align*}
for each $r>0$, which in turn yields
\begin{align}
\lim_{r \to 0} \mathcal{V}((\bar{y},0),r^2) \leq 1-2\delta_0.
\label{eqn:SL14_7}
\end{align}
Then (\ref{eqn:SC30_4}) follows from the combination of (\ref{eqn:SL14_6}) and (\ref{eqn:SL14_7}).
\end{proof}
\begin{theorem}[\textbf{Metric structure of a blowup limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A;1)$ satisfies (\ref{eqn:SL06_1}), $x_i \in M_i$.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i, x_i, g_i(0))$. Then $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{thm:SC04_1}
\end{theorem}
\begin{proof}
We only need to check $\bar{M}$ satisfies all the 6 properties required in the definition of $\widetilde{\mathscr{KS}}(n,\kappa)$. In fact,
the 1st property is implied by Theorem~\ref{thm:HE11_1}.
The 2nd property follows from the fact that $\mathcal{R}$ is scalar flat and satisfies K\"ahler Ricci flow equation.
The 3rd property, weak convexity of $\mathcal{R}$ is shown in Proposition~\ref{prn:SC30_1}.
The 4th property, codimension estimate of singularity follows from Proposition~\ref{prn:SL15_1}.
The 5th property, gap estimate, follows from Proposition~\ref{prn:SC17_7}.
The 6th property, asymptotic volume ratio estimate can be obtained by the condition $Vol(M_i) \to \infty$, Sobolev constant uniformly bounded, and the volume convergence, Proposition~\ref{prn:SC12_1}.
So we have checked all the properties needed to define $\widetilde{\mathscr{KS}}(n,\kappa)$ are satisfied by $\bar{M}$.
In other words, $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\end{proof}
Since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, it is clear that $\mathbf{cr}(\bar{x})=\infty$. Therefore, we have $\mathbf{vr}(\bar{x})=\mathbf{cvr}(\bar{x})$
by definition.
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}. Let $\displaystyle \bar{r}=\lim_{i \to \infty} \mathbf{cr}(x_i)$.
Then we have
\begin{align}
\min\{\bar{r}, \mathbf{vr}(\bar{x})\} = \lim_{i \to \infty} \mathbf{cvr}(x_i).
\label{eqn:SC06_9}
\end{align}
\label{prn:SC06_5}
\end{proposition}
\begin{proof}
We divide the proof in three cases according to the value of $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}$. \\
\textit{Case 1. $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}=0$.}
Otherwise, there exists a positive number $\rho_0$ such that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) \geq \rho_0$. Therefore,
$\bar{x} \in \mathcal{R}_{\rho_0} \subset \mathcal{R}$, which in turn implies that $\mathbf{vr}(\bar{x})>0$.
Consequently, we have $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}>0$. Contradiction.
\textit{Case 2. $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}=\infty$.}
In this case, $\bar{r}=\infty$.
By the gap theorem in the space $\widetilde{\mathscr{KS}}(n,\kappa)$, we see that $\bar{M}$ is the Euclidean space $\ensuremath{\mathbb{C}}^{n}$.
Therefore, for each $H>0$, we have $\omega_{2n}^{-1}H^{-2n}|B(x_i,H)|$ converges to $1$, the normalized volume ratio of $\ensuremath{\mathbb{C}}^{n}$.
Since $\displaystyle \bar{r}=\lim_{i \to \infty} \mathbf{cr}(x_i)=\infty$, this means that $\mathbf{cvr}(x_i) \geq H$ for large $i$ by the volume convergence.
Since $H$ is chosen arbitrarily, we obtain $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i)=\infty$.\\
So the remainder case is that $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}$ is a finite positive number.
Two more subcases can be divided.\\
\textit{Case 3(a). $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}<\bar{r}$.}
Let $H=\mathbf{vr}(\bar{x})$, a finite number in this case. Clearly, $\bar{x}$ is a regular point and the normalized volume ratio of the ball $B(\bar{x},H)$
is $1-\delta_0$. Clearly, $B(\bar{x},H)$ cannot be a isometric to a Euclidean ball.
Therefore, by the rigidity of $\widetilde{\mathscr{KS}}(n,\kappa)$(c.f. Proposition~\ref{prn:SC17_3}), we see that
\begin{align*}
&\omega_{2n}^{-1}r^{-2n}|B(\bar{x},r)|>1-\delta_0, \quad \forall \; r \in (0,H), \\
&\omega_{2n}^{-1}r^{-2n}|B(\bar{x},r)|<1-\delta_0, \quad \forall \; r \in (H,\bar{r}).
\end{align*}
Then the volume convergence implies that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i)=H$. \\
\textit{Case 3(b). $\min\{\bar{r}, \mathbf{vr}(\bar{x})\}=\bar{r}$.}
In this case, we see that the normalized volume ratio of $B(\bar{x},\bar{r})$ is at least $1-\delta_0$. Also, we see that $\bar{x}$ is a regular point.
Same argument as in the previous case, we see that
\begin{align*}
\omega_{2n}^{-1}r^{-2n}|B(\bar{x},r)|>1-\delta_0, \quad \forall \; r \in (0,\bar{r}).
\end{align*}
Therefore, for every fixed $r \in (0, \bar{r})$, the volume convergence implies that $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) \geq r$. Consequently,
we have $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) \geq \bar{r}$ by the arbitrariness of $r$. On the other hand, the definition of $\mathbf{cvr}(x_i)$
implies that
\begin{align*}
\lim_{i \to \infty} \mathbf{cvr}(x_i) \leq \lim_{i \to \infty} \mathbf{cr}(x_i)=\bar{r}.
\end{align*}
Therefore, we obtain $\displaystyle \lim_{i \to \infty} \mathbf{cvr}(x_i) =\bar{r}$.
\end{proof}
\begin{corollary}
Same conditions as in Theorem~\ref{thm:SC04_1}. Then for each $r \in (0,1)$, we have
\begin{align}
\mathcal{F}_{r}(\bar{M})=\mathcal{R}_{r}(\bar{M}).
\label{eqn:SC06_11}
\end{align}
In particular, for each $0<r<1<H<\infty$, we have
\begin{align*}
B(x_i,H) \cap \mathcal{F}_{r}(M_i) \longright{G.H.} B(\bar{x},H) \cap \mathcal{F}_{r}(\bar{M}).
\end{align*}
Moreover, this convergence can be improved to take place in $C^{\infty}$-topology, i.e.,
\begin{align}
B(x_i,H) \cap \mathcal{F}_{r}(M_i) \longright{C^{\infty}} B(\bar{x},H) \cap \mathcal{F}_{r}(\bar{M}). \label{eqn:SC06_12}
\end{align}
\label{cly:SC06_1}
\end{corollary}
\begin{corollary}
Same conditions as in Theorem~\ref{thm:SC04_1}, $0<H\leq 3$.
Then we have
\begin{align}
\lim_{i \to \infty} \int_{B(x_i,H)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy \leq H^{2n-2p_0}\mathbf{E}.
\label{eqn:SC06_13}
\end{align}
\label{cly:SC06_2}
\end{corollary}
\begin{proof}
Fix two positive scales $r_1,r_2$ such that $0<r_2<r_1<1$.
\begin{align}
\int_{B(x_i,H) \cap \mathcal{F}_{r_1}} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\leq r_1^{-2p_0} |B(x_i,H) \cap \mathcal{F}_{r_1}| \leq r_1^{-2p_0}|B(x_i,H)|.
\label{eqn:SC17_6}
\end{align}
Fix arbitrary $r \in (0,1)$, then we have
\begin{align*}
\lim_{i \to \infty} \int_{B(x_i,H) \cap (\mathcal{F}_{r_2} \backslash \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy
=\int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy.
\end{align*}
Note that
\begin{align*}
\int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy
&\leq \int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \min\{\mathbf{vr}, 1\}^{-2p_0} dy\\
&<\int_{B(\bar{x},H) \cap (\mathcal{F}_{r_2} \cap \mathcal{F}_{r_1})} \left\{1+\mathbf{vr}(y)^{-2p_0} \right\} dy\\
&<\int_{B(\bar{x},H)} \left\{1+\mathbf{vr}(y)^{-2p_0} \right\} dy\\
&<|B(\bar{x},H)| + H^{2n-2p_0}E(n,\kappa,p_0).
\end{align*}
It follows that
\begin{align}
\lim_{i \to \infty} \int_{B(x_i,H) \cap (\mathcal{F}_{r_2} \backslash \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0} dy
\leq |B(\bar{x},H)| + H^{2n-2p_0}E(n,\kappa,p_0).
\label{eqn:SC17_7}
\end{align}
Note that $\mathcal{S} \cap \overline{B(\bar{x},H)}$ is a compact set with Hausdorff dimension at most
$2n-4$, which is strictly less than $2n-2p_0$. By the definition of Hausdorff dimension, for every small number $\xi$,
we can find finite cover
$\cup_{j=1}^{N_{\xi}} B(\bar{y}_j, \rho_j)$ of $\mathcal{S} \cap \overline{B(\bar{x},H)}$ , such that
\begin{align*}
\sum_{j=1}^{N_{\xi}} |\rho_{j}|^{2n-2p_0} < \xi.
\end{align*}
By the finiteness of this cover, we can choose an $r_2$ very small such that $\cup_{j=1}^{N_{\xi}} B(\bar{y}_j, \rho_j)$
is a cover of $\mathcal{D}_{r_2} \cap \overline{B(\bar{x},H)}$.
Therefore, for large $i$, we have a finite cover $\cup_{j=1}^{N_{\xi}} B(y_{i,j}, \rho_j)$ of the set
$\mathcal{D}_{r_2}(M_i) \cap \overline{B(x_i,H)}$ such that
\begin{align*}
\sum_{j=1}^{N_i} |\rho_{i,j}|^{2n-2p_0} < \xi.
\end{align*}
Combining this with the canonical radius density estimate, we have
\begin{align}
\int_{B(x_i,H) \cap \mathcal{D}_{r_2}} \mathbf{vr}^{(1)}(y)^{-2p_0}dy
\leq \sum_{j=1}^{N_i} \int_{B(y_{i,j},\rho_{i,j})} \mathbf{vr}^{(\rho_{i,j})}(y)^{-2p_0} dy
\leq 2\mathbf{E} \sum_{j=1}^{N_i} |\rho_{i,j}|^{2n-2p_0} < 2\mathbf{E}\xi.
\label{eqn:SC17_8}
\end{align}
Putting (\ref{eqn:SC17_6}), (\ref{eqn:SC17_7}) and (\ref{eqn:SC17_8}) together, we have
\begin{align*}
&\qquad \int_{B(x_i,H)} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\\
&\leq \int_{B(x_i,H) \cap \mathcal{F}_{r_1}} \mathbf{vr}^{(1)}(y)^{-2p_0}dy
+\int_{B(x_i,H) \cap (\mathcal{F}_{r_2} \backslash \mathcal{F}_{r_1})} \mathbf{vr}^{(1)}(y)^{-2p_0}dy
+\int_{B(x_i,H) \cap \mathcal{D}_{r_2}} \mathbf{vr}^{(1)}(y)^{-2p_0} dy\\
&\leq r_1^{-2p_0}|B(x_i,H)|+ |B(\bar{x},H)| +H^{2n-2p_0} E(n,\kappa,p_0) +2\mathbf{E}\xi.
\end{align*}
Take limit on both sides and then let $\xi \to 0, r_1 \to 1$, then we have
\begin{align*}
\lim_{i \to \infty} \int_{B(x_i,H)} \mathbf{vr}^{(1)}(y)^{-2p_0} &\leq 2|B(\bar{x},H)| + H^{2n-2p_0} E(n,\kappa,p_0)\\
&\leq \left( 2 \omega_{2n} H^{2p_0} + E(n,\kappa,p_0) \right) H^{2n-2p_0}\\
&\leq \left( 2 \cdot 9^{p_0}\omega_{2n} + E(n,\kappa,p_0) \right) H^{2n-2p_0},
\end{align*}
where we used the fact that $H\leq 3$ in the last step. Then (\ref{eqn:SC06_13}) follows from the definition of $\mathbf{E}$.
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $1 \leq H <\infty$. Then we have
\begin{align}
\lim_{i \to \infty} \sup_{1 \leq \rho \leq H} \omega_{2n}^{-1}\rho^{-2n}|B(x_i,\rho)| <\kappa^{-1}, \label{eqn:SC06_4}
\end{align}
where $g_i(0)$ is the default metric. In particular, for every large $i$, the volume ratio estimate holds
on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_1}
\end{proposition}
\begin{proof}
We argue by contradiction. If (\ref{eqn:SC06_4}) were false, by taking subsequence if necessary, one can assume that there exists $\rho_i \in [1,H]$ such that
\begin{align*}
\omega_{2n}^{-1}\rho_i^{-2n}|B(x_i,\rho_i)| >\kappa^{-1}.
\end{align*}
Let $\bar{\rho}$ be the limit of $\rho_i$, then by the volume continuity in the Cheeger-Gromov convergence, we see that
\begin{align}
\omega_{2n}^{-1}\bar{\rho}^{-2n}|B(\bar{x},\bar{\rho})| \geq \kappa^{-1}.
\label{eqn:SC06_5}
\end{align}
However, since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, we know $ \omega_{2n}^{-1}\bar{\rho}^{-2n}|B(\bar{x},\bar{\rho})| \leq 1$,
which contradicts (\ref{eqn:SC06_5}).
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $0<H<\infty$. For every large $i$, the regularity estimate holds
on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_2}
\end{proposition}
\begin{proof}
If the statement were false, then by taking subsequence if necessary, we can assume there exists $\rho_i \in (0,H]$ such that the regularity estimates fail on
the scale $\rho_i$, i.e., the following two inequalities hold simultaneously.
\begin{align}
&\omega_{2n}^{-1}\rho_i^{-2n}|B(x_i,\rho_i)|>1-\delta_0, \label{eqn:SC06_6}\\
&\max_{0 \leq k \leq 5} \left\{\rho_i^{2+k} \sup_{B(x_i,\frac{1}{2}c_a \rho_i)} |\nabla^k Rm| \right\} > 4c_a^{-2}. \label{eqn:SC06_7}
\end{align}
Clearly, $\rho_i \in [1,H]$ by the fact $\mathbf{cr}(x_i,0) \geq 1$. Let $\bar{\rho}$ be the limit of $\rho_i$. Then we have
\begin{align}
\omega_{2n}^{-1}\bar{\rho}^{-2n}|B(\bar{x},\bar{\rho})| \geq 1-\delta_0. \label{eqn:SC06_8}
\end{align}
Since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, (\ref{eqn:SC06_8}) implies
\begin{align*}
\max_{0 \leq k \leq 5} \left\{\bar{\rho}^{2+k} \sup_{B(\bar{x}, c_a \bar{\rho})} |\nabla^k Rm| \right\} < c_a^{-2},
\end{align*}
which contradicts (\ref{eqn:SC06_7}) in light of the smooth convergence(c.f. Proposition~\ref{prn:SL04_1}).
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $1 \leq H \leq 2$. Then we have
\begin{align}
\lim_{i \to \infty} \sup_{1 \leq \rho \leq H} \rho^{2p_0-2n} \int_{B(x_i,\rho)} \mathbf{vr}^{(\rho)}(y)^{-2p_0}dy \leq \frac{3}{2}\mathbf{E}. \label{eqn:SC06_16}
\end{align}
In particular, for every large $i$, the density estimate holds on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_3}
\end{proposition}
\begin{proof}
Since $\mathbf{vr}^{(\rho)} \geq \mathbf{vr}^{(1)}$ whenever $\rho \geq 1$, in order to show (\ref{eqn:SC06_16}), it suffices to show
\begin{align}
\lim_{i \to \infty} \sup_{1 \leq \rho
\leq H} \rho^{2p_0-2n} \int_{B(x_i,\rho)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy \leq \frac{3}{2}\mathbf{E}.
\label{eqn:SL23_9}
\end{align}
We argue by contradiction. If (\ref{eqn:SL23_9}) were false, by taking subsequence if necessary, one can assume that there exists $\rho_i \in [1,H]$ such that
\begin{align*}
\rho_i^{2p_0-2n} \int_{B(x_i,\rho_i)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy > \frac{3}{2}\mathbf{E}, \quad
\Rightarrow \quad \int_{B(x_i,\rho_i)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy \geq \frac{3}{2}\mathbf{E} \rho_i^{2n-2p_0}.
\end{align*}
Let $\bar{\rho}$ be the limit of $\rho_i$. Fix $\epsilon$ arbitrary small positive number, then we have
\begin{align}
\int_{B(x_i,\bar{\rho}+\epsilon)} \mathbf{vr}^{(1)}(y)^{-2p_0}dy> \frac{3}{2}\mathbf{E} \rho_i^{2n-2p_0}> \frac{5}{4} \mathbf{E} (\bar{\rho}+\epsilon)^{2n-2p_0}
\label{eqn:SC07_1}
\end{align}
for large $i$. Note that $\bar{\rho}+\epsilon<3$, so (\ref{eqn:SC07_1}) contradicts (\ref{eqn:SC06_13}).
\end{proof}
\begin{proposition}
Same conditions as in Theorem~\ref{thm:SC04_1}, $0<H\leq 2$.
Then for every large $i$, the connectivity estimate holds on $(M_i,x_i,g_i(0))$ for every scale $\rho \in (0,H]$.
\label{prn:SC06_4}
\end{proposition}
\begin{proof}
By the canonical radius assumption, we know the connectivity estimate holds for every scale $\rho \in (0,1]$.
If the statement were false, then by taking subsequence if necessary, we can assume that for each $i$,
there is a scale $\rho_i \in [1,H]$ such that the connectivity estimate fails on the scale $\rho_i$.
In other words, $\mathcal{F}_{\frac{1}{50}c_b\rho_i} \cap B(x_i,\rho_i)$ is not $\frac{1}{2}\epsilon_b \rho_i$-regular-connected.
So there exist points $y_i,z_i \in \mathcal{F}_{\frac{1}{50}c_b\rho_i} \cap B(x_i,\rho_i)$ which cannot be connected by a curve
$\gamma \subset \mathcal{F}_{\frac{1}{2}\epsilon_b \rho_i}$ satisfying $|\gamma|\leq 2d(y_i,z_i)$.
By the canonical radius assumption, it is clear that $\rho_i \in [1,H]$, $d(y_i,z_i) \in [1,2H]$.
Let $\bar{\rho}$ be the limit of $\rho_i$, $\bar{y}$ and $\bar{z}$ be the limit of $y_i$ and $z_i$ respectively.
Clearly, we have $\bar{y}, \bar{z} \in \mathcal{R}_{\frac{1}{50}c_b \bar{\rho}} \subset \mathcal{F}_{\frac{1}{100}c_b \bar{\rho}}(\bar{M})$.
Since $\bar{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$, we can find a shortest geodesic
$\bar{\gamma}$ connecting $\bar{y}$ and $\bar{z}$ such that $\bar{\gamma} \subset \mathcal{F}_{\epsilon_b \bar{\rho}}$.
Note that the limit set of $\mathcal{F}_{\frac{1}{50}c_b \rho_i}\cap B(x_i,\rho_i)$ falls into $\mathcal{F}_{\frac{1}{100}c_b \bar{\rho}}$.
Moreover, this convergence takes place in the smooth topology(c.f.~Corollary~\ref{cly:SC06_1}).
So by deforming $\bar{\gamma}$ if necessary, we can construct a curve $\gamma_i$ which locates in $\mathcal{F}_{\frac{1}{2}\epsilon_b \rho_i}$
and $|\gamma_i| <\frac{3}{2}d(\bar{y},\bar{z})<3d(y_i,z_i)$. The existence of such a curve contradicts the choice of the points
$y_i$ and $z_i$.
\end{proof}
Combining Proposition~\ref{prn:SC06_1} to~\ref{prn:SC06_4}, we obtain a weak-semi-continuity of canonical radius.
\begin{theorem}[\textbf{Weak continuity of canonical radius}]
Same conditions as in Theorem~\ref{thm:SC04_1}. Then we have
\begin{align*}
\lim_{i \to \infty} \mathbf{cr}(\mathcal{M}_i^{0})=\infty.
\end{align*}
\label{thm:SC03_1}
\end{theorem}
\begin{proof}
If the statement were wrong, then we can find a sequence of polarized K\"ahler Ricci flow solutions
$\mathcal{LM}_i \in \mathscr{K}(n,A;1)$ satisfying (\ref{eqn:SL06_1}) and
\begin{align}
\lim_{i \to \infty} \mathbf{cr}(\mathcal{M}_i^{0})=H<\infty.
\label{eqn:SC06_17}
\end{align}
For each $\mathcal{M}_i$, we can find a point $x_i$ such that $\mathbf{cr}(x_i,0) \leq \frac{3}{2} \mathbf{cr}(\mathcal{M}_i^{0})$ by definition. So we have
\begin{align}
\lim_{i \to \infty} \mathbf{cr}(x_i,0) \leq \frac{3}{2}H<\infty. \label{eqn:SC04_3}
\end{align}
Note that $T_i \to \infty$, so the first property holds trivially on the scale $2H$ for large $i$.
By Propositions listed before, we see that there exists an $N=N(H)$ such that for every $i>N$, we have volume ratio estimate, regularity estimate,
density estimate and connectivity estimate hold on each scale $\rho \in (0,2H]$. Therefore, by definition, we obtain that
\begin{align*}
\lim_{i \to \infty} \mathbf{cr}(x_i,0) \geq 2H,
\end{align*}
which contradicts (\ref{eqn:SC04_3}).
\end{proof}
\begin{corollary}[\textbf{Weak continuity of canonical volume radius}]
Same conditions as in Theorem~\ref{thm:SC04_1}. Then we have
\begin{align*}
\mathbf{vr}(\bar{x})=\lim_{i \to \infty} \mathbf{cvr}(x_i).
\end{align*}
\label{cly:SL27_1}
\end{corollary}
\begin{proof}
It follows from the combination of Proposition~\ref{prn:SC06_5} and Theorem~\ref{thm:SC03_1}.
\end{proof}
\begin{theorem}[\textbf{Weak continuity of polarized canonical radius}]
Same conditions as in Theorem~\ref{thm:SC04_1}. Then $\mathbf{pcr}(\mathcal{M}_i^{0}) \geq 1$ for $i$ large enough.
\label{thm:SK27_2}
\end{theorem}
\begin{proof}
It follows from the combination of Theorem~\ref{thm:SC04_1}, Theorem~\ref{thm:SC03_1} and Corollary~\ref{cly:SK27_1}.
\end{proof}
\subsection{A priori bound of polarized canonical radius}
We shall use a maximum principle type argument to show that the polarized canonical radius cannot be too small.
The technique used in the following proof is inspired by the proof of Theorem 10.1 of~\cite{Pe1}.
\begin{proposition}[\textbf{A priori bound of $\textbf{pcr}$}]
There is a uniform integer constant $j_0=j_0(n,A)$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, then
\begin{align}
\mathbf{pcr}(\mathcal{M}^{t}) \geq \frac{1}{j_0}
\label{eqn:SB27_4}
\end{align}
for every $t \in [-1, 1]$.
\label{prn:SC28_2}
\end{proposition}
\begin{proof}
Suppose for some positive integer $j_0$, (\ref{eqn:SB27_4}) fails at time $t_0 \in [-1,1]$.
Then we check whether $\mathbf{pcr}(\mathcal{M}^t) \geq \frac{1}{2j_0}$ on the interval $[t_0-\frac{1}{2j_0}, t_0+\frac{1}{2j_0}]$.
If so, stop. Otherwise, choose $t_1$ to be such a time and continue to check if $\mathbf{pcr}(\mathcal{M}^t) \geq \frac{1}{4j_0}$
on the interval $[t_1-\frac{1}{4j_0}, t_1+\frac{1}{4j_0}]$. In each step, we shrink the scale to one half of the scale in the previous step.
Note this process will never escape the time interval $[-2,2]$ since
\begin{align*}
|t_k-t_0|<\frac{1}{j_0} \left( \frac{1}{2} + \frac{1}{4} + \cdot +\frac{1}{2^k}\right)<\frac{1}{j_0}<1, \quad |t_k|<|t_0|+1 \leq 2.
\end{align*}
By compactness of the underlying manifold,
it is clear that the process stops after finite steps. So we can find $t_k$ such that $\frac{1}{2^{k+1}j_0} \leq \mathbf{pcr}(\mathcal{M}^{t_k}) < \frac{1}{2^kj_0}$ and
$\mathbf{pcr}(\mathcal{M}^{t}) \geq \frac{1}{2^{k+1}j_0}$ for every
$t \in [t_k-\frac{1}{2^{k+1} j_0}, t_k + \frac{1}{2^{k+1} j_0}]$.
Translate the flow and rescale by constant $4^{k}j_0^2$,
we obtain a new polarized K\"ahler Ricci flow $\widetilde{\mathcal{LM}} \in \mathscr{K}(n,A)$ such that
\begin{align}
\begin{cases}
&\mathbf{pcr}(\widetilde{\mathcal{M}}^0) <1, \\
&\mathbf{pcr}(\widetilde{\mathcal{M}}^{t}) \geq \frac{1}{2}, \quad \forall \; t \in [-2^{k-1}j_0,2^{k-1}j_0], \\
&|R|+|\lambda|<\frac{A}{4^{k}j_0^2}<\frac{A}{j_0^2}, \quad \textrm{on} \; \widetilde{\mathcal{M}}, \\
&\frac{1}{T}+\frac{1}{\Vol(M)}<\frac{1}{2^{k-1}j_0} + \frac{A}{j_0^2}, \quad \textrm{on} \; \widetilde{\mathcal{M}}.
\end{cases}
\label{eqn:SL04_3}
\end{align}
In other words, $\widetilde{\mathcal{LM}} \in \mathscr{K}(n,A;0.5)$ and $|R|+|\lambda|+\frac{1}{T}+\frac{1}{\Vol(M)}$ very small.
Now we return to the main proof. If the statement fails, after adjusting, translating and rescaling, we can find a sequence of polarized
K\"ahler Ricci flow $\widetilde{\mathcal{LM}}_i \in \mathscr{K}(n,A;0.5)$ satisfying
\begin{align*}
\begin{cases}
&\mathbf{pcr}\left( \mathcal{M}_i^{0} \right) <1, \\
&\frac{1}{T_i}+\frac{1}{\Vol(M_i)}+\sup_{\widetilde{\mathcal{M}}_i} (|R|+|\lambda|) \to 0,
\end{cases}
\end{align*}
which contradicts Theorem~\ref{thm:SK27_2}. Note that Theorem~\ref{thm:SK27_2} holds even if $\mathscr{K}(n,A;1)$ is replaced by $\mathscr{K}(n,A;0.5)$.
\end{proof}
Let $\hslash=\frac{1}{j_0}$. Then we have the following fact.
\begin{theorem}[\textbf{Homogeneity on small scales}]
For some small positive number $\hslash=\hslash(n,A)$, we have
\begin{align}
\mathscr{K}(n,A)=\mathscr{K}\left(n,A;\hslash \right). \label{eqn:SL04_4}
\end{align}
\label{thm:SC28_1}
\end{theorem}
\section{Structure of polarized K\"ahler Ricci flows in $\mathscr{K}(n,A)$}
Because of Theorem~\ref{thm:SC28_1}, $\mathscr{K}(n,A)=\mathscr{K}\left(n,A;\hslash \right)$.
We do have a uniform lower bound for polarized canonical radius.
\subsection{Local metric structure, flow structure, and line bundle structure}
The purpose of this subsection is to set up estimates related to the local metric structure, flow structure and
line bundle structure of every flow in $\mathscr{K}(n,A)$. In particular, we shall prove
Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1}.
\begin{proposition}[\textbf{K\"ahler tangent cone}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ is a sequence of polarized K\"ahler Ricci flows.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i,x_i, g_i(0))$.
Then for each $\bar{y} \in \bar{M}$, every tangent space of $\bar{M}$ at $\bar{y}$ is an irreducible metric cone.
Moreover, this metric cone can be extended as an eternal, possibly singular Ricci flow solution.
\label{prn:HA07_2}
\end{proposition}
\begin{proof}
It follows from Theorem~\ref{thm:SC28_1} and Theorem~\ref{thm:SC09_1} that every tangent space is an irreducible metric cone. From the proof of Theorem~\ref{thm:SC09_1}, it is clear that the tangent cone can be extended as
an eternal, static Ricci flow solution.
\end{proof}
\begin{proposition}[\textbf{Regularity equivalence}]
Same condition as in Proposition~\ref{prn:HA07_2}, $\bar{y} \in \bar{M}$. Then the following statements are equivalent.
\begin{enumerate}
\item One tangent space of $\bar{y}$ is $\ensuremath{\mathbb{C}}^n$.
\item Every tangent space of $\bar{y}$ is $\ensuremath{\mathbb{C}}^n$.
\item $\bar{y}$ has a neighborhood with $C^{4}$-manifold structure.
\item $\bar{y}$ has a neighborhood with $C^{\infty}$-manifold structure.
\item $\bar{y}$ has a neighborhood with $C^{\omega}$-manifold (real analytic manifold) structure.
\end{enumerate}
\label{prn:HA08_1}
\end{proposition}
\begin{proof}
It is obvious that $5 \Rightarrow 4 \Rightarrow 3 \Rightarrow 2 \Rightarrow 1$. So it suffices to show $1 \Rightarrow 5$
to close the circle. Suppose $\bar{y}$ has a tangent space which is isometric to $\ensuremath{\mathbb{C}}^n$. So we can find a sequence $r_k \to 0$ such that
\begin{align*}
(\bar{M}, \bar{y}, r_k^{-2}\bar{g}) \stackrel{G.H.}{\longrightarrow} (\ensuremath{\mathbb{C}}^n, 0, g_{Euc}).
\end{align*}
So for large $k$, the unit ball $B_{r_k^{-2}\bar{g}}(\bar{y},1)$ has volume ratio almost the Euclidean one.
Fix such a large $k$, we see that $B_{\bar{g}}(\bar{y}, r_k)$ has almost Euclidean volume ratio. It follows from volume convergence that $\mathbf{cvr}(y_i,0) \geq r_k$ for large $i$, where $y_i \in M_i$ and $y_i \to \bar{y}$ as
$(M_i,x_i,g_i(0))$ converges to $(\bar{M}, \bar{x}, \bar{g})$. By the regularity improving property of
canonical volume radius, there is a uniform small constant $c$ such that $B(y_i,cr_k)$ is diffeomorphic to the
same radius Euclidean ball in $\ensuremath{\mathbb{C}}^n$ and the metrics on $B(y_i, c r_k)$ is $C^2$-close to the Euclidean metric.
Then one can apply the backward pseudolocality(c.f. Theorem~\ref{thm:SL27_2}) to obtain higher order derivative estimate for the metrics. Therefore, $B(y_i, \frac{1}{2}c r_k)$ will converge in smooth topology to a limit smooth geodesic ball $B(\bar{y}, \frac{1}{2}cr_k)$. Moreover, it is clear that geometry is uniformly bounded in a space-time neighborhood containing
$B(y_i, \frac{1}{2} cr_k) \times [-c^2 r_k^2, 0]$, by shrinking $c$ if necessary. So we obtain a limit K\"ahler Ricci flow
solution on $B(\bar{y}, \frac{1}{4}c r_k) \times [-\frac{1}{4}c^2 r_k^2, 0]$.
It follows from the result of Kotschwar(c.f.~\cite{Kotsch}), that $B(\bar{y}, \frac{1}{4}cr_k)$ is actually an analytic manifold, which is the desired neighborhood of $\bar{y}$.
So we finish the proof of $1 \Rightarrow 5$ and close the circle.
\end{proof}
\begin{remark}
By Proposition~\ref{prn:HA08_1}, our initial non-classical definition of regularity is proved to be the same as the classical
one (c.f.~Remark~\ref{rmk:HA01_1}).
\label{rmk:HA07_1}
\end{remark}
\begin{proposition}[\textbf{Volume density gap}]
Same condition as in Proposition~\ref{prn:HA07_2}, $\bar{y} \in \bar{M}$.
Then $\bar{y}$ is singular if and only if
\begin{align}
\limsup_{r \to 0} \frac{|B(\bar{y}, r)|}{\omega_{2n}r^{2n}} \leq 1-2\delta_0.
\label{eqn:HA07_3}
\end{align}
\label{prn:HA07_1}
\end{proposition}
\begin{proof}
If (\ref{eqn:HA07_3}) holds, then every tangent cone of $\bar{y}$ cannot be $\ensuremath{\mathbb{C}}^n$, so $\bar{y}$ is singular.
If $\bar{y}$ is singular, then every tangent space of $\bar{y}$ is an irreducible metric cone in the model space
$\widetilde{\mathscr{KS}}(n,\kappa)$ with vertex a singular point, it follows from the gap property of
$\widetilde{\mathscr{KS}}(n,\kappa)$ that asymptotic volume ratio of such a metric cone must be at most $1-2\delta_0$.
Then (\ref{eqn:HA07_3}) follows from the volume convergence and a scaling argument.
\end{proof}
\begin{proposition}[\textbf{Regular-Singular decomposition}]
Same condition as in Proposition~\ref{prn:HA07_2}, $\bar{M}$ has the regular-singular decomposition
$\bar{M}=\mathcal{R} \cup \mathcal{S}$. Then the regular part $\mathcal{R}$ admits a natural K\"ahler structure $\bar{J}$.
The singular part $\mathcal{S}$ satisfies the estimate $\dim_{\mathcal{H}} \mathcal{S} \leq 2n-4$.
\label{prn:HA08_2}
\end{proposition}
\begin{proof}
The existence of $\bar{J}$ on $\mathcal{R}$ follows from smooth convergence,
due to the backward pseudolocality(c.f. Theorem~\ref{thm:SL27_2}) and Shi's estimate.
The Hausdorff dimension estimate of $\mathcal{S}$ follows from the
combination of Proposition~\ref{prn:SL15_1} and Theorem~\ref{thm:SC28_1}.
\end{proof}
Therefore, Theorem~\ref{thmin:SC24_1} follows from
the combinations from Proposition~\ref{prn:HA07_2} to Proposition~\ref{prn:HA08_2}.
Now we are going to discuss more delicate properties of the moduli space
$\widetilde{\mathscr{K}}(n,A)$.
\begin{proposition}[\textbf{Improve regularity in two time directions}]
There is a small positive constant $c=c(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$. Let $r_0=\min\{\mathbf{cvr}(x_0,0), 1\}$. Then we have
\begin{align*}
r^{2+k}|\nabla^k Rm|(x,t) \leq \frac{C_k}{c^{2+k}}, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{+}, \quad x \in B_{g(0)}(x_0, cr_0),
\quad t \in [-c^2 r^2, c^2 r^2],
\end{align*}
where $C_k$ is a constant depending on $n,A$ and $k$.
\label{prn:HA08_3}
\end{proposition}
\begin{proof}
Otherwise, there exists a fixed positive integer $k_0$ and a sequence of $c_i \to 0$ such that
\begin{align}
(c_i r_i)^{2+k_0}|\nabla^{k_0} Rm|(y_i,t_i) \to \infty
\label{eqn:HA08_4}
\end{align}
for some $y_i \in B_{g_i(0)}(x_i, r_i)$, $t_i \in [-c_i r_i^2, c_i r_i^2]$, where $r_i=\min\{\mathbf{cvr}(x_i,0), 1\}$.
Let
$\tilde{g}_i(t)=(c_i r_i)^{-2}g_i((c_ir_i)^2t+t_i)$.
Then we have $\mathbf{cvr}_{\tilde{g}_i}(y_i, 0)=(c_i r_i)^{-1} \to \infty$. Note that
$\mathbf{pcr}_{\tilde{g}_i}(y_i,0) \geq \min\{\hslash (c_i r_i)^{-1}, 1\} \geq 1$.
It is also clear that for the flows $\tilde{g}_i$, $|R|+|\lambda| \to 0$.
Therefore, Proposition~\ref{prn:SL04_1} can be applied to obtain
\begin{align}
(M_i, y_i, \tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{M},\hat{y},\hat{g}).
\label{eqn:HA08_5}
\end{align}
However, it follows from Theorem~\ref{thm:SC04_1} and Corollary~\ref{cly:SL27_1} that
\begin{align*}
(\hat{M},\hat{y},\hat{g}) \in \widetilde{\mathscr{KS}}(n,\kappa), \quad
\mathbf{cvr}(\hat{y})=\infty.
\end{align*}
In light of the gap property, Proposition~\ref{prn:SC17_1},
we know that $\hat{M}$ is isometric to $\ensuremath{\mathbb{C}}^n$.
So the convergence (\ref{eqn:HA08_5}) can be rewritten as
\begin{align*}
(M_i, y_i, \tilde{g}_i(0)) \stackrel{C^{\infty}}{\longrightarrow} (\ensuremath{\mathbb{C}}^n, 0, g_{Euc}).
\end{align*}
In particular, $|\nabla^{k_0} Rm|_{\tilde{g}_i}(y_i,0) \to 0$, which is the same as
\begin{align*}
(c_i r_i)^{2+k_0}|\nabla^{k_0} Rm|(y_i,t_i) \to 0.
\end{align*}
This contradicts the assumption (\ref{eqn:HA08_4}).
\end{proof}
Perelman's pseudolocality theorem says that an almost Euclidean domain cannot become very singular in a short time.
His almost Euclidean condition is explained as isoperimetric constant close to that of the Euclidean one. In our special setting,
we can reverse this theorem, i.e., an almost Euclidean domain cannot become very singular in the reverse time direction for a short time period.
\begin{theorem}[\textbf{Two-sided pseudolocality}]
There is a small positive constant $\xi=\xi(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$.
Let $\Omega=B_{g(0)}(x_0,r)$, $\Omega'=B_{g(0)}(x_0, \frac{r}{2})$ for some $0<r\leq 1$.
Suppose $\mathbf{I}(\Omega) \geq (1-\delta_0) \mathbf{I}(\ensuremath{\mathbb{C}}^n)$ at time $t=0$, then
\begin{align*}
(\xi r)^{2+k}|\nabla^k Rm|(x,t) \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0}, \quad x \in \Omega',
\quad t \in [-\xi^2 r^2, \xi^2 r^2],
\end{align*}
where $C_k$ is a constant depending on $n,A$ and $k$.
\label{thm:SL27_2}
\end{theorem}
\begin{proof}
Note that each geodesic ball contained in $\Omega$ has volume ratio at least $(1-\delta_0)\omega_{2n}$.
Then the theorem follows from directly from Proposition~\ref{prn:HA08_3}.
\end{proof}
After we obtain the bound of geometry, we can go further to study the evolution of potential functions.
\begin{theorem}[\textbf{Two-sided pseudolocality on the potential level}]
Same condition as in Theorem~\ref{thm:SL27_2}.
Let $\omega_B$ be a smooth metric form in $2\pi c_1(M,J)$ and denote
$\omega_t$ by $\omega_B + \sqrt{-1} \partial \bar{\partial} \varphi(\cdot, t)$.
Suppose $\varphi(x_0,0)=0$ and $Osc_{\Omega} \varphi(\cdot, 0) \leq H$.
Let $\Omega''=B_{g(0)}(x_0,\frac{r}{4})$. Then we have
\begin{align}
(\xi r)^{-2+k}\norm{\varphi(\cdot, t)}{C^k(\Omega'', \omega_t)} \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0},
\quad t \in \left[-\frac{\xi^2}{2} r^2, \frac{\xi^2}{2} r^2 \right],
\label{eqn:HA05_2}
\end{align}
where $C_k$ depends on $k,n,A,\xi$ and $\frac{H}{r^2}$.
\label{thm:HA03_5}
\end{theorem}
\begin{proof}
Up to rescaling, we may assume $\xi r=1$.
Note that $\varphi$ and $\dot{\varphi}$ satisfy the equations
\begin{align*}
\begin{cases}
&\dot{\varphi}=\log \frac{\omega_t^n}{\omega_B^n}+\varphi +\dot{\varphi}(\cdot, 0), \\
&-\sqrt{-1} \partial \bar{\partial} \dot{\varphi}=Ric-\lambda g.
\end{cases}
\end{align*}
It follows from Theorem~\ref{thm:SL27_2} that geometry is uniformly bounded in $\Omega' \times [-\xi r^2, \xi r^2]$.
The trace form of the second equation in the above list is $-\Delta \dot{\varphi}=R-n\lambda$.
Therefore, the regularity theory
of Laplacian operator applies and we have uniform bound of $\norm{\dot{\varphi}}{C^k}$ in a neighborhood of
$\Omega'' \times [-\frac{\xi}{2}r^2, \frac{\xi}{2}r^2]$.
Up to a normalization, we can rewrite the first equation as
\begin{align*}
\log \frac{(\omega_t-\sqrt{-1}\partial \bar{\partial} \varphi)^n}{\omega_t^n}=\varphi - \dot{\varphi}+\dot{\varphi}(\cdot, 0).
\end{align*}
On $\Omega'$, the metric $g(0)$ and $g(t)$ are uniformly equivalent in each $C^k$-topology. So it is clear that
$\norm{\dot{\varphi}-\dot{\varphi}(\cdot, 0)}{C^k(\Omega')}$ are uniformly bounded, for each $k$, with respect to
metric $g(t)$.
Since all higher derivatives of curvature are uniformly bounded on $\Omega'$, (\ref{eqn:HA05_2})
follows from standard Monge-Ampere equation theory and bootstrapping argument.
\end{proof}
\begin{theorem}[\textbf{Improving regularity of potentials}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $\mathbf{cvr}(M,0)=r_0$.
Let $\omega_B$ be a smooth metric in $[\omega_0]$ such that
\begin{align}
\frac{1}{2}\omega_B \leq \omega_0 \leq 2\omega_B.
\label{eqn:HA06_1}
\end{align}
Let $\omega_0=\omega_B + \sqrt{-1} \partial \bar{\partial} \varphi$.
Suppose $\int_M \varphi \omega_0^n=0$ and $Osc_M \varphi \leq H$. Then we have
\begin{align}
\norm{\varphi}{C^k(M,\omega_B)} \leq C_k, \quad \forall \; k \in \ensuremath{\mathbb{Z}}^{\geq 0},
\label{eqn:HA06_2}
\end{align}
where $C_k$ depends on $k,\omega_B, n,A,r_0$ and $H$.
\label{thm:HA06_1}
\end{theorem}
\begin{proof}
Since $\mathbf{cvr}(M,0)=r_0>0$, we see that all the possible $\omega_0$'s form a compact set under the smooth topology.
In other words, $\omega_0$ has uniformly bounded geometry in each regularity level. Fix a positive integer $k_0 \geq 4$.
Therefore, around each point $x \in M$, one can find a coordinate chart $\Omega$, with uniform size, such that
\begin{align*}
\omega_0=\omega_{Euc} + \sqrt{-1} \partial \bar{\partial} f, \quad
\norm{f}{C^{k_0}(\Omega, \omega_{Euc})} \leq 0.01.
\end{align*}
Note that in $\Omega$, the connection terms of the metric $\omega_0$ are pure derivatives $f_{i\bar{j}l}$,
which are uniformly bounded. Similarly, all derivatives of connection terms can be expressed as high order
pure derivatives of $f$. Therefore, up to order $k_0-3$, the derivatives of connections are uniformly bounded.
It is clear that the metric $\omega_0$ and $\omega_{Euc}$ are uniformly equivalent.
By the covariant derivatives' bounds $\norm{\varphi}{C^k(M, \omega_0)} \leq C_k$,
the bounds of connection derivatives yield that
\begin{align}
\norm{\varphi}{C^k(\Omega, \omega_{Euc})} \leq C_k, \quad \forall \; 0 \leq k \leq k_0-1.
\label{eqn:HA08_3}
\end{align}
In other words, we have uniform bound for every order pure derivatives of $\varphi$, up to order $k_0-1$.
Together with the choice assumption of $\Omega$, we have
\begin{align*}
\norm{f-\varphi}{C^k(\Omega, \omega_{Euc})} \leq C_k, \quad \forall \; 0 \leq k \leq k_0-1.
\end{align*}
Therefore, the connection derivatives of metric $\omega_B$ in $\Omega$ are uniformly bounded, up to order $k_0-4$.
Consequently, the pure derivative bound (\ref{eqn:HA08_3}) implies
\begin{align*}
\norm{\varphi}{C^k(\Omega, \omega_B)} \leq C_k, \quad \forall \; 0 \leq k \leq k_0-1,
\end{align*}
since $\omega_B$ is a fixed smooth, compact metric with every level of regularity. Clearly, the above constant $C_k$ depends on
$k,n,A,r_0,\omega_B$ and $H$. Recall that the size of $\Omega$ is uniformly bounded from below, $(M, \omega_B)$ is a compact
manifold. Consequently, a standard covering argument implies (\ref{eqn:HA06_2}) for each $k \leq k_0-1$.
In the end, we free $k_0$ and finish the proof.
\end{proof}
In Ricci-flat theory, a version of Anderson's gap theorem says that regularity can be improved in the center of a ball
if the volume ratio of the unit ball is very close to the Euclidean one.
In our special setting, this gap theorem has a reduced volume version.
\begin{theorem}[\textbf{Gap of reduced volume}]
There is a constant $\delta_0' \in (0, \delta_0]$ and a small constant $\eta$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$, $0<r \leq 1$.
If $\mathcal{V}((x_0,0),r^2) \geq 1-\delta_0'$, then we have
\begin{align}
\mathbf{cvr}(x_0,0) \geq \eta r.
\label{eqn:SL27_2}
\end{align}
\label{thm:SL27_3}
\end{theorem}
\begin{proof}
If $\lambda=0$, reduced volume is monotone. If $\lambda$ is bounded, then reduced volume is almost monotone.
A simple calculation shows that $\mathcal{V}((x_0,0), \rho^2) \geq 1-\delta_0$ for all $0<\rho \leq r^2$
whenever $\mathcal{V}((x_0,0),r^2) \geq 1-\delta_0'$ for some $0<r\leq 1$. Therefore, without loss of generality,
we may assume $\lambda=0$ and $\delta_0'=\delta_0$ in the proof.
If the statement was wrong, there exists a sequence of $\eta_i \to 0$, $0<r_i \leq 1$ and $x_i \in M_i$, and corresponding K\"ahler Ricci flows satisfying
\begin{align*}
\begin{cases}
&\mathcal{V}((x_i,0),r_i^2) \geq 1-\delta_0, \\
&\mathbf{cvr}(x_i,0) <\eta_i r_i.
\end{cases}
\end{align*}
By the monotonicity of reduced volume, we have
\begin{align*}
\begin{cases}
&\mathcal{V}((x_i,0),H\eta_i^2 r_i^2) \geq 1-\delta_0, \\
&\mathbf{cvr}(x_i,0) <\eta_i r_i,
\end{cases}
\end{align*}
for each fixed $H$ and large $i$.
Let $\tilde{g}_i(t)=(\eta_i r_i)^{-2}g((\eta_i r_i)^2t)$. It is clear that
\begin{align}
\mathbf{cvr}_{\tilde{g}_i}(x_i,0)=1. \label{eqn:HA09_1}
\end{align}
The canonical radius of $\tilde{g}_i$ tends to infinity, $|R|+|\lambda| \to 0$.
Similar to the proof of Proposition~\ref{prn:HA08_3}, we have the convergence.
\begin{align*}
(M_i, x_i, \tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{M}, \hat{x}, \hat{g}) \in \widetilde{\mathscr{KS}}(n,\kappa).
\end{align*}
The limit space $\hat{M}$ can be extended to a static eternal K\"ahler Ricci flow solution.
Moreover, Proposition~\ref{prn:SL14_2} can be applied here and guarantees the reduced volume convergence.
\begin{align*}
\mathcal{V}((\hat{x},0), H)=\lim_{i \to \infty} \mathcal{V}_{\tilde{g}_i}((x_i,0), H)
=\lim_{i \to \infty} \mathcal{V}((x_i,0), H(\eta_i r_i)^2) \geq 1-\delta_0.
\end{align*}
Note that $H$ is arbitrary. By the homogeneity
of reduced volume at infinity, Theorem~\ref{thm:SL25_1}, we see that
\begin{align*}
\mathrm{avr}(\hat{M})=\lim_{H \to \infty} \mathcal{V}((\bar{x},0),H) \geq 1-\delta_0 \geq 1-\delta_0.
\end{align*}
So Proposition~\ref{prn:SC17_1} applies to force $\hat{M}$ to be isometric to be $\ensuremath{\mathbb{C}}^n$.
In particular, $\mathbf{vr}(\hat{x})=\infty$.
It follows from Corollary~\ref{cly:SL27_1} that
\begin{align*}
\lim_{i \to \infty} \mathbf{cvr}_{\tilde{g}_i}(x_i,0)=\infty,
\end{align*}
which contradicts (\ref{eqn:HA09_1}).
\end{proof}
According to Theorem~\ref{thm:SL27_3}, one can define a concept of reduced volume radius for the purpose of
improving regularity. Clearly, other regularity radius can also be defined. However, it seems all of them are equivalent.
For simplicity, we shall not compare all of them, but only prove an example case:
the equivalence of harmonic radius and canonical volume radius. The proof of other cases are verbatim.
\begin{proposition}[\textbf{Equivalence of regularity radii}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$.
Suppose $\max\{\mathbf{hr}(x,0), \mathbf{cvr}(x,0)\} \leq 1$,
then we have
\begin{align*}
\frac{1}{C} \mathbf{hr}(x,0) \leq \mathbf{cvr}(x,0) \leq C \mathbf{hr}(x,0)
\end{align*}
for some uniform constant $C=C(n,A)$.
\label{prn:HA09_2}
\end{proposition}
\begin{proof}
Clearly, $\mathbf{cvr}(x,0) \leq C \mathbf{hr}(x,0)$ follows from the $C^5$-regularity property of canonical volume radius. It suffices to show $\frac{1}{C} \mathbf{hr}(x,0) \leq \mathbf{cvr}(x,0)$. However, since $\mathbf{cr}(x,0) \geq \hslash$,
it is clear from definition that
\begin{align*}
\mathbf{cvr}(x,0) \geq \frac{1}{C} \min\{\mathbf{hr}(x,0), \hslash\}.
\end{align*}
If $\mathbf{hr}(x,0)\leq \hslash$, then we are done. Otherwise, we have $\hslash <\mathbf{hr}(x,0) \leq 1$. It follows that
\begin{align*}
\mathbf{cvr}(x,0) \geq \frac{1}{C}\hslash \geq \frac{\hslash}{C} \mathbf{hr}(x,0) \geq \frac{1}{C'} \mathbf{hr}(x,0).
\end{align*}
So we finish the proof.
\end{proof}
\begin{theorem}[\textbf{Improved density estimate}]
For arbitrary small $\epsilon$, arbitrary $0 \leq p<2$, there is a constant $\delta=\delta(n,A,p)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$. Then under the metric $g(0)$, we have
\begin{align}
\log \frac{\int_{B(x,r)} \mathbf{cvr}^{-2p}dv}{E(n,\kappa,p) r^{2n-2p} } < \epsilon
\label{eqn:SC13_2}
\end{align}
whenever $r<\delta$.
Here the number $E(n,\kappa,p)$ is defined in Proposition~\ref{prn:SB25_2}.
\label{thm:SC12_2}
\end{theorem}
\begin{proof}
We argue by contradiction.
Note that every blowup limit is in $\widetilde{\mathscr{KS}}(n,\kappa)$(c.f. Theorem~\ref{thm:SC04_1}).
Then a contradiction can be obtained by the weak continuity of $\mathbf{cvr}$(c.f. Corollary~\ref{cly:SL27_1})
if the statement of this theorem does not hold.
\end{proof}
Note that $E(n,\kappa,0)=\omega_{2n}$. So we are led to the volume ratio estimate immediately.
\begin{corollary}[\textbf{Volume-ratio estimate}]
For arbitrary small $\epsilon$, there is a constant $\delta=\delta(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$. Then under the metric $g(0)$, we have
\begin{align}
\log \frac{|B(x,r)|}{\omega_{2n} r^{2n}} < \epsilon
\label{eqn:SC13_1}
\end{align}
whenever $r<\delta$.
\label{cly:SC13_1}
\end{corollary}
In the K\"ahler Ricci flow setting, Corollary~\ref{cly:SC13_1} improves the volume ratio estimates in~\cite{Zhq3}
and~\cite{CW5} (c.f.~Remark 1.1 of \cite{CW5}). Note that the integral (\ref{eqn:SC13_2}) can be used to show that
for every $p \in (0,2)$, there is a $C=C(n,A,p)$ such that the $r$-neighborhood of $\mathcal{S}$ in a unit ball is
bounded by $C r^{2p}$(c.f. Theorem~\ref{thm:HE11_1}), where $\mathcal{S}$ is the singular part of a limit space.
By the definition of Minkowski dimension(c.f.~Definition~\ref{dfn:HE08_1}),
we can improve Proposition~\ref{prn:HA08_2} as follows.
\begin{corollary}[\textbf{Minkowski dimension of singular set}]
Same condition as in Proposition~\ref{prn:HA07_2}, $\bar{M}$ has the regular-singular decomposition
$\bar{M}=\mathcal{R} \cup \mathcal{S}$. Then $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$.
\label{cly:HE25_2}
\end{corollary}
In \cite{Wa2}, the second author developed an estimate of the type $|Ric| \leq \sqrt{|Rm||R|}$, where $\sqrt{|Rm|}$ should be understood as the reciprocal of a regular scale. Due to the improving regularity property of canonical volume radius, it induces the estimate
$|Ric| \leq \frac{\sqrt{|R|}}{\mathbf{cvr}}$ pointwisely.
By the uniform bound of scalar curvature and Theorem~\ref{thm:SC12_2}, the following estimate is clear now.
\begin{corollary}[\textbf{Ricci curvature estimate}]
There is a constant $C=C(n,A,r_0)$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x_0 \in M$, $0<r \leq r_0$, $0<p<2$.
Then under the metric $g(0)$, we have
\begin{align}
r^{2p-2n}\int_{B(x_0,r)} |Ric|^{2p} dv<C.
\label{eqn:SL27_3}
\end{align}
\label{cly:SL27_4}
\end{corollary}
Corollary~\ref{cly:SL27_4} localizes the $L^{2p}$-curvature estimate of \cite{TZZ2} in a weak sense, since
(\ref{eqn:SL27_3}) only holds for $p<2$. If $n=2$, (\ref{eqn:SL27_3}) also holds for $p=2$,
since the finiteness of singularity guarantees that one can choose good cutoff functions.
We believe that the same localization result hold for $p=2$ even if $n>2$.\\
We return to the canonical neighborhood theorems in the introduction,
Theorem~\ref{thmin:SC24_1}, Theorem~\ref{thmin:HC08_1} and Theorem~\ref{thmin:HC06_1}.
However, Theorem~\ref{thmin:HC06_1} is not completely local.
Actually, Theorem~\ref{thm:SL27_2} is enough to show
the local flow structure of $\mathscr{K}(n,A)$ can be approximated by $\mathscr{KS}(n,\kappa)$.
In light of its global properties, the proof of Theorem~\ref{thmin:HC06_1} is harder and is postponed to section 5.5.
On the other hand, Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1} are local.
We now close this subsection by proving Theorem~\ref{thmin:SC24_1} and Theorem~\ref{thmin:HC08_1}.
\begin{proof}[Proof of Theorem~\ref{thmin:SC24_1}]
It follows from the combination of Proposition~\ref{prn:HA07_2}, Proposition~\ref{prn:HA08_1}, Proposition~\ref{prn:HA07_1},
and Proposition~\ref{prn:HA08_2}.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thmin:HC08_1}]
It follows from Theorem~\ref{thm:SC28_1}, Definition~\ref{dfn:SK27_1} and a scaling argument.
\end{proof}
\subsection{Local variety structure}
We focus on the variety structure of the limit space in this subsection.
We essentially follow the argument in~\cite{DS}, with slight modification.
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Let $(\bar{M}, \bar{x},\bar{g})$
be a pointed-Gromov-Hausdorff limit of $(M_i,x_i,g_i(0))$.
Since $\bar{M}$ may be non-compact, the limit line bundle $\bar{L}$ may have infinitely many orthogonal holomorphic sections.
Therefore, in general, we cannot expect to embed $\bar{M}$ into a projective space of finite dimension
by the complete linear system of $\bar{L}$. However, when
we focus our attention to the unit geodesic ball $B(\bar{x},1)$,
we can choose some holomorphic sections of $\bar{L}$, peaked around $\bar{x}$, to embed $B(\bar{x},1)$ into $\ensuremath{\mathbb{CP}}^N$ for a
finite $N$.
Actually, for every $\epsilon>0$, we can find an $\epsilon$-net of $B(\bar{x},2)$ such that every point in this net has canonical volume radius
at least $c_0 \epsilon$. For each point $y$ in this $\epsilon$-net, we have a peak section $s_y$, which is a holomorphic section such that
$\norm{s(y)}{}$ achieves the maximum among all unit $L^2$-norm holomorphic sections $s \in H^0(\bar{M},\bar{L})$. By the partial-$C^0$-estimate
argument, we can assume that $\norm{s_y}{}^2$ is uniformly bounded below in $B(y,2\epsilon)$.
On the other hand, by the choice of $y$, $B(y, \eta \epsilon)$ has a smooth manifold structure for some $\eta=\eta(n)$.
Therefore, we can choose $n$ holomorphic sections of $\bar{L}^k$ such that these sections are the local deformation of $ z_1, z_2, \cdots, z_n$.
Here $k$ is a positive integer proportional to $\epsilon^{-2}$. Put these holomorphic sections together with $s_y^k$,
we obtain $(n+1)$-holomorphic sections of $\bar{L}^k$ based at the point $y$. Let $y$ run through all points in the $\epsilon$-net and collect all the holomorphic sections based at $y$, we obtain a set of holomorphic sections $\{s_i\}_{i=0}^N$ of $\bar{L}^{k}$. Let $\{\tilde{s}_i\}_{i=0}^N$
be the orthonormal basis of $span\{s_0,s_1, \cdots, s_N\}$. We define the Kodaira map $\iota$ as follows.
\begin{align*}
\iota: B(0,2) &\mapsto \ensuremath{\mathbb{CP}}^N, \\
x &\mapsto [\tilde{s}_0(x) : \tilde{s}_1(x): \cdots:\tilde{s}_N(x)].
\end{align*}
This map is well defined. In fact, for every $z \in B(\bar{x},1)$, we can find a point $y$ in the $\epsilon$-net and $z \in B(y,2\epsilon)$, then
$\norm{s_y}{}^2(z)>0$ by the partial-$C^0$-estimate. It forces that $\tilde{s}_j(z) \neq 0$ for some $j$.
Since $k$ is proportional to $\epsilon^{-2}$, we can just let $\epsilon=\frac{1}{\sqrt{k}}$ without loss of generality.
In the following argument, by saying ``raise the power of line bundle" from $k_1$ to $k_2$,
we simultaneously means the underlying $\epsilon$-net is
strengthened from a $\frac{1}{\sqrt{k_1}}$-net to a $\frac{1}{\sqrt{k_2}}$-net.
\begin{lemma}
Suppose $w \in \iota(B(\bar{x},1))$, then $ \iota^{-1}(w) \cap \overline{B(\bar{x},1)}$ is a finite set.
\label{lma:HB09_1}
\end{lemma}
\begin{proof}
Let $y \in \iota^{-1}(w) \cap \overline{B(\bar{x},1)}$. It is clear that $\iota^{-1}(w)$ is contained in a ball centered at $y$ with fixed radius, say $10\epsilon$.
Therefore, $\iota^{-1}(w)$ is a bounded, closed set and therefore compact. Let $F$ be a connected component of $\iota^{-1}(w)$. Then
$\iota(F)$ is a connected, compact subvariety of $\ensuremath{\mathbb{C}}^N$, and consequently is a point. Note that $\iota(F)$ is always a connected set no matter how do we raise the power of $\iota$. On the other hand, $\iota(F)$ will contain more than one point if $F$ is not a single point, after we raise power high enough.
These force that $F$ can only be a point. Since $\iota^{-1}(w) \cap \overline{B(\bar{x},1)}$ is compact, it must be union of finite points.
\end{proof}
Denote $\iota(\overline{B(\bar{x},1)})$ by $W$. Then $W$ is a compact set and locally can be extended as an analytic variety.
By dividing $W$ into different components, one can apply induction argument as that in~\cite{DS}.
Following verbatim the argument of Proposition 4.10, Lemma 4.11 of~\cite{DS}, one can show that $\iota$ is an injective, non-degenerate embedding
map on $B(\bar{x},1)$, by raising power of $\bar{L}$ if necessary. Furthermore, since being normal is a local property, one can
improve Lemma 4.12 of~\cite{DS} as follows.
\begin{lemma}
By raising power if necessary, $W$ is normal at the point $\iota(y)$ for every $y \in B(\bar{x}, \frac{1}{2})$.
\end{lemma}
Under the help of parabolic Schwartz lemma and heat flow localization technique(c.f. Section 4.1 and Proposition~\ref{prn:HB05_1}),
we can parallelly generalize Proposition 4.14 of~\cite{DS} as follows.
\begin{lemma}
Suppose $y \in B(\bar{x},\frac{1}{2})\cap \mathcal{S}$, then $\iota(y)$ is a singular point of $W$.
\end{lemma}
It follows from the proof of Proposition 4.15 of~\cite{DS} that there always exist a holomorphic form $\Theta$ on $\mathcal{R} \cap B(\bar{x},1)$ such that
$\int_{\mathcal{R} \cap B(\bar{x},1)} \Theta \wedge \bar{\Theta}<\infty$. This means that every singular point $y \in \iota(B(\bar{x},\frac{1}{2})) \cap W$ is
log-terminal.
Combine all the previous lemmas, we have the following structure theorem.
\begin{theorem}[\textbf{Analytic variety structure}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$,
$(\bar{M}, \bar{x},\bar{g})$ is a pointed Gromov-Hausdorff limit of $(M_i,x_i,g_i(0))$.
Then $\bar{M}$ is an analytic space with normal, log terminal singularities.
\label{thm:HE19_1}
\end{theorem}
\subsection{Distance estimates}
In this subsection, we shall develop the distance estimate along polarized K\"ahler Ricci flow in terms of the estimates
from line bundle.
\begin{lemma}
Suppose $(M,L)$ is a polarized K\"ahler manifold satisfying the following conditions
\begin{itemize}
\item $|B(x,r)| \geq \kappa \omega_{2n} r^{2n}, \quad \forall x \in M, 0<r<1$.
\item $|\mathbf{b}|\leq 2c_0$ where $\mathbf{b}$ is the Bergman function.
\item $\norm{\nabla S}{} \leq C_1$ for every $L^2$-unit section $S \in H^0(M,L)$.
\end{itemize}
For every positive number $a$, define
\begin{align}
\Omega(x,a)
\triangleq \left\{z \left| \norm{S}{}^2(z) \geq e^{-2a-2c_0},
\norm{S}{}^2(x)=e^{\mathbf{b}(x)}, \int_M \norm{S}{}^2dv=1 \right.\right\}.
\label{eqn:HA15_2}
\end{align}
Then we have
\begin{align}
B\left(x,r\right) \subset \Omega(x,a) \subset B(x, \rho)
\label{eqn:HA15_1}
\end{align}
for some $r=r(n,\kappa,c_0,C_1,a)$ and $\rho=\rho(n,\kappa,c_0,C_1,a)$.
\label{lma:HA15_1}
\end{lemma}
\begin{proof}
Define
\begin{align*}
r \triangleq \frac{1-e^{-a}}{C_1 e^{a+c_0}}.
\end{align*}
Recall that $\norm{S}{}(x) \geq e^{-c_0}$. By the gradient bound of $S$, it is clear that every point in
$B(x,r)$ satisfies $\norm{S}{} \geq e^{-a-c_0}$. In other words, we have
\begin{align*}
B(x,r) \subset \Omega(x,a).
\end{align*}
On the other hand, we can cover $\Omega(x,a)$ by finite balls $B(x_i, 2r)$ such that each $x_i \in \Omega(x,a)$
and different $B(x_i,r)$'s are disjoint to each other. Again, the gradient bound of $S$ implies that
$\norm{S}{} \geq e^{-2a-c_0}$ in each $B(x_i,r)$. Then we have
\begin{align*}
N \kappa \omega_{2n} r^{2n} \leq \sum_{i=1}^{N} |B(x_i,r)| \leq |\Omega(x, 2a)| \leq e^{4a+2c_0}.
\end{align*}
For every $z \in \Omega(x,a)$, we have
\begin{align*}
d(x,z) \leq 4Nr \leq \frac{4e^{4a+2c_0}}{\kappa \omega_{2n}r^{2n-1}}
=\frac{4e^{4a+2c_0}}{\kappa \omega_{2n}} \cdot \frac{C_1^{2n-1} e^{(2n-1)(a+c_0)}}{(1-e^{-a})^{2n-1}}.
\end{align*}
Let $\rho$ be the number on the right hand side of the above inequality. Then it is clear that
\begin{align*}
\Omega(x,a) \subset B(x,\rho).
\end{align*}
So we finish the proof.
\end{proof}
Lemma~\ref{lma:HA15_1} implies that the level sets of peak holomorphic sections are comparable to
geodesic balls. However, the norm of peak holomorphic section has stability under the K\"ahler Ricci flows
in $\mathscr{K}(n,A)$. Therefore, one can compare distances at different time slices in terms the values
of norms of a same holomorphic sections.
\begin{lemma}
There exists a small constant $\epsilon_0=\epsilon_0(n,A)$ such that
the following properties are satisfied.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, then we have
\begin{align}
B_{g(t_1)}(x,\epsilon_0) \subset B_{g(t_2)}(x,\epsilon_0^{-1}) \label{eqn:HA14_1}
\end{align}
whenever $t_1, t_2 \in [-1,1]$.
\label{lma:HA14_1}
\end{lemma}
\begin{proof}
Without loss of generality, we only need to show (\ref{eqn:HA14_1}) for time $t_1=0, t_2=1$.
Because of Theorem~\ref{thmin:HC08_1} and Moser iteration, we can assume $|\mathbf{b}| \leq 2c_0$ for some $c_0=c_0(n,A)$.
By standard elliptic theory, we can also assume $\norm{\nabla S}{} \leq C_1$ for every unit $L^2$-norm holomorphic
section of $L$. Note that $e^{\mathbf{b}(x)}$ is the maximum value of $\norm{S}{}^2$ among all
unit $L^2$-norm holomorphic sections of $L$. So we can choose $\epsilon$ small enough such that
\begin{align*}
\norm{S}{}(z) \geq \frac{1}{2}e^{-c_0}, \quad \forall \; z \in B(x,2\epsilon)
\end{align*}
for some unit holomorphic section $S$. Note $\epsilon$ can be chosen uniformly, say $\epsilon=\frac{e^{-c_0}}{4C_1}$.
Fix $S$ and define
\begin{align*}
&\Omega \triangleq \left\{z\left| \norm{S}{0}(z)\geq \frac{1}{2}e^{-c_0} \right. \right\}, \\
&\tilde{\Omega} \triangleq \left\{z\left| \norm{S}{1}(z) \geq \frac{1}{4}e^{-c_0-A} \right. \right\}.
\end{align*}
It follows from definition that $B(x,\epsilon) \subset \Omega$.
In view of the volume element evolution equation,
it is also clear that $\Omega \subset \tilde{\Omega}$.
Note that $S$ is a unit section at time $t=0$.
At time $t=1$, its $L^2$-norm locates in $[e^{-2A}, e^{2A}]$. So we have
\begin{align*}
e^{2A} \geq \int_M \norm{S}{1}^2 dv>|\tilde{\Omega}|_{1} \frac{1}{16}e^{-2c_0-2A}, \quad
\Rightarrow \quad |\tilde{\Omega}|_{1}<16 e^{2c_0+4A}.
\end{align*}
Now we can follow the covering argument in the previous lemma to show a diameter bound of $\tilde{\Omega}$
under the metric $g(1)$. In fact, we can cover $\tilde{\Omega}$ by finite geodesic balls $B(x_i,2\epsilon)$ such that $x_i \in \tilde{\Omega}$
and all different $B(x_i,\epsilon)$'s are disjoint to each other. Clearly, each geodesic ball
$B(x_i,\epsilon)$
has volume at least $\kappa \omega_{2n} \epsilon^{2n}$, where $\kappa=\kappa(n,A)$. Let
$N$ be the number of balls, then
\begin{align*}
N \kappa \omega_{2n} \epsilon^{2n} \leq \sum_{i=1}^{N}|B(x_i,1)| \leq |\tilde{\Omega}| \leq 16e^{2c_0+4A}.
\end{align*}
Therefore, under metric $g(1)$, we obtain
\begin{align*}
\diam \Omega \leq \diam \tilde{\Omega} \leq 4N\epsilon \leq \frac{64e^{2c_0+4A}}{\kappa \omega_{2n}\epsilon^{2n-1}}.
\end{align*}
Recall that $B_{g(0)}(x,\epsilon) \subset \Omega \subset \tilde{\Omega}$, $\epsilon=\frac{e^{-c_0}}{4C_1}$.
So under metric $g(1)$, we have
\begin{align*}
\diam B_{g(0)}(x,\epsilon) \leq \diam \tilde{\Omega} \leq \frac{4^{2n+2}e^{(2n+3)c_0+4A}C_1^{2n-1}}{\kappa \omega_{2n}}.
\end{align*}
Define
\begin{align}
\epsilon_0 \triangleq
\min\left\{\frac{e^{-c_0}}{4C_1}, \frac{\kappa \omega_{2n}}{4^{2n+2}e^{(2n+3)c_0+4A}C_1^{2n-1}} \right\}.
\label{eqn:HA15_3}
\end{align}
Note that $\epsilon_0$ depends only on $n,A$.
Then we have
\begin{align*}
\diam B_{g(0)}(x,\epsilon_0) \leq \epsilon_0^{-1},
\end{align*}
which implies
\begin{align*}
B_{g(0)}(x, \epsilon_0) \subset B_{g(1)}(x,\epsilon_0^{-1}).
\end{align*}
So we finish the proof.
\end{proof}
\begin{lemma}
For every $r$ small, there is a $\delta$ with the following property.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Suppose $|R|+|\lambda|<\delta$ on $M \times [-1,1]$,
then we have
\begin{align}
B_{g(t_1)}(x,\epsilon_0 r) \subset B_{g(t_2)}(x,\epsilon_0^{-1}r)
\label{eqn:HA15_4}
\end{align}
for every $t_1,t_2 \in [-1,1]$. Here $\epsilon_0$ is the constant in Lemma~\ref{lma:HA14_1}.
\label{lma:HA14_2}
\end{lemma}
\begin{proof}
We proceed by a contradiction argument.
Again, it suffices to show (\ref{eqn:HA15_4}) for $t_1=0$ and $t_2=1$. By adjusting $r$ if necessary, we can
also make a rescaling by integer factor. Up to rescaling, (\ref{eqn:HA15_4}) is the same as
\begin{align}
B_{g(0)}(x,\epsilon_0) \subset B_{g(r^{-2})}(x, \epsilon_0^{-1}). \label{eqn:HA15_5}
\end{align}
Suppose the statement of this lemma was wrong. Then there is an $r_0>0$ and a sequence of
points $x_i \in M_i$ such that
\begin{align*}
B_{g_i(0)}(x_i,\epsilon_0) \not\subset B_{g_i(r_0^{-2})}(x_i,\epsilon_0^{-1}).
\end{align*}
However, $|R|+|\lambda| \to 0$ in $C^0$-norm as $i \to \infty$. So we can take a limit
\begin{align*}
(M_i,x_i,g_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M},\bar{x},\bar{g}).
\end{align*}
As usual, we can find a regular point $\bar{z} \in \bar{M}$ near $\bar{x}$. Let $z_i \in M_i$ and $z_i \to \bar{z}$
as the above convergence happens. Then we can extend the above convergence to each time slice.
\begin{align*}
(M_i,z_i,g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M},\bar{z},\bar{g}), \quad \forall \; t \in [0,r_0^{-2}].
\end{align*}
Note that $\bar{x}$ may be a singular point of $\bar{M}$. So in the above convergence, we only have $x_i$
converges to $\bar{x}(t)$, which may depends on time $t$. Lemma~\ref{lma:HA14_1} guarantees that $\bar{x}(t)$
is not at infinity.
It is clear that $\dot{\varphi}$ converges to a limit harmonic function on the regular part of $\bar{M}$,
which must be a constant function by Corollary~\ref{cly:HE08_1}. Apply normalization condition, then the limit function must be
zero on $\bar{M} \times [0,r_0^{-2}]$. Therefore, the limit line bundle $\bar{L}$
admits a limit metric which does not evolve along time. Therefore, for a fixed holomorphic section $\bar{S}$ and a
fixed level value, the level sets of $\norm{\bar{S}}{}^2$ does not depend on time.
Choose $S_i$ be the peak section of $L_i$ at $x_i$, with respect to the metrics at time $t=0$. By the choice of
$\epsilon_0$, it is clear that $\norm{S_i}{}\geq \frac{1}{2}e^{-c_0}$ on the ball $B(x_i,\epsilon_0)$.
In other words, we have
\begin{align*}
B(x_i, \epsilon_0) \subset
\Omega_{i,t} \triangleq \left\{z\left| \norm{S_i}{t}(z)\geq \frac{1}{2}e^{-c_0} \right. \right\}.
\end{align*}
Clearly, each $\Omega_{i,t}$ has uniformly bounded diameter, due to Lemma~\ref{lma:HA15_1}.
Let $\bar{\Omega}$ be the limit set of $\Omega_{i,0}$. Clearly, $\bar{z} \in \bar{\Omega}$.
Then the above discussion implies that $\bar{\Omega}$ is actually the limit set of each $\Omega_{i,t}$.
Let $\bar{y}$ be the limit point of $y_i$, which is a point in $B_{g_i(0)}(x_i, \epsilon_0)$ and start to escape
$B(x_i, \epsilon_0^{-1})$ at time $t_i$, which converges to $\bar{t}$. So we obtain
\begin{align}
\bar{y} \in B_{\bar{g}(0)}(\bar{x},\epsilon_0),
\quad
d(\bar{y}, \bar{x}(\bar{t}))=\epsilon_0^{-1}.
\label{eqn:HA15_6}
\end{align}
Since volume element of the underlying manifold and the line bundle metric are all almost static when time evolves,
it is easy to see that $y_i$ can never escape $\Omega_{i,t}$. So $\bar{y} \in \bar{\Omega}$.
Similarly, we know $\bar{x}(\bar{t}) \in \bar{\Omega}$. Therefore, at time $\bar{t}$, we have
\begin{align*}
d(\bar{y}, \bar{x}(\bar{t})) \leq \diam \bar{\Omega}.
\end{align*}
Note that the argument in the proof of Lemma~\ref{lma:HA15_1} holds for the polarized singular manifold
$(\bar{M}, \bar{L})$, due to the high codimension of $\mathcal{S}(M)$ and the gradient bound of each $S_i$.
Since $\int_{\bar{M}} \norm{\bar{S}}{}^2 dv$ is uniformly bounded from above by $1$, we can follow the proof of
Lemma~\ref{lma:HA15_1} to show that
\begin{align*}
\diam(\bar{\Omega}) \leq \rho(n,\kappa,c_0,C_1,\log 2)<\epsilon_0^{-1}
\end{align*}
by the choice of $\epsilon_0$ in (\ref{eqn:HA15_3}). Consequently, we have $d(\bar{y},\bar{x}(\bar{t}))<\epsilon_0^{-1}$,
which contradicts (\ref{eqn:HA15_6}).
\end{proof}
Based on Lemma~\ref{lma:HA14_2}, we can improve Proposition~\ref{prn:SL04_1}.
Namely, under the condition $|R|+|\lambda| \to 0$, the limit flow is static, even on the singular part.
Clearly, due to Theorem~\ref{thm:SC28_1}, we do not need the assumption of lower bound of polarized
canonical radius anymore.
\begin{proposition}[\textbf{Static limit space-time}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies
\begin{align*}
\lim_{i \to \infty} \sup_{\mathcal{M}_i} (|R|+|\lambda|)=0.
\end{align*}
Suppose $x_i \in M_i$. Then
\begin{align*}
(M_i,x_i, g_i(0)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g}).
\end{align*}
Moreover, we have
\begin{align*}
(M_i,x_i, g_i(t)) \longright{\hat{C}^{\infty}} (\bar{M},\bar{x}, \bar{g})
\end{align*}
for every $t \in (-\bar{T}, \bar{T})$, where $\displaystyle \bar{T}=\lim_{i \to \infty} T_i>0$.
In other words, the identity maps between different time slices converge to the limit identity map.
\label{prn:HE15_1}
\end{proposition}
As a direct application, we obtain the bubble structure of a given family of polarized K\"ahler Ricci flows.
\begin{theorem}[\textbf{Space-time structure of a bubble}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$, $t_i \in (-T_i, T_i)$, and $r_i \to 0$.
Suppose $\widetilde{\mathcal{M}}_i$ is the adjusting of $\mathcal{M}_i$ by shifting time $t_i$ to $0$ and then
rescaling the space-time by the factors $r_i^{-2}$, i.e., $\tilde{g}_i(t)=r_i^{-2}g(r_i^2 t + t_i)$.
Suppose $r_i^{-2} \max \{|t_i-T_i|, |t_i+T_i|\}=\infty$. Then we have
\begin{align*}
(M_i,x_i, \tilde{g}_i(t)) \longright{\hat{C}^{\infty}} (\hat{M}, \hat{x}, \hat{g})
\end{align*}
for each time $t \in (-\infty, \infty)$ with $\hat{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$.
\label{thm:HE26_1}
\end{theorem}
Theorem~\ref{thm:HE26_1} means that the space-time structure of
$\hat{M} \in \widetilde{\mathscr{KS}}(n,\kappa)$ is the model for the space-time structures around $(x_i, t_i)$,
up to proper rescaling. Therefore, Theorem~\ref{thm:HE26_1} is an improvement of Theorem~\ref{thm:SC04_1},
where we only concern the metric structure.
In view of Proposition~\ref{prn:HE15_1},
it is not hard to see that distance is a uniform continuous function of time in $\mathscr{K}(n,A)$.
\begin{theorem}[\textbf{Uniform continuity of distance function}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x,y \in M$.
Suppose $d_{g(0)}(x,y)<1$. Then for every small $\epsilon$, there is a
$\delta=\delta(n,A,\epsilon)$ such that
\begin{align*}
|d_{g(t)}(x,y)-d_{g(0)}(x,y)|<\epsilon
\end{align*}
whenever $|t|<\delta$.
\label{thm:HE15_1}
\end{theorem}
\begin{proof}
We argue by contradiction. Suppose the statement was wrong, we can find an $\bar{\epsilon}>0$ and a sequence of
flows violating the statement for time $|t_i| \to 0$. Around $x_i$,
in the ball $B_{g_i(0)}\left(x_i,\frac{\epsilon_0^2\bar{\epsilon}}{10} \right)$,
we can find $x_i'$ which are uniform regular at time $t=0$, where $\epsilon_0$
is the same constant in Lemma~\ref{lma:HA14_2} and Lemma~\ref{lma:HA14_1}.
By two-sided pseudolocality, Theorem~\ref{thm:SL27_2}, it is clear that $x_i'$ is also
uniform regular at time $t=t_i$. Similarly, we can choose $y_i'$.
By virtue of triangle inequality and Lemma~\ref{lma:HA14_2}, we obtain
\begin{align*}
& d_{g_i(0)}(x_i', y_i') -\frac{\epsilon_0^2 \bar{\epsilon}}{5}
\leq d_{g_i(0)}(x_i, y_i) \leq d_{g_i(0)}(x_i', y_i') + \frac{\epsilon_0^2 \bar{\epsilon}}{5},\\
& d_{g_i(t_i)}(x_i', y_i') -\frac{\bar{\epsilon}}{5}
\leq d_{g_i(t_i)}(x_i, y_i) \leq d_{g_i(t_i)}(x_i', y_i') + \frac{\bar{\epsilon}}{5}.
\end{align*}
By argument similar to that in Proposition~\ref{prn:SC17_4}, it is clear that
\begin{align*}
\lim_{i \to \infty} d_{g_i(t_i)}(x_i', y_i')= \lim_{i \to \infty} d_{g_i(0)}(x_i', y_i').
\end{align*}
Then it follows that
\begin{align*}
\lim_{i \to \infty} d_{g_i(0)}(x_i,y_i) -\frac{(1+\epsilon_0^2)}{5}\bar{\epsilon}
\leq
\lim_{i \to \infty} d_{g_i(t_i)}(x_i, y_i) \leq \lim_{i \to \infty} d_{g_i(0)}(x_i,y_i) + \frac{(1+\epsilon_0^2)}{5} \bar{\epsilon}.
\end{align*}
In particular, for large $i$, we have
\begin{align*}
\left| d_{g_i(0)}(x_i,y_i) - d_{g_i(t_i)}(x_i, y_i) \right|<\frac{(1+\epsilon_0^2)}{5} \bar{\epsilon}<\bar{\epsilon},
\end{align*}
which contradicts our assumption.
\end{proof}
\subsection{Volume estimate for high curvature neighborhood}
In this subsection, we shall develop the flow version of the volume estimate of
Donaldson and the first author(c.f.~\cite{CD1},~\cite{CD2}, see also~\cite{CN}).
\begin{proposition}[\textbf{K\"ahler cone complex splitting}]
Same condition as in Proposition~\ref{prn:HA07_2}, $\bar{y} \in \bar{M}$. Suppose
$\hat{Y}$ is a tangent cone of $\bar{M}$ at $\bar{y}$, then there is a fixed nonnegative integer $k$
such that
\begin{align}
\hat{Y}=C(Z) \times \ensuremath{\mathbb{C}}^{n-k},
\label{eqn:HA05_1}
\end{align}
where $C(Z)$ is a metric cone without straight line. A point in $\bar{M}$ is regular if and only if one
of the tangent cone is $\ensuremath{\mathbb{C}}^{n}$.
\label{prn:SL26_1}
\end{proposition}
\begin{proof}
By definition of tangent cone, one can find a sequence of numbers $r_i \to 0$. Taking subsequence if necessary,
let $\tilde{g}_i(t)=r_i^{-2} g_i(r_i^2t)$, then we have
\begin{align*}
(M_i, y_i, \tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{Y}, \hat{y}, \hat{g}).
\end{align*}
By compactness, we see that $(\hat{Y},\hat{y},\hat{g}) \in \widetilde{\mathscr{KS}}(n,\kappa)$.
On the other hand, it is a metric cone, which is the tangent space of itself at the origin.
So $\hat{Y}$ has the decomposition (\ref{eqn:HA05_1}),
by Theorem~\ref{thm:HE08_1} or Lemma~\ref{lma:HD30_1}.
\end{proof}
\begin{proposition}[\textbf{K\"ahler tangent cone rigidity}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$.
Suppose $x_i \in M_i$ and $(\bar{M}, \bar{x}, \bar{g})$ is a limit space of $(M_i, x_i, g_i(0))$.
Let $\hat{Y}$ be
a tangent space of $\bar{M}$. Then $\hat{Y}$ satisfies the splitting (\ref{eqn:HA05_1}) for $k=2$ or $k=0$.
\label{prn:HA10_2}
\end{proposition}
\begin{proof}
Clearly, $k=0$ if and only if the base point is regular. So it suffices to show that for every singular tangent space we have
$k=2$. By Proposition~\ref{prn:SL26_1}, we only need to rule out the case $k \geq 3$. However, this follows from the rigidity of complex structure on the smooth annulus in $\ensuremath{\mathbb{C}}^k /\Gamma$, where $\Gamma$ is a finite group of holomorphic isometry
of $\ensuremath{\mathbb{C}}^k$, when $k \geq 3$. Note that $[\omega_i]=c_1(L_i)$, which is an integer class.
Therefore, the proof follows verbatim as that in~\cite{CD2}.
\end{proof}
\begin{proposition}[\textbf{Existence of holomorphic slicing}]
Suppose $\hat{Y} \in \widetilde{\mathscr{KS}}(n,\kappa)$ is a metric cone satisfying the splitting (\ref{eqn:HA05_1}).
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in M$.
If $(M,x,g(0))$ is very close to $(\hat{Y}, \hat{y}, \hat{g})$, i.e.,
\begin{align*}
d_{GH}((M,x,g(0)), (\hat{Y}, \hat{y}, \hat{g}))<\epsilon
\end{align*}
for sufficiently small $\epsilon$, which depends on $n,A,\hat{Y}$, then there exists a holomorphic map
\begin{align*}
\Psi=(u_{k+1}, u_{k+2}, \cdots, u_n): B(x, 10) \mapsto \ensuremath{\mathbb{C}}^{n-k}
\end{align*}
satisfying
\begin{align}
& |\nabla \Psi| \leq C(n,A), \label{eqn:HA10_1}\\
& \sum_{k+1 \leq i,j\leq n}
\int_{B(x,10)} \left| \delta_{ij} - \langle \nabla u_i, \nabla u_j \rangle\right|dv
\leq \eta(n,A,\epsilon), \label{eqn:HA10_2}
\end{align}
where $\eta$ is a small number such that $\displaystyle \lim_{\epsilon \to 0} \eta=0$.
\label{prn:HA10_1}
\end{proposition}
\begin{proof}
It follows from the argument in~\cite{DS} that the constant section $1$ of the trivial bundle over $\hat{Y}$ can be
``pulled back" as a non vanishing holomorphic section of $L$ over $B(x,10)$, up to a finite lifting of power of $L$.
Therefore, we can regard $L$ as a trivial bundle over $B(x,10)$ without loss of generality.
Let $S_0$ be the pull-back of the constant $1$ section. In particular, $S_0$ is a non-vanishing
holomorphic section on $B(x,10)$.
On $B(x,10)$, every holomorphic section $S$ of $L$ can be written as $S=u S_0$ for a holomorphic function $u$ and
$\norm{S}{h}^2= \snorm{u}{}^2 \norm{S_0}{h}^2$.
From the splitting (\ref{eqn:HA05_1}), there exist natural coordinate holomorphic functions
$\{z_j\}_{j=k+1}^{n}$ on $\hat{Y}$.
Same as~\cite{DS}, one can apply H\"ormander's estimate
to construct $\{S_j\}_{j=k+1}^{n}$, which are holomorphic sections of $L$.
Each $S_j$ can be regarded as an ``approximation" of $z_j$, although they have different base spaces.
Let $u_j=\frac{S_j}{S_0}$ for each $j \in \{k+1, \cdots, n\}$.
Then we can define a holomorphic map $\Psi$ from $B(x,10)$ to $\ensuremath{\mathbb{C}}^{n-k}$ as follows
\begin{align*}
\Psi(y) \triangleq \left( u_{k+1}, u_{k+2}, \cdots, u_n \right).
\end{align*}
Note that each $S_i$ is a holomorphic section of $L$ with $L^2$-norm bounded from two sides, according to its
construction. It is easy to check that $\norm{\nabla S_i}{h}^2$ satisfies a sub-elliptic system. So there exists a uniform
bound $\norm{\nabla S_i}{h}^2 \leq C(n,A)$, which implies (\ref{eqn:HA10_1}) when restricted on $B(x,10)$.
Moreover, on $B(x,10)$, by smooth convergence, it is not hard to see that $ \langle \nabla u_i, \nabla u_j \rangle$ can pointwisely
approximate $\delta_{ij}$ away from singularities of $\hat{Y}$, in any accuracy level when $\epsilon \to 0$.
This approximation together with (\ref{eqn:HA10_1}) yields (\ref{eqn:HA10_2}).
\end{proof}
\begin{theorem}[\textbf{Weak monotonicity of curvature integral}]
There exists a small constant $\epsilon=\epsilon(n,A)$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Suppose $x \in M$, $0<r \leq 1$.
Then under the metric $g(0)$, we have
\begin{align}
\sup_{B(x, \frac{1}{2}r)} |Rm| \leq r^{-2}
\label{eqn:SC13_3}
\end{align}
whenever $r^{4-2n}\int_{B(x,r)} |Rm|^2 dv \leq \epsilon$.
\label{thm:SC12_3}
\end{theorem}
\begin{proof}
Up to rescaling, we can assume $r=1$ without loss of generality. If the statement was wrong, we can find a sequence of
points $x_i \in M_i$ such that
\begin{align*}
\int_{B(x_i,1)} |Rm|^2 dv \to 0, \quad
\sup_{B(x_i,\frac{1}{2})} |Rm| \geq 1,
\end{align*}
where the default metric is $g_i(0)$, the time zero metric of a flow $g_i$, in the moduli space $\mathscr{K}(n,A)$.
By the smooth convergence at places when curvature uniformly bounded, it is clear that the above conditions imply that
\begin{align*}
\int_{B(x_i,1)} |Rm|^2 dv \to 0, \quad
\sup_{B(x_i,\frac{3}{4})} |Rm| \to \infty.
\end{align*}
Let $(\bar{M},\bar{x},\bar{g})$ be the limit space of $(M_i,x_i,g_i(0))$. Then $B(\bar{x}, \frac{3}{4})$ contains at least one
singularity $\bar{y}$. Without loss of generality, we can assume $\bar{x}$ is a singular point.
Note that $B(\bar{x}, \frac{1}{4})$ is a flat manifold away from singularities. So every tangent space
of $\bar{M}$ at $\bar{x}$ is a flat metric cone. Let $\hat{Y}$ be one of such a flat metric cone. By taking subsequence if necessary, we can assume
\begin{align*}
(M_i,x_i,\tilde{g}_i(0)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\hat{Y}, \hat{x},\hat{g}),
\end{align*}
for some flow metrics $\tilde{g}_i$ satisfying $\displaystyle \tilde{g}_i(t)=r_i^{-2} g_i(r_i^2 t), \; r_i \to 0$. Since $\hat{Y}$ is
a flat metric cone, in light of Proposition~\ref{prn:HA10_2}, we have the splitting
\begin{align*}
\hat{Y}= (\ensuremath{\mathbb{C}}^2/\Gamma) \times \ensuremath{\mathbb{C}}^{n-2}.
\end{align*}
Let $(M,x,\tilde{g})$ be one of $(M_i,x_i,\tilde{g}_i(0))$ for some large $i$. Because of Proposition~\ref{prn:HA10_1}, we can construct a holomorphic map $\Psi:B(x,10) \to \ensuremath{\mathbb{C}}^{n-2}$
satisfying (\ref{eqn:HA10_1}) and (\ref{eqn:HA10_2}).
Then we can follow the slice argument as in~\cite{CCT} and~\cite{Cheeger03}.
Our argument will be simpler since our slice functions are holomorphic rather than harmonic.
Actually, for generic $\vec{z}=(z_{3}, z_{4}, \cdots, z_n)$ satisfying $|\vec{z}|<0.1$,
we know $\Psi^{-1}(\vec{z}) \cap B(x,5)$ is a complex surface with boundary.
Clearly, $\Psi^{-1}((S^3/\Gamma) \times \{\vec{z}\})$ is close to $(S^3/\Gamma) \times \vec{z}$,
if we regard $S^3/\Gamma$ as the unit sphere in $\ensuremath{\mathbb{C}}^2/\Gamma$. Deform the preimage a little bit if necessary, we can obtain
a $\partial \Omega$ which bounds a complex surface $\Omega$. By coarea formula and the bound of
$|\nabla \Psi|$, it is clear that for generic $\Omega$ obtained in this way, we have
\begin{align*}
\int_{\Omega} |Rm|^2 d\sigma\to 0.
\end{align*}
Consider the restriction of $TM$ on $\Omega$. Let $c_2$ be a form representing the second Chern class of the tangent bundle $TM$, obtained from the K\"ahler metric $\tilde{g}(0)$ from the classical way.
Let $\hat{c}_2$ be the corresponding differential character with value in $\ensuremath{\mathbb{R}}/\ensuremath{\mathbb{Z}}$. Since the pointwise norm of $c_2$ is bounded by $|Rm|^2$, it is clear that
\begin{align}
\hat{c}_2(\partial \Omega)=\int_{\Omega} c_2 \quad (mod \; \ensuremath{\mathbb{Z}}) \to 0. \label{eqn:HA10_3}
\end{align}
On the other hand, since $\partial \Omega$ converges to $S^3/\Gamma$, we have
\begin{align}
\hat{c}_2(\Omega) \to \frac{1}{|\Gamma|}. \label{eqn:HA10_4}
\end{align}
Therefore, the combination of (\ref{eqn:HA10_3}) and (\ref{eqn:HA10_4}) forces that $|\Gamma|=1$. This is impossible
since $|\Gamma| \geq 2$ by our assumption that $\hat{Y}$ is a singular metric cone.
\end{proof}
From now on to the end of this subsection, we use $g(0)$ as the default metric.
Similar to the definition in~\cite{CD1}, for any small $r$, let $\mathcal{Z}_r$ be the $r$-neighborhood of the points where $|Rm| > r^{-2}$.
Recall the definition equation (\ref{eqn:SL27_1}), we denote $\mathcal{F}_r$ as the collection of points
whose canonical volume radii are greater than $r$, $\mathcal{D}_r$ as the component of $\mathcal{F}_r$.
Under these notations, we have the following property.
\begin{proposition}[\textbf{Equivalence of singular neighborhoods}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $0<r<\hslash$.
Then at time zero, we have
\begin{align}
\mathcal{D}_{cr} \subset \mathcal{Z}_r \subset \mathcal{D}_{\frac{1}{c}r}
\label{eqn:HA11_1}
\end{align}
for some small constant $c=c(n,A)$.
\label{prn:HA11_1}
\end{proposition}
\begin{proof}
Let us first prove $\mathcal{D}_{cr} \subset \mathcal{Z}_r$. Suppose the statement was wrong, we can find a sequence $c_i \to 0$
and flows in $\mathscr{K}(n,A)$ such that $\mathcal{D}_{c_ir_i} \not\subset Z_{r_i}$ for some $r_i <\hslash$.
Choose $x_i \in \mathcal{D}_{c_i r_i} \cap \mathcal{Z}_{r_i}^{c}$. Let $\rho_i$ be the canonical volume radius of $x_i$. Rescale
the flow such that the canonical volume radius at $x_i$ becomes $1$. Take limit, we will obtain a smooth flat space in
$\widetilde{\mathscr{KS}}(n,\kappa)$, which is nothing but $\ensuremath{\mathbb{C}}^n$. Therefore, the canonical volume radii
of the base points $x_i$ should tend to infinity, which is a contradiction.
Then we prove $\mathcal{Z}_r \subset \mathcal{D}_{\frac{1}{c}r}$. Suppose $x \in \mathcal{Z}_r$, then $|Rm|(y) \geq r^{-2}$ for some $y \in B(x,r)$. By the regularity improving property of canonical volume radius, it is clear
that $\mathbf{cvr}(x) \leq \frac{2}{c_a} r$. In other words, $x \in \mathcal{D}_{\frac{2}{c_a}r}$.
\end{proof}
\begin{theorem}[\textbf{Volume estimates of high curvature neighborhood}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Under the metric $g(0)$, we have
\begin{align*}
|\mathcal{Z}_r| \leq C r^4,
\end{align*}
where $C$ depends on $n,A$ and the upper bound of $\int_M |Rm|^2 dv$.
\label{thm:HA11_1}
\end{theorem}
\begin{proof}
Because of Proposition~\ref{prn:HA11_1}, it suffices to show $|\mathcal{D}_{cr}| \leq C r^4$.
In light of Theorem~\ref{thm:SC12_3}, if $r^{4-2n}\int_{B(x,r)} |Rm|^2 dv<\epsilon$ for some
$r<\hslash$, then $x \in \mathcal{F}_{cr}$. In other words, if $x \in \mathcal{D}_{cr}(M,0)$, then it is forced that
\begin{align*}
r^{4-2n} \int_{B(x,r)} |Rm|^2 dv \geq \epsilon.
\end{align*}
Let $\bigcup_{i=1}^{N} B(x_i,2r)$ be a finite cover of $\mathcal{D}_{cr}$ such that
\begin{itemize}
\item $x_i \in \mathcal{D}_{cr}$.
\item $B(x_i,r)$ are disjoint to each other.
\end{itemize}
Then we can bound $N$ as follows.
\begin{align*}
N\epsilon r^{2n-4} \leq \sum_{i=1}^{N} \int_{B(x_i,r)} |Rm|^2 dv \leq \int_M |Rm|^2 dv \leq H.
\end{align*}
Consequently, we have
\begin{align*}
|\mathcal{D}_{cr}(M)| \leq \sum_{i=1}^{N} |B(x_i, 2r)| \leq \frac{H}{\epsilon} r^{4-2n} \kappa^{-1} \omega_{2n} (2r)^{2n}
\leq C r^4.
\end{align*}
Since both $\kappa$ and $\epsilon$ depends only on $n$ and $A$. It is clear that $C=C(n,A,H)$ where $H$
is the upper bound of $\int_M |Rm|^2 dv$.
\end{proof}
\begin{corollary}[\textbf{Volume estimates of singular neighborhood}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$.
Suppose $\int_{M_i} |Rm|^2 dv \leq H$ uniformly under the metric $g_i(0)$. Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit
space of $(M_i,x_i,g_i(0))$. Let $\mathcal{S}_r$ be the set defined in (\ref{eqn:HA11_4}), then we have
\begin{align*}
|\mathcal{S}_{r}| \leq C r^4
\end{align*}
for each small $r$ and some constant $C=C(n,A,H)$.
In particular, we have the estimate of Minkowski dimension of the singularity
\begin{align*}
\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4.
\end{align*}
\label{cly:HA11_1}
\end{corollary}
Following~\cite{CW4}, the space $\bar{M}=\mathcal{R} \cup \mathcal{S}$
is called a metric-normal $Q$-Fano variety if there exists
a homeomorphic map $\varphi: \bar{M} \to Z$ for some $Q$-Fano normal variety $Z$ such that $\varphi|_{\mathcal{R}}$
is a biholomorphic map. Moreover, $\dim_{\mathcal{M}} \mathcal{S} \leq 2n-4$.
\begin{theorem}[\textbf{Limit structure}]
Suppose that $\mathcal{LM}_i \in \mathscr{K}(n,A)$.
Under the metric $g_i(0)$, suppose
\begin{align}
\Vol(M_i) + \int_{M_i} |Rm|^2 dv \leq H \label{eqn:HA11_2}
\end{align}
for some uniform constant $H$.
Let $(\bar{M}, \bar{x}, \bar{g})$ be the limit space of $(M_i,x_i,g_i(0))$.
Then $\bar{M}$ is a compact metric-normal $Q$-Fano variety.
\label{thm:SC13_4}
\end{theorem}
\begin{proof}
It follows from (\ref{eqn:HA11_2}) and the non-collapsing that $diam(M_i)$ is uniformly bounded.
So the limit space $\bar{M}$ is compact. Due to Theorem~\ref{thmin:HC08_1}, the partial $C^0$-estimate,
one can follow the argument in~\cite{DS} to show that $\bar{M}$ is a $Q$-Fano, normal variety.
The metric-normal property follows from Corollary~\ref{cly:HA11_1}.
\end{proof}
Based on the estimates developed in this subsection, we can easily prove Corollary~\ref{clyin:SC24_4}
and Corollary~\ref{clyin:SC24_5} in the introduction.
\begin{proof}[Proof of Corollary~\ref{clyin:SC24_4} and Corollary~\ref{clyin:SC24_5}]
It follows from the combination of Theorem~\ref{thm:SC13_4}, Corollary~\ref{cly:HA11_1} of this paper and
main results in~\cite{CW4}. Note that the line bundle metric choice in this paper is
equivalent to that in~\cite{CW4}, due to the bound of $\dot{\varphi}$.
\end{proof}
\subsection{Singular K\"ahler Ricci flow solution}
In this subsection, we shall relate the different limit time slices, without the assumption of $|R|+|\lambda| \to 0$.
We shall further improve regularity, by estimates essentially arising from complex analysis of holomorphic sections.
We want to compare $\omega_t$, the K\"ahler Ricci flow metrics, and $\tilde{\omega}_t$, the evolving Bergman metrics.
We first show that $\tilde{\omega}_t$ is very stable when $t$ evolves.
\begin{lemma}
Suppose $G(t)$ is a family of $(N+1)\times (N+1)$ matrices parameterized by $t \in [-1,1]$. Suppose $G(0)=Id$, $\dot{G}(0)=B$.
Let $\lambda_0\leq \lambda_1 \leq \cdots \leq \lambda_N$ be the real eigenvalues of the Hermitian matrix $B+\bar{B}^{\tau}$.
If we regard $G$ as a holomorphic map from $\ensuremath{\mathbb{CP}}^N$ to $\ensuremath{\mathbb{CP}}^N$, then we have
\begin{align}
(\lambda_0-\lambda_N) \omega_{FS} \leq
\left. \frac{d}{dt}G(t)^*(\omega_{FS}) \right|_{t=0} \leq (\lambda_N-\lambda_0) \omega_{FS}.
\label{eqn:HB02_1}
\end{align}
\label{lma:HB02_1}
\end{lemma}
\begin{proof}
Let $\{z_i\}_{i=0}^N$ be the homogeneous coordinate of $\ensuremath{\mathbb{CP}}^N$. Let $G=G(t)$. Then we have
\begin{align*}
&\omega_{FS}=\sqrt{-1} \partial \bar{\partial} \log (|z_0|^2+|z_1|^2+\cdots |z_N|^2)
=\sqrt{-1} \left\{\frac{\partial z_i \wedge \bar{\partial} \bar{z}_i}{|z|^2}
+\frac{(z_i \bar{\partial} \bar{z}_i) \wedge (\bar{z}_j \partial z_j)}{|z|^4}\right\}, \\
&G^*(\omega_{FS})=\sqrt{-1} \partial \bar{\partial} \log (|\tilde{z}_0|^2+|\tilde{z}_1|^2+\cdots |\tilde{z}_N|^2)
=\sqrt{-1} \left\{\frac{\partial \tilde{z}_i \wedge \bar{\partial} \bar{\tilde{z}}_i}{|\tilde{z}|^2}
+\frac{(\tilde{z}_i \bar{\partial} \bar{\tilde{z}}_i) \wedge (\bar{\tilde{z}}_j \partial \tilde{z}_j)}{|\tilde{z}|^4}\right\},
\end{align*}
where $\tilde{z}_i=G_{ij}z_j$. Let $\{ w_1, \cdots, w_N\}$ be local coordinate.
At point $z$, the matrix of $\omega_{FS}$ is
\begin{align*}
E_0=J \left(\frac{Id}{|z|^2}-\frac{\bar{z}^{\tau}z}{|z|^4}\right) \bar{J}^{\tau}=J F_0 \bar{J}^{\tau},
\end{align*}
where $J$ is an $N \times (N+1)$ matrix which is the Jacobi matrix $\left(\D{z_j}{w_{\alpha}}\right)$.
The matrix of
$\omega_{G^*\omega_{FS}}$ is
\begin{align*}
E_t=J G\left(\frac{Id}{|\tilde{z}|^2}-\frac{\bar{\tilde{z}}^{\tau} \tilde{z}}{|\tilde{z}|^4}\right) \bar{G}^{\tau} \bar{J}^{\tau}
=JF_t \bar{J}^{\tau}.
\end{align*}
Clearly, we have
\begin{align*}
\left. \frac{d}{dt} F_t \right|_{t=0}&=B\left(\frac{Id}{|z|^2}-\frac{\bar{z}^{\tau}z}{|z|^4}\right)
+\left(\frac{Id}{|z|^2}-\frac{\bar{z}^{\tau}z}{|z|^4}\right) \bar{B}^{\tau}
-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} Id +\frac{2z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^6} \bar{z}^{\tau}z
-\frac{\bar{B}^{\tau}\bar{z}^{\tau}z+\bar{z}^{\tau}zB}{|z|^4}\\
&=\left\{ \frac{B+\bar{B}^{\tau}}{|z|^2} -\frac{(B+\bar{B}^{\tau})\bar{z}^{\tau}z + \bar{z}^{\tau}z(B+\bar{B}^{\tau})}{|z|^4}
+\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^6} \bar{z}^{\tau}z \right\}
-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} F_0\\
&\triangleq M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} F_0.
\end{align*}
It follows that
\begin{align*}
\left. \frac{d}{dt} E_t \right|_{t=0}=
\left. \frac{d}{dt}
\left\{J G\left(\frac{Id}{|\tilde{z}|^2}-\frac{\bar{\tilde{z}}^{\tau} \tilde{z}}{|\tilde{z}|^4}\right) \bar{G}^{\tau} \bar{J}^{\tau} \right\}
\right|_{t=0}=J\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{J}^{\tau}.
\end{align*}
It is easy to check that
\begin{align*}
zM\bar{z}^{\tau}=0, \quad zF_0\bar{z}^{\tau}=0, \quad
z\left( M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^4} F_0\right)\bar{z}^{\tau}=0.
\end{align*}
Without loss of generality, we can assume $B+\bar{B}^{\tau}$ is a diagonal matrix
$\diag(\lambda_0, \lambda_1, \cdots, \lambda_N)$.
Let $v=(v_0, v_1, \cdots, v_N)$ be a vector in $\ensuremath{\mathbb{C}}^{N+1}$ satisfying
\begin{align*}
z\bar{v}^{\tau}=\bar{v}_0z_0 +\bar{v}_1z_1 + \cdots + \bar{v}_N z_N=0.
\end{align*}
Then it is clear that
\begin{align*}
vM\bar{v}^{\tau}=\frac{v(B+\bar{B}^{\tau})\bar{v}^{\tau}}{|z|^2}, \quad
vF_0\bar{v}^{\tau}=\frac{|v|^2}{|z|^2}.
\end{align*}
Therefore, we have
\begin{align*}
&\quad v\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{v}^{\tau}\\
&=\frac{1}{|z|^4}\left\{ (\lambda_0 |v_0|^2+\cdots+\lambda_N |v_N|^2)(|z_0|^2+\cdots+|z_N|^2)
-(\lambda_0|z_0|^2+\cdots \lambda_N|z_N|^2)(|v_0|^2+\cdots+|v_N|^2)\right\}\\
&=\frac{1}{|z|^4} \left\{ \left[(\lambda_0-\lambda_0)|z_0|^2 +(\lambda_0-\lambda_1)|z_1|^2+\cdots
+(\lambda_0-\lambda_N)|z_N|^2 \right]|v_0|^2 \right.\\
&\quad + \left[(\lambda_1-\lambda_0)|z_0|^2 +(\lambda_1-\lambda_1)|z_1|^2+\cdots
+(\lambda_1-\lambda_N)|z_N|^2 \right]|v_1|^2 \\
&\quad +\cdots\\
&\quad \left.+\left[(\lambda_N-\lambda_0)|z_0|^2 +(\lambda_N-\lambda_1)|z_1|^2+\cdots
+(\lambda_N-\lambda_N)|z_N|^2 \right]|v_N|^2 \right\}\\
&\leq (\lambda_N-\lambda_0) \frac{|v|^2}{|z|^2}.
\end{align*}
Similarly, we have
\begin{align*}
v\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{v}^{\tau} \geq (\lambda_0-\lambda_N)\frac{|v|^2}{|z|^2}.
\end{align*}
Note that $z F_0 \bar{v}^{\tau}=0$. Therefore, we can apply the orthogonal decomposition with respect to $F_0$ to obtain that
for every vector $f=(f_0,f_1,\cdots, f_N) \in \ensuremath{\mathbb{C}}^{N+1}$, we have
\begin{align*}
\left| f\left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\bar{f}^{\tau} \right|
\leq (\lambda_N-\lambda_0) \frac{|f|^2}{|z|^2}=(\lambda_N-\lambda_0)fF_0\bar{f}^{\tau}.
\end{align*}
Let $u \in T_z^{(1,0)}\ensuremath{\mathbb{CP}}^N$. Then we have
\begin{align*}
\langle u,u \rangle_{\omega_{FS}}&=(uJ) F_0 \overline{(uJ)}^{\tau}\\
\left| \langle u, u \rangle_{\left.\frac{d}{dt}G^*(\omega_{FS}) \right|_{t=0}} \right|
&=\left|(uJ) \left(M-\frac{z(B+\bar{B}^{\tau})\bar{z}^{\tau}}{|z|^2} F_0 \right)\overline{uJ}^{\tau} \right|\\
&\leq (\lambda_N-\lambda_0) (uJ) F_0 \overline{(uJ)}^{\tau}\\
&\leq (\lambda_N-\lambda_0)\langle u,u \rangle_{\omega_{FS}}.
\end{align*}
By the arbitrary choice of $u$, then (\ref{eqn:HB02_1}) follows directly from the above inequality.
\end{proof}
\begin{lemma}
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$.
Let $\tilde{\omega}_t$ be the pull back of the Fubini-Study metric by orthonormal basis of $L$ with respect to $\omega_t$ and $h_t$.
Then we have the evolution inequality of $\tilde{\omega}_t$:
\begin{align}
-2A\tilde{\omega}_t \leq \frac{d}{dt} \tilde{\omega_t} \leq 2A \tilde{\omega}_t.
\label{eqn:HB03_1}
\end{align}
\label{lma:HB05_1}
\end{lemma}
\begin{proof}
Without loss of generality, it suffices to show (\ref{eqn:HB03_1}) at time $t=0$. \\
Suppose $\left\{ s_i \right\}_{i=0}^{N}$ is an orthonormal basis at time $0$, $\left\{ \tilde{s}_i \right\}_{i=0}^{N}$
is an orthonormal basis at time $t$. They are related by $\tilde{s}_i=s_jG_{ji}$. Fix $e_L$ a local representation of the line bundle $L$ around a point $x$ so that locally we have $s_j=z_j e_L$ and $\tilde{s}_j=\tilde{z}_j e_L=z_j G_{ji}e_L$. Then we have
\begin{align*}
& \tilde{\omega}_0=\sqrt{-1} \partial \bar{\partial} \log (|z_0|^2+|z_1|^2+\cdots +|z_N|^2), \\
& \tilde{\omega}_t=\sqrt{-1} \partial \bar{\partial} \log (|\tilde{z}_0^2|+|\tilde{z}_1|^2+\cdots +|\tilde{z}_N|^2).
\end{align*}
Let $\iota$ be the Kodaira embedding map induced by $\left\{ s_i \right\}_{i=0}^{N}$ at time $0$. Then it is clear that
\begin{align*}
\tilde{\omega}_0=\iota^* \omega_{FS}, \quad \tilde{\omega}_t=\iota^* (G^*\omega_{FS}).
\end{align*}
Therefore, we have
\begin{align*}
\left. \frac{d}{dt} \tilde{\omega}_t \right|_{t=0}=\iota^*\left( \left. \frac{d}{dt}G^*(\omega_{FS}) \right|_{t=0} \right).
\end{align*}
So (\ref{eqn:HB03_1}) is reduced to the estimate
\begin{align}
-2A \omega_{FS} \leq \left. \frac{d}{dt}G(t)^*(\omega_{FS}) \right|_{t=0} \leq 2A \omega_{FS}.
\label{eqn:HB05_5}
\end{align}
However, note that
\begin{align*}
\delta_{ik}=G_{ij}\bar{G}_{kl} \int_{M} \langle s_j, s_l\rangle_{h_t} \frac{\omega_t^n}{n!}.
\end{align*}
Take derivative on both sides at time $0$ and denote $\dot{G}$ by $B$, we obtain
\begin{align*}
0=B_{ik}+\bar{B}_{ki} + \int_M (-\dot{\varphi}+n\lambda-R)\langle s_i,s_k\rangle_{h_0} \frac{\omega_0^n}{n!}.
\end{align*}
Therefore, for every $v\in \ensuremath{\mathbb{C}}^{N+1}$, the following inequality holds.
\begin{align}
\left| v_i (B_{ij} + \bar{B}_{ji}) \bar{v}_j \right| =\left|-v_i\bar{v}_j \int_M (-\dot{\varphi}+n\lambda-R)\langle s_i,s_j\rangle_{h_0} \frac{\omega_0^n}{n!} \right|\leq A|v|^2.
\label{eqn:HB05_1}
\end{align}
In particular, each eigenvalue of the Hermitian matrix $B+\bar{B}^{\tau}$ has absolute value bounded by $A$. Then (\ref{eqn:HB05_5}) follows from
Lemma~\ref{lma:HB02_1}.
\end{proof}
In view of Lemma~\ref{lma:HB05_1}, the following property is obvious now.
\begin{proposition}[\textbf{Bergman metric equivalence along time}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$. Then we have
\begin{align}
e^{-2A|t|} \tilde{\omega}_0 \leq \tilde{\omega}_t \leq e^{2A|t|} \tilde{\omega}_0.
\label{eqn:HB03_2}
\end{align}
\label{prn:HB05_2}
\end{proposition}
In general, we cannot hope a powerful estimate like (\ref{eqn:HB03_2}) holds for metrics $\omega_t$, since such an estimate will imply the Ricci curvature is uniformly bounded by $A$. However, if we only focus on points regular enough, then we do have a similar weaker estimate.
\begin{proposition}[\textbf{Flow metric equivalence along time}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $x \in \mathcal{F}_r(M,0)$. Then we have
\begin{align}
\frac{1}{C} \omega_0(x) \leq \omega_t(x) \leq C\omega_0(x) \label{eqn:HB05_2}
\end{align}
for every $t \in [-1,1]$. Here $C$ is a constant depending only on $n,A$ and $r$.
\label{prn:HB05_1}
\end{proposition}
\begin{proof}
Without loss of generality, we assume $\mathbf{b}$ is uniformly bounded.
By short time two-sided pseudolocality, Theorem~\ref{thm:SL27_2} and rescaling, it suffices to show (\ref{eqn:HB05_2}) for $t=-1$ and $t=1$.
At time $0$, it is clear that $\omega_0(x)$ and $\tilde{\omega}_0(x)$ are uniformly equivalent. The volume form
$\omega_0^n$ is uniformly equivalent to $\omega_t^n$. By the stability of $\tilde{\omega}$, inequality (\ref{eqn:HB03_2}), it suffices to prove
the following two inequalities hold at point $x$.
\begin{align}
& \Lambda_{\omega_1} \tilde{\omega}_0 \leq C, \label{eqn:HB05_3}\\
& \Lambda_{\omega_0} \tilde{\omega}_{-1} \leq C. \label{eqn:HB05_4}
\end{align}
We shall prove the above two inequalities separately.
Let $w_0$ be defined as that before Lemma~\ref{lma:SK23_1}.
Let $w$ be the solution of $\square w=0$, initiating from $w_0$. By the heat kernel estimate and the uniform upper
bound of diameter of $B_{g(0)}(x,r)$ under metric $g(t)$(c.f. Lemma~\ref{lma:HA14_1}), we see that $w(x,1)$ is uniformly bounded away from $0$.
Then Lemma~\ref{lma:J02_1} applies and we obtain that
\begin{align*}
\Lambda_{\omega_1(x)} \tilde{\omega}_0(x) =F(x,1) \leq \frac{C}{w(x,1)} <C.
\end{align*}
So we finish the proof of (\ref{eqn:HB05_3}). The proof of (\ref{eqn:HB05_4}) is similar. Modular time shifting, the only difference is that we do not know
whether $x$ is very regular at time $t=-1$, so the construction of initial value of a heat equation may be a problem.
However, due to Proposition~\ref{prn:SC02_2}, we can always find a point $y_0 \in \mathcal{F}_{c_b \hslash}(M,-1) \cap B_{g(-1)}(x,\hslash)$. Consider the heat equation $w'$, starting from a cutoff function supported around $y_0$ at time $t=-1$.
In light of uniform diameter bound of $B_{g(-1)}(y_0,\hslash)$ under the metric $g(0)$, $w'(x,0)$ is uniformly bounded away from $0$. So we can follow the proof of
Lemma~\ref{lma:J02_1} to obtain that
\begin{align*}
\Lambda_{\omega_0(x)} \tilde{\omega}_{-1}(x) <\frac{C}{w'(x,0)}<C.
\end{align*}
Therefore, (\ref{eqn:HB05_4}) is proved.
\end{proof}
Note that due to the two-sided pseudolocality, Theorem~\ref{thm:SL27_2}, we now can use blowup argument, taking for granted that every convergence in regular part takes place in
smooth topology. Therefore, we can use the blowup argument in the proof of Proposition~\ref{prn:HE11_2},
based on the Liouville type theorem, Lemma~\ref{lma:HE11_1}.
Then the following corollary follows directly from Proposition~\ref{prn:HB05_1}.
\begin{corollary}[\textbf{Long-time regularity improvement in two time directions}]
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$, $r>0$, then
\begin{align*}
\mathcal{F}_r(M,0) \subset \bigcap_{-1 \leq t \leq 1} \mathcal{F}_{\delta}(M,t)
\end{align*}
for some $\delta=\delta(n,A,r)$.
\label{cly:HB05_1}
\end{corollary}
Now we are ready to prove Theorem~\ref{thmin:HC06_1}, the long-time, two-sided pseudolocality theorem.
\begin{proof}[Proof of Theorem~\ref{thmin:HC06_1}]
It follows from the combination of Corollary~\ref{cly:HB05_1} and Proposition~\ref{prn:HA08_3}.
\end{proof}
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Then for each time $t \in [-1,1]$ we have
\begin{align}
(M_i,x_i,g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M}(t),\bar{x}(t),\bar{g}(t)).
\label{eqn:HE15_1}
\end{align}
Let us see how are the two time slice limits $\bar{M}(0)$ and $\bar{M}(1)$ related.
Clearly, by Theorem~\ref{thmin:HC06_1}, the regular parts of $\bar{M}(0)$ and $\bar{M}(1)$ can be identified.
The relations among the singular parts at different time slices are more delicate.
For simplicity of notations, we denote $(\bar{M}(0), \bar{x}(0), \bar{g}(0))$ by $(\bar{M},\bar{x},\bar{g})$,
denote $(\bar{M}(1), \bar{x}(1), \bar{g}(1))$ by $(\bar{M}', \bar{x}', \bar{g}')$.
Let us also assume $\Vol(M_i)$ is uniformly bounded. Then it is clear that both $\bar{M}$
and $\bar{M}'$ are compact by the uniform non-collapsing caused by Sobolev constant bound.
In light of the uniform partial-$C^0$-estimate along the flow, without loss of generality,
we can assume that the Bergman function $\mathbf{b}$ is uniformly bounded below. By the fundamental estimates
in~\cite{DS}, we obtain that the map
\begin{align*}
Id_{0}: \quad (\bar{M}, \bar{x}, \bar{g}) \to (\bar{M}, \bar{x}, \tilde{\bar{g}})
\end{align*}
is a homeomorphism. Recall that $(\bar{M}, \bar{x}, \tilde{\bar{g}})$ is the limit of $(M_i, x_i, \tilde{g}_i(0))$, where
$\tilde{g}_i$ is the pull-back of Fubini-Study metric. Similarly, we have another homeomorphism map at time
$t=1$.
\begin{align*}
Id_{1}: \quad (\bar{M}', \bar{x}', \bar{g}') \to (\bar{M}', \bar{x}', \tilde{\bar{g}}').
\end{align*}
By Proposition~\ref{prn:HB05_2}, the pulled back Fubini-Study metrics $\tilde{g}_i(t)$ are uniformly equivalent
for $t \in [-1,1]$. It follows that there is a Lipschitz map $Id_{01}$ between two time slices, for the pulled back
Fubini-Study metrics:
\begin{align*}
Id_{01}: \quad (\bar{M}, \bar{x}, \tilde{\bar{g}}) \to (\bar{M}', \bar{x}', \tilde{\bar{g}}').
\end{align*}
Combine the previous steps, let $\Psi=Id_{1}^{-1} \circ Id_{01} \circ Id_{0}$, we obtain that the map
\begin{align*}
\Psi: (\bar{M}, \bar{x}, \bar{g}) \to (\bar{M}', \bar{x}', \bar{g}')
\end{align*}
is a homeomorphism.
By analyzing each component identity map,
it is clear that $\Psi|_{\mathcal{R}(\bar{M})}$, where $\mathcal{R}(\bar{M})$ is the regular part of $\bar{M}$,
maps $\mathcal{R}(\bar{M})$ to $\mathcal{R}(\bar{M}')$, as a biholomorphic map.
Similarly, $\Psi|_{\mathcal{S}(\bar{M})}$ is a homeomorphism to $\mathcal{S}(\bar{M}')$.
Therefore, the variety structure of the $\bar{M}(t)$ does not depend on time.
We remark that the compactness of $\bar{M}$ is not essentially used here.
If $\bar{M}$ is noncompact, the above argument go through formally
if we replace the target embedding space $\ensuremath{\mathbb{CP}}^{N}$ by $\ensuremath{\mathbb{CP}}^{\infty}$.
This formal argument can be made rigorous by applying delicate localization technique.
However, in our applications, $\bar{M}$ is always compact except it is a bubble, i.e., a blowup limit.
In this situation, we have the extra condition $|R|+|\lambda| \to 0$,
then $\Psi$ can be easily chosen as identity map, due to Proposition~\ref{prn:HE15_1}.
From the above discussion, it is clear that the topology structure and variety structure of $\bar{M}(t)$ does not depend on time.
So we just denote $\bar{M}(t)$ by $\bar{M}$. Then we can denote the convergence (\ref{eqn:HE15_1}) by
\begin{align*}
(M_i,x_i,g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M},\bar{x}(t),\bar{g}(t))
\end{align*}
for each $t$.
Hence, the limit family of metric spaces can be regarded as a family of evolving metrics on the limit
variety. Therefore, the above convergence at each time $t$ can be glued together to obtain a global convergence
\begin{align*}
\left\{ (M_i,x_i,g_i(t)), -T_i < t < T_i\right\} \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left\{ (\bar{M}, \bar{x}, \bar{g}(t)), -\bar{T} <t <\bar{T}\right\}
\end{align*}
where $\displaystyle \bar{T}=\lim_{i \to \infty} t_i$. Clearly, $\bar{g}(t)$ satisfies the K\"ahler Ricci flow equation on the regular part of $\bar{M}$.
Recall that we typically denote the K\"ahler Ricci flow $\left\{ (M_i,x_i,g_i(t)), -T_i < t < T_i\right\}$ by $\mathcal{M}_i$. Then we obtain the convergence of
K\"ahler Ricci flows (with base points):
\begin{align}
(\mathcal{M}_i, x_i) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{\mathcal{M}}, \bar{x}).
\label{eqn:HE15_2}
\end{align}
If we further know the underlying space $\bar{M}$ is compact, then the notation can be even simplified as
\begin{align*}
\mathcal{M}_i \stackrel{\hat{C}^{\infty}}{\longrightarrow} \bar{\mathcal{M}}.
\end{align*}
\begin{remark}
The limit flow $\bar{\mathcal{M}}$ can be regarded as an intrinsic K\"ahler Ricci flow on the normal variety $\bar{M}$.
Actually, it is already clear that $\bar{\mathcal{M}}$ is at least a
weak super solution of Ricci flow, in the sense of R.J. Mccann and P.M. Topping(\cite{MccTop}).
From the point of view of K\"ahler geometry, when restricted to the potential level,
the flow $\bar{\mathcal{M}}$ coincides with the weak K\"ahler Ricci flow solution defined by
Song and Tian(\cite{SongTian}), if $\bar{M}$ is compact.
\label{rmk:HE15_1}
\end{remark}
If we also consider the convergence of the line bundle structure, we can obviously generalize the convergence in (\ref{eqn:HE15_2})
as
\begin{align*}
& \left(\mathcal{LM}_i, x_i \right) \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left(\overline{\mathcal{LM}}, \bar{x}\right),
\quad \textrm{if $\bar{M}$ is non-compact}. \\
& \mathcal{LM}_i \stackrel{\hat{C}^{\infty}}{\longrightarrow} \overline{\mathcal{LM}}, \quad \textrm{if $\bar{M}$ is compact}.
\end{align*}
With these notations, we can formulate our compactness theorem as follows.
\begin{theorem}[\textbf{Polarized flow limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$, $x_i \in M_i$. Then we have
\begin{align*}
\left(\mathcal{LM}_i, x_i \right) \stackrel{\hat{C}^{\infty}}{\longrightarrow} \left(\overline{\mathcal{LM}}, \bar{x}\right),
\end{align*}
where $\overline{\mathcal{LM}}$ is a polarized K\"ahler Ricci flow solution on an analytic normal variety $\bar{M}$.
Moreover, if $\bar{M}$ is compact, then it is a projective normal variety.
\label{thm:HB07_1}
\end{theorem}
Notice that we have already proved Theorem~\ref{thmin:HC06_2} now.
\begin{proof}[Proof of Theorem~\ref{thmin:HC06_2}]
The limit polarized flow on variety follows from the combination of Theorem~\ref{thm:HB07_1} and Theorem~\ref{thm:HE19_1}.
The Minkowski dimension estimate of the singular set follows from Corollary~\ref{cly:HE25_2}.
\end{proof}
The properties of the limit spaces can be improved if extra conditions are available.
\begin{proposition}[\textbf{KE limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies
\begin{align}
\int_{-T_i}^{T_i}\int_{M_i} |R-n\lambda| dv dt \to 0.
\label{eqn:HA03_1}
\end{align}
Then $\overline{\mathcal{LM}}$ is a static, polarized K\"ahler Ricci flow solution. In other words, $\bar{g}(t) \equiv \bar{g}(0)$
and consequently are K\"ahler Einstein metric.
\label{prn:HB07_1}
\end{proposition}
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$ and $\lambda>0$. Then it is clear that $c_1(M)>0$, or $M$ is Fano.
Note that for every Fano manifold, we have a uniform bound $c_1^n(M) \leq C(n)$(c.f.~\cite{Deba}). This implies that
\begin{align*}
\frac{1}{A} \leq \Vol(M)=c_1^n(L)=\lambda^{-n} c_1^n(M) \leq C \lambda^{-n}.
\end{align*}
So $\lambda$ is bounded away from above. If we assume $\lambda$ is bounded away from zero, then
$\Vol(M)=c_1^{n}(L)$ is uniformly bounded. Consequently, $\diam(M)$ is uniformly bounded by non-collapsing, due to the Sobolev constant bound. Therefore, if we have a sequence of $\mathcal{LM}_i \in \mathscr{K}(n,A)$
with $\lambda_i>\lambda_0>0$, we can always assume
\begin{align*}
\lambda_i \to \bar{\lambda}>0, \quad \mathcal{LM} \stackrel{\hat{C}^{\infty}}{\longrightarrow} \overline{\mathcal{LM}}
\end{align*}
without considering the base points.
\begin{proposition}[\textbf{KRS limit}]
Suppose $\mathcal{LM}_i \in \mathscr{K}(n,A)$ satisfies
\begin{align}
\lambda_i>\lambda_0>0, \quad
\mu\left(M_i, g(T_i), \frac{\lambda_i}{2} \right)-\mu\left(M_i, g(-T_i), \frac{\lambda_i}{2} \right) \to 0,
\label{eqn:HA03_2}
\end{align}
where $\mu$ is Perelman's $W$-functional. Suppose $\overline{\mathcal{LM}}$
is the limit of $\mathcal{LM}$.
Then $\overline{\mathcal{M}}$ is a gradient shrinking K\"ahler Ricci soliton. In other words, there is a smooth real valued function $\hat{f}$
defined on $\mathcal{R}(\bar{M}) \times (-\bar{T},\bar{T})$ such that
\begin{align}
\hat{f}_{jk}=\hat{f}_{\bar{j}\bar{k}}=0, \quad R_{j\bar{k}} + \hat{f}_{j\bar{k}}-\hat{g}_{j\bar{k}}=0.
\label{eqn:HB07_1}
\end{align}
\label{prn:HB07_2}
\end{proposition}
\begin{proof}
Without loss of generality, we may assume $\lambda_i=1$.
Let $\mathcal{LM} \in \mathscr{K}(n,A)$. At time $t=1$, let $u$ be the minimizer of Perelman's $\mu$-functional.
Then solve the backward heat equation $\square^* u= (-\partial_{t}-\Delta + R-n\lambda) u=0$.
Let $f$ be the function such that $(2\pi)^{-n}e^{-f}=u$. Then we have
\begin{align*}
&\quad \int_{-1}^{1}\int_M (2\pi)^{-n} \left\{\left| R_{j\bar{k}}+f_{j\bar{k}}-g_{j\bar{k}}\right|^2 + |f_{jk}|^2+|f_{\bar{j}\bar{k}}|^2\right\} e^{-f} dv\\
&\leq \mu\left(M,g(1),\frac{1}{2} \right)-\mu\left(M,g(-1),\frac{1}{2}\right)\\
&\leq \mu\left(M,g(T),\frac{1}{2} \right)-\mu\left(M,g(-T),\frac{1}{2}\right) \to 0.
\end{align*}
At time $t=1$, $f$ has good regularity estimate for it is a solution of an elliptic equation.
For $t \in (-1,1)$, we have estimate of $f$ from heat kernel estimate.
It is not hard to see that, on the space-time domain $\mathcal{R} \times (-1,1)$, $f$ converges to a limit function $\hat{f}$ satisfying (\ref{eqn:HB07_1}).
Clearly, the time interval of $(-1,1)$ can be replaced by $(-a,a)$ for every $a \in (1,\bar{T})$. For each $a$, we have a limit function
$\hat{f}^{(a)}$, which satisfies equation (\ref{eqn:HB07_1}) and therefore has enough a priori estimates.
Then let $a \to \bar{T}$ and take diagonal sequence limit, we obtain a limit function $\hat{f}^{(\bar{T})}$ which satisfies (\ref{eqn:HB07_1})
on $\mathcal{R} \times (-\bar{T}, \bar{T})$. Without loss of generality, we still denote $\hat{f}^{(\bar{T})}$ by $\hat{f}$.
Then $\hat{f}$ satisfies (\ref{eqn:HB07_1}) on $\mathcal{R} \times (-\bar{T}, \bar{T})$.
\end{proof}
\begin{remark}
It is an interesting problem to see whether $(\bar{M}, \bar{g}(0))$ is a conifold in Theorem~\ref{thm:HB07_1}.
This question has affirmative answer when we know $(\bar{M}, \bar{g}(0))$ has Einstein regular part, following the proof of
Theorem~\ref{thm:HE08_1} and Proposition~\ref{prn:SC30_1}.
In particular, the limit spaces in Proposition~\ref{prn:HB07_1} and Proposition~\ref{prn:HE15_1} are K\"ahler Einstein conifolds.
\label{rmk:HE20_1}
\end{remark}
\section{Applications}
In this section, we will focus on the applications of our structure theory to the study
of anti-canonical K\"ahler Ricci flows.
\subsection{Convergence of anti-canonical K\"ahler Ricci flows at time infinity}
Based on the structure theory, Theorem~\ref{thmin:SC24_3} can be easily proved.
\begin{proof}[Proof of Theorem~\ref{thmin:SC24_3}]
In view of the fundamental estimate of Perelman (c.f.~\cite{SeT}), in order (\ref{eqn:SK20_1}) to hold, we only need a Sobolev constant bound,
which was proved by Q. Zhang (c.f.\cite{Zhq1}) and R. Ye (c.f. \cite{Ye}).
Therefore, the truncated flow sequences locate in $\mathscr{K}(n,A)$ for a uniform $A$.
It follows from Theorem~\ref{thmin:HC06_2} that the limit K\"ahler Ricci flow exists on a compact projective normal variety.
The limit normal variety is $Q$-Fano since it has a limit anti-canonical polarization.
According to Proposition~\ref{prn:HB07_2},
the boundedness and monotonicity of Perelman's $\mu$-functional force the limit flow to be a K\"ahler Ricci soliton.
The volume estimate of $r$-neighborhood of $\mathcal{S}$ follows from Corollary~\ref{cly:HA11_1}
and estimate (\ref{eqn:SB13_1}).
\end{proof}
We continue to discuss applications beyond Theorem~\ref{thmin:SC24_3}.
The following property is well known to experts, we write it down here for the convenience of the readers.
\begin{proposition}[\textbf{Connectivity of limit moduli}]
Suppose $\mathcal{M}=\left\{ (M^n, g(t)), 0 \leq t <\infty \right\}$ is an anti-canonical K\"ahler Ricci flows on Fano manifold
$(M, J)$. Let $\mathscr{M}$ be the collection of all the possible limit space along this flow. Then $\mathscr{M}$ is connected.
\label{prn:SL06_1}
\end{proposition}
\begin{proof}
If the statement was wrong, we have two limit spaces $\bar{M}_a$ and $\bar{M}_b$, locating in different
connected components of $\mathscr{M}$. Let $\mathscr{M}_a$ be the connected component containing
$\bar{M}_a$. Since $\mathscr{M}_a$ is a connected component, it is open and closed.
So its closure $\overline{\mathscr{M}_a}$ is the
same as $\mathscr{M}_a$. Clearly, $\mathscr{M}_a$ is compact under the Gromov-Hausdorff topology.
Define
\begin{align}
&d(X,\mathscr{M}_a) \triangleq \inf_{Y \in \mathscr{M}_a} d_{GH}(X,Y), \label{eqn:SL06_7}\\
&\eta_a \triangleq \inf_{X \in \mathscr{M} \backslash \mathscr{M}_a} d(X, \mathscr{M}_a). \label{eqn:SL06_8}
\end{align}
Clearly, $\eta_a>0$ by the compactness of $\mathscr{M}_a$ and the fact that $\mathscr{M}_a$ is a connected component.
Without loss of generality, we can assume $(M, g(t_i))$ converges to $\bar{M}_a$,
$(M, g(s_i))$ converges to $\bar{M}_b$, for $t_i \to \infty$ and $s_i >t_i$.
For simplicity of notation, we denote $(M,g(t_i))$ by $M_{t_i}$, $(M,g(s_i))$ by $M_{s_i}$.
For large $i$, we have
\begin{align}
d_{GH}(M_{t_i}, \bar{M}_a)<\frac{\eta_a}{100}, \quad
d_{GH}(M_{s_i}, \bar{M}_b) < \frac{\eta_a}{100}. \label{eqn:SL06_6}
\end{align}
In particular, the above inequalities imply that
\begin{align*}
d(M_{t_i}, \mathscr{M}_a)<\frac{\eta_a}{100}, \quad d(M_{s_i}, \mathscr{M}_a)>\frac{99}{100}\eta_a.
\end{align*}
By continuity of the flow, we can find $\theta_i \in (t_i, s_i)$ such that
$d(M_{\theta_i}, \mathscr{M}_a)=\frac{1}{2}\eta_a$, whose limit form is
\begin{align}
d(\bar{M}_c, \mathscr{M}_a)=\frac{1}{2}\eta_a,
\label{eqn:SL06_10}
\end{align}
where $\bar{M}_c$ is the limit of $M_{\theta_i}$. However, (\ref{eqn:SL06_10}) contradicts with
(\ref{eqn:SL06_8}) and the fact $\eta_a>0$.
\end{proof}
Proposition~\ref{prn:SL06_1} can be generalized as follows.
\begin{proposition}[\textbf{KRS limit moduli}]
Suppose $\mathcal{M}_s=\left\{ (M_s^n, g_s(t)), 0 \leq t <\infty, s \in X \right\}$
is a smooth family of anti-canonical K\"ahler Ricci flows on Fano manifolds $(M_s, J_s)$, where
$X$ is a connected parameter space.
We call $(\bar{M}, \bar{g})$ as a limit space if $(\bar{M},\bar{g})$ is the Gromov-Hausdorff limit of $(M, g_{s_i}(t_i))$ for some
$t_i \to \infty$ and $s_i \to \bar{s} \in X$.
Suppose $\displaystyle f(s)=\lim_{t \to \infty} \mu \left(g_s(t), \frac{1}{2} \right)$ is an upper semi-continuous function on $X$.
Then we have the following properties.
\begin{itemize}
\item Every limit space is a K\"ahler Ricci soliton.
\item Let $\widetilde{\mathscr{M}}$ be the collection of all the limit spaces.
Then $\widetilde{\mathscr{M}}$ is connected under the Gromov-Hausdorff topology.
\end{itemize}
\label{prn:SL07_2}
\end{proposition}
\begin{proof}
We shall only show that every limit space is a K\"ahler Ricci soliton.
The connectedness of $\widetilde{\mathscr{M}}$ can be proved almost the same as Proposition~\ref{prn:SL06_1}.
So we leave the details to the readers.
Suppose $s_i \to \bar{s}$. Fix $\epsilon$, we can choose $T_{\epsilon}$ such that
\begin{align*}
\mu\left(g_{\bar{s}}(T_{\epsilon}), \frac{1}{2} \right)>f_{\bar{s}}-\epsilon.
\end{align*}
By the smooth convergence of $g_{s_i}(T_{\epsilon})$ and the upper semi-continuity of $f$, we have
\begin{align*}
\mu\left(g_{s_i}(T_{\epsilon}), \frac{1}{2} \right) >f_{s_i}-\epsilon
\end{align*}
for large $i$. Recall that $t_i \to \infty$. Therefore, it follows from the monotonicity of Perelman's functional that
\begin{align*}
\mu\left(g_{s_i}(T_{\epsilon}), \frac{1}{2} \right)<\mu\left(g_{s_i}(t_i-1), \frac{1}{2} \right)<\lim_{t \to \infty}\mu\left(g_{s_i}(t), \frac{1}{2} \right)=f_{s_i}.
\end{align*}
Hence, we have
\begin{align*}
0\leq \mu\left(g_{s_i}(t_i+1), \frac{1}{2} \right)-\mu\left(g_{s_i}(t_i-1), \frac{1}{2} \right) <\epsilon
\end{align*}
for large $i$. By the arbitrary choice of $\epsilon$, we obtain
\begin{align*}
\lim_{i \to \infty} \left\{ \mu\left(g_{s_i}(t_i+1), \frac{1}{2} \right)-\mu\left(g_{s_i}(t_i-1), \frac{1}{2} \right) \right\}=0.
\end{align*}
Therefore, $(M, g_{s_i}(t_i))$ converges to a K\"ahler Ricci soliton, in light of Proposition~\ref{prn:HB07_2}.
\end{proof}
The gap between singularity and regularity in Theorem~\ref{thmin:SC24_1} has a global version as follows.
\begin{proposition}[\textbf{Gap around smooth KE}]
Suppose $(\tilde{M},\tilde{g},\tilde{J})$ is a compact, smooth K\"ahler Einstein manifold which belongs to $\mathscr{K}(n,A)$
when regarded as a trivial polarized K\"ahler Ricci flow solution.
Then there exists an $\epsilon=\epsilon(n,A,\tilde{g})$ with the following properties.
Suppose $\mathcal{LM} \in \mathscr{K}(n,A)$ and $ d_{GH}((\tilde{M},\tilde{g}), (M,g(0))) < \epsilon$, then we have
\begin{align*}
\mathbf{vcr}(M, g(0)) >\frac{1}{2}\mathbf{vcr}(\tilde{M},\tilde{g}).
\end{align*}
\label{prn:SL07_3}
\end{proposition}
\begin{proof}
It follows from the continuity of canonical volume radius under the Cheeger-Gromov convergence.
\end{proof}
Proposition~\ref{prn:SL07_3} means that there is no singular limit space around any given smooth K\"ahler Einstein manifold.
Clearly, the single smooth K\"ahler Einstein manifold in this Proposition
can be replaced by a family of smooth K\"ahler Einstein manifolds with bounded geometry.
The gap between smooth and singular K\"ahler Einstein metrics can be conveniently used to carry out topology argument.
\begin{theorem}[\textbf{Convergence of KRF family}]
Suppose $\mathcal{M}_s=\left\{ (M_s^n, g_s(t), J_s), 0 \leq t <\infty, s \in X \right\}$
is a smooth family of anti-canonical K\"ahler Ricci flows on Fano manifolds $(M_s, J_s)$, where
$X$ is a connected parameter space. Moreover, we assume that
\begin{itemize}
\item The Mabuchi's K-energy is bounded from below along each flow.
\item Smooth K\"ahler Einstein metrics in all adjacent complex structures(c.f. Defintion 1.4 of~\cite{CS1}) have uniformly bounded Riemannian curvature.
\end{itemize}
Let $\Omega$ be the collection of $s$ such that the flow
$g_s$ has bounded Riemannian curvature. Then $\Omega=\emptyset$ or $\Omega=X$.
\label{thm:HC03_2}
\end{theorem}
\begin{proof}
It suffices to show that $\Omega$ is both open and closed in $X$.
The openness follows from the stability of K\"ahler Ricci flow around a given smooth K\"ahler Einstein metric, due to Sun and Wang (c.f.~\cite{SW}).
Suppose $s \in \Omega$, then the flow $g_s$ converges to some K\"ahler Einstein manifold $(M',g',J')$, which is the unique
K\"ahler Einstein metric in its small smooth neighborhood. By continuous dependence of flow on the initial data, and the stability of K\"ahler Ricci flow in a
very small neighborhood of $(M',g',J')$,
it is clear that $s$ has a neighborhood consisting of points in $\Omega$. Therefore, $\Omega$ is an open subset of $X$.
The closedness follows from Proposition~\ref{prn:SL07_2}.
Suppose $s_i \in \Omega$ and $s_i \to \bar{s} \in X$.
Due to the fact that the Mabuchi's K-energy is bounded from below along each K\"ahler Ricci flow we are concerning now, the
limit Perelman functional is always the same(c.f.\cite{CS1}). Therefore, we can apply Proposition~\ref{prn:SL07_2} to show
that every limit space is a possibly singular K\"ahler Einstein. However, along every $g_{s_i}$, we obtain a smooth
limit K\"ahler Einstein manifold
$(M',g',J')$, which has uniformly bounded curvature, as a K\"ahler Einstein manifold in an adjacent complex structure.
Note that the diameter of $M'$ is uniformly bounded by Myers theorem.
The volume of $M'$ is a topological constant. Therefore, the geometry of $(M',g')$ are uniformly bounded.
By a generalized version of Proposition~\ref{prn:SL07_3}, $(M',g',J')$ is uniformly bounded away from singular K\"ahler Einstein metrics.
Due to Proposition~\ref{prn:SL07_2}, the connectedness of $\mathscr{M}$ forces that the flow
$g_{\bar{s}}$ must converge to a smooth $(M',g',J')$. In particular, $g_{\bar{s}}$ has bounded curvature.
Threrefore, $\bar{s} \in \Omega$ and $\Omega$ is closed.
\end{proof}
The two assumptions in Theorem~\ref{thm:HC03_2} seem to be artificial. However, if $J_s$ is a trivial family or a test configuration family, by the unique
degeneration theorem of Chen-Sun(c.f.~\cite{CS1}), all the smooth K\"ahler Einstein metrics form an isolated family,
then the second condition is satisfied automatically. On the other hand, by the existence of K\"ahler Einstein metrics in the weak sense,
one can also obtain the lower bound of Mabuchi's $K$-energy(c.f.~\cite{BM},~\cite{DT},~\cite{C08}).
Consequently, Theorem~\ref{thm:HC03_2} can be applied to these special cases and obtain the following corollaries.
\begin{corollary}[\textbf{Convergence to given KE}]
Suppose $(M,J)$ is a Fano manifold with a K\"ahler Einstein metric $g_{KE}$. Then every anti-canonical K\"ahler Ricci flow on $(M,J)$ converges to
$(M,g_{KE},J)$.
\label{cly:HC04_0}
\end{corollary}
\begin{proof}
Let $\omega_{KE}$ be the K\"ahler Einstein metric form. Then every metric form $\omega$ can be written as $\omega_{KE}+\sqrt{-1} \partial \bar{\partial}\varphi$
for some smooth function $\varphi$. Define
\begin{align*}
\omega_s= \omega_{KE} + s \sqrt{-1} \partial \bar{\partial} \varphi, \quad s \in [0,1].
\end{align*}
It follows from Theorem~\ref{thm:HC03_2} that the K\"ahler Ricci flow from every $\omega_s$ has bounded curvature, and consequently
converges to $\omega_{KE}$, by the uniqueness theorem of Chen-Sun(c.f.~\cite{CS1}). In particular, the flow
start from $\omega$ converges to $\omega_{KE}$.
\end{proof}
\begin{corollary}[\textbf{Convergence of a test configuration}]
Suppose $\mathbf{M}$ is a smooth test configuration, i.e., a family of Fano manifolds $(M_s,J_s)$
parameterized by $s$ in unit disk $D \subset \ensuremath{\mathbb{C}}^1$ with a natural $C^*$-action.
Suppose each fiber is smooth and the central fiber $(M_0,g_0,J_0)$
admits K\"ahler Einstein metric $(M_0,g_{KE},J_0)$. Then each K\"ahler Ricci flow starting from $(M_s, g_s,J_s)$ for arbitrary $s \in D$
converges to $(M_0, g_{KE},J_0)$.
\label{cly:HC04_1}
\end{corollary}
\begin{proof}
Theorem~\ref{thm:HC03_2} can be applied for $X=D$.
The central K\"ahler Ricci flow converges by Corollary~\ref{cly:HC04_0}. Therefore, the K\"ahler Ricci flow on each fiber has bounded curvature and converge to
some smooth K\"ahler Einstein metric, which can only be $(M_0,g_{KE},J_0)$, due to the uniqueness theorem of Chen-Sun again.
\end{proof}
\begin{remark}
Corollary~\ref{cly:HC04_0} was announced by G.Perelman. The first written proof was given by Tian-Zhu in~\cite{TZ1}.
The strategy of Corollary~\ref{cly:HC04_0} was first used in~\cite{TZ2}.
Corollary~\ref{cly:HC04_0}-Corollary~\ref{cly:HC04_1} have the corresponding K\"ahler Ricci soliton versions. These generalizations will be discussed in a separate paper.
\label{rmk:HC09_1}
\end{remark}
\subsection{Degeneration of anti-canonical K\"ahler Ricci flows}
In this subsection, we shall prove Theorem~\ref{thmin:HA01_1} and related corollaries.
The following Theorem is due to Jiang(c.f.~\cite{Jiang}).
\begin{theorem}[\textbf{Jiang's estimate}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying \begin{align}
\norm{Ric^{-}}{C^0(M)} + |\log \Vol(M)| + C_S(M, g(0)) \leq F
\label{eqn:HA01_1}
\end{align}
at time $t=0$. Then we have
\begin{align}
|R|+ |\nabla \dot{\varphi}|^2 \leq \frac{C}{t^{n+1}}
\label{eqn:HA01_2}
\end{align}
for some constant $C=C(n,F)$.
\label{thm:HA01_1}
\end{theorem}
Note that (\ref{eqn:HA01_2}) implies a uniform bound of diameter at each time $t>0$, by the uniform bound of
Perelman's functional. Then one can easily deduce a uniform bound (depending on $t$) of $\norm{\dot{\varphi}}{C^1(M)}$.
Combing this with the Sobolev constant estimate along the flow(c.f.~\cite{Zhq1},~\cite{Ye}), we see that
\begin{align}
\norm{R}{C^0(M)} + \norm{\dot{\varphi}}{C^1(M)} + C_S(M, g(t)) \leq C(n,F,t)
\label{eqn:HA01_3}
\end{align}
for each $t>0$. Therefore, away from the initial time, we can always apply our structure theory.
\begin{theorem}[\textbf{Weak convergence with initial time}]
Suppose $\mathcal{M}_i=\{(M_i^n,g_i(t), J_i), 0 \leq t <\infty\}$ is a sequence of anti-canonical K\"ahler Ricci flow
solutions, whose initial time slices satisfy estimate (\ref{eqn:HA01_1}) uniformly. Then we have
\begin{align}
(\mathcal{M}_i, g_i) \stackrel{G.H.}{\longrightarrow} (\bar{\mathcal{M}}, \bar{g}),
\end{align}
where the limit is a weak K\"ahler Ricci flow solution on a $Q$-Fano normal variety $\bar{M}$, for time $t>0$.
Moreover, the convergence can be improved to
be in the $\hat{C}^{\infty}$-Cheeger-Gromov topology for each $t>0$, i.e.,
\begin{align}
(M_i, g_i(t)) \stackrel{\hat{C}^{\infty}}{\longrightarrow} (\bar{M}(t), \bar{g}(t))
\end{align}
for each $t>0$.
\label{thm:HA01_2}
\end{theorem}
Clearly, if $(M_i,g_i)$ is a sequence of almost K\"ahler Einstein metrics(c.f.~\cite{TW2}) in the anti-canonical classes, then
$(\bar{M}(0), \bar{g}(0))$ and $(\bar{M}(1), \bar{g}(1))$ are isometric to each other, due to Proposition~\ref{prn:HB07_1} and the estimate in~\cite{TW2}.
In this particular case, it is easy to see that partial-$C^0$-estimate holds uniformly at time $t=0$ for each $i$, at least intuitively.
Actually, by the work Jiang~\cite{Jiang}, it is now clear that partial-$C^0$-estimate at time $t=0$ only requires a uniform Ricci lower bound.
Note that the evolution equation of the anti-canonical K\"ahler Ricci flow is
\begin{align}
\dot{\varphi}=\log \frac{\omega_{\varphi}^n}{\omega^n} + \varphi - u_{\omega},
\label{eqn:HA01_5}
\end{align}
where $u_{\omega}$ is the Ricci potential satisfying the normalization condition
$\int_M e^{-u_{\omega}} \frac{\omega^n}{n!}=(2\pi)^n$.
By maximum principle and Green function argument, we have the following property(c.f.~\cite{Jiang}).
\begin{proposition}[\textbf{Potential equivalence}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying
(\ref{eqn:HA01_1}). At time $t=0$, let $\varphi=0$ and $u_{\omega}$ satisfy the normalization condition. Then we have
\begin{align}
C(1-e^{t}) \leq \varphi \leq C e^t
\label{eqn:HA01_4}
\end{align}
for a constant $C=C(n,F)$.
\label{prn:HA01_3}
\end{proposition}
Let $\mathbf{b}(\cdot, t)$ be the Bergman function at time $t$. By definition, at point $x \in M$ and time $t=0$,
we can find a holomorphic section $S \in H^0(M, K_M^{-1})$ such that
\begin{align*}
\int_{M} \norm{S}{h(0)}^2 \frac{\omega^n}{n!}=1, \quad\mathbf{b}(x,0)=\log \norm{S}{h(0)}^2(x).
\end{align*}
Note that $\norm{S}{h(1)}^2=\norm{S}{h(0)}^2e^{-\varphi(1)}$.
By (\ref{eqn:HA01_4}), it is clear that $\norm{S}{h(1)}^2$ and $\norm{S}{h(0)}^2$ are uniformly equivalent.
On the other hand, $\Delta \norm{S}{}^2 \geq -n \norm{S}{}^2$. At time $t=0$, applying Moser iteration implies that $\norm{S}{h(0)}^2 \leq C$.
Hence we obtain $\norm{S}{h(1)}^2 \leq C$.
At time $t=1$, let $\tilde{S}$ be the normalization of $S$, i.e., $\tilde{S}=\lambda S$ such that
$\int_{M} \norm{\tilde{S}}{h(1)}^2 \frac{\omega_1^n}{n!}=1$. Then we have
\begin{align*}
\lambda^{-2}= \int_{M} \norm{S}{h(1)}^2 \frac{\omega_1^n}{n!} \leq C.
\end{align*}
It follows that
\begin{align*}
\mathbf{b}(x,1) &\geq \log \norm{\tilde{S}}{h(1)}^2(x)= \log \norm{S}{h(1)}^2(x)+ \log \lambda^2\\
&=\log \norm{S}{h(0)}^2(x) -\varphi(1) + \log \lambda^2\\
&=\mathbf{b}(x,0)-\varphi(1) + \log \lambda^2\\
&\geq \mathbf{b}(x,0)-C.
\end{align*}
By reversing time, we can obtain a similar inequality with reverse direction.
Same analysis applies to $\mathbf{b}^{(k)}$ for each positive integer $k$.
So we have the following property.
\begin{proposition}[\textbf{Bergman function equivalence}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying
(\ref{eqn:HA01_1}). For each positive integer $k$, there exists $C=C(n,F,k)$ such that
\begin{align}
\mathbf{b}^{(k)}(x,0) -C \leq \mathbf{b}^{(k)}(x,1) \leq \mathbf{b}^{(k)}(x,0)+C
\label{eqn:HE27_1}
\end{align}
for all $x \in M$.
\label{prn:HA01_4}
\end{proposition}
In view of Theorem~\ref{thm:HA01_2}, partial-$C^0$-estimate holds at time $t=1$,
which induces the partial-$C^0$-estimate at time $t=0$, by Proposition~\ref{prn:HA01_4}.
Therefore, the following theorem is clear now.
\begin{theorem}[\textbf{Partial-$C^0$-estimate at initial time}]
Suppose $\mathcal{M}=\{(M^n, g(t), J), 0 \leq t <\infty\}$ is an anti-canonical K\"ahler Ricci flow solution satisfying
(\ref{eqn:HA01_1}). Then
\begin{align*}
\inf_{x \in M} \mathbf{b}^{(k_0)}(x,0) \geq -c_0
\end{align*}
for some positive integer $k_0=k_0(n,F)$ and positive number $c_0=c_0(n,F)$.
\label{thm:HA01_3}
\end{theorem}
By the Sobolev constant estimates for manifolds with uniform positive Ricci curvature, it is clear that
Theorem~\ref{thmin:HA01_1} follows from Theorem~\ref{thm:HA01_3} directly.
It is also clear that Corollary~\ref{clyin:HA01_2} follows from Theorem~\ref{thm:HA01_3}.
The proof of Corollary~\ref{clyin:HA03_1} is known in literature, provided the partial-$C^0$-estimate along
the K\"ahler Ricci flow. We shall be sketchy here.
The more detailed proof will appear in~\cite{CW6}.
Actually, due to the work of S. Paul(\cite{Paul1209},\cite{Paul1308}) and the argument in section 6 of Tian and Zhang(\cite{TZZ2}), one obtains that the $I$-functional is bounded along the flow.
Then the K\"ahler Ricci flow converges to a K\"ahler Einstein metric, on the same Fano manifold.\\
Our structure theory has many other applications. However, in order to make this article self contained,
we prefer to discuss them in separate papers.
|
\section{Introduction}\label{sec1}
We consider the long-standing problem of estimating the variance of the
difference-in-means estimator as applied to a completely randomized
experiment performed on a random sample of size $n$ selected without
replacement from a population of size $N$ under a nonparametric model
of deterministic potential outcomes. It has been known since Neyman
\cite
{neyman23} that neither unbiased nor consistent variance estimation is
generally possible in this setting, due to the fact that the joint
distribution of the potential outcomes can never be fully recovered
from data.
In this paper, we propose an interval estimator that is consistent for
sharp bounds, defined as the smallest interval containing all values of
the variance that are compatible with the observable information. The
upper bound is never larger than and often smaller than conventional
approximations. Our estimator is also applicable to all possible cases
of $N$ and $n$ ($n=N<\infty$, $n<N<\infty$, and $N=\infty$), thus
providing a unified treatment of the problem. In the case where the
outcomes are dichotomous and $n=N<\infty$, our estimator reproduces
Robins \cite{robins88} results. The case $n < N < \infty$
generalizes the
settings considered by prior researchers. Unbiased variance estimation
is not generally possible when $N<\infty$, but our estimator produces
asymptotically sharp bounds. When the population size $N$ is infinite,
our estimator recovers the standard variance point estimator for mean
differences between independent groups \cite{neyman23}.
A practical implication of our work is that it facilitates confidence
intervals that are often narrower than intervals produced by
conventional methods: our upper bound variance estimator may be used to
construct conservative Wald-type confidence intervals for the average
treatment effect. Asymptotically, these intervals are the narrowest
Wald-type intervals that are assured to have at least the nominal
coverage. We illustrate empirical performance using data from an
randomized controlled trial, discuss extensions and provide R code
implementing our estimator.
An implementation in Stata is also available from the authors.
\section{Setting}
Consider a population $U_N$ consisting of $N \geq4$ units. From $U_N$,
$n$ units are randomly sampled into the experimental sample, and the
remaining $N - n$ units are left unsampled. Of the $n$ sampled units,
$m \geq2$ units are randomly assigned to the treatment condition, and
$n - m \geq2$ units are randomly assigned to the control condition.
Let the indicator variable $X^T_i$ be one if unit $i$ is assigned to
the treatment condition, and let the indicator $X^C_i$ be one if unit
$i$ is assigned to the control condition. If $X^T_i = X^C_i = 0$, then
the unit is unsampled. Since units are sampled without replacement,
$X^T_i + X^C_i \leq1$. Without loss of generality, assume an index
ordering $i=1,\ldots,N$ such that those assigned to treatment come
first, $X^T_1,\ldots,X^T_m=1$, and those assigned to control come
after, $X^C_{m+1},\ldots,X^C_n = 1$, and the remaining $N - n$
unsampled units, if any, come last.
Associated with each unit $i$ are two potential outcomes \cite
{neyman23,rubin78} under control and treatment, respectively: $y_{0i}$
and $y_{1i}$. For each unit $i$, the analyst then observes $y_{0i}$
when $X^C_i = 1$ and $y_{1i}$ when $X^T_i = 1$. Given elements $v_i$,
$w_i$ for $i = 1,\ldots,N$, we define the finite population mean $\mu
_N(v)$, finite population variance $\S_N^2(v)$ and finite population
covariance $\S_N(v,w)$, respectively, as
\begin{eqnarray*}
\mu_N(v) &=& \frac{1}{N}\sum_{i=1}^N
v_i,\qquad \S_N^2(v) = \frac
{1}{N}\sum
_{i=1}^N \bigl\{v_i -
\mu_N(v)\bigr\}^2,
\\
\S_N(v,w) &=& \frac{1}{N}\sum_{i=1}^N
\bigl\{v_i - \mu_N(v)\bigr\} \bigl\{w_i -
\mu _N(w)\bigr\}.
\end{eqnarray*}
The average treatment effect for the population $U_N$ is $\tau_N = \mu
_N(y_1) - \mu_N(y_0)$. The difference-in-means estimator of $\tau_N$ is
\begin{equation}
\label{eq:est-diffinmeans} \hat\tau_N = \hat\mu_N(y_1) -
\hat\mu_N(y_0) = \frac{1}{m}\sum
_{i=1}^m y_{1i} - \frac{1}{n-m}\sum
_{i=m+1}^n y_{0i},
\end{equation}
with $\E_X (\hat\tau_N) = \tau_N$, where the expectation operator
$\E
_X$ averages over all ${{N}\choose{n}} {{n}\choose{m}}$ possible
treatment assignments.
Our inferential target is the variance of $\hat\tau_N$. Adapting
Freedman \cite{freedman08b}, Proposition~1, the variance is
\begin{equation}
\label{eq:varcalc} V_N = \frac{1}{N-1} \biggl\{ \frac{N-m}{m}
\S_N^2(y_{1}) + \frac{N -
(n-m)}{n-m}
\S_N^2(y_{0}) + 2\S_N(y_1,y_0)
\biggr\}.
\end{equation}
The unknown quantities in this expression are $\S_N^2(y_{1})$, $\S
_N^2(y_{0})$ and $\S_N(y_1,y_0)$.
By Cochran \cite{cochran77}, Theorem~2.4, unbiased estimators of $\S
_N^2(y_{1})$ and $\S_N^2(y_{0})$ are
\begin{eqnarray*}
\hat{\S}_N^{2}(y_{1}) &= &\frac{N-1}{N(m-1)}\sum
_{i=1}^m \bigl\{ y_{1i} - \hat
\mu_N(y_1) \bigr\}^2,
\\
\hat{\S}_N^{2}(y_{0}) & =& \frac
{N-1}{N(n-m-1)}
\sum_{i=m+1}^n \bigl\{y_{0i} - \hat
\mu_N(y_0) \bigr\}^2.
\end{eqnarray*}
Since both potential outcomes $y_{0i}$ and $y_{1i}$ for the same unit
can never be observed simultaneously, consistent estimators do not
generally exist for $\S_N(y_1,y_0)$ or for $V_N$ when the population
size $N$ is finite. However, when the population being sampled from is
infinite ($N=\infty$), Neyman \cite{neyman23} noted that the control and
treatment units are effectively sampled independently from their
respective distributions. Hence, the covariance term vanishes, and
$V_N$ is point identified. To see this, let $N\rightarrow\infty$ while
holding $m$ and $n$ fixed so that $V_N \rightarrow\frac{1}{m} \S
_N^2(y_{1}) + \frac{1}{n-m} \S_N^2(y_{0})$, the sampling variance for
the difference of independent means.
\subsection{Neyman \texorpdfstring{\cite{neyman23}}{[13]} approximations when $n = N$}
When $n = N$, the sampling variance of the difference-in-means
estimator reduces to
\begin{equation}
\label{eq:true_variance} V_n = \frac{1}{n-1} \biggl\{ \frac{n-m}{m}
\S_n^2(y_{1}) + \frac
{m}{n-m}
\S_n^2(y_{0}) + 2\S_n(y_1,y_0)
\biggr\}.
\end{equation}
Neyman \cite{neyman23} proposed an estimator of $V_n$ that uses the
inequality $2 \S_n(y_1,\break y_0) \leq2 \{ \S_n^2(y_{1})\S
_n^2(y_{0}) \}^{1/2} \leq\S_n^2(y_{1}) + \S_n^2(y_{0})$, by
application of the Cauchy--Schwarz inequality and the inequality of
arithmetic and geometric means. An upper bound estimate for $V_n$ is
obtained by setting $2\S_n(y_1,y_0) = \S_n^2(y_{1}) + \S_n^2(y_{0})$
and substituting $\hat{\S}_n^2(y_{1})$ and $\hat{\S}_n^2(y_{0})$
for $\S
_n^2(y_{1})$ and $\S_n^2(y_{0})$, respectively:
\begin{equation}
\hat{V}_n^{a} = \frac{n}{n-1} \biggl\{
\frac{ \hat{\S}_n^2(y_{1})}{m} + \frac{\hat{\S}_n^2(y_{0})}{n-m} \biggr\}.
\end{equation}
Since $\E_X \{\hat{\S}_n^2(y_{1}) \}=\S_n^2(y_{1})$ and
$\E
_X \{\hat{\S}_n^2(y_{0}) \}=\S_n^2(y_{0})$, $\hat{V}_n^{a}$
is conservative as its bias is nonnegative:
\begin{equation}
\label{eq:bias} \E_X \bigl(\hat{V}_n^{a} -
V_n \bigr)= (n-1)^{-1} \bigl\{ \S_n^2(y_{1})
+ \S _n^2(y_{0}) - 2 \S_n(y_1,y_0)
\bigr\} \geq0.
\end{equation}
The estimate $\hat{V}_n^{a}$ is also produced by common estimators that
presuppose sampling from an infinite superpopulation, including
heteroskedasticity-\break robust variance estimators \cite{lin,samiiaronow}
and the standard variance estimate for mean differences between
independent groups \cite{neyman23}. Furthermore, $\hat{V}_n^{a}$ is
known to be unbiased for $V_n$ when effects are constant, as would hold
when there exist no treatment effects whatsoever \cite{gadburyamstat}.
For these reasons, the estimate $\hat{V}_n^{a}$ is often recommended
for the analysis of experimental data \cite{freedmanpp,gerbergreen}.
Neyman \cite{neyman23} also proposed a method for computing bounds on
$V_n$.\break
Given only knowledge of the second moments $\S_n^2(y_1)$ and $\S
_n^2(y_0)$, the\break sharpest bound on $\S_n(y_1,y_0)$ is given by the
Cauchy--Schwarz inequality:\break $- \{ \S_n^2(y_1) \S_n^2(y_0)
\}
^{1/2} \leq\S_n(y_1,y_0) \leq \{ \S_n^2(y_1) \S
_n^2(y_0) \}
^{1/2}$. By substituting $\hat{\S}_n^2(y_{1})$ and $\hat{\S
}_n^2(y_{0})$ for $\S_n^2(y_{1})$ and $\S_n^2(y_{0})$, Neyman's bound
estimator is
\begin{equation}
\hat{V}_n^{b\pm} = \frac{1}{n-1} \biggl[
\frac{n-m}{m}\hat\S_n^2(y_{1}) +
\frac{m}{n-m}\hat\S_n^2(y_{0}) \pm2 \bigl
\{ \hat{\S}_n^2(y_1) \hat{\S
}_n^2(y_0) \bigr\}^{1/2} \biggr].
\end{equation}
The plus or minus sign is chosen depending on whether an upper or a
lower bound estimate is desired. Neyman recommended choosing $\hat
{V}_n^{b+}$ as a conservative approximation to the true variance, and
suggested that it is ``necessary'' (page 471) to assume that the upper
bound given by the Cauchy--Schwarz inequality holds.
\section{Sharp bounds on $V_N$ given marginal distributions of outcomes}
Under the setting considered, estimates for the marginal distributions
of $y_1$ and $y_0$ exist and can be used to obtain asymptotically sharp
bounds on $V_N$ given the information available. Let $G_N(y) =
N^{-1}\sum_{i=1}^N I(y_{1i} \leq y)$ and $F_N(y) = N^{-1}\sum_{i=1}^N
I(y_{0i} \leq y)$ be the marginal distribution functions of $y_1$ and
$y_0$, respectively. Define their left-continuous inverses as
$G^{-1}_N(u) = \inf\{y\dvtx G_N(y) \geq u \}$ and $F^{-1}_N(u) = \inf\{y\dvtx F_N(y) \geq u \}$. Define also
\begin{eqnarray}
\label{eq:exactcovbound}
\S_N^H
(y_1,y_0) & = & \int_0^1
G_N^{-1}(u) F_N^{-1}(u) \,du -
\mu_N(y_1) \mu_N(y_0),
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\S_N^L (y_1,y_0) & = & \int
_0^1 G_N^{-1}(u)
F_N^{-1}(1-u) \,du - \mu _N(y_1)
\mu_N(y_0).
\end{eqnarray}
\begin{lemma}[(Hoeffding)] \label{lemma:covbound} Given only $G_N$ and
$F_N$ and no other information on the joint distribution of
$(y_1,y_0)$, the bound
\[
\S_N^L(y_1,y_0) \leq
\S_N(y_1,y_0) \leq\S_N^H(y_1,y_0)
\]
is sharp. The upper bound is attained if $y_1$ and $y_0$ are
comonotonic, that is, $(y_1,y_0) \sim\{ G_N^{-1}(U),
F_N^{-1}(U)\}$ for a uniform random variable $U$ on $[0,1]$. The lower
bound is attained if $y_1$ and $y_0$ are countermonotonic, \textit{that
is,} $(y_1,y_0) \sim\{ G_N^{-1}(U), F_N^{-1}(1-U)\}$.
\end{lemma}
Lemma~\ref{lemma:covbound} implies that $[V_N^L, V_N^H]$ is the
sharpest interval bound for $V_N$:
\begin{eqnarray*}
V_N^H & = &
\frac{1}{N-1} \biggl\{ \frac{N-m}{m}\S_N^2(y_{1})
+ \frac{N
- (n-m)}{n-m}\S_N^2(y_{0}) + 2
\S_N^H(y_1,y_0) \biggr\},
\\
V_N^L & = & \frac{1}{N-1} \biggl\{ \frac{N-m}{m}
\S_N^2(y_{1}) + \frac{N
- (n-m)}{n-m}
\S_N^2(y_{0}) + 2\S_N^L(y_1,y_0)
\biggr\}.
\end{eqnarray*}
In practice, we observe neither $G_N$ nor $F_N$, but rather their
estimates $\hat G_N(y) = m^{-1}\sum_{i=1}^N X^T_i I(y_{1i} \leq y)$,
$\hat F_N(y) = (n-m)^{-1}\sum_{i=1}^{N} X^C_i I(y_{0i} \leq y)$ and
left-continuous inverses
\begin{eqnarray*}
\hat G^{-1}_N(u)
&=& \inf\bigl\{y\dvtx \hat G_N(y) \geq u \bigr\} =
y_{1(\lceil m
u \rceil)},
\\
\hat F^{-1}_N(u) &=& \inf\bigl\{y\dvtx \hat
F_N(y) \geq u \bigr\} = y_{0(m +\lceil
(n-m)u \rceil)},
\end{eqnarray*}
where $y_{1(1)} \leq\cdots\leq y_{1(m)}$ and $y_{0(m+1)} \leq\cdots
\leq y_{0(n)}$ are the ordered observed outcomes, and $\lceil x \rceil$
denotes the smallest integer greater than or equal to $x$. Substituting
$(\hat G_N,\hat F_N)$ for $(G_N,F_N)$ in (\ref{eq:exactcovbound})
yields an interval estimator $ [ \hat\S_N^L(y_1,y_0), \hat\S
_N^H(y_1, y_0) ]$ for $\S_N(y_1,y_0)$:
\begin{eqnarray*}
\hat\S_N^H
(y_1,y_0) & = & \int_0^1
\hat G_N^{-1}(u) \hat F_N^{-1}(u) \,du
- \hat\mu_N(y_1) \hat\mu_N(y_0),
\\
\hat\S_N^L (y_1,y_0) & = & \int
_0^1 \hat G_N^{-1}(u)
\hat F_N^{-1}(1-u) \,du - \hat\mu_N(y_1)
\hat\mu_N(y_0).
\end{eqnarray*}
Let the $[0,1]$-partition $\mathcal{P}_{m,n} = \{p_0, p_1,\ldots, p_{P}
\}$ be the ordered distinct elements of $\{0,1/m,2/m,\ldots,1\} \cup\{
0,1/ ( n-m ),2/ ( n-m ),\ldots,1 \}$. Let
$y_{1[i]} = y_{1(\lceil mp_i \rceil)}$ and $y_{0[i]} = y_{0 \{
m+\lceil(n-m) p_i \rceil \}}$. The inverses $\hat G^{-1}_N$ and
$\hat F^{-1}_N$ are piecewise constant since $\hat G^{-1}_N(u) =
y_{1[i]}$ and $\hat F^{-1}_N(u) = y_{0[i]}$ for $u \in(p_{i-1},p_i]$.
In addition, the symmetry $p_i = 1 - p_{P-i}$ implies that $p_i -
p_{i-1} = p_{P+1-i} - p_{P-i}$. Thus, $ [\hat\S_N^L(y_1,y_0),
\hat
\S_N^H(y_1,y_0) ]$ reduces to
\begin{eqnarray}
\label{eq:est-covbound}
\hat\S_N^H
(y_1,y_0) & = & \sum_{i=1}^P
(p_i-p_{i-1}) y_{1[i]} y_{0[i]} - \hat
\mu_N(y_1) \hat\mu_N(y_0),
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\hat\S_N^L (y_1,y_0) & = & \sum
_{i=1}^P (p_i-p_{i-1})
y_{1[i]} y_{0[P+1-i]} - \hat\mu_N(y_1) \hat
\mu_N(y_0),
\end{eqnarray}
where $\hat\mu_N(y_1)$ and $\hat\mu_N(y_0)$ are as defined in (\ref
{eq:est-diffinmeans}).
Substituting $\hat\S_N^2(y_1)$, $\hat\S_N^2(y_0)$, and (\ref
{eq:est-covbound}) for $\{\S_N^2(y_1), \S_N^2(y_0), \S_N(y_1,y_0)\}$ in
the expressions for $V_N^L$ and $V_N^H$, we obtain the interval
estimator $[\hat V_N^L, \hat V_N^H]$ for $V_N$:
\begin{eqnarray}
\label{eq:varbound}
\hat V_N^H
& = & \frac{1}{N-1} \biggl\{ \frac{N-m}{m}\hat\S_N^2(y_{1})
+ \frac{N - (n-m)}{n-m}\hat\S_N^2(y_{0}) + 2\hat
\S_N^H(y_1,y_0) \biggr\},
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\hat V_N^L & = & \frac{1}{N-1} \biggl\{
\frac{N-m}{m}\hat\S_N^2(y_{1}) +
\frac{N - (n-m)}{n-m}\hat\S_N^2(y_{0}) + 2\hat
\S_N^L(y_1,y_0) \biggr\}.
\end{eqnarray}
Since Lemma~\ref{lemma:covbound} applies to the sample populations as
well, it follows that $\hat V_N^H$ is never greater than $\hat
{V}_N^{b+}$, and $\hat V_N^L$ is never smaller than $\hat{V}_N^{b-}$. R
code to implement the estimators $\hat V_N^H $ and $\hat V_N^L$ is
presented in Appendix \ref{rcode}.
It is possible to demonstrate that, when outcomes are dichotomous and
$n=N$, our estimator essentially reproduces the estimator proposed by
Robins \cite{robins88}, equation (3), with a slight difference due to finite
population corrections. See Copas~\cite{copas}, Gadbury, Iyer and
Albert \cite{gadburyjspi}, Heckman, Smith and
Clements \cite
{heckmanetal} and Zhang et al.~\cite{zhang} for additional details on
identification of the joint distribution of potential outcomes when
outcomes are dichotomous.
\section{Asymptotic sharpness of interval estimator}
Let $\{ U_N \}_N$ be a nested sequence of finite populations. The
potential outcomes $y_1$ and $y_0$ of each unit are fixed, and hence
the population grows deterministically. As in Isaki and Fuller~\cite
{isakifuller}, we
do not assume that the sequences of treatment assignments are nested;
instead, each $U_N$ hosts its own random assignment. Let $H_N(\cdot
,\cdot)$ be the joint distribution function of $(y_1,y_0)$ for $U_N$.
Under mild conditions on the scaling of $U_N$, the interval estimator
$[\hat V_N^L, \hat V_N^H]$ converges to sharp bounds on $V_N$.
\begin{proposition}\label{prop:asymptotics}
Suppose the following conditions hold as $N\rightarrow\infty$:
\begin{longlist}[1.]
\item[1.]$(m/N,n/N) \rightarrow(\theta\rho,\theta)$ for $\theta\in
(0,1]$ and $\rho\in(0,1)$;
\item[2.]$H_N$ converges weakly to a limit distribution $H$ with marginals
$G(y) = H(y,\infty)$ and $F(y) = H(\infty,y)$;
\item[3.]$G_N(y) \rightarrow G(y)$ at any discontinuity point of $G$, and
$F_N(y) \rightarrow F(y)$ at any discontinuity point of $F$;
\item[4.] The sequences of distributions represented by $\{G_N \}_N$ and $\{
F_N \}_N$ are uniformly square-integrable. That is, as $\beta
\rightarrow\infty$,
\[
\sup_N \Biggl\{ \frac{1}{N} \sum
_{i:y_{1i}^2\geq\beta}^N y_{1i}^2 \Biggr\},\qquad
\sup_N \Biggl\{ \frac{1}{N} \sum
_{i:y_{0i}^2\geq\beta}^N y_{0i}^2 \Biggr\}
\rightarrow0.
\]
\end{longlist}
Then for the collection $\mathcal{H}$ of all bivariate distributions
with marginals $G$ and $F$, the moments of each $h\in\mathcal{H}$
exist up to second order and
\begin{eqnarray*}
NV_N^H &\rightarrow&\frac{1-\theta\rho}{\theta\rho}
\Var_H(y_1) + \frac
{1 - \theta(1-\rho)}{\theta(1-\rho)} \Var_H(y_0)
+ 2 \sup_{h \in
\mathcal{H}} \Cov_h (y_1,y_0),
\\
NV_N^L &\rightarrow&\frac{1-\theta\rho}{\theta\rho}
\Var_H(y_1) + \frac
{1 - \theta(1-\rho)}{\theta(1-\rho)} \Var_H(y_0)
+ 2 \inf_{h \in
\mathcal{H}} \Cov_h (y_1,y_0).
\end{eqnarray*}
Moreover, $(\hat V_N^H - V_N^H, \hat V_N^L - V_N^L) = o_P(1/N)$.
\end{proposition}
\begin{remark}\label{re1}
Condition 3 is used to
establish the
functional convergence of $(G_N,F_N)$ to $(G,F)$. When the units of
$U_N$ are independent and identically distributed samples from a
superpopulation, the condition holds with probability one because of
the strong law of large numbers. The condition is also satisfied if $G$
and $F$ are continuous, regardless of whether or not the units come
from a superpopulation. We thank Professor A.~W. van der Vaart for
suggesting the latter as an alternate sufficient condition for
convergence, which subsequently inspired condition~3.
\end{remark}
\begin{remark}\label{re2} Given condition 2, any convergence
of the marginal second moments of $H_N$ to those of $H$ (should they
exist) necessarily implies condition 4. Thus, the condition
is the weakest possible complement to conditions 1--3.
\end{remark}
\begin{remark}\label{re3} If condition 4 of Proposition~\ref
{prop:asymptotics} is strengthened to require that $y_1$ and $y_0$ be
bounded, then higher order rates of convergence can be obtained, namely
that $\pr( N| \hat V_N^H - V_N^H | > \varepsilon$) and $\pr( N| \hat
V_N^L - V_N^L | > \varepsilon$) are both of order $\mathcal{O}(1/N)$.
Interested readers are referred to Proposition~\ref{prop:rate} in the \hyperref[app]{Appendix}.
\end{remark}
\textit{Outline of proof}. The random treatment assignment process can
be expressed as a triangular array $\mathcal{X}$ where the $N$th row
$(\mathcal{X}_{N,1},\ldots,\mathcal{X}_{N,N}) = \{
(X^T_1,X^C_1),\ldots
, (X^T_N,X^C_N) \}$ is the treatment/control assignment for population
$U_N$. Since the treatment/control assignment for $U_{N+1}$ is not
related to that for $U_N$, each row of $\mathcal{X}$ is a random vector
of a different probability space. As a result, the sequence of random
distribution functions $(\hat G_N,\hat F_N)$ do not share a common
probability space. However, by treating $(\hat G_N,\hat F_N)$ as random
elements taking values in the product space of \textit{c\`adl\`ag}
functions $D([-\infty,\infty],\mathbb{R})^2$ endowed with the uniform
metric, we show that $(\hat G_N,\hat F_N) \rightarrow(G,F)$ in
probability. It then follows from the Skorohod representation that
there exists a sequence of random elements $(\hat G_N',\hat F_N')$
defined on a common probability space that has the same law as $(\hat
G_N,\hat F_N)$. Moreover, $(\hat G_N',\hat F_N')$ converges to $(G,F)$
almost everywhere. Pathwise convergence of the moments of $(\hat
G_N',\hat F_N')$ then implies probabilistic convergence of the moments
of $(\hat G_N,\hat F_N)$ to the desired result. We refer the reader to
the \hyperref[app]{Appendix} for details of the formal argument.
\section{Confidence intervals for \texorpdfstring{$\tau_N$}{tau N}}
The upper bound estimator $\hat{V}_N^H$ may be used as a basis for
Wald-type confidence intervals for the average treatment effect. The
proof of the following corollary follows directly from Freedman
\cite{freedman08b}, Theorem~1, and associated remarks.
\begin{corollary}
Suppose that the support of $H$ is nonsingular and that conditions 1--3 of Proposition~\ref
{prop:asymptotics} hold. Suppose in\vadjust{\goodbreak} addition that condition 4 is strengthened to require uniformly bounded third moments:
\[
\sup_N \Biggl\{ \frac{1}{N}\sum
_{i=1}^N | y_{1i} |^3 \Biggr
\},\qquad \sup_N \Biggl\{ \frac{1}{N}\sum
_{i=1}^N | y_{0i} |^3 \Biggr
\} < \infty.
\]
Then
\[
\frac{\hat{\tau}_N - \tau_N}{ (\gamma\hat V_N^H )^{1/2}}
\]
converges weakly to the standard normal distribution where $\gamma=
\lim_N (NV_N)/\break \lim_N (N V_N^H) \leq1$.
\end{corollary}
\begin{remark}\label{re4}
As $\hat V_N^H$ is consistent for the sharp upper bound
on $V_N$, then given large $N$, a confidence interval constructed as
$\hat\tau_N \pm z_{1-\alpha/2} (\hat V_N^H)^{1/2}$ is asymptotically
the narrowest Wald-type confidence interval assured to have at least
the nominal coverage.
\end{remark}
\section{Application}
We consider the randomized controlled trial reported by Harrison and
Michelson \cite
{michelson}, which assessed the intention-to-treat effects of an
experimental phone call on donations to a nonprofit gay rights
organization. The control phone call script contained a standard
appeal. The experimental phone call script included an additional
sentence that revealed the sexual orientation of the volunteer caller.
The finite population $U_N$, which was not selected from any broader
population, contains $N = n = 1561$ subjects, $m = 781$ of whom were
randomly assigned to receive the experimental phone call. Outcomes were
measured in terms of US dollars (USD) received per subject, ranging
from $\$0$ to $\$150$. The mean donation given by subjects assigned to
control was $\hat\mu_N(y_0) = \$1.397$, and the mean donation given by
subjects assigned to treatment was $\hat\mu_N(y_1) = \$0.715$, yielding
the difference-in-means estimate $\hat\tau_N = -\$0.682$.
\begin{table}[b]
\caption{Variance estimates and confidence intervals for Harrison and
Michelson \protect\cite{michelson}}\label{tab1}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcc@{}}
\hline
& \textbf{Variance} & \multicolumn{1}{c@{}}{\textbf{95\% confidence}} \\
& \textbf{estimate (USD$^2$)} & \multicolumn{1}{c@{}}{\textbf{interval for} $\bolds{\tau_N}$} \\
\hline
Conventional ($\hat{V}_N^{a}$) & 0.199 & $(-\$1.555, \$0.192)$ \\
Neyman upper bound ($\hat{V}_N^{b+}$) & 0.196 & $(-\$1.548, \$0.185)$ \\[2pt]
Neyman lower bound ($\hat{V}_N^{b-}$) & 0.003 & N$/$A \\[2pt]
Sharp upper bound ($\widehat V_N^H$) & 0.186 & $(-\$1.528, \$0.165$) \\[2pt]
Sharp lower bound ($\widehat V_N^L$) & 0.098 & N$/$A \\
\hline
\end{tabular*}
\end{table}
In Table~\ref{tab1}, we report the variance estimates and confidence intervals
associated with Neyman's\vadjust{\goodbreak} approximations and our proposed estimator. We
find, as expected, that our estimates are sharper than Neyman's
approximations. Compared to the conventional variance estimator $\hat
{V}_N^{a}$, we find that our upper bound estimator yields a 7\%
reduction in the nominal variance. Importantly, if using $\hat
{V}_N^{a}$ as a basis for conservative inference, one would need over
100 additional subjects in order to achieve the same nominal variance
as that of our proposed upper bound estimate $\hat V_N^H$, all else
equal. Similarly, if using $\hat{V}_N^{b+}$, one would need over 75
additional subjects to achieve the nominal variance of $\hat V_N^H$.
\subsection{Simulations}
We use the data from \cite{michelson} to assess the operating
characteristics of the upper bound estimators and associated Wald-type
confidence intervals. These characteristics depend on the underlying
joint distribution of potential outcomes, which cannot be directly
observed and are instead hypothesized as part of these simulations. We
thus impute the missing potential outcomes (potential control outcomes
for treatment subjects, and potential treatment outcomes for control
subjects) by asserting varying hypotheses about treatment effects. We
simulate 25 million random assignments and, for each of these random
assignments, compute the upper bound variance estimates $\hat
{V}_N^{a}$, $\hat{V}_N^{b+}$ and $\hat V_N^H$, and associated
confidence intervals that would have been obtained. For the collection
of 25 million simulations, we calculate the mean variance estimate, the
mean width of the associated 95\% confidence intervals for $\tau_N$ and
the fraction of simulated confidence intervals covering $\tau_N$.
The first hypothesis that we evaluate is the sharp null hypothesis of
no effect whatsoever. This hypothesis, denoted ``Sharp Null,'' assumes
that $y_{0i} = y_{1i}$ for all~$i$. Under the Sharp Null, the treatment
effect estimator variance is $0.199$ USD$^2$. As can be seen in Table~\ref{tab2}, Neyman's estimators predictably perform well since they implicitly
assume that the outcomes are perfectly correlated: the bias \eqref
{eq:bias} for $\hat{V}_N^{a}$ is zero because $\S_N^2(y_1) = \S
_N^2(y_0) = \S_N(y_1,y_0)$. Due to the nonlinearity of the square root
function, the Cauchy--Schwarz inequality implies that $\hat{V}_N^{b+}$
has nonpositive bias ($-0.007$ USD$^2$). The 95\% confidence intervals
associated with $\hat{V}_N^{a}$ and $\hat{V}_N^{b+}$ have coverage of
95.2\% and 94.1\%, respectively (the former is not exactly 95\% because
the sampling distribution of $\tau_N$ is not perfectly normal). Because
$\hat V_N^H \leq\hat V_N^{b+}$, $\hat V_N^H$ is slightly more
negatively biased ($-0.010$ USD$^2$) and has lower coverage ($93.7\%$)
than $\hat V_N^{b+}$.
\begin{table}
\caption{Simulated variance estimator properties under varying
treatment effect hypotheses for Harrison and
Michelson \protect\cite{michelson}, using 25 million
simulated random assignments each}\label{tab2}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccc@{}}
\hline
& & \textbf{Mean var.} &
\textbf{Mean 95\%} &
\textbf{Coverage} \\
\textbf{Effect hypothesis} &\textbf{Variance estimator}& \textbf{estimate} & \textbf{CI width} & \multicolumn{1}{c@{}}{\textbf{for} $\bolds{\tau_N}$} \\
\hline
Sharp Null & Conventional ($\hat{V}_N^{a}$) & 0.199 & 1.747 & 95.2\% \\
(True Var.: 0.199) &Neyman upper bound ($\hat{V}_N^{b+}$) & 0.193 &
1.724 & 94.1\% \\[2pt]
& Sharp upper bound ($\hat V_N^H$) & 0.189 & 1.703 & 93.7\% \\[2pt]
Heterogeneity A & Conventional ($\hat{V}_N^{a}$) & 0.279 & 2.067 &
96.7\% \\[2pt]
(True Var.: 0.238) &Neyman upper bound ($\hat{V}_N^{b+}$) & 0.268 &
2.028 & 95.9\% \\[2pt]
& Sharp upper bound ($\hat V_N^H$) &0.258 & 1.987 & 95.4\% \\[2pt]
Heterogeneity B & Conventional ($\hat{V}_N^{a}$) & 0.244 & 1.933 &
97.4\% \\[2pt]
(True Var.: 0.186) &Neyman upper bound ($\hat{V}_N^{b+}$) &0.226 &
1.860 & 96.5\% \\[2pt]
& Sharp upper bound ($\hat V_N^H$) & 0.214 & 1.809 & 96.0\%\\
\hline
\end{tabular*}
\end{table}
We next consider two hypotheses that embed treatment effect
heterogeneity, denoted ``Heterogeneity A'' and ``Heterogeneity B.'' Under
Heterogeneity~A, we assume that the sharp null hypothesis holds, with
the exception of 10 subjects who had an observed $y_{0i} = 0$ USD under
control. For these 10 subjects, we assume that $y_{1i} = 100$ USD.
Under Heterogeneity A, the treatment effect estimator variance is
$0.238$ USD$^2$\vadjust{\goodbreak} and, as expected, all variance estimators are
conservative (positively biased). However, the bias, confidence
interval widths, and coverage for $\tau_N$ are all improved when $\hat
V_N^H$ is used in place of either of Neyman's estimators. In
formulating the Heterogeneity B hypothesis, we assume that
Heterogeneity A holds, but, in addition, for all 6 subjects under
treatment with an observed $y_{1i} \geq50$ USD, we assume that $y_{0i}
= 0$ USD. Under Heterogeneity B, the treatment effect estimator
variance is $0.186$ USD$^2$ and, again, while all estimators are
conservative, $\hat V_N^H$ improves over Neyman's estimators.
In Appendix \ref{illustrative}, we further explore the relative
performance of the upper bound estimates under varying assumptions
about the distribution of potential outcomes. Using the Beta
distribution family as an example to represent varying shapes of
marginal treatment and control distributions, we show that it is
possible for $\hat V_N^H$ to materially outperform $\hat V_N^a$ and
$\hat V_N^{b+}$ as the two marginals diverge in shape. Our simulations
therefore illustrate how $\hat V_N^H$ can improve upon Neyman's bounds
under effect heterogeneity.
\section{Discussion}
The proposed variance estimator may also be extended to alternative
designs. For block-randomized designs where the number of units per
block grows asymptotically large, Proposition~\ref{prop:asymptotics}
holds within each block, and thus calculation of the overall variance
is straightforward. In cluster-randomized designs with equally-sized
clusters, the proposed estimator may be used with the unit of analysis
being the cluster and the outcome being the cluster mean. It is also
straightforward to adapt the estimator to completely randomized
experiments with multiple treatments, which may be shown to be
logically equivalent to sampling from a broader population. In
addition, we note that our result can be generalized to characterize
estimation error for arbitrary target populations within the sampling
frame (e.g., unsampled units).
Finally, we remark on the scope of our findings, as our results
presuppose deterministic potential outcomes. When the potential
outcomes are stochastic, the total variance is greater than the
conditional variance (\ref{eq:varcalc}) because of the additional
stochasticity. If one sought to estimate the total variance or bounds
thereof, additional structure would need to be imposed on the
stochastic process (e.g., independence across units and finite
variances); otherwise it is possible for the identification set to be unbounded.
\begin{appendix}\label{app}
\section{Proofs}\label{proofs}
\begin{pf*}{Proof of Lemma~\ref{lemma:covbound}}
Let $H_N(y_1,y_0)$
be the joint distribution function of $(y_1,y_0)$, and define two other
distributions $H_N^H(y_1,y_0) = \min\{G_N(y_1),\break F_N(y_0)\}$ and
$H_N^L(y_1,y_0) = \max\{0,G_N(y_1)+F_N(y_0)-1\}$. All three
distributions have the same marginals $G_N$ and $F_N$. Defining $\E_Q$
as the expectation operator with respect to a distribution $Q$, a
result by Hoeffding, recounted in Tchen \cite{tchen1980}, shows that
\[
\E_{H_N^L} (y_1 y_0) \leq\E_{H_N}
(y_1y_0) \leq\E_{H_N^H} (y_1
y_0).
\]
Since $\{G_N^{-1}(U), F_N^{-1}(U)\} \sim H_N^H$ and $\{G_N^{-1}(U),
F_N^{-1}(1-U)\} \sim H_N^L$, the lower and upper bounds are equivalent to
\begin{eqnarray*}
\E_{H_N^H} (y_1 y_0) &=& \int
_0^1 G_N^{-1}(u)
F_N^{-1}(u) \,du,
\\
\E_{H_N^L} (y_1 y_0) &= &\int
_0^1 G_N^{-1}(u)
F_N^{-1}(1-u) \,du.
\end{eqnarray*}
The integrals exist because \mbox{$|G_N^{-1}(u)|, |F_N^{-1}(u)| \leq\max_{i=1}^N \max(|y_{1i}|, |y_{0i}|) < \infty$}.
\end{pf*}
Lemma~\ref{lemma:glivenko} below will be required in the proofs of
Propositions \ref{prop:asymptotics} and \ref{prop:rate}. In the special
case where the units of $U_N$ are independent and identically
distributed samples from a superpopulation, the first part of the lemma
reduces to the classical Glivenko--Cantelli theorem, and the
convergence implied by the second part follows from the conditional
bootstrap convergence results in van der Vaart and
Wellner \cite{VW1996}, Example~3.6.14. We
thank an anonymous reviewer for suggesting a more elegant way for
bounding (\ref{eq:Gvarbound}) and (\ref{eq:Fvarbound}) than our
original approach.
\begin{lemma}\label{lemma:glivenko}
Suppose conditions 1--3 of
Proposition~\ref{prop:asymptotics} hold. Then
\[
\sup_y \bigl|G(y)-G_N(y)\bigr| \rightarrow0\quad \mbox{and}\quad
\sup_y \bigl|F(y)-F_N(y)\bigr| \rightarrow0.
\]
In addition, given $\eta_1,\eta_0>0$, there exist two positive integers
$K_1(\eta_1)$ and $K_0(\eta_0)$ such that
\begin{eqnarray*}
\limsup_N \Bigl\{ N\pr\Bigl(\sup_y
\bigl|G(y)-\hat G_N(y)\bigr| \geq\eta_1\Bigr) \Bigr\}& \leq&
\frac{(1-\theta\rho)K_1(\eta_1)}{\theta\rho\eta_1^2},
\\
\limsup_N \Bigl\{ N\pr\Bigl(\sup_y
\bigl|F(y)-\hat F_N(y)\bigr| \geq\eta_0\Bigr) \Bigr\}& \leq&
\frac{ \{ 1 - \theta(1-\rho) \} K_0(\eta_0)}{\theta(1-\rho
)\eta
_0^2}.
\end{eqnarray*}
The integers are nonincreasing in $\eta$, and depend also on the
limiting distribution $H$ of $(y_1,y_0)$.
\end{lemma}
\begin{pf}
For the first part of the lemma, we follow the argument used in the
Glivenko--Cantelli theorem. Given $\eta_1>0$, there exists a partition
$-\infty= s_0 < s_1 < \cdots< s_{K_1(\eta_1)} = \infty$ such that
$G(s_i-) < G(s_{i-1}) + \eta_1/2$. For any $1 \leq i \leq K_1(\eta_1)$
and $s_{i-1} \leq s < s_i$,
\[
G(s_i-) - G_N(s_i-) - \eta_1/2
< G(s) - G_N(s) < G(s_{i-1}) - G_N(s_{i-1})
+ \eta_1/2,
\]
hence $\sup_y |G(y) - G_N(y)| < \eta_1$ if $|G(s_{i-1}) - G_N(s_{i-1})|
< \eta_1/2$ and $| G(s_i-) - G_N(s_i-) | < \eta_1/2$ for all $i$. By
conditions 2 and 3, this is
satisfied for all $N$ sufficiently large. The uniform convergence of
$F_N$ follows in the same way.
To establish the second part of the lemma, note that $\sup_y |G(y) -
\hat G_N(y) | < \eta_1$ on the set
\[
\bigcap_{i=1}^{K_1(\eta_1)} \bigl\{
\bigl|G(s_{i-1}) - \hat G_N(s_{i-1})\bigr|, \bigl|
G(s_i-) - \hat G_N(s_i-) \bigr| <
\eta_1/2 \bigr\}.
\]
Since $\pr\{ (\bigcap_i A_i)^c \} = \pr(\bigcup_i A_i^c) \leq\sum_i \pr
(A_i^c)$, we have
\begin{eqnarray}\label{eq:supCDFpbound}
&& \pr \Bigl\{ \sup_y \bigl|G(y)-\hat G_N(y)\bigr| \geq
\eta_1 \Bigr\}
\nonumber
\\
&&\qquad\leq \sum_{i=1}^{K_1(\eta_1)} \pr \bigl
\{\bigl|G(s_{i-1}) - \hat G_N(s_{i-1})\bigr| \geq
\eta_1/2 \bigr\}
\nonumber
\\
&&\qquad\quad{}+ \sum_{i=1}^{K_1(\eta_1)} \pr \bigl\{ \bigl|
G(s_i-) - \hat G_N(s_i-) \bigr| \geq
\eta_1/2 \bigr\}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\leq \sum_{i=1}^{K_1(\eta_1)} \pr\bigl\{\bigl |\hat
G_N(s_{i-1}) - G_N(s_{i-1})\bigr| \geq
\eta_1/2 - o(1) \bigr\}
\\
&&\qquad\quad{}+ \sum_{i=1}^{K_1(\eta_1)} \pr\bigl\{\bigl | \hat
G_N(s_i-) - G_N(s_i-)\bigr | \geq
\eta_1/2 - o(1) \bigr\}
\nonumber
\\
&&\qquad\leq \sum_{i=1}^{K_1(\eta_1)} \frac{ \Var_X \{ \hat G_N(s_{i-1})
\} +
\Var_X \{ \hat G_N(s_i-) \} }{ \{ \eta_1/2 - o(1) \}^2 },\nonumber
\end{eqnarray}
where the second inequality follows from $|G(y) - G_N(y)| = o(1)$, and
the last inequality from Chebyshev's inequality and the fact that $\E_X
\hat G_N(y) = G_N(y)$.
The argument used to derive (\ref{eq:varcalc}) can also be used to
bound the variances in (\ref{eq:supCDFpbound}). Noting that $\sigma
_N^2(I\{ y_1 \leq y \}) = G_N(y)\{ 1-G_N(y) \} \leq1/4$ and similarly
$\sigma_N^2(I\{ y_0 \leq y \}) \leq1/4$, we have for all $y$,
\begin{eqnarray}
\Var_X \hat G_N(y) &=& \frac{N-m}{(N-1)m}
\sigma_N^2\bigl(I\{ y_1 \leq y \}\bigr) =
\frac{N-m}{4(N-1)m}, \label{eq:Gvarbound}
\\
\Var_X \hat F_N(y) &=& \frac{N-(n-m)}{(N-1)(n-m)}
\sigma_N^2\bigl(I\{ y_0 \leq y \}\bigr) \leq
\frac{N-(n-m)}{4(N-1)(n-m)}. \label{eq:Fvarbound}
\end{eqnarray}
Plugging (\ref{eq:Gvarbound}) into (\ref{eq:supCDFpbound}) and taking
limits yields the desired result for $\hat G_N$, after absorbing the
factor of 2 into $K_1(\eta_1)$. The result for $\hat F_N$ can be
obtained in the same manner.
\end{pf}
\begin{pf*}{Proof of Proposition~\ref{prop:asymptotics}}
As indicated
in the proof outline, we proceed in several stages.
(i) \textit{Functional convergence of random distribution functions.}
Let $D([-\infty,\break \infty], \mathbb{R})^2$ be the Cartesian product of the
space of \textit{c\`adl\`ag} functions with itself, endowed with the
uniform metric induced by the norm $\| (v,u) \| = \max\{ \sup_y | v(y)
|, \sup_y | u(y) | \}$. Thus, $D([-\infty,\infty],\mathbb{R})^2$
is a
nonseparable metric space. Lemma~\ref{lemma:glivenko} shows that the
distribution functions $(\hat G_N, \hat F_N)$ converge in probability
to $(G,F)$ in $D([-\infty,\infty],\mathbb{R})^2$. That is, $\pr(\|
(\hat G_N-G,\hat F_N-F) \| \geq\varepsilon) \rightarrow0$ for every
$\varepsilon>0$. As is the case with the lemma, the statement does not
require the use of outer measures because for each $N$, $(\hat G_N,
\hat F_N)$ can take on at most ${{N}\choose{n}} {{n}\choose{m}}$
distinct values in $D([-\infty,\infty],\mathbb{R})^2$; therefore, $\|
(\hat G_N-G,\hat F_N-F) \|$ is finite discrete valued.
(ii) \textit{Existence of random distributions $(\hat G_N',\hat F_N')$
defined on a common probability space.} Since the limit $(G,F)$ is
deterministic, the support of the limiting probability measure on
$D([-\infty,\infty],\mathbb{R})^2$ is a singleton. Applying the
Skorohod representation \cite{VW1996},\vspace*{1pt} Theorem~1.10.3, to $(\hat
G_N,\hat F_N)$ yields new random elements $(\hat G_N',\hat F_N')$ on
$D([-\infty,\infty],\mathbb{R})^2$ that have the same law as $(\hat
G_N,\hat F_N)$. Furthermore, $(\hat G_N',\hat F_N')$ converges to
$(G,F)$ almost everywhere, in the sense that along each sample path
$\omega'$ (in a set of measure one), the distribution functions
converge uniformly:
\[
\sup_y \bigl| \hat G_N'\bigl(y;
\omega'\bigr)-G(y) \bigr| \rightarrow0\quad \mbox{and}\quad \sup
_y\bigl | \hat F_N'\bigl(y;
\omega'\bigr)-F(y) \bigr| \rightarrow0.
\]
(iii) \textit{Convergence of $\E_{\hat G_N} (y_1^p)$ and $\E_{\hat F_N}
(y_0^p)$ for \textit{p} = 1, 2.} Define $\E_Q$ as the expectation
operator with respect to a distribution $Q$. Under condition 1, there exists $N_0$ such that $1/m \leq2/(\theta\rho
N)$ and $1/(n-m) \leq2/\{\theta(1-\rho) N\}$ for $N\geq N_0$. Then for
each $N \geq N_0$ and every realization of $(\hat G_N, \hat F_N)$,
condition 4 implies that as $\beta\rightarrow\infty$,
\begin{eqnarray*}
\E_{\hat G_N} \bigl( y_1^2I\bigl
\{y_1^2\geq\beta\bigr\} \bigr)& =& \sum
_{i:y_{1i}^2\geq
\beta
}^N \frac{ X^T_i y_{1i}^2 }{m} \leq\frac{2}{\theta\rho}
\sup_{N
\geq
N_0} \Biggl\{ \sum_{i:y_{1i}^2\geq\beta}^N
\frac{y_{1i}^2}{N} \Biggr\} \rightarrow0,
\\
\E_{\hat F_N} \bigl( y_0^2I\bigl
\{y_0^2\geq\beta\bigr\} \bigr) &=& \sum
_{i:y_{0i}^2\geq
\beta}^N \frac{ X^C_i y_{0i}^2 }{n-m} \leq\frac{2}{\theta(1-\rho)}
\sup_{N \geq N_0} \Biggl\{ \sum_{i:y_{0i}^2\geq\beta}^N
\frac
{y_{0i}^2}{N} \Biggr\} \rightarrow0.
\end{eqnarray*}
Recall that both ($\hat G_N',\hat F_N'$) and ($\hat G_N,\hat F_N$)
share the same finite discrete distribution. Thus, for almost all
sample paths $\omega'$ in the probability space of $(\hat G_N',\hat
F_N')$, the sequences of distributions represented by $\{ \hat
G_N'(\cdot;\omega') \}_N$ and $\{ \hat F_N'(\cdot;\omega') \}_N$ are
uniformly square-integrable. Moreover, since $\hat G_N'(\cdot;\omega')
\rightarrow G(\cdot)$, the random moments $\{ \E_{\hat G_N'} (y_1),
\E
_{\hat G_N'} (y_1^2) \}$ converge to $\{ \E_H (y_1),\break \E_H (y_1^2) \}$
almost everywhere, with the limits being finite. Similarly, $\{ \E
_{\hat F_N'} (y_0),\break \E_{\hat F_N'} (y_0^2) \} \rightarrow\{ \E_H
(y_0), \E_H (y_0^2) \}$ almost everywhere as well. Translating this
back into convergence in probability for the first two random moments
of $\hat G_N$ and $\hat F_N$, we have
\begin{eqnarray}
\hat\S_N^2(y_1) & =& \frac{N-1}{N}
\frac{m}{m-1} \bigl[ \E_{\hat G_N} \bigl(y_1^2
\bigr) - \bigl\{\E_{\hat G_N} (y_1) \bigr\}^2 \bigr]
\rightarrow \Var_H(y_1), \label{eq:y1varplim}
\\
\hat\S_N^2(y_0) & = &\frac{N-1}{N}
\frac{n-m}{n-m-1} \bigl[ \E _{\hat
F_N} \bigl(y_0^2
\bigr) - \bigl\{\E_{\hat F_N} (y_0) \bigr\}^2 \bigr]
\rightarrow\Var_H(y_0) \label{eq:y0varplim}
\end{eqnarray}
in probability.
(iv) \textit{Convergence of $\hat\S_N^H(y_1,y_0)$ and $\hat\S
_N^L(y_1,y_0)$.} Define the distributions
$H^H(y_1,y_0) = \min\{ G(y_1),F(y_0) \}$ and $H^L(y_1,y_0) = \max\{
0,G(y_1)+F(y_0)-1 \}$, both of which have marginals $G$ and $F$. Using
Hoeffding's result from the proof of Lemma~\ref{lemma:covbound}, we
have that
\begin{eqnarray*}
\E_{H^H} (y_1y_0) &=& \sup_{h\in\mathcal{H}}
\E_h(y_1y_0),
\\
\E_{H^L} (y_1y_0) &=& \inf_{h\in\mathcal{H}}
\E_h(y_1y_0).
\end{eqnarray*}
Now fix a sample path and define two sequences of distributions $\hat
H_N^{H'}(y_1,y_0;\break \omega') = \min\{ \hat G_N'(y_1;\omega'),\hat
F_N'(y_0;\omega') \}$ and $\hat H_N^{L'}(y_1,y_0;\omega') = \max\{
0,\hat G_N'(y_1;\omega')+\hat F_N'(y_0; \omega')-1 \}$. It is clear that
$\hat H_N^{H'}(\cdot,\cdot; \omega')$ converges to $H^H(\cdot,\cdot)$
and $\hat H_N^{L'}(\cdot,\cdot;\break\omega')$ converges to $H^{L}(\cdot
,\cdot
)$ pointwise. Given that the product $y_1y_0$ is also uniformly
integrable with respect to almost all sequences $\{ \hat H_N^{H'}(\cdot
,\cdot;\omega') \}_N$ and\vadjust{\goodbreak} $\{ \hat H_N^{L'}(\cdot,\cdot;\omega') \}_N$
because $\{|XY|\geq\beta^2 \} \subset\{|X|\geq\beta\} \cup\{
|Y|\geq\beta\}$, it follows that $\E_{\hat H_N^{H'}} (y_1 y_0)
\rightarrow\sup_{h\in\mathcal{H}} \E_h(y_1y_0)$ and $\E_{\hat
H_N^{L'}} (y_1 y_0) \rightarrow\inf_{h\in\mathcal{H}} \E_h(y_1y_0)$
almost everywhere. Thus,
\begin{eqnarray}
\hat\S_N^H(y_1,y_0)& =&
\E_{\hat H_N^H} (y_1 y_0) - \E_{\hat G_N}
(y_1) \E_{\hat F_N} (y_0) \rightarrow\sup
_{h\in\mathcal{H}} \Cov _h(y_1,y_0),
\label{eq:covUBplim}
\\
\hat\S_N^L (y_1,y_0) &=&
\E_{\hat H_N^L} (y_1 y_0) - \E_{\hat G_N}
(y_1) \E_{\hat F_N} (y_0) \rightarrow\inf
_{h\in\mathcal{H}} \Cov_h(y_1,y_0)
\label{eq:covLBplim}
\end{eqnarray}
in probability. Plugging (\ref{eq:y1varplim})--(\ref{eq:covLBplim})
into (\ref{eq:varbound}) then yields the proposition.
\end{pf*}
\begin{proposition}\label{prop:rate}
Suppose conditions 1--3 of
Proposition~\ref{prop:asymptotics} hold, and that $y_1$ and $y_0$ are
bounded: $|y_{1i}|,|y_{0i}| \leq C < \infty$ for all $i$. Given
$\varepsilon>0$, for any $\varepsilon_1,\varepsilon_2,\varepsilon_3>0$
such that $\sum_i \varepsilon_i = \varepsilon$,
\begin{eqnarray*}
\pr\bigl( N\bigl| \hat V^H_N - V^H_N
\bigr| \geq\varepsilon\bigr) &\leq&\mathcal{O} \biggl( \frac{C^4}{N}
\kappa_1(\varepsilon_1) \biggr),
\\
\pr\bigl( N\bigl| \hat V^L_N - V^L_N
\bigr| \geq\varepsilon\bigr) &\leq&\mathcal{O} \biggl( \frac{C^4}{N} \bigl\{ 1/
\varepsilon_1^2 + \kappa_2(
\varepsilon_2) + \kappa _3(\varepsilon_3)
\bigr\} \biggr),
\end{eqnarray*}
where $\kappa_1(\varepsilon_1)$, $\kappa_2(\varepsilon_2)$ and
$\kappa
_3(\varepsilon_3)$ depend on the limiting distribution $H$.
\end{proposition}
\begin{pf} Define the bivariate distribution functions
$H^H_N(y_1,y_0) =\break \min\{ G_N(y_1), F_N(y_0) \}$, $H^L_N(y_1,y_0) \vadjust{\goodbreak}= \max
( 0,G_N(y_1)+F_N(y_0)-1 )$,\break $\hat H^H_N(y_1, y_0) = \min\{ \hat
G_N(y_1), \hat F_N(y_0) \}$, and $\hat H^L_N(y_1,y_0) = \max( 0, \hat
G_N(y_1)+\break \hat F_N(y_0)-1 )$. Let $\E_Q$ be the expectation operator
with respect to a distribution $Q$. Using another result by Hoeffding
as recounted in Lehmann \cite{lehmann}, Lemma~2, the following covariances can
be expressed as
\begin{eqnarray*}
\hat\S_N^H(y_1,y_0) & = &\int
_{-C}^C\int_{-C}^C
\hat H_N^H(y_1,y_0)
\,dy_1 \,dy_0 \\
&&{}- \int_{-C}^C
\hat G_N(y_1) \,dy_1 \int_{-C}^C
\hat F_N(y_0) \,dy_0
\\
& =&\int_{-C}^C\int_{-C}^C
\hat H_N^H(y_1,y_0)
\,dy_1 \,dy_0 - C^2 + C \E _{\hat G_N}
(y_1)
\\
&&{} + C \E_{\hat F_N} (y_0) - \E_{\hat G_N} (y_1)
\E_{\hat F_N} (y_0),
\\
\S_N^H(y_1,y_0) & = &\int
_{-C}^C \int_{-C}^C
H_N^H(y_1,y_0) \,dy_1
\,dy_0 \\
&&{}- \int_{-C}^C
G_N(y_1) \,dy_1 \int_{-C}^C
F_N(y_0) \,dy_0
\\
& =& \int_{-C}^C \int_{-C}^C
H_N^H(y_1,y_0) \,dy_1
\,dy_0 - C^2 + C \E_{G_N} (y_1)
\\
&&{} + C \E_{F_N} (y_0) - \E_{G_N} (y_1)
\E_{F_N} (y_0),
\end{eqnarray*}
where the second equality follows from the identity $\E(W) = C - \int_{-C}^C \pr(W\leq w)\,dw$ for any random variable $W$ bounded by $C$. Then
\begin{eqnarray*}
N\bigl(\hat V_N^H - V_N^H\bigr)
& = &\frac{N-m}{m-1} \biggl\{ \E_{\hat G_N} \bigl(y_1^2
\bigr) - \frac{N(m-1)}{m(N-1)} \E_{G_N} \bigl(y_1^2
\bigr) \biggr\}
\\
&&{} - \frac{N-m}{m-1} \biggl[ \bigl\{ \E_{\hat G_N} (y_1) \bigr
\}^2 - \frac
{N(m-1)}{m(N-1)} \bigl\{ \E_{G_N} (y_1)
\bigr\}^2 \biggr]
\\
&&{} + \frac{N-(n-m)}{n-m-1} \biggl\{ \E_{\hat F_N} \bigl(y_0^2
\bigr) - \frac
{N(n-m-1)}{(N-1)(n-m)} \E_{F_N} \bigl(y_0^2
\bigr) \biggr\}
\\
& &{}- \frac{N-(n-m)}{n-m-1} \biggl[ \bigl\{ \E_{\hat F_N} (y_0)
\bigr\}^2 - \frac
{N(n-m-1)}{(N-1)(n-m)} \bigl\{ \E_{F_N}
(y_0) \bigr\}^2 \biggr]
\\
&&{} + \frac{2N}{N-1} C \bigl\{ \E_{\hat G_N} (y_1) -
\E_{G_N} (y_1) + \E _{\hat F_N}(y_0) -
\E_{F_N} (y_0) \bigr\}
\\
& &{}- \frac{2N}{N-1} \bigl\{ \E_{\hat G_N} (y_1)
\E_{\hat F_N}(y_0) - \E _{G_N} (y_1)
\E_{F_N} (y_0) \bigr\}
\\
&&{} + \frac{2N}{N-1} \int_{[-C,C]^2} \bigl\{ \hat
H_N^H(y_1,y_0) -
H_N^H(y_1,y_0) \bigr\}
\,dy_1 \,dy_0.
\end{eqnarray*}
To obtain the desired result, we proceed by bounding the probability
that each of the seven terms are large. Let $\nu_1,\ldots,\nu_8 > 0$ be
a tuple whose sum is $\varepsilon_1$. For the first term,
\begin{eqnarray*}
&& \pr \biggl\{ \frac{N-m}{m-1} \biggl\llvert \E_{\hat G_N}
\bigl(y_1^2\bigr) - \frac
{N(m-1)}{m(N-1)} \E_{G_N}
\bigl(y_1^2\bigr) \biggr\rrvert \geq\nu_1
\biggr\}
\\
&&\qquad\leq \pr \biggl\{ \bigl\llvert \E_{\hat G_N} \bigl(y_1^2
\bigr) - \E_{G_N} \bigl(y_1^2\bigr) \bigr\rrvert
\geq\frac{(m-1)\nu_1}{N-m} - \frac{(N-m)C^2}{m(N-1)} \biggr\}
\\
&&\qquad\leq \Var_X \bigl\{\E_{\hat G_N} \bigl(y_1^2
\bigr) \bigr\}\Big/ \biggl\{ \frac
{(m-1)\nu_1}{N-m} - o(1) \biggr\}^2
\\
&&\qquad\leq \frac{(N-m) C^4}{(N-1)m} \Big/ \biggl\{ \frac{(m-1)\nu_1}{N-m} - o(1) \biggr
\}^2,
\end{eqnarray*}
where the first inequality follows from $\llvert \E_{\hat G_N} (y_1^2) -
\beta_N \E_{G_N} (y_1^2) \rrvert \leq\llvert \E_{\hat G_N} (y_1^2)
- \E
_{G_N} (y_1^2) \rrvert + \llvert (1 - \beta_N) \rrvert \E_{G_N}
(y_1^2)$, and the second inequality from Chebyshev's inequality and the
fact that $\E_X \E_{\hat G_N} (y_1^p) = \E_{G_N} (y^p)$. The bound on
the variance is obtained in the same way as (\ref{eq:Gvarbound}). Thus,
\begin{eqnarray}
\label{eq:pbound1} &&\limsup_N N \pr \biggl\{ \frac{N-m}{m-1}
\biggl\llvert \E_{\hat G_N} \bigl(y_1^2\bigr) -
\frac{N(m-1)}{m(N-1)} \E_{G_N} \bigl(y_1^2\bigr)
\biggr\rrvert \geq\nu_1 \biggr\}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\leq \frac{(1-\theta\rho)^3 C^4}{\theta^3 \rho^3 \nu_1^2}.
\end{eqnarray}
For the second term,
\begin{eqnarray*}
&& \pr \biggl\{ \frac{N-m}{m-1} \biggl\llvert \bigl\{ \E_{\hat G_N}
(y_1) \bigr\}^2 - \frac{N(m-1)}{m(N-1)} \bigl\{
\E_{G_N} (y_1) \bigr\}^2 \biggr\rrvert \geq
\nu_2 \biggr\}
\\
&&\qquad\leq \pr \biggl\{ \biggl\llvert \E_{\hat G_N} (y_1) - \biggl\{
\frac
{N(m-1)}{m(N-1)} \biggr\}^{1/2} \E_{G_N} (y_1)
\biggr\rrvert \\
&&\hspace*{55pt}{}\geq\frac
{(m-1)\nu_2/ \{ C(N-m) \} }{ 1 + \{
{N(m-1)}/{(m(N-1))} \}^{1/2} } \biggr\}
\\
&&\qquad\leq \pr \biggl\{ \bigl\llvert \E_{\hat G_N} (y_1) -
\E_{G_N} (y_1) \bigr\rrvert \geq\frac{(m-1)\nu_2}{ \{ 2+o(1) \}
(N-m) C} - o(1)
\biggr\}
\\
&&\qquad\leq \Var_X \bigl\{ \E_{\hat G_N} (y_1) \bigr\} \Big/
\biggl[ \frac
{(m-1)\nu_2 }{ \{2+o(1)\} (N-m) C} - o(1) \biggr]^2
\\
&&\qquad \leq\frac{(N-m) C^2}{(N-1)m} \Big/ \biggl[ \frac{(m-1)\nu_2 }{ \{
2+o(1)\}
(N-m) C} - o(1)
\biggr]^2,
\end{eqnarray*}
where the first inequality follows from the identity $u^2-v^2 =
(u+v)(u-v)$. Hence,
\begin{eqnarray}
\label{eq:pbound2} &&\limsup_N N \pr \biggl\{ \frac{N-m}{m-1}
\biggl\llvert \bigl\{ \E_{\hat G_N} (y_1) \bigr\}^2 -
\frac{N(m-1)}{m(N-1)} \bigl\{ \E_{G_N} (y_1) \bigr
\}^2 \biggr\rrvert \geq \nu_2 \biggr\}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\leq \frac{4(1-\theta\rho)^3 C^4 }{\theta^3\rho^3 \nu_2^2}.
\end{eqnarray}
The same arguments apply to the third and fourth terms:
\begin{eqnarray}\label{eq:pbound3}
&&\limsup_N N \pr \biggl\{ \frac{N-(n-m)}{n-m-1} \biggl\llvert
\E_{\hat F_N} \bigl(y_0^2\bigr) - \frac{N(n-m-1)}{(N-1)(n-m)}
\E_{F_N} \bigl(y_0^2\bigr) \biggr\rrvert \geq
\nu _3 \biggr\}
\nonumber\hspace*{-25pt}
\\[-8pt]
\\[-8pt]
\nonumber
&&\quad\leq \frac{ \{ 1-\theta(1-\rho) \}^3 C^4}{\theta^3 (1-\rho)^3
\nu_3^2},\hspace*{-25pt}\\
\label{eq:pbound4}
&&\limsup_N N \pr\biggl\{ \frac{N-(n-m)}{n-m-1} \biggl\llvert
\bigl\{ \E_{\hat
F_N} (y_0) \bigr\}^2 -
\frac{N(n-m-1)}{(N-1)(n-m)} \bigl\{ \E_{F_N} (y_0) \bigr
\}^2 \biggr\rrvert \geq\nu_4 \biggr\}
\nonumber\hspace*{-25pt}
\\[-8pt]
\\[-8pt]
\nonumber
&&\quad\leq \frac{4 \{ 1-\theta(1-\rho) \}^3 C^4 }{\theta^3(1-\rho)^3
\nu_4^2}.\hspace*{-25pt}
\end{eqnarray}
For the fifth term,
\begin{eqnarray*}
&& \pr \biggl[ \frac{2NC}{N-1} \bigl\llvert \E_{\hat G_N}
(y_1) - \E_{G_N} (y_1) + \E_{\hat F_N}(y_0)
- \E_{F_N} (y_0) \bigr\rrvert < \nu_5 +
\nu_6 \biggr]
\\
&&\qquad\geq \pr \biggl\{ \bigl\llvert \E_{\hat G_N} (y_1) -
\E_{G_N} (y_1) \bigr\rrvert < \frac{(N-1)\nu_5}{2NC}, \\
&&\hspace*{50pt}\bigl
\llvert \E_{\hat F_N}( y_0) - \E_{F_N}
(y_0) \bigr\rrvert < \frac{(N-1)\nu_6}{2NC} \biggr\}
\\
&&\qquad\geq 1 - \pr \biggl\{ \bigl\llvert \E_{\hat G_N} (y_1) -
\E_{G_N} (y_1) \bigr\rrvert \geq\frac{(N-1)\nu_5}{2NC} \biggr\}
\\
&&\qquad\quad - \pr \biggl\{ \bigl\llvert \E_{\hat F_N} (y_0) -
\E_{F_N} (y_0) \bigr\rrvert \geq\frac{(N-1)\nu_6}{2NC} \biggr\}
\\
&&\qquad\geq 1 - \Var_X \bigl\{ \E_{\hat G_N} (y_1) \bigr
\}\Big/ \biggl\{ \frac
{(N-1)\nu_5}{2NC} \biggr\}^2 - \Var_X \bigl
\{ \E_{\hat F_N} (y_0) \bigr\} \Big/ \biggl\{ \frac{(N-1)\nu_6}{2NC}
\biggr\}^2,
\end{eqnarray*}
so we have
\begin{eqnarray}\quad
\label{eq:pbound5}&& \limsup_N N \pr \biggl[ \frac{2NC}{N-1}
\bigl\llvert \E_{\hat G_N} (y_1) - \E _{G_N}
(y_1) + \E_{\hat F_N}(y_0) - \E_{F_N}
(y_0) \bigr\rrvert \geq\nu _5 + \nu_6
\biggr]
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\leq \frac{4(1-\theta\rho) C^4}{\theta\rho\nu_5^2} + \frac{4 \{
1-\theta(1-\rho) \} C^4}{\theta(1-\rho)\nu_6^2}.
\end{eqnarray}
For the sixth term, we use the fact that $|uv-u'v'| \leq|uv-u'v| +
|u'v-u'v'|$ to obtain
\begin{eqnarray*}
&& \pr \biggl\{ \frac{2N}{N-1} \bigl\llvert \E_{\hat G_N}(
y_1) \E_{\hat
F_N}(y_0) - \E_{G_N}
(y_1) \E_{F_N} (y_0) \bigr\rrvert <
\nu_7 + \nu_8 \biggr\}
\\
&&\qquad\geq 1 - \pr \biggl\{ \bigl\llvert \E_{\hat G_N} (y_1) -
\E_{G_N} (y_1) \bigr\rrvert \geq\frac{(N-1)\nu_7}{2NC} \biggr\}
\\
&&\qquad\quad{} - \pr \biggl\{ \bigl\llvert \E_{\hat F_N} (y_0) -
\E_{F_N} (y_0) \bigr\rrvert \geq\frac{(N-1)\nu_8}{2NC} \biggr\}.
\end{eqnarray*}
Following the rest of the derivation of (\ref{eq:pbound5}) gives
\begin{eqnarray}
\label{eq:pbound6}&& \limsup_N N \pr \biggl\{ \frac{2N}{N-1}
\bigl\llvert \E_{\hat G_N} (y_1) \E _{\hat F_N}(y_0)
- \E_{G_N} (y_1) \E_{F_N} (y_0) \bigr
\rrvert \geq\nu_7 + \nu_8 \biggr\}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\leq \frac{4(1-\theta\rho) C^4}{\theta\rho\nu_7^2} + \frac{4 \{
1-\theta(1-\rho) \} C^4 }{\theta(1-\rho)\nu_8^2}.
\end{eqnarray}
To bound the probability that the last term exceeds $1-\varepsilon_1$,
first note that $| u - u' | < \eta$ and $| v - v' | < \eta$ implies $|
\min(u,v) - \min(u',v') | < \eta$. This gives the third inequality below:
\begin{eqnarray*}
&& \pr \biggl[ \frac{2N}{N-1} \biggl\llvert \int_{[-C,C]^2}
\bigl\{ \hat H_N^H(y_1,y_0) -
H_N^H(y_1,y_0) \bigr\}
\,dy_1 \,dy_0 \biggr\rrvert < 1-\varepsilon_1
\biggr]
\\
&&\qquad\geq \pr \biggl\{ \sup_{y_1,y_0} \bigl\llvert \hat
H_N^H(y_1,y_0) -
H_N^H(y_1,y_0) \bigr\rrvert <
\frac{(N-1)(1-\varepsilon_1)}{8NC^2} \biggr\}
\\
&&\qquad\geq \pr \biggl[ \sup_{y_1,y_0} \bigl\llvert \min\bigl\{\hat
G_N(y_1), \hat F_N(y_0) \bigr\}
- \min\bigl\{G_N(y_1), F_N(y_0)
\bigr\} \bigr\rrvert \\
&&\hspace*{194pt}{}< \frac
{(N-1)(1-\varepsilon_1)}{8NC^2} \biggr]
\\
&&\qquad\geq \pr \biggl\{ \sup_y\bigl | G_N(y) - \hat
G_N(y) \bigr|, \sup_y \bigl| F_N(y) - \hat
F_N(y) \bigr| < \frac{(N-1)(1-\varepsilon_1)}{8NC^2} \biggr\}
\\
&&\qquad\geq \pr \biggl\{ \sup_y \bigl| G(y) - \hat G_N(y)
\bigr|, \sup_y \bigl| F(y) - \hat F_N(y)\bigr | <
\frac{(N-1)(1-\varepsilon_1)}{8NC^2} - o(1) \biggr\}
\\
&&\qquad\geq 1 - \pr \biggl\{ \sup_y\bigl | G(y) - \hat
G_N(y) \bigr| \geq\frac
{(N-1)(1-\varepsilon_1)}{8NC^2} - o(1) \biggr\}
\\
&&\qquad\quad{}- \pr \biggl\{ \sup_y \bigl| F(y) - \hat F_N(y) \bigr|
\geq\frac
{(N-1)(1-\varepsilon_1)}{8NC^2} - o(1) \biggr\}.
\end{eqnarray*}
The fourth inequality follows from Lemma~\ref{lemma:glivenko} which
shows that $\sup_y |G(y)-G_N(y)| = o(1)$ and $\sup_y |F(y)-F_N(y)| =
o(1)$. We can now apply the second part of Lemma~\ref{lemma:glivenko}
to bound the probability above. Given $\xi> 0$,
\begin{eqnarray*}
&& \limsup_N N \pr \biggl[ \frac{2N}{N-1} \biggl\llvert
\int_{[-C,C]^2} \bigl\{ \hat H_N^H(y_1,y_0)
- H_N^H(y_1,y_0) \bigr\}
\,dy_1 \,dy_0 \biggr\rrvert \geq 1-\varepsilon_1
\biggr]
\\
&&\qquad\leq \limsup_N N \pr \biggl\{ \sup_y
\bigl| G(y) - \hat G_N(y) \bigr| \geq \frac
{1-\varepsilon_1}{8C^2} - \xi \biggr\}
\\
&&\qquad\quad{}+ \limsup_N N \pr \biggl\{ \sup_y \bigl|
F(y) - \hat F_N(y) \bigr| \geq\frac
{1-\varepsilon_1}{8C^2} - \xi \biggr\}
\\
&&\qquad\leq \frac{(1-\theta\rho) K_1 (
({(1-\varepsilon_1)}/{(8C^2)}) -
\xi ) }{ \theta\rho \{ ({(1-\varepsilon_1)}/{(8C^2)}) -
\xi
\}^2 } \\
&&\qquad\quad{}+ \frac{ \{ 1-\theta(1-\rho) \} K_0
( ({(1-\varepsilon_1)}/{(8C^2)}) - \xi ) }{ \theta(1-\rho) \{
({(1-\varepsilon_1)}/{(8C^2)}) - \xi \}^2 }.
\end{eqnarray*}\eject\noindent
Since $\xi$ is arbitrary and both $K_1(\cdot)$ and $K_0(\cdot)$ are
nonincreasing, there exists $\kappa_1(\varepsilon_1)$ such that
\begin{eqnarray}
\label{eq:pbound7} \quad&& \limsup_N N \pr \biggl[ \frac{2N}{N-1}
\biggl\llvert \int_{[-C,C]^2} \bigl\{ \hat H_N^H(y_1,y_0)
- H_N^H(y_1,y_0) \bigr\}
\,dy_1 \,dy_0 \biggr\rrvert \geq 1-\varepsilon_1
\biggr]
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad\leq C^4 \kappa_1(\varepsilon_1).
\end{eqnarray}
The bounds (\ref{eq:pbound1})--(\ref{eq:pbound7}) imply that
\[
\limsup_N N\pr\bigl(N\bigl|\hat V_N^H-V_N\bigr|
\geq\varepsilon\bigr) \leq C^4 \Biggl( \sum
_{i=1}^8 \frac{c_i}{\nu_i^2} + \kappa_1(
\varepsilon_1) \Biggr).
\]
By minimizing the right-hand side over $\nu_1,\ldots,\nu_8 > 0$ subject
to the constraint $\nu_1+\cdots+\nu_8= \varepsilon_1$, the sum in the
parenthesis can be absorbed into $\kappa_1(\varepsilon_1)$, yielding
the desired convergence rate for $N\hat V_N^H$. To get the rate for
$N\hat V_N^L$, we repeat the argument used to derive (\ref
{eq:pbound7}). First note that $| u - u' | < \eta$ and $| v - v' | <
\zeta$ implies $| \max(0,u+v-1) - \max(0,u'+v'-1) | < \eta+ \zeta$.
This gives the second inequality below:
\begin{eqnarray*}
&& \pr \biggl[ \frac{2N}{N-1} \biggl\llvert \int_{[-C,C]^2}
\bigl\{ \hat H_N^L(y_1,y_0) -
H_N^L(y_1,y_0) \bigr\}
\,dy_1 \,dy_0 \biggr\rrvert < \varepsilon_2 +
\varepsilon _3 \biggr]
\\
&&\qquad\geq \pr \biggl[ \sup_{y_1,y_0} \bigl\llvert \max \bigl\{ 0,\hat
G_N(y_1)+\hat F_N(y_0)-1 \bigr\}
\\
&&\hspace*{65pt}{}- \max \bigl\{ 0,G_N(y_1)+F_N(y_0)-1
\bigr\} \bigr\rrvert
\\
&&\hspace*{97pt}\qquad\quad < \frac{(N-1)(\varepsilon_2 + \varepsilon_3)}{8NC^2} \biggr]
\\
&&\qquad\geq \pr \biggl\{ \sup_y \bigl| G(y) - \hat G_N(y)
\bigr| < \frac
{(N-1)\varepsilon_2}{8NC^2} - o(1),
\\
&&\hspace*{17pt}\qquad\quad \sup_y \bigl| F(y) - \hat F_N(y) \bigr| <
\frac{(N-1)\varepsilon
_3}{8NC^2} - o(1) \biggr\}
\\[-1pt]
&&\qquad\geq 1 - \pr \biggl\{ \sup_y \bigl| G(y) - \hat
G_N(y) \bigr| \geq\frac
{(N-1)\varepsilon_2}{8NC^2} - o(1) \biggr\}
\\[-1pt]
&&\qquad\quad{}- \pr \biggl\{ \sup_y \bigl| F(y) - \hat F_N(y)\bigr |
\geq\frac
{(N-1)\varepsilon_3}{8NC^2} - o(1) \biggr\}.\vspace*{-1pt}
\end{eqnarray*}
Thus there exist $\kappa_2(\varepsilon_2)$ and $\kappa_3(\varepsilon
_3)$ such that
\begin{eqnarray}
\label{eq:pbound8} && \limsup_N N \pr \biggl[ \frac{2N}{N-1}
\biggl\llvert \int_{[-C,C]^2} \bigl\{ \hat H_N^L(y_1,y_0)
- H_N^L(y_1,y_0) \bigr\}
\,dy_1 \,dy_0 \biggr\rrvert \geq\varepsilon_2 +
\varepsilon_3 \biggr]
\nonumber
\\[-10pt]
\\[-10pt]
\nonumber
&&\qquad\leq C^4 \bigl\{ \kappa_2(\varepsilon_2) +
\kappa_3(\varepsilon_3) \bigr\}.
\end{eqnarray}
\upqed\end{pf}
\section{R code for implementing estimator} \label{rcode}
Here, we present\vspace*{1pt} R code for the function \texttt{sharp.var}, which
outputs the bound estimates $\hat V^H_N$ (given input \texttt
{upper=TRUE}) and $\hat V^L_N$ (given input\break \texttt{upper=FALSE}). The
other inputs are \texttt{yt} (the observed outcomes under treatment),
\texttt{yc} (the observed outcomes under control) and \texttt{N} (the
total number of units in the population)
\mbox{}
{\footnotesize{
\begin{verbatim}
sharp.var <- function(yt,yc,N=length(c(yt,yc)),upper=TRUE) {
m <- length(yt)
n <- m + length(yc)
FPvar <- function(x,N) (N-1)/(N*(length(x)-1))
* sum((x - mean(x))^2)
yt <- sort(yt)
if(upper == TRUE) yc <- sort(yc) else
yc <- sort(yc,decreasing=TRUE)
p_i <- unique(sort(c(seq(0,n-m,1)/(n-m),seq(0,m,1)/m))) -
.Machine$double.eps^.5
p_i[1] <- .Machine$double.eps^.5
yti <- yt[ceiling(p_i*m)]
yci <- yc[ceiling(p_i*(n-m))]
p_i_minus <- c(NA,p_i[1: (length(p_i)-1)])
return(((N-m)/m * FPvar(yt,N) + (N-(n-m))/(n-m) * FPvar(yc,N)
+ 2*sum(((p_i-p_i_minus)*yti*yci)[2:length(p_i)])
- 2*mean(yt)*mean(yc))/(N-1))
}
\end{verbatim}
}}
\section{Illustrative upper bound improvements} \label{illustrative}
In Table~\ref{tab3}, we present illustrations of the improvements in the
variance upper bounds by varying the marginal distributions of
potential outcomes over the Beta distribution family: the control
potential outcomes are assumed to be distributed according to
$\operatorname
{Beta}(\alpha_0,\beta_0)$, and the treatment potential outcomes
according to $\operatorname{Beta}(\alpha_1,\beta_1)$. Strictly
speaking, since
finite populations cannot have continuous marginals, the Beta\vadjust{\goodbreak}
distributions represent approximations to plausible marginals when $N$
is large. We report the ratios ${ V^H_N}/{ V^{a}_N}$ and ${ V^H_N}/ {
V^{b+}_N}$ (the limits of ${ \hat V^H_N}/{ \hat V^{a}_N}$ and ${ \hat
V^H_N}/ { \hat V^{b+}_N}$) under different values of ($\alpha_0,\beta
_0,\alpha_1,\beta_1$) while holding $m = n/2$ and $n = N$ fixed.
\begin{table}
\caption{Illustrative upper bound ratios given Beta distributed
potential outcomes}\label{tab3}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lllllcc@{}}
\hline
& \multicolumn{1}{c}{$\bolds{\alpha_0}$} &
\multicolumn{1}{c}{$\bolds{\beta_0}$} & \multicolumn{1}{c}{$\bolds{\alpha_1}$} &
\multicolumn{1}{c}{$\bolds{\beta_1}$} & \multicolumn{1}{c}{$\bolds{{ V^H_N}/{
V^{a}_N}}$} & \multicolumn{1}{c@{}}{$\bolds{{ V^H_N}/ { V^{b+}_N}}$} \\
\hline
\phantom{0}1 & 0.1 & 0.1 & 0.1 & 0.1 & 1.00 & 1.00 \\
\phantom{0}2 & 0.1 & 0.1 & 0.1 & 1 & 0.68 & 0.79 \\
\phantom{0}3 & 0.1 & 0.1 & 0.1 & 2 & 0.61 & 0.81 \\
\phantom{0}4 & 0.1 & 0.1 & 1 & 1 & 0.92 & 0.97 \\
\phantom{0}5 & 0.1 & 0.1 & 1 & 2 & 0.86 & 0.95 \\
\phantom{0}6 & 0.1 & 0.1 & 2 & 2 & 0.86 & 0.96 \\
\phantom{0}7 & 1 & 1 & 0.1 & 0.1 & 0.92 & 0.97 \\
\phantom{0}8 & 1 & 1 & 0.1 & 1 & 0.81 & 0.84 \\
\phantom{0}9 & 1 & 1 & 0.1 & 2 & 0.71 & 0.83 \\
10 & 1 & 1 & 1 & 1 & 1.00 & 1.00 \\
11 & 1 & 1 & 1 & 2 & 0.98 & 0.99 \\
12 & 1 & 1 & 2 & 2 & 0.98 & 1.00 \\
13 & 2 & 2 & 0.1 & 0.1 & 0.86 & 0.96 \\
14 & 2 & 2 & 0.1 & 1 & 0.85 & 0.85 \\
15 & 2 & 2 & 0.1 & 2 & 0.76 & 0.83 \\
16 & 2 & 2 & 1 & 1 & 0.98 & 1.00 \\
17 & 2 & 2 & 1 & 2 & 0.99 & 0.99 \\
18 & 2 & 2 & 2 & 2 & 1.00 & 1.00 \\
\hline
\end{tabular*}\vspace*{-9pt}
\end{table}
Table~\ref{tab3} presents 18 scenarios, wherein $(\alpha_0,\beta_0) \in\{
(0.1,0.1),(1,1),(2,2)\}$, and $\alpha_1, \beta_1 \in\{0.1,1,2\}$. The
results are identical for $\operatorname{Beta}(\alpha_1,\beta_1)$
and\break $\operatorname
{Beta}(\beta_1,\alpha_1)$; thus, we omit redundant results. The ratios
were computed via numerical quadrature using the \texttt{NIntegrate}
command in Mathematica 7.0.1.0 under the default settings.
Our results illustrate that when the marginal distributions are
identical (i.e., cases 1, 10 and 18), all upper bounds are identical,
since the Cauchy--Schwarz and AM-GM inequalities hold exactly. However,
as the marginal distributions diverge in shape (e.g., cases 3, 9 and
15), our proposed upper bound $ V^H_N$ materially outperforms Neyman's
bounds $ V^a_N$ and $ V^{b+}_N$.
\end{appendix}
\section*{Acknowledgements}
The authors thank Allison Carnegie, Ed Kaplan, Winston Lin, Cyrus
Samii, Aad van der Vaart and the review team for helpful comments.
|
\section{Introduction}
In~\cite{Masina:2011aa,Masina:2011un, Masina:2012yd} we have proposed that Inflation in the Early Universe can be realized for a narrow band of values of the top quark and Higgs boson masses, for which
the Standard Model (SM) Higgs potential develops a second local minimum~\cite{CERN-TH-2683, hep-ph/0104016, strumia2, Bezrukov:2012sa, strumia3} at energy scales of about $10^{16}$ GeV.
Such a scenario is viable if a successful transition to a radiation-dominated era can be obtained, which we have shown to be possible in ref.~\cite{Masina:2011aa} using an explicit model in the framework of a scalar-tensor theory of gravity developed in~\cite{hep-ph/0511207,astro-ph/0511396} and in~\cite{Masina:2012yd} using a hybrid inflationary model with an additional weakly coupled scalar, in standard gravity. Such a generic scenario could be be realized only for a Higgs mass $m_H$ in the range $m_H=(126.0 \pm 3.5)$ GeV, the error being mainly due to the theoretical uncertainty of the 2-loops Renormalization Group Equations (RGE) used in that calculation. Such a prediction for the Higgs mass range has turned out to be surprisingly compatible with the measurement of a Higgs mass of $125.9\pm 0.4$ GeV by the LHC~\cite{HCP11,TEV}.
As it is well known, Inflation can generate quantum mechanically tensor (gravity wave) modes, usually parameterized
through $r$, the ratio of tensor-to-scalar perturbation spectra at large scales. In~\cite{Masina:2011aa, Masina:2011un} we had predicted a relatively large amplitude for $r$ in our scenario, which is set by the overall height of the second minimum, which can be computed up to the precision of the RGE calculations and of the experimental input for the Standard Model parameters: $m_H$, the top mass $m_t$ and the strong coupling constant $\alpha_s$. However the situation has recently changed, since more refined theoretical calculations of the Standard Model potential have been performed, therefore making more precise the correspondence between the input parameters and the scale of the second minimum.
Moreover the tensor modes have now been claimed to be measured by the BICEP2 collaboration~\cite{BICEP2} with a large amplitude, which corresponds to a scale of about $2\times 10^{16}$ GeV.
In this paper we update our results, using the results of~\cite{strumia2,strumia3} for the Higgs effective potential, and show that if the theoretical errors are under control the scale of Inflation is generically larger than $2\times 10^{16}$ GeV and can be reconciled with experiments only if $m_H\lesssim 124$ GeV (plus a theoretical error, at present estimated to be of order $0.3$ GeV).
Of course such conclusion can be changed if extra ingredients are added to the Standard Model. We show the results that can be obtained by introducing a non-minimal coupling of the Higgs to gravity. Of course less minimal modifications, such as a direct coupling of the Higgs to an extra scalar, can also easily shift the scale of Inflation.
Finally, provided the scale of Inflation due to the Higgs potential is low enough, we also reanalyze the specific model based on scalar tensor theories~\cite{Masina:2011aa}, which realizes the exit to a radiation dominated era, showing that it could reproduce observations on $r$ and on the scalar spectral index $n_s$.
As in~\cite{Masina:2011aa,Masina:2011un,Masina:2012yd} we consider the potential for the Higgs field $\chi$ in the SM of particle physics.
For very large values of the Higgs field such potential can be written as
\begin{equation}
V(\chi) \simeq \lambda_{\rm eff}(\chi)/4 \, \chi^4\,\,.
\end{equation}
where $\lambda_{\rm eff}$ is some effective quartic coupling~\cite{strumia2} which is very close, but not identical, to the quartic coupling $\lambda$. Due to the running of couplings $\lambda_{\rm eff}$ can become very small at high energies, while keeping always positive values (stability regime). In such conditions the potential can have an additional minimum at very high field values. While this is at present disfavored by the current calculations and measurements in the SM it still not ruled out. In particular this can still happen if the top quark mass has a low value compared to the present best fit~\cite{PDG,ATLAS:2014wva}.
If the Higgs field starts in the false minimum at $\chi_0$ and dominates the energy density of the Universe, the Friedmann equation leads to a stage of inflationary expansion
\begin{equation}
H^2 \simeq \frac{V(\chi_0)}{3 M^2} \equiv H_I^2 \,\,\,,\,\,\, \,\,a(t)\propto e^{H_I t}
\label{eq-M}
\end{equation}
where $a(t)$ is the scale factor, $H \equiv \dot a /a$ is the Hubble rate and $M$ is the reduced Planck mass ($M\equiv 2.435 \times 10^{18} GeV$).
A nontrivial model-dependent ingredient is a mechanism to achieve a graceful exit from Inflation, that is a transition to a radiation-dominated era, in a nearly flat Universe at a sufficiently high-temperature.
In~\cite{Masina:2011aa} we have proposed that the Higgs field can tunnel to the other side of the potential barrier by nucleating bubbles~\cite{Coleman}
that can successfully collide and percolate, in the presence of a scalar tensor theory of gravity. Alternatively in~\cite{Masina:2012yd} we had proposed that a smooth transition may happen also in standard gravity, in the presence of an extra scalar field very weakly coupled to the Higgs field. The main point of both mechanism is that they do {\it not} affect appreciably the SM runnings of couplings, since in the first model there is only a gravitational coupling to the Standard Model, and in the second case the dimensionless coupling constant is extremely small, of order $10^{-11}$. For this reason precise connections with low energy parameters still hold even in the presence of such new fields. Of course it is possible that such a new scalar can also interact with the Higgs field with a large direct coupling and in this case this could also affect the RGE equations and we briefly discuss later such possibilities.
Subsequently the Higgs field could roll down the potential, reheat the Universe and finally relax in the present true vacuum with $v=246$ GeV.
In the scalar-tensor theory of gravity model an extra scalar field $\phi$ is introduced, the Brans-Dicke scalar or dilaton, which has an interaction term of the form
$f(\phi) R$, where $R$ is the Ricci scalar and $f(\phi)>0$ thus sets the value of the Planck mass. The presence of such field makes the Planck mass time-dependent, and therefore also $H$ during Inflation: this slows down the expansion sufficiently enough if $f(\phi)$ grows faster than $\phi^2$ for large $\phi$ values, since a stage of quasi exponential expansion is followed by a stage of power-law (even decelerated) expansion ~\cite{hep-ph/0511207,astro-ph/0511396}.
During the exponential phase the quantum fluctuations in $\phi$ lead to a nearly scale-invariant spectrum of perturbations.
During the subsequent decelerated phase, $H$ decreases rapidly, and therefore the expansion is sufficiently slowed down, so that many bubbles can be nucleated in a Hubble patch and subsequently collide and reheat the Universe, leading to a spatially flat radiation dominated Universe.
As discussed in~\cite{Masina:2011aa}, after tunneling we require the field $\phi$ to relax to zero if a suitable potential $U(\phi)$ is present, which allows us to identify the present reduced Planck mass with the quantity $M$
and, at the same time, to satisfy constraints from fifth-force experiments and time-dependence of the Newton constant $G_N=1/(8 \pi M^2)$ ~\cite{Will:2005va}.
An alternative to scalar-tensor theories is given by models~\cite{Masina:2012yd}
with a direct tiny coupling of the Higgs field to an additional scalar field,
which induces a time-dependence directly into $\Gamma$ by flattening the barrier in the potential.
It is crucial to notice that a graceful exit can be realized in the above models only if there is a very shallow false minimum, otherwise the tunneling rate would be negligibly small in the scalar-tensor model, since the probability is exponentially sensitive to the barrier~\cite{Coleman}. Such shallowness is required also in the hybrid model since a change in the barrier by a large amount would introduce a departure from an almost flat scalar spectral index.
So, the shape of the potential is very close to the case in which there is just an inflection point and thus we have a generic prediction for the scale of Inflation and for $r$, while the specific model only affects the prediction for the spectral index of cosmological density perturbations $n_S$.
Using the recent state-of-the-art theoretical calculations given in~\cite{strumia2,strumia3}, we show the specific values of the top and Higgs masses
which allow for the presence of a false minimum. Such calculations are now so accurate that also the precise value of $\alpha_3(m_Z)$ inside the present allowed experimental range $\alpha_3=0.1184\pm0.007$~\cite{PDG} is a relevant parameter.
As an example, in fig.\ref{fig-Vmin} we also show the Higgs potential for very specific values of $m_t$ and $m_H$, showing agreement with fig.7 of~\cite{strumia2}.
The extremely precise values shown in the caption are not to be taken sharply, because of a theoretical
uncertainty, estimated to be of about $0.3$ GeV on $m_H$ and about $0.15$ GeV on $m_t$.
\begin{figure}[t!]
\centering
\includegraphics[width=8.5cm]{grafpot.pdf}
\caption{Higgs potential as a function of the Higgs field $\chi$ in units of $m_{Pl}=1.22\times 10^{19}$ GeV. We fixed $\alpha_3(m_Z)=0.1184$. The upper curve correspond to $m_H =125, m_t= 171.0305$, which shows good agreement with the results of~\cite{strumia2}, fig.~7. The lower curve corresponds to $m_H =123.2$, $m_t= 170.1228$, and a value of $r=0.28$.
As mentioned in the text and discussed in~\cite{Masina:2011aa}, in order to have a sizable tunneling probability through the left side, the barrier must be very low, almost as in an inflection point.}
\label{fig-Vmin}
\vskip .2 cm
\end{figure}
Increasing (decreasing) $m_t$, one has also to increase (decrease) $m_H$ in order to develop the shallow false minimum;
accordingly, the value of both $V(\chi_0)$ and $\chi_0$ increase (decrease).
The solid blue lines in fig.\ref{fig-mtmh} show the points in the $m_t-m_H$ plane where the shallow SM false minima exists, for different values of $
\alpha_3(m_Z)$ in the 3-sigma range.
The red dashed lines show the regions in ($m_H, m_t$) which corresponds to a value of $r$ compatible with observations, directly given by the height of the potential via the equation
$r=\frac{V(\chi_0)}{(2.24\times 10^{16} {\rm GeV})^4} 0.2$ and it turns out that these lines point to a too small value of $m_H$ compared to the present experimental values. Of course this is a valid conclusion under the assumption that the theoretical errors on $m_H$ are now estimated to be of about $0.3$ GeV by~\cite{strumia2,strumia3}. This scenario could still be compatible with data only if the theoretical errors in the RGE and the matching conditions were larger, say of about $1$ GeV.
Discarding such possibility we show that this could be cured by introducing a non-minimal coupling $\xi \chi^2 R$ also for the Higgs field. This transforms the potential in the slow roll phase in the Einstein frame in the following form (see~\cite{hep-ph/0511207} and citations therein):
$$
V(\chi)= \frac{\lambda_{\rm eff}/4 \chi^4}{(1+\xi (\chi^2/M^2))^2}
$$
which suppresses the potential at large field values and we show in fig.~\ref{plotxi} that a moderately large value of $\xi$ allows for the potential to fit CMB observations, also for larger $m_H$, compatible with the LHC measurements. However let us comment that this is based only on a tree level analysis and when having a value of the extra coupling constant $\xi$ larger than ${\cal O}(1)$ we have to be careful about the regime of validity of the theory, similarly to the models proposed in~\cite{Bezrukov:2007ep}.
\begin{figure*}[t!]
\centering
\includegraphics[width=9cm]{mhmtr.pdf}
\caption{ The solid lines indicate the $m_t-m_H$ values compatible with a shallow Higgs false minimum, taking a 3$\sigma$ variation of $\alpha_s(m_Z)$ around its central value
of 0.1184. Here the Higgs is assumed to be minimally coupled ($\xi$=0).
There is a (vertical) uncertainty of about $0.15$ GeV in $m_t$ and a (horizontal) one of $0.3$ GeV in $m_H$ due to the present~\cite{strumia2,strumia3} theoretical uncertainties in the RGE. The red dashed lines represent the values of the tensor-to-scalar ratio associated to the energy scale of the minimum, where we displayed the $r=0.2^{+0.07}_{0.09}$ values which correspond to the claimed recent detection by BICEP2~\cite{BICEP2}.}
\label{fig-mtmh}
\vskip .22 cm
\end{figure*}
\begin{figure*}[t!]
\includegraphics[width=7.5cm,height=6cm]{plotxi2.pdf}
\includegraphics[width=7.5cm,height=6cm]{plotxi.pdf}
\caption{ The contours indicate different values of the tensor-to-scalar ratio $r$ associated to the energy scale of the minimum when introducing a the nonminimal coupling $\xi$; we show it as a function of the Higgs and top mass for the central value of $\alpha_3$ (left plot) and for a larger value (3$\sigma$ deviation, right plot). There is a (vertical) uncertainty of about $0.15$ GeV in $m_t$ and a (horizontal) one of $0.3$ GeV in $m_H$ due to the present~\cite{strumia2,strumia3} theoretical uncertainties in the RGE.}
\label{plotxi}
\vskip .22 cm
\end{figure*}
Finally we also show the results for the $n_s-r$ plot for the scalar tensor model proposed in~\cite{Masina:2011aa} in figure~\ref{nsr}. Such a model is based on the introduction of an additional non-minimally coupled field $\phi$ with the following action:
\begin{equation}
-S=\int d^4 x \sqrt{-g} \left[ {\cal L}_{SM} + \frac{(\partial_\mu\phi\partial^\mu\phi)}{2} -\frac{M^2}{2} f(\phi) R -U(\phi) \right]
\label{azione}
\end{equation}
where $ {{\cal L}}_{SM}$ is the Standard Model Lagrangian and
the potential $U$ is not specified and assumed to be relevant only after Inflation to stabilize the field. The coupling function is given by:
\begin{equation}
f( \phi)\simeq 1 +\beta \left( \frac{ \phi}{M} \right)^2 +\sum_{n \ge 4} \gamma_n \left( \frac{ \phi}{M} \right)^n \ \,\,.
\label{eq-ftot}
\end{equation}
with the requirement that $f(\phi)>\phi^2$, which is guaranteed for instance if the couplings $\gamma_n $ are positive. Under these conditions $H(t)$ starts almost constant and after a sufficient umber of efolds starts decreasing as a power law, allowing for a tunneling transition when $H^4$ becomes equal to the tunneling rate per unit volume $\Gamma$. As a simple example we will study the case in which $\gamma_4>0$ and all other couplings are vanishing.
If the Higgs fields is at the minimum and it is not evolving during Inflation the only time dependent quantity is $\phi$ and so we can go the Einstein frame with a metric defined by $\bar{g}_{\mu \nu}\equiv f(\phi) g_{\mu \nu}$ and study the evolution of a canonically normalized field $\Phi$ defined through $d\Phi=d\phi \sqrt{K(\phi)}$, where
$$K(\phi)\equiv {2f(\phi)+3 M^2 f^{'2}(\phi)\over 2f^2(\phi)}
$$
In this frame the action becomes:
\begin{equation}
S_{E}= {1\over 2}\int d^4x\ \sqrt{-\bar{g}}[M^2 \bar{R}-(\bar{\partial}\Phi)^2- 2 \bar{{\cal L}}_{SM} ]
\label{e-action-2} \, .
\end{equation}
where the bar represents quantities in the Einstein frame. The Higgs potential contained in $\bar{{\cal L}}_{SM}$ is now $ V(\chi)/f(\Phi)^{2}$, so that the potential energy at the false Higgs minimum $\chi_0$ gives rise to a potential term for $\Phi$
\begin{equation}
S_{E}^{\chi_0}=\int d^4x\ \sqrt{-\bar{g}} \bar{V} \,\,\,\,, \,\,\, \bar{V}\equiv \frac{V(\chi_0)}{f(\Phi)^{2}} \,\,.
\label{eq-Vb0}
\end{equation}
This acts as a hill-top potential for the $\Phi$ field and we assume that $\Phi$ rolls down the potential from small to high values.
Given the potential we define as usual the slow-roll parameters
\begin{equation}
\epsilon(N) = \frac{1}{2} \left| \frac{1}{\bar V} \frac{d \bar V}{d(\Phi/M) }\right|^2
\qquad
\eta(N)= \frac{1}{\bar V} \frac{d^2 \bar V}{d(\Phi/M)^2 }
\end{equation}
We solve numerically the Klein-Gordon equation for $\Phi$ under such a potential assuming that it starts as close to zero as possible ({\it i.e.} with an initial value $\Phi_I$ given by the quantum fluctuations $\Phi_I=\bar{H}/(2\pi)$) together with the Friedmann equation, which gives us a numerical solution for $\Phi(N)$, where N is the number of efolds at some value of the scale factor $\bar{a}$, defined as $N \equiv \ln(\bar{a}_F/\bar{a})$, where $\bar{a}_F$ is the scale factor at the end of the Inflationary stage, defined by either $\epsilon$ or $\eta$ becoming of order 1.
Finally we evaluate the quantites
$$r \equiv P_T/P_S = 16 \epsilon_{\bar N} \, \qquad
n_S=1 -6 \epsilon_{\bar N} +2 \eta_{\bar N} \, $$
where ${\bar N}$ is the number of efolds that corresponds to a certain cosmological scale. We take $\bar{N}=60$ although the precise number depends on the history of the evolution subsequent to Inflation. The results are shown in fig.\ref{nsr}.\footnote{Note that in ref.\cite{Masina:2011aa} the calculations were obtained in an expansion for $\Phi\ll M$ for the slow-roll parameters, which does not hold for very small $\gamma_4$, therefore leading to incorrect results for $\gamma_4\ll 10^{-4}$, which corresponds to a sizable $r$ in fig.\ref{nsr}.}
In conclusion we have shown that a mechanism in which the energy density for Inflation is provided by the Higgs field starting in a shallow false minimum can be compatible with observations only if $m_H\lesssim 124$ GeV (and also $m_t$ has to be smaller than about $171$ GeV ). This is significantly lower than experimental observations unless the theoretical error on the RGE, of about $0.3$ GeV on $m_H$ according to~\cite{strumia2,strumia3}, is largely underestimated. We have also shown that introducing a non-minimal coupling of the Higgs of about ${\cal O}(10)$ can push this upper bound on larger values, closer to present observations.
Of course such bounds can be also evaded by modifying the shape of the potential in a model-dependent way. For instance a possibility would be to introduce direct couplings of the Higgs with the extra scalar as in~\cite{strumiascalar}, which would change the potential either at tree level or in the RGE, or perhaps to allow a faster change of the barrier in hybrid models during Inflation.
\begin{figure}[t!]
\centering
\includegraphics[width=9cm]{graf-eps-converted-to.pdf}
\caption{Spectral index $n_S$ and tensor-to-scalar ratio $r$ for the scalar-tensor model of~\cite{Masina:2011aa}. Here the number of efolds is assumed to be $\bar{N}=60$. Only one nonvanishing parameter $\gamma_4$ was assumed to be present in eq.~\ref{eq-ftot} and the various points along the black lines correspond to different values of $\gamma_4$. }
\label{nsr}
\vskip .2 cm
\end{figure}
\section*{Acknowledgements}
We thank A. Strumia and F.~Bezrukov for useful discussions.
|
Subsets and Splits